Sunteți pe pagina 1din 114

Geotechnical Engineering, Part I

Adam Bezuijen
Ghent University

The pile drivers, Maximilien Luce (1902-1903)

Geotechnical Engineering

ii

iii

Geotechnical Engineeing

Table of contents
1

Table of contents
Introduction
1.1 What is special about Geotechnical Engineering
1.1 Contents of the course and why this syllabus
Soil investigation
2.1 Introduction
2.2 Available information
2.3 Soil investigation tools
2.3.1
In-situ tests versus laboratory tests
2.3.2
The cone penetration test
2.4 The standard penetration test (SPT)
2.5 Other in-situ tests
2.6 Soil sampling
Pile foundations vertical loads
3.1 Introduction
3.2 Shallow foundations (see also the course soil mechanics)
3.3 Calculation methods for pile foundations
3.3.1
Calculation principle tip resistance
3.3.2
Koppejans method
3.3.3
De Beers method
3.3.4
Shaft friction
3.3.5
Negative skin friction
3.3.6
Problems with pile design calculation methods
From pile resistance to pile design
4.1 Design approaches
4.2 Necessary steps to calculate the design load on a pile
Pile systems
5.1 Introduction
5.2 Soil displacement pile systems
5.3 Systems with limited displacement or stress release during
construction
5.4 Soil replacing pile systems
5.5 Differences between pile systems
5.6 Pile factors
Theory for cast-in-situ piles
6.1 Cavity expansion and pile installation
6.1.1
Application
6.2 Water loss from the concrete
6.3 Advantages and disadvantages of various pile types
6.4 Developments
Pile driving
7.1 Different blocks
7.2 Pile driving theory
7.3 Stresses in piles and impedance
7.4 Impact force
7.5 Testing piles

iv

iv
1
1
1
2
2
3
4
4
5
9
11
13
15
15
15
17
17
18
21
23
24
25
27
27
28
29
29
29
30
30
31
32
34
34
37
38
41
41
42
42
42
47
49
50

Table of contents
7.6
7.7
7.8
Pile
8.1
8.2

Pile driving by vibration


Influence on surroundings
Group effects during installation
8
settlement and group effects after installation
Negative skin friction
Settlement
8.2.1
Pile shortening due to loading
8.2.2
Settlement of tip and shaft
8.3 Group effects after installation
8.3.1
Elastic deformation
8.3.2
Plastic deformation
8.4 Horizontal loading due to soil movement
9 Pile load tests
9.1 Type of tests
9.2 The dynamic loading test
9.3 The static loading test
9.4 The statnamic test
10 Physical modelling in geotechnics
10.1 Why physical modelling
10.2 1-g and N-g models
11 SLOPE STABILITY
11.1 Circular slip surface
11.2 Fellenius
11.3 Bishop
11.4 Other mechanism that lead to slope instability
11.4.1 Squeezing of soft layers underneath a slope
11.4.2 Liquefaction
11.4.3 Retrogressive flow slide or Breaching
11.4.4 Breaching or liquefaction
11.4.5 Groundwater flow within a slope
11.4.6 Monitoring pore water pressures
12 Trenchless Technologies
12.1 Introduction
12.2 Horizontal Directional Drilling.
12.2.1 What is HDD?
12.2.2 Mud properties
12.2.3 Measuring viscosity and yield strength of drilling
fluid
12.2.4 Stability of bore hole
12.3 Pipe jacking
12.4 TBM tunnelling, geotechnical aspects
12.4.1 Definitions
12.5 Stability of the tunnel face
12.5.1 Introduction stability of tunnel face
12.5.2 Stability tunnel face in clay
12.5.3 Stability tunnel face in sand
12.5.4 Influence of pore pressures
12.6 Tail void grouting
12.7 Buoyancy forces

51
52
53
56
56
58
58
59
61
61
62
62
66
66
66
67
67
70
70
70
73
73
74
76
77
77
78
79
81
81
82
84
84
84
84
86
89
91
93
94
96
98
98
98
99
101
102
105

Geotechnical Engineeing

vi

Introduction

1 Introduction
1.1

What is special about Geotechnical Engineering

Construction materials are usually made according to specifications. The strength, elastic and plastic
properties of iron, concrete, bricks and wood is reasonable well known. For soil this is different, it is a
highly variable construction material that consists of different layers of material and phases with quite
different mechanical properties. This makes soil as construction material a challenge for the engineer.
This challenge is not always taken up in the right way: up to 50% of the failure costs in construction
projects are related to the soil. Excessive settlement, leakage of underground structures, dam failure: it is
not only from the past, but still happens regularly. Lives can be at risk and costs to repair can be
enormous. In the most extreme cases, the damage is not repaired at all. This is also the interesting part of
geotechnical engineering. Each job has different soil conditions and consequently each job is different.

1.1

Contents of the course and why this syllabus

This course is the follow up of the bachelor course on soil mechanics.


This course deals with how soil mechanical knowledge is used in construction. The following topics will
be dealt with:
- Pile foundations
driven piles, bored piles, calculations methods, different pile types
- Pile driving
- Field investigation
- Physical modelling
- Ground improvement
temporarily, permanent without additives, permanent with additives.
- Piled embankments
A textbook was used in the bachelor course. The main reason not to use a textbook for this course is the
variety of calculation methods for the axial pile loading capacity. As we shall see these are different for in
different European countries and most (English) text books explain the methods used in England and the
USA and not the Belgian and Dutch methods.
Not using a text book also allows studying some new developments in the field.
This syllabus must be seen as back ground information to what will be told during the lectures.
What is dealt with during the lectures will be examined.

Geotechnical Engineering

2 Soil investigation
2.1

Introduction

Nowadays it is unlikely that a major building project starts without a soil investigation. However, around
100 years ago, building without soil investigation was the common situation, because there were no
adequate soil investigation techniques. Figure 2-1 shows what may happen in such a situation. It was
known that in Amsterdam most pile foundations are founded on a Pleistocene sand layer, the first sand
layer. This layer is present from 11 till 14 m below N.A.P. (Nieuw Amsterdams Peil, the reference height,
comparable to the average sea water level). When the Beurs van Berlage was built between 1898 and
1903, it was assumed that this layer was also present underneath the Beurs. However, as can be seen in
the figure it was not. An old branch of a river (the Oer-IJ) had made an incision in the sand that was
later filled with soft material. This results in settlements and cracks in the Beurs.

Figure 2-1. Soil layers underneath the "Beurs van Berlage" in Amsterdam.

This is just one of the countless examples that show the importance of a soil investigation. What has to be
investigated depends on the type of structure to be built. For a foundation this will be the sequence of the
layers and the strength and compressibility of various layers. For a retaining structure, it will also be the
different layers in the soil but now stiffness and permeability are more of importance. Strength and
stiffness are of importance in case the stability is critical.
The critical questions are:
What do I need to know more specifically about the subsoil and until what depth?
What information is available?
What can I expect around the building location and what is the variation that can be expected in
the parameters?
How do I collect the information and determine the parameters?

Soil investigation

The last question suggests that there are a large number of possibilities. In fact there are, but it appears
that only a few are used regularly. The methods used also depend on tradition. In Belgium and the
Netherlands, the cone penetration test is very popular whereas in England and the USA the standard
penetration test (SPT) is the test used most often.
In this chapter, some of these tests are described as well as the procedure to derive parameters from the
results.

2.2

Available information

The first step is: What is available? A first impression of what can be expected comes from geological
maps. A very general map as shown in Figure 2-2 already shows where soft and stiff soil layers can be
expected and much more detailed maps are available.

Figure 2-2. Overview geological conditions Belgium.

Apart from maps there is also a database available with the soil investigations performed in the past.
Figure 2-3 shows what tests have been performed in and around the university complex of Zwijnaarde
and were stored in the DOV (Database Ondergrond Vlaanderen), the Database for the subsoil of Flanders.
The outcome of these tests can be obtained and studied.
With the results of the existing tests and some calculations it is possible to decide whether for example a
shallow foundation is possible and consequently detailed information of the upper layers is necessary, or
that it will be a piled foundation and also the information of larger depths is of importance.

Geotechnical Engineering

Figure 2-3. Soil investigations (cone penetration tests) around university complex Zwijnaarde collected in the DOV.

2.3

Soil investigation tools

2.3.1

In-situ tests versus laboratory tests

Soil can be tested, as it is by pushing a device into the soil, or inflating a device in the soil (this will be
discussed later). Or a soil sample can be taken to the laboratory and tested there. The advantage of the
first method is that only a minimum disturbance of the sample will occur.

Figure 2-4. Different laboratory tests for geotechnical investigation (Mayne et al. 2009).

Soil investigation
The advantage of the last method is that more and more complicated tests are possible. Laboratory tests
are shown in Figure 2-4. The test most often used are dealt with during other courses, some in-situ testing
methods will be dealt with in the coming sections.

2.3.2

The cone penetration test

The most common in-situ test is the cone penetration test. The original test was developed by Barentsen
in 1932.This test was simple: a cone was pushed into the soil. To be able to measure the cone resistance
separately from the total penetration resistance, the cone was pushed with an inner rod in a tube as shown
in Figure 2.5. The cone and tube were penetrated into the soil together in steps of 0.2 m, then only the
cone was pushed another 65 mm to measure the cone resistance.
The cone has been improved after its invention. The shape was changed to avoid soil blocking the
opening between the tube and the cone. The first electrical cone was developed in 1948, thus avoiding the
friction between the tube and the rod. In the 60-ties the mechanical friction cone was developed further
and in the seventies the electrical cone measuring tip force and sleeve friction was developed. Mechanical
cones are still used as there is a risk that boulders damage the electrical ones.
Nowadays most electrical cones also have a pore pressure transducer to measure the pore pressure at
various depths (cone (2013) see Figure 2-6). The pore pressure is measured with a pressure gauge in the
cone with a filter of ceramic or porous metallic material in front of it. This material is saturated with
glycerine to have a quick reaction curve when the pore water changes. The saturation is quite critical and
special equipment is required to reach a good saturation. The influence of saturation of the pore filter
stone is shown in Figure 2-8. The results shown in the left plot show a sharp increase and decrease of pore
pressure allowing to determine a transition in soil layers accurately. In the right plot that is more difficult.
An inclinometer is normally included to measure whether the cone is still pushed down vertically.

Figure 2-5. Original cone as invented by Barentsen


(1932).

Figure 2-6. Modern electrical cone

Geotechnical Engineering

Figure 2-7. Cone with partly removed friction sleeve to see the interior.

Latest developments are the use of drilling mud to reduce the friction on the tube which allows
penetrating up to 100 m and a digital cone that transforms the measurement data in the cone to digital data
to avoid electrical noise in the data transmission from the cone to the data acquisition system.
The electrical cone results in a more or less continuous data profile, taking measurements every 0.02 m,
which corresponds with one sample a second for the standard penetration speed of 0.02m/s.
The friction and especially the ratio between the sleeve friction and the tip resistance appear to be an
important parameter to distinguish various types of soil.

C171

C175.2

pore press (MPa)


0.75

0.50

0.25

pore press (MPa)


0.00

0.75

0
cone res. (MPa)
pore press (MPa)

0.25

0.00

cone res. (MPa)


pore press (MPa)

8
depth (m)

depth (m)

0.50

10
12

10
12

14

14

16

16

18

18

20

20

22

22
0

10

cone res. (MPa)

15

10

15

cone res. (MPa)

Figure 2-8. Results of piezocone tests with a well saturated filter (left) and
with an insufficient saturated filter (right).

Soil investigation
The results of a CPT are used to determine quite a number of soil parameters. Numerous methods are
available for this purpose.
Table 2.1 shows for example quite simple rules. A bit more complicated are the rules suggested by
Begemann (1963, Figure 2.7), or Robertson (1990, Figure 2-10).

Table 2.1. Simple soil classification system


based on CPT.
Cone
Friction
Soil type Resistance
ratio
(MPa)
(%)
Sand
2 - 30
1
Clay
0.5 2
35
Peat
0-1
>8

Figure 2-9. Begemann profiling chart for a mechanical CPT.

This last system, proposed by Robertson (1990), uses dimensionless numbers. The vertical stress ( v0) is
subtracted from the cone resistance, corrected for the pore water pressure (qt) and this is divided by the
effective vertical stress (v0) to find the number Qt, thus: Qt=(qt-v0)/v0.
The correction of the cone resistance (qc) for the pore water pressure, to find qt is:

qt qc (1 a)u2

(2.1)

Geotechnical Engineering
Where u2 is the pore water pressure measured at U2 (see Figure 2-10 in the graph on the right) and a is the
ratio between shoulder area (cone base, AN in Figure 2-12) unaffected by the pore water pressure to total
shoulder area (AT).
Fr is slightly different from the normal friction number as shown in cone penetration test results. The
normal friction number (Rf) is defined as:

Rf

100 f s
qt

(2.2)

While Fr is defined as:

Fr

100 f s
qt v 0

(2.3)

The difference between these 2 numbers will only be significant in very soft soils.

Figure 2-10. Soil classification system developed by Robertson (1990).

The number Bq , along the horizontal axis of the right hand side figure of Figure 2-10, shows whether the
pore pressure u2 measured at the cone is higher or lower than the hydrostatic pore pressure. It should be
noted that the chart is valid for the pore pressure measured at the location U2. This is the usual position to
measure the pore pressure but sometimes it is measured on the cone itself (half way on the cone, U 1) or
behind the friction sleeve (U3). These different positions results in different pore pressures measured.
The Robertson chart is frequently used. However, Fellenius and Eslami (2000) correctly noticed that the
chart is not presented optimally. A drawback is that the axes of both figures are not independent (as they

Soil investigation
are in the Begemann chart). Both axes have the term qt - vo, so the cone resistance is plotted against its
own inverse value. Furthermore, they argue that the normalisation is unnecessary complicated and
sometimes difficult to achieve in practice. They suggested a chart as shown in Figure 2-11 and define the
effective cone resistance (qe) as:

qe qt u2 qc au2

Figure 2-11. The Eslami-Fellenius profiling chart.

2.4

(2.4)

Figure 2-12. Definition sketch AN and


AT

The standard penetration test (SPT)

In Anglo Saxon countries the standard penetration test is a popular test for soil classification and
parameter determination. It is one of the oldest ways for soil exploration and started in 1902.
The test uses a thick-walled sample tube, with an outside diameter of 50 mm and an inside diameter of
35 mm, and a length of around 650 mm. This is driven into the ground at the bottom of a borehole by
blows from a slide hammer with a weight of 63.5 kg (140 lb) falling over a distance of 760 mm (30 inch).
The sample tube is driven 150 mm into the ground and then the number of blows needed for the tube to
penetrate each 150 mm (6 inch) up to a depth of 450 mm (18 inch) is recorded. The sum of the number of
blows required for the second and third 6 inch of penetration is termed the "standard penetration
resistance" or the "N-value". In cases where 50 blows are insufficient to advance it through a 150 mm (6
inch) interval, the penetration after 50 blows is recorded. The blow count provides an indication of
the density of the ground, and is used in many empirical geotechnical engineering equations. (alinea
taken from Wikipedia).
The SPT test has some advantages in comparison to the CPT:
No heavy equipment is needed, because it is a dynamic test.
Soil samples are obtained, thus the classification shown above is not necessary. The type of soil is
visible.

Geotechnical Engineering
However, the method is not as accurate as a CPT and does not give results with the high resolution of a
CPT. Furthermore, it takes more time than pushing a device into the ground as for example a CPT. A
casing has to be used for soft soil to prevent closing of the borehole.

According to Wikipedia:
The usefulness of SPT results depends on the soil type, with fine-grained sands giving the most useful
results, with coarser sands and silty sands giving reasonably useful results, and clays and gravelly soils
yielding results which may be very poorly representative of the true soil conditions. Soils in arid areas,
such as the Western United States, may exhibit natural cementation. This condition will often increase the
standard penetration value.
[]
Despite its many flaws, it is usual practice to correlate SPT results with soil properties relevant for
geotechnical engineering design. The reason being that SPT results are often the only test results
available, therefore the use of direct correlations has become common practice in many countries.
Different correlations are proposed for granular and cohesive soils.
A soil classification system based on the SPT is proposed by Terzaghi and Peck (1948), see Table 2.2.

Table 2.2. Interpretation SPT Terzaghi and Peck (1948)

Sand

10

Clay

Density

Consistency

<4
4 - 10
10 - 30
30 - 50
> 50

Very loose
Loose
Normal
Dense
Very dense

<2
2-4
4-8
8 - 15
15 - 30
> 30

Very soft
Soft
Normal
Stiff
Very stiff
Hard.

Soil investigation

2.5

Other in-situ tests

The tests mentioned above are by far the most widely used in-situ tests. Other tests are used when special
conditions are expected or special parameters are required. Quite a lot of probes are developed, see Figure
2-13.

Figure 2-13. In-situ field probes for evaluating soil parameters (Mayne et al. 2009).

Vane tests, T-bars and ball-penetrometers are used in very soft clay. The larger area that is rotated or
pushed into the clay allows a more accurate measurement. The vane is less robust, but allows measuring
the peak and residual strength, See Figure 2-14. The seismic cone allows measuring the velocity v of the
vibrations in the soil and based on that, the shear modulus (G) for low deformations can be determined,
using the equation:

(2.5)

The dilatometer and pressure meter test allow measuring elastic deformation at larger deformations. The
dilatometer test produces values for the elasticity and the K0 value of the soil (K0 is the lateral earth
pressure coefficient at rest: the ratio between horizontal and vertical pressure).
The electrical resistivity and nuclear probes allow measuring the density of the soil. The electrical
resistivity probe gives a relative measurement; the electrical resistivity in the soil is compared to the
electrical resistivity of the pore water itself. The nuclear resistivity probe sends fast neutrons and -rays
and measures the back scatter of slow neutrons and -rays to determine water content and the density of
the soil respectively. It is used quite often to test the densification of the sand embankment for a road.
The CPT with pore pressure gauge (CPTu) can also be used for a dissipation test. The cone is pushed over
a certain distance, for example 1m.Then the penetration stops and excess pore pressures are allowed to
dissipate. The time needed for dissipation is a measure for the consolidation coefficient of the soil.

11

Geotechnical Engineering

Figure 2-14. Vane and result of vane test (Mayne 2009).

Mayne et al. (2009) argues that one in-situ test is insufficient to find all parameters needed. He suggests
that a probe pressed into the soil, should be used to measure more than just one parameter. He suggests to
use SCPT (Seismic Cone Penetration Tests with Pore pressure and Pore pressure dissipation tests) or a
SDMTa (Seismic Dilato Meter Test, with waiting time to estimate dissipation). Results of these tests are
shown in Figure 2-15 respectively.

Figure 2-15. Seismic piezocone test with dissipation (SCPT) in soft varved clays at Amherst national geotechnical
test site, Massachusetts (Mayne et al., 2009).

12

Soil investigation

Figure 2-16. Seismic flat dilatometer test with dissipations (SDMT) in stratified sediments of Venetian lagoon at
Treporti test embankment. (Mayne et al., 2009)

2.6

Soil sampling

For sampling it is of importance whether an undisturbed sample is needed. A disturbed sample shows the
type of soil that is present, but does not give any information about the density or stress-situation of the
soil. Purely undisturbed samples do not exist. Taking a sample will always result in some disturbance.
The influence of sample disturbance is shown in Figure 2-17. It shows triaxial tests on 3 different samples
of the same clay. The failure line is the same for all 3 samples, but the way the failure line is reached is
quite different. The dilatancy that is present in the block sample is absent in the tube samples. All tube
samplers have this disturbance because friction develops along the tube walls when the tube is pushed
into the soil. To overcome this influence of friction, the Begemann (or Delft continuous) sampler can be
used. This sampler uses a stocking tube with a stocking and a lubricant between stocking and tube, that
reduces the friction between tube and soil considerably. The sampler uses bentonite as lubricant and
prefabricated nylon coated stocking and is therefore much more complicated (and thus more expensive)
than a tube sampler.

Figure 2-17. Sample disturbance effects for CAUC triaxial compression tests on Lierstranda clay showing
effective stress paths (plot from Mayne et al. 2009). The block sample is closest to undisturbed;
the 54 mm tube has most disturbance.

13

Geotechnical Engineering

An advantage is that due to the absence of side wall friction continuous samples are possible and thus a
detailed soil profile can be obtained (Figure 2-18).

Figure 2-18. Soil profile obtained with a Begemann sampler.

14

Pile foundations vertical loads

3
3.1

Pile foundations vertical loads


Introduction

In this chapter we will deal with the calculation methods to determine the maximum resistance of a
foundation pile against vertical load. The allowable load on a pile from a building will be less, because
safety factors have to be taken into account as will be dealt with in Chapter 4. Sometimes the horizontal
loading on a foundation pile is decisive and the moments in the pile due to horizontal loading have to be
calculated. This will be dealt with briefly in Section 8.4.

3.2

Shallow foundations (see also the course soil mechanics)

We start with the calculation of the bearing capacity Pc of a shallow foundation of a strip footing. The
subsoil is a rather theoretical clay, that is weightless and has an undrained shear strength su. This
calculation is based on a lower bound and an upper bound solution. The lower bound solution assumes
that there is no plasticity in the clay at all, see Figure 3-1. The upper bound solution calculates a possible
continues slip plane through the clay.

Figure 3-1. Lower bound solution of maximum loading on shallow foundation.


Equilibrium according to Drucker (Figure Verruijt & Broere, 2011. They use c instead of su).

We assume three zones as shown in Figure 3-1 left-hand drawing. The pressure next to the loading on the
surface is zero. According to the Mohr circle, Figure 3-1 right-hand plot, the maximum possible
horizontal stress is 2 su and the maximum possible vertical stress underneath the load is 4c. Since this a
lower bound solution, the maximum possible pressure loading on the foundation (Pc) in this lower bound
solution can be written as:
4

(3.1)

In an upper bound solution, there must be a continuous failure plane. A simple failure plane is a half
circle, see Figure 3-2. The shear strength su is mobilized along this circle.

Figure 3-2. Half circle as failure plane.


(Figure Verruijt & Broere, 2011. They use c instead of su)

15

Geotechnical Engineering

Momentum equilibrium is obtained when:

0.5r P r su

(3.2)

Thus since this is an upper bound solution:

Pc 2 su

(3.3)

However, there are failure planes that result in lower values. For a circular plane with an origin above the
surface it can be calculated that the minimum value of the pressure Pc is:

Pc 5.52su

(3.4)

Prandtl found in 1920 a solution that fulfills both the lower and upper bound solution, that solution should
result in the real failure load. He uses the slip surfaces shown in Figure 3.3.

Figure 3-3. Failure surface in Prandtls solution. (Figure Verruijt & Broere 2011)

With these slip surfaces he came to the well-known solution for a strip footing:
Pc = (2 + ) su = 5.14 su

(3.5)

For real soils with a certain density, a cohesion and a friction angle there is no closed form solution.
Normally the following relation is used to determine the maximum possible pressure before failure
occured:

1
BN
2

(3.6)

1 sin
exp( tan )
1 sin
N c ( N q 1) cot

(3.7)

Pc su N c qN q
Where:

Nq

N 2( N q 1) tan
Where:
q

: the surcharge load, see Figure 3-4

16

Pile foundations vertical loads

: the volumetric weight of the soil (effective weight when submerged)


: the friction angle
: dimensionless factors
: the width of the footing

Nq, Nc, N
B

Figure 3-4. Definition sketch and possible failure surfaces. (Figure Verruijt & Broere, 2011)

The bearing capacity strongly depends on the friction angle as can be seen in Figure 3-5 that shows
results of calculations with the formulas (3.6) and (3.7).
350

7,000
Nq
Nc
Ny
p

N (-)

250

6,000
5,000

200

4,000

150

3,000

100

2,000

50

1,000

0
30.0

p (kpa)

300

0
32.5

35.0

37.5

40.0

42.5

45.0

Figure 3-5. Results of example calculation for c=1 kPa, q = 10 kPa, B=2 m and =18 kN/m3

3.3
3.3.1

Calculation methods for pile foundations


Calculation principle tip resistance

Calculation methods following the same principle as in the previous section can be used to calculate the
end bearing capacity for pile foundations. Again no closed form solutions are available for real soils with
a certain density, a cohesion and a friction angle and the solutions that do exist, tend to be quite
complicated. Budhu (2007) mentioned the solutions of Janbu (1976) for the end bearing of a pile in sand:

P 'v N q

(3.8)

where:

17

Geotechnical Engineering

N q (tan ' 1 tan 2 ' ) 2 exp( 2 ptan ' )

(3.9)

In this equation is:

the friction angle (the accent is added by Budhu to indicate that it is the friction angle for the effective
stress)
p the angle of pastification.
The angle of pastification is /3 for soft, fine-grained soils and 0.58 for dense, coarse grained soils
and overconsolidated fine-grained soils.
Another relation mentioned in Budhu is the relation from Vesic (1975), from theoretical considerations
based on the cavity expansion theory for a cylindrical cavity:
4 / 3 sin '

2 ' 1sin '
3

Nq
exp ' tan ' tan I rr

3 sin ' 2

4 2

(3.10)

where Irr is the rigidity index expressed as:

I rr

Ir

1 p Ir

(3.11)

G
' z b tan '

(3.12)

and

Ir

With G the shear modulus, p the volumetric strain around the pile and (z)b the vertical effective stress at
the base.
In case the soil properties are determined from laboratory tests or standard penetration tests (SPT) (see
Section 2.4), these formulas have to be used. In countries where the cone penetration test (CPT) is used,
more straightforward methods have been developed. In a cone penetration test, a cone of 36 mm diameter
is pushed into the ground (see also Chapter 2). The penetration resistance of the soil is measured at the tip
of the cone. In some cases also the sleeve friction is measured. The resulting cone resistance over depth
can be seen as the tip resistance of a miniature pile. Calculation methods have been developed that
directly calculate the pile resistance using the results of a cone penetration test. Since such methods are
mostly used in Belgium and the Netherlands, this syllabus focuses on these methods.

3.3.2

Koppejans method

Koppejan assumed the failure surface around a soil penetrating pile shown in Figure 3.6.

18

Pile foundations vertical loads

Figure 3-6. Failure surfaces according to Koppejan (Van Mierlo en Koppejan, 1952).

If this failure surface occurs in homogeneous soil, the pile resistance is comparable to the cone resistance.
If the failure surface intersects with weak soil layers, the pile resistance will be lower than the cone
resistance, because the smaller cone has a smaller failure surface and therefore measures the local
strength with much less influence of different layers.
The tip resistance of a pile (p, with the dimension pressure) at a certain depth is therefore calculated using
a normative value for the cone resistance. This is calculated from a CPT with the relation:

0.5( p1 p2 ) p3
2

(3.13)

Figure 3-7. Determination of p1, p2 en p3

19

Geotechnical Engineering

Where:
p2 :
is the average value of the cone resistance in trajectory I (Figure 3-6) from the pile tip level to a
level in between 0.7 Deq and 4 Deq. The depth has to be chosen in a way that the average value of
the cone resistance is at minimum and falls within the limits mentioned (0.7Deq and 4Deq). The
equivalent diameter Deq of the pile is used to account for hollow piles and non-circular piles as
will be explained later.
p1 :
the average value of the cone resistance in trajectory II (Figure 3-6). This is determined nearly the
same as p2, but has the restriction that the cone resistance used to calculate the average in a
certain layer must always be equal or lower than the value used in the underlain layers.
p3 :
the average value of the cone resistance over trajectory III (Figure 3-6) starting at the pile tip
level to a level 8 Deq higher. The cone resistance used to calculate the average in a certain layer
must always be equal or lower than the value used in the underlain layers, just like for p1. The
starting value of p3, at the tip of the pile, is the end value of the cone resistance used to calculate
p1.
The procedure is explained in Figure 3-7 and Figure 3-8.

Where qc,avg = 0.5 [0.5 (qcI + qcII) + qcIII] and D is the pile diameter
p2 qcI: arithmetic average of qc values below pile tip over a depth which may vary between
0.7D to 4D as shown in Figure 3-6;
p1 qcII: arithmetic average of qc values following a minimum path rule recorded below the pile
Tip over the same depth of 0.7D to 4D ;
p3 qcIII: arithmetic average of the minimum qc values following a minimum path rule recorded
above the pile tip over a height of 8D.
Figure 3-8. Cone penetration resistance as a function of depth for the situation that only 0.7 D below the tip of a pile
has to be taken into account (left) and for the situation that 4 D has to be used (right).

According to Xu and Lehane (2005), the method as described above is better than just averaging qc over
1.5 D, as used in France and used as the base of the Imperial College Method (Tomlinson, 2001).
Another advantage of Koppejan is that the method can easily be carried out by hand.

20

Pile foundations vertical loads

3.3.3

De Beers method

Prof. De Beer
Prof. De Beer was the founding father of soil mechanics and geotechnics in Belgium. He was professor
at Ghent University and director of the Rijksinstituut voor Grondmechanica (1941-1978). He was
involved in all major infrastructure projects in Belgium. It was mentioned at the 60th anniversary of the
Expert group Geotechnics of the Royal Flemish Society of Engineers: By that time (in the seventies) we
didnt have standards, there was no need for, because Prof. De Beer was the standard. The reality of that
time was: when he approved a construction that was sufficient. His influence is still present today, for
example in the De Beers method to calculate a pile foundation.

Description of method
De Beers method to determine the tip resistance p of a pile is especially popular in Belgium, where it is
the standard method Paalfunderingen, De Beer (1971-1972). It does not assume a certain failure surface
like Koppejan, but considers the CPT as a kind of scale model test of a real pile test and assumes that the
scale model relations apply. For example consider the case where the CPT encounters a strong layer with
a high CTP value. The resistance will increase rapidly with depth. For a larger diameter pile this increase
in strength will be much slower, because it has to be pushed further into the strong layer, before the same
failure pattern can develop, see Figure 3-9.

Figure 3-9. De Beer's method, scale effect principle.

However, when this failure pattern has developed, the pile tip is lower than the cone tip and the stresses
will be higher.
To correct for this higher stress, the following relation is used:

21

Geotechnical Engineering

1
qr , crit

' h'crit

2. p0
q
' hcrit c , crit
1
2. p0

(3.14)

And to correct for the difference in diameter between the cone and the pile:

qr , j 1 qr , j

' h'crit

1 2. p
d
0

qc , j 1 qr , j
1 ' hcrit
D

2. p0

(3.15)

The parameters used in these formulae are:


p0
: overburden pressure at the depth where the increase in cone resistance starts.
hcrit
: critical depth for the cone.
hcrit
: critical depth for the pile
qr,crit : ultimate bearing resistance for the cone
qc,crit : cone resistance at critical depth hcrit
d
: diameter of the cone
D
: diameter of pile
qr,j
: pile resistance at location j.
qc,j
: cone resistance at location j.
This last equation has the problem that hcrit is not known. In the Beers method it is assumed that this is
0.2 m (the distance between 2 separate readings in the now obsolete mechanical cone penetration test) and
hcrit = hcrit.D/d, with hcrit=0.2 m.
The method mentioned above works well when the CPT resistance of the soil increases with depth. It still
works when the CPT value locally decreases, but the calculated pile resistance still increases. The
calculated pile resistance may, however, never become higher than the local CPT value. When
encountering a layer with a low CPT resistance, the method would predict a higher value for the pile
resistance than the actual CPT resistance in that layer. For that situation we therefore work bottom up to
achieve the right pile resistance. This makes it possible to use the values calculated with the formula
above. The values obtained with that equation (qr,j+1) are used to reduce the gradient in qr,i where i
increases going upwards, using the following the formula:

qr ,i 1 qr ,i (qr , j 1 )i 1 qr ,i

Dd

(3.16)

Van Impe (1986) has shown that the original equation is a bit conservative and can be changed to:

qr ,i 1 qr ,i 2. (qr , j 1 ) i 1 qr ,i

Dd

(3.17)

Allowing twice larger gradients in the part where the pile resistance decreases with increasing depth. The
influence of Van Impes change is shown in figure 3.10.

22

Pile foundations vertical loads

Figure 3-10. Comparison upward calculation according to De Beer and Van Impe et al.

Now a thin weak layer can still lead to a locally significant decrease in pile resistance because it was
assumed that the pile resistance in that situation cannot be higher than the local minimum cone resistance.
In reality the pile resistance will be higher because of the influence of the stronger layers. Therefore the
final pile tip resistance is calculated by taking the values obtained by the procedure given above and
taking the moving average of these values over a distance equal to the pile diameter.
The various steps make the calculation method not very suitable for hand calculations. but calculations
can easily be performed with the various computer programs available.

3.3.4

Shaft friction

The total shaft friction Rs is calculated with the formula:

Rs s ( s ,i .hi . *p,i .qc,m,i )


Where:
s
s,i
hi
*p,i
qc,m,i

(3.18)

: equivalent circumference of a pile (see WTCB 2009)


: installation factor depending on the pile type (see WTCB 2009)
: thickness of layer i
: empirical factor presenting the relation between cone resistance and friction factor
: average cone resistance of layer i

The values used for the empirical installation factor depend on the type of pile, the soil and the way it is
installed. Driven concrete piles have a value of 1.0, but 0.9 in Tertiary clay; bored piles have values of 0.5
depending on the installation method, for other values, see WTCB 2009.
The relation between cone resistance and friction factor depends on the soil type and the cone resistance,
see also WTCB 2009.
The values presented in WTCB 2009, are values used in Belgium. Other countries have the same formula
to calculate the shaft friction, but use different values for the empirical factors.

23

Geotechnical Engineering
It should be noted that even if the sleeve friction is measured in a CPT, this value is not used, but
Equation 3.18 is used. The reason is that the sleeve friction may be inaccurate if the cone has some wear,
resulting in a slightly smaller diameter. Therefore the sleeve friction is used to determine the soil layers
but the tip resistance is used to calculate the shaft friction.
The shaft friction is only calculated for layers where qc 1 MPa. Negative skin friction can occur when
the upper soil layers are compressible \ and thus have relatively low cone resistances. This will be dealt
with in the next section.
Even when the sleeve friction is measured directly during a cone penetration test, this measured friction is
used for soil classification but not to calculate the friction along the pile. In all cases Eq. (3.18) is used to
calculate the friction along the pile. The idea is that the measured sleeve friction is not very accurate due
to small differences in diameter between the cone and the pile and that Eq. (3.18) results in a more
accurate estimation of the friction.

3.3.5

Negative skin friction

Negative skin friction may occur when the soil settles more than the pile. This is quite common when the
upper layers consist of soft Holocene material and sand is added on top of this material to have a better
working area. This sand causes settlement of the soft layers. It can also be caused by a later raising of the
soil surface leading to settlement. Due to negative skin friction the load on the pile increases and thus the
ultimate bearing capacity of a pile decreases. When not taken into account in the design of the foundation
this may lead to excessive settlements, see Figure 3-11.

Figure 3-11. Example of settlement due to negative skin friction

If there is negative skin friction, the force distribution on the pile will be as sketched in Figure 3.12

24

Pile foundations vertical loads

force in pile

soil settles
more than pile

depth

soil settles
less than pile

a. CPT and pile position

b. forces in pile

Figure 3-12 (a) Typical CPT where negative skin friction can be expected
(b) forces on pile if negative skin friction occurs.

3.3.6

Problems with pile design calculation methods

Although millions of pile foundations are installed and problems with these foundations are very rare,
there are still some problems with respect to the calculation methods as described above.
The first is that the standardization of these calculation methods has up to now only been possible at the
level of individual countries. Harmonization of methods on a European or world wide scale has not been
possible. Conditions differ and local experience is assumed to be of importance. For an open European
market, harmonization within Europe would be helpful. The methods and their results really differ in
various countries.
Another problem is that most methods seem to over-predict the tip resistance for piles driven in sand for
more than several pile diameters. Results from an Australian PhD study are shown in Figure 3.13. It
shows the ratio between measured and calculated tip resistance, using the method of Koppejan, as a
function of the penetration depth of the pile into the sand. When the pile tip is less than 2 diameters into
the sand the ratio is 1.0, which means that there is agreement between measurements and theory.
However, at larger depths the ratio decreases until 0.6, thus a design for these piles that are driven deeply
into the stiff layer may be unsafe.
Although there is always a safety factor applied in the design, the discrepancy between calculated and
measured resistance is so big that problems with pile foundations should be expected regularly. Since this
is not the case, it is assumed that there is some hidden safety in the design. Such a hidden safety can be
the aging effect. The pile shaft resistance increases in time and this is not taken into account. Research is
going on to clarify this hidden safety. However, due to the uncertainty the NEN, the Dutch
standardisation institute, wants to lower the pile factors in 2016 to be sure that safety can be guaranteed.

25

Geotechnical Engineering
1.6
1.4
1.2

measured/calculated

perfect fit
0.8
0.6
0.4
0.2
0.0
0

10

15

20

25

Length/Diameter in sand

Figure 3-13. Ratio between measured and calculated (Koppejan) pile tip resistance as a function of the
penetration in the bearing sand layer (Stoevelaar et al., 2011).

26

From pile resistance to pile design

From pile resistance to pile design

The previous chapter described methods to determine the tip resistance and the shaft friction of a pile. If
these are known and the corresponding pile factors are also known, there are still some steps to go to
come to a design value for the bearing capacity of the pile.
The Eurocode gives some guidelines about how to carry out these steps, but allows the individual
countries to have, within certain limits, their own interpretation. The procedure to be used in a Belgian
design for axially loaded piles is written down in detail in the WTCB report Richtlijnen voor de
toepassing van de Eurocode 7 in Belgi, Deel 1 and will not be repeated here, but some principles of this
method will be dealt with.

4.1

Design approaches

Eurocode 7 allows 3 design approaches (DAs), to be chosen by National standardization committees. The
difference between these design approaches is the way the partial safety factors are applied. The safety
factors can be applied on:
1. The action (A, the loading on the pile)
2. The material parameters (M, for example friction angle , cohesion c)
3. The resistance (R, the resistance calculated with for example method De Beer).
The first design approach has 2 possibilities, 1.1 and 1.2. The various design approaches are shown in
Table 4.1, showing that Design Approach 1.1 puts all safety in the action (A). The others have a
combination of safety factors.
Table 4.1. Possible design approaches in Eurocode 7. (g means statistical load, q is dynamic load)
(Van Tol, lecture notes)

Design
Approach

Action /
Action effect
A - Action
> 1.0

MaterialFactor
M- Material
= 1.0

Factor on
Resistance
R - Resistance
= 1.0

DA 1.2

= 1.0 (g)
> 1.0 (q)

> 1.0

= 1.0

DA 2

> 1.0

= 1.0

> 1.0

DA 3

> 1.0

> 1.0

= 1.0

DA 1.1

When Design Approach 1 is followed, it is necessary to fulfil DA 1.1 and DA 1.2. While in geotechnics
the uncertainty in the material properties is often dominant on the overall uncertainty, DA 1.2 is often the
relevant design approach. DA 1.1 is dominant when it comes to calculate the strength of the pile itself.
For piles most countries have chosen DA 2, but Belgium chose DA 1 and The Netherlands DA 3.

27

Geotechnical Engineering

4.2

Necessary steps to calculate the design load on a pile

To come from a CPT to a geotechnical design load of a pile, the following steps are necessary:
Step 1.
The maximum pile load has to be calculated from the CPT using a method described in Chapter 2 and
using the right installation factors for the tip resistance and shaft friction for the pile type used. See the
WTCB report for more detailed information on these factors for applications in Belgium.
Step 2.
A model factor is applied to take into account the uncertainty in the calculation model. If the contractor
decides to perform loading tests, it is possible to reduce the model factor, giving the possibility to reduce
the number of piles or the size of the piles. This results in the calibration value of the bearing capacity.
Step 3.
Correlation factors have to be used to account for the variation and uncertainty in the soil conditions.
More CPTs available? A lower factor may be applied (resulting in a higher pile capacity). Different
factors are applied on the average (3) and minimum (4) bearing capacity. The lowest value is determines
the characteristic value of the bearing capacity (Rc,k):

(R )
(R )
Rc,K min c,cal mean ; c,cal min
4
3

(4.1)

The values of these factors depend on the type of foundation. When the piles are used underneath a stiff
foundation, a pile with a lower bearing capacity will not directly lead to failure of the structure since the
load can be re-distributed to the other piles. For a foundation with limited stiffness this is more critical.
Step 4.
After calculating the bearing resistance of a pile, the design load can be determined (rekenwaarde van de
belasting in Dutch). The design load should be lower than the design resistance in 95% of the cases.
Therefore some safety factors are added. These safety factors again depend on the type of pile and on the
quality of the installation. With a quality guarantee, the factors to be applied may be lower than without
such a guarantee.
Actual values for the factors described above and other details of the procedure to be followed can be
found in the WTCB guideline (WTCB, 2009).

28

Pile systems

Pile systems

5.1

Introduction

The oldest pile foundations found were made 6000 years ago by a group now called the Swiss Lake
Dwellers in Switzerland (Tritech, 2016). It is remarkable that in 5820 from these 6000 years there was
only one pile system: the timber pile. The screw pile was developed in the mid of the 19th century. In
1909 came the Franki pile, that was constructed into the soil with concrete, by that time also the first steel
sheet pile was developed. In the 50-ties of the 20th century came the pre-stressed reinforced concrete pile,
the present standard. Nowadays there are more than a dozen pile systems, each with special advantages
and disadvantages. A first rough division can be made: soil displacing systems and soil removing systems
with in between the systems that create limited or no soil displacement. This division is important, since
in general a soil displacing pile system reacts stiffer and has a higher loading capacity than a soil
replacing system. The pile factors of a pile system are dependent on its category.
Advantages of screwed and bored pile foundations compared to drilled foundations or pre-cast concrete
piles and vice versa are presented below.
Advantages pre-fabricated piles
Pile has constant quality
Special reinforcement designs
constructed (when ordered in time)
Optimal soil displacement
Relatively cheap

5.2

can

Advantages screwed and bored piles


No vibrations in the soil (no damage to nearby
be structures)
Limited working height necessary (some systems)
Larger diameters
Easy logistics
Limited heavy transport
Availability is not an issue

Soil displacement pile systems

Drilled pile systems are for example: pre-fabricated concrete piles, timber piles, closed tube piles and
open tube piles with plug formation.
Furthermore, some cast-in-place piles where soil displacement is created and screwed piles are a part of
this category. For these systems it is very important that there is no stress-release during the construction
of the pile. The pressure on the concrete should preferably be kept high enough throughout the entire
casting process. Such a stress release can significantly reduce shaft and/or tip resistance.

Figure 5-1. Examples driven piles (reinforced concrete, steel and timber piles).

29

Geotechnical Engineering

5.3

Systems with limited displacement or stress release during construction

Systems with limited displacement or stress release during construction are for example drilled thinwalled systems like open steel pipes without plug formation and steel sheet piles. Furthermore, there are
the CFA systems (CFA stands for Continuous Flight Auger) with special precautions to avoid stress
release along the pile, for example because the process of removing the soil with the auger and replacing
it with concrete is performed under a pressure or a tube system is used.

Figure 5-2. Examples of piles with limited displacement (screwed piles with casing, H-beam and hollow tubes).

5.4

Soil replacing pile systems

Soil replacing pile systems are for example CFA systems without any precautions to avoid stress release
along the pile and bored piles, see Figure 5-3.

Figure 5-3. Examples soil replacing piles (CFA, and 2 pictures of bored piles).

30

Pile systems

5.5

Differences between pile systems

The systems mentioned are, for clarity, presented as systems clearly within one category. For driven piles
this is more or less correct. However, for the systems with limited stress relieve and the replacement
systems, there is quite a grey area. For these pile types, the execution has a great influence. The time is
important, see Figure 5-4.

Figure 5-4. Influence of time between drilling and concreting on cone resistance. The pile on the left hand side plot
is conreted one hour after excavation on the right hand site two weeks later (De Beer 1988).

Figure 5-5 shows the stress release due to the installation of an auger pile and the casting of the concrete.
When the tip of the auger is lifted for concrete filling and passes the measurement device, there is again a
decrease in pressure. It should be noted that the pressures measured with the DMT (Dilatometer Test) are
much higher than the hydrostatic pressures, thus there has been some disturbance by the installation of the
DMT in the loose clayey sand.

Figure 5-5. Soil conditions and soil pressure measurements for an auger pile. (Van Impe, 1993).

31

Geotechnical Engineering

5.6

Pile factors

The standards give pile factors for each pile system. When a pile supplier has developed a new pile
system, the supplier should carry out a number of pile loading tests to determine the associated pile
factors. Often the supplier tries to prevent to perform the quite expensive static pile loading tests by
arguing that his system is better than (another system).
The 2 installation factors are b and s: factor b presents the ratio between the measured cone resistance
and the tip resistance; factor s presents the ratio between shaft friction, determined from a CPT, see
above, and the shaft friction on the pile. These factors are close or equal to 1.0 for a prefabricated
concrete pile, but lower for systems with less displacement or stress release. The lowest values are 0.5 in
the usual soils, but can be as low as 0.3 for the shaft friction of soil replacing piles in tertiary clay.
Therefore a lot of pile systems are designed to minimize the stress release during pile installation. The
factors differ in different countries. This may look odd, because the soil is the same in different countries,
but can be somewhat justified by the fact that also the calculation methods differ in various countries.
Van Impes graphs (Figure 5.6 and Figure 5.7) for the pile factors b and s show that there is still room
for international harmonisation.
Installation factor b in granular soil
0

0.2

0.4

0.6

0.8

1.2

1.4

BE
Char.-10%

FR

Char.-10%

DE
IT
NL

Critical
5%

Ultimate
Large dia.

CFA with qc3<2 MPa

Increased crit. depth

PL

Displacement piles

Non-Displacement piles

Figure 5-6. Pile factors b in various European countries. (Van Impe, lecture notes)

To use the subdivision between displacement piles, piles with limited displacement and piles without
displacement to determine the pile factors is under discussion. Details in the execution, of screwed and
bored piles especially, appear to be more important than just the pile system.

32

Pile systems

160

sand

IT/CFA (Viggiani)
IT

140
DE

Unit shaft friction (kPa)

120

Van Impe/

100

max/uncased

NL/CFA &
80

NL/BPcased
FR/CFA

60

Van Impe
/min

40
FR/Large dia
20
0
0

10

20

30
40
Cone resistance (MPa)

Figure 5-7. Relation between cone resistance and shaft friction in various
European countries. (Van Impe, lecture notes)

Due to different pile factors and subjective interpretation of soil parameters the calculated ultimate
loading capacity has quite some variation. This appears when parties were asked to make a prediction of
the ultimate loading capacity for a square 0.25 m pile to be installed in Amsterdam at 14.1 m depth 1.5 m
in a sand layer. The predicted and measured values of the 27 parties are shown in Figure 5-8. Also the
predicted load settlement curves, to be dealt with later varied quite a lot, Figure 5-9.

(dynamic probing)

Figure 5-8: Results of ESOPII, 1982, Predicted and measured ultimate load capacity.
Different colours present the soil investigation used.

33

Geotechnical Engineering

Figure 5-9. ESOPII 1982, Predicted and measured load settlement curves
Different lines present the soil investigation method on which the prediction was based.

Theory for cast-in-situ piles

In this chapter some theory is presented that helps understanding some aspects of cast-in-situ piles. Cavity
expansion for cohesive soils and pressure filtration of the concrete mortar are described.

6.1

Cavity expansion and pile installation

Cavity expansion theory can be used to calculate the allowable pressure that can be applied on the liquid
water-cement mixture of cast-in-situ piles. Low pressures in the concrete will lead to unloading the soil
around a cast-in-situ pile. This has to be avoided since it will reduce both the stiffness and the bearing
capacity of the pile. Too high pressures are not desirable, since this will lead to uncontrolled deformation
of the pile because there will be on-going plastic deformation of the soil around the pile. The cavity
expansion theory is useful to understand the pile-soil interaction for both driven piles and cast-in-situ
piles. In this chapter, the theory is derived for a soil that is characterised with a Youngs modulus,
Poisson ratio and an undrained shear strength. This is the case for pure clay. More general descriptions
can be found in Cavity expansion methods in Geomechanics Hai-Sui, Yu (2000).
Figure 6.1 shows an element of material in a cylindrical coordinate system. It is assumed that the
displacement field is cylindrically symmetrical, so that there are no shear stresses acting upon the
element, and the tangential stress tt is independent of the orientation of the plane. The figure shows the

34

Theory for cast-in-situ piles


stresses acting on the element. Stresses and strains are assumed to be positive for compression.
Equilibrium of forces leads to (see also Figure 6-1):


rr r ( rr d rr )(r dr ) 2 tt dr sin 0

(6.1)

Neglecting the term with drr.dr leads to the only non-trivial equation of equilibrium, the one in radial
direction:


rr .r . rr .r . rr . dr . d rr . r . d rr . dr. 2. tt dr sin 0
2
rr
d

rr 2 sin . tt
r
dr
2 r
d rr tt rr
d rr rr tt

or

0
dr
r
dr
r

d rr rr tt

0
dr
r

(6.2)

y
tt

rr+drr

rr
tt

dr

x
Figure 6-1: Element in circular coordinates with principal stresses for radial deformation

In case of cavity expansion there is a plastic zone and an elastic zone, See Figure 6-2.
For an undrained material with an undrained shear strength su, there is a constant difference between the
major and minor principal stresses: rr -tt = 2.su.
In the plastic zone, therefore, equation (6.1) can be written as:

d rr 2su

0
dr
r

(6.3)

This differential equation can be solved easily, leading to:

rr 2.su .ln ri c

(6.4)

35

Geotechnical Engineering
Where r0 < ri < rp; and c is an integration constant.

rp

r0
ri

Figure 6-2. Definition sketch cavity expansion. The cavity expands from diameter 0 to r0.
This creates a plastic zone between r0 and rp. Outside the plastic zone the deformation is elastic.

In the elastic zone, for undrained conditions, the following relations between stresses and displacements
apply:

u
r
u
tt 2G
r

rr 2G

(6.5)

Where u is the displacement.


At the boundary (rp) between the elastic and plastic zone the stresses should be the same, thus:

rr tt 4G

u
2 su
rp

(6.6)

Eq. (6.4) shows that rr=-tt for elastic circumstances and thus that at the boundary (rp) we find: rr= su
and tt=- su. This means that the integration constant c = 2.su ln rp+ su in equation (6.4). Resulting in:

rr 2.su . ln

rp
ri

su

(6.7)

There is no volume change in undrained material, this means that when the cavity increases from 0 to r0,
the displacement u at rp should get such a value that the same volume is displaced as in the centre:

36

Theory for cast-in-situ piles

2urp ro2

ro2
2rp

(6.8)

Filling the result of Eq. (6.6) in Eq. (6.5) leads to:

4G

r02
2 rp
rp

2 su

rp

G
r0
su

(6.9)

ro2
Su
rp2

This result can be used in Eq.(6.5a), leading to:

rr ,r 2su ln
0

6.1.1

G
su
su

su ln

G
su
su

(6.10)

Application

The equations presented in the previous section allow calculating the pressure that can be applied on the
liquid concrete (before hardening) of a cast-in-situ pile.
In its simplest form, the density of the concrete mixture results in a pressure that increases with depth and
since the density of the mixture is higher than the density of the surrounding soil, there will be an excess
pressure on the soil. If this excess pressure is higher than the result of Eq. (6.8) there will be uncontrolled
deformation of the pile that still consists of liquid concrete. An example of such a calculation is shown in
Figure 6-3.
pressure difference (kPa)
0

100

200

0.0
press. diff.
su required
su from clay

2.5
5.0

depth (m)

7.5
10.0
12.5
15.0
17.5
20.0
22.5
25.0
0

25

50

75

su (kPa)

Figure 6-3. Calculation of allowable pressure in water-cement mixture of cast-in-situ pile.

37

Geotechnical Engineering
In this calculation the density of the soil is assumed to be 1400 kg/m3, that of the concrete mixture 2200
kg/m3. The ratio between shear-modulus G and undrained shear strength su is 50 and the ratio between the
vertical stress and su is 0.4. The graph shows that the calculated su of the soil just equals the necessary su.
As mentioned this calculation is a simplification. It is assumed that the concrete level is equal to ground
level. As can be seen in the sketch the level of the concrete is normally higher leading to an even higher
loading. However, it is also assumed that the water level is at the surface. In reality the water level is
often lower, leading to extra vertical stresses in the first meters. Furthermore, as will be explained later
on, when the pile is placed in sand, it will lose some water from the concrete mixture. As a result, the
excess pore water pressure in the concrete sludge decreases. The concrete hangs partly on the side walls
and the pressure in the concrete will be lower than the hydrostatic stress and the loading on the soil is also
less.
Shape discontinuities of cast-in-situ piles may occur as a result of soil deformation, as shown in Figure
6.4.

Figure 6-4. Shape discontinuities of cast-in-situ piles.

6.2

Water loss from the concrete

Just after installation, a cast-in-situ pile consists of steel reinforcement and a liquid cement mortar.
Hardening of the cement will occur. But before hardening starts, the cement loses water due to what is
called pressure filtration for piles that are installed in a permeable subsoil like sand. The pressure in the
mortar will be higher than the pore water pressure and water will flow from the mortar into the soil.
To calculate this water flow we assume, for simplicity a one dimensional situation, see Figure 6-5.

sand

non
liquid
mortar

pile

dx dx
i

liquid
mortar

38

Theory for cast-in-situ piles

Figure 6-5. Water loss in a cast-in-situ pile installed in


sand.

During the water loss, there are 3 zones going from the center of the pile to the boundary sand/pile see
also Figure 6-5.
1. The centre of the pile that still consists of liquid mortar.
2. A transition zone, where the liquid mortar loses its water and changes to the non-liquid mortar.
3. A zone of mortar that has lost some pore water and is not liquid anymore.
The velocity with which the mortar transforms from a liquid to a non-liquid state depends on the initial
porosity of the liquid mortar, the final porosity after losing water and the permeability of the non-liquid
mortar.
According to Darcy:

qk

Where:
q
: water flow
k
: permeability of the non-liquid mortar

: difference in piezometric head between the centre of the pile and the sand
x
: thickness of the non-liquid zone

(6.11)

(m/s)
(m/s)
(m)
(m)

When the mortar loses its water, the porosity of the mortar decreases from ni to a lower porosity ne. The
thickness of that mortar layer decreases accordingly:

dxi

1 ne
dx
1 ni

(6.12)

where dx and dxi are defined in Figure 6-5.


Continuity demands:

n n
qdt dxi dx i e dx
1 ni

(6.13)

With Darcys equation this leads to the differential equation:

1 ni
dx

k
dt
n

n
i e

(6.14)

With the boundary condition x=0 at t=0, the solution of this equation is:

1 ni
t
x 2k
n

n
i
e

(6.15)

39

Geotechnical Engineering
This leads to the results shown in Figure 6-6 for the parameters mentioned in that figure. This example
shows that an in-situ-cast pile can lose its water quite quickly when in contact with a permeable layer.
thickness non-liquid zone (m)

0.50
0.45
0.40
0.35
0.30

k
ni
ne

0.25
0.20
0.15
0.10

=5m
= 2 e-6 m/s
= 0.31
= 0.27

0.05
0.00
0

100

200

300

400

500

600

time (s)

Figure 6-6. Development of pressure filtration.

The mechanism described above sometimes leads to damage to in-situ-cast piles, see Figure 6-7. The left
hand side picture shows the original water pressure and the pore pressure distribution in the mortar
directly after the construction of the pile. The pressure in the mortar is approximately 0.8 times the
hydrostatic pressure due to the wall friction between the mortar and the soil, as mentioned before. The
soil profile is shown at the left. This is mainly sand, but there is a layer of impermeable clay. After some
time, the situation is as sketched in the right hand side picture, the mortar in contact with the sand has lost
water due to pressure filtration and the mortar pressure in the pile decreases. The mortar in contact with
the clay cannot lose water and here the pressure remains high. For a well-constructed pile there is still no
problem, but when there are openings between the mortar and the reinforcement, a flow of water from the
cement mortar along the reinforcement is possible, resulting in pile damage at the upper part of the pile,
because the cement has been washed out partly. Creating some extra flow resistance in the reinforcement
may help to avoid this problem.

100

200 300

400

[kPa]

-5

-5

-10

-10

hydrostatic
hydrostatische
morteldruk **0,8
pressure
0.8

-15

-20
vers gestorte paal

freshly poured pile

-15

100

200 300

400

maximale drukhoogte ligt


Max.
pressure height is
boven paalkop, waardoor
above pile top, this may
stroming van cement water
lead
langsto
debleeding
wapeningalong
staven
reinforcement
kan optreden = bleeding

hydrostatic
hydrostatische
morteldruk
pressure **0,8
0.8

-20
enige
tijdafter
na storten
liquidvloeistof
pressuredruk
some
time
pouring

Figure 6-7. Pore pressure distribution in-situ-cast pile just after installation and after pressure filtration of
some part of the concrete (pictures Van Tol, lecture notes)

40

Theory for cast-in-situ piles

6.3

Advantages and disadvantages of various pile types

The strong and weak points of different pile types are summarized in the table below. This is just an
indication. For all pile systems the installation is very important.
Table 6.1. Advantages/disadvantages of various pile types (van Tol lecture notes).

aspect
aspect

6.4

prefab.
prefab vibro
Vibropile
pile
paal
paal

CFA
avegaar Fundex
Fundex Tubex
Tubex bored
boorpile
pilepaal pile
paal pilepaal

diaphragm
diepwall
wand

price/
prijs/
bearing
draagkr.c.

++

++

++

bearing
draagcapacity
kracht

++

++

+ ++

+ ++

belenother
dingen
buildings

+/0

+/0

+/-

++

++

noise
geluid/
trillingen
vibrations

++

++

++

++

++

working
werkheight
hoogte

++

grondinfluence
++
opbouw
soil
strength

+/-

+/-

Developments

Especially the cast-in-situ pile systems are still developing. Their pile factors have smaller values at the
moment than the values for driven piles. Improvement of the systems may lead to higher pile factors.
The following lines are followed:
Systems are developed that prevent unloading of the soil during installation
Since the installation is very important, some contractors want a certification system for the
installation procedure. Piles installed by a certified contractor may have higher pile factors than
piles from an uncertified contractor.
In case tubes are used, grout mortar is injected around the tubes to increase the friction between
the tube and the soil.
Swelling concrete is used to increase the horizontal pressure.
Continuous measurement of settlement, pump pressure and pull down force (on the
reinforcement) can prove that no unloading has occurred during installation.

41

Geotechnical Engineering

Pile driving

7.1

Different blocks

Although the number of in-situ cast piles is increasing, most piles are still driven pre-fabricated piles.
Until the middle of the 19th century pile driving was done by hand. The first mechanical driving
installations were steam engines and in the thirties of the 20th century the diesel block was developed and
later also the hydraulic block. Nowadays all types of blocks are still used with the diesel block as the most
popular block.
The principle of all blocks is the same: a weight is lifted and then dropped on the pile. This lifting can be
done by hand, with steam, diesel or by hydraulic equipment. The diesel block is a bit different in the way
that when the weight hits the pile, this is also the moment the diesel-air mixture explodes. This explosion
extends the loading period of the pile. The loading sequence for a diesel block is shown in Figure 7-1.

piston = falling weight

diesel
reservoir

fuel pump
outlet ports
cylinder
combustion
space
rubber ring

Figure 7-1. Loading sequence for a diesel block.

The diesel block is in fact a 2-stroke diesel engine. It produces more pollution than a usual 4-stroke
engine and this can be seen as a disadvantage of the diesel block.

7.2

Pile driving theory

The principle of pile driving is that the pile is hit for a very short period (order of milliseconds). This
creates a longitudinal wave into the pile. The pile is thus not moving as one piece but the top moves first
and this movement goes through the pile to the bottom. The short impact time is useful because that
means that there is a huge deceleration in the falling block and since F=ma, the larger the deceleration (a)
for a given mass of the block (m), the larger the force (F), the force that drives the pile into the soil.
The impact velocity is around 4 m/s, the impact duration is difficult to determine and depends on the
cushion to be discussed later, but will be in the order of 2 ms. This means that the deceleration is 2000
m/s2, thus 200 times more than the acceleration of gravity. Consequently the maximum force on the pile

42

Pile driving
tip is 200 times the weight of the falling weight. There is of course a limitation: the force exerted by the
falling weight has to be lower than the strength of the concrete otherwise the pile will fail.
When the block hits the pile the longitudinal wave will start in the pile. First a simple case without
friction between the pile and the soil is assumed.

N+N
Figure 7-2. Element of an axially loaded pile.

According to Newton (F=m.a), the equation of motion for an element, see Figure 7-2, can be written as:

N
2w
A 2
z
t

(7.1)

From:

N A

(7.2)

z
follows:

N EA

w
z

(7.3)

Which leads to the wave equation:

2w
2w
E 2 2
z
t

(7.4)

In these formulas:
N
z

A
w
t

: normal force at a location in the pile (kN)


: vertical coordinate (m)
: density of the pile (kg/m3)
: cross-sectional area of the pile (m2)
: displacement of the pile (m)
: time (s)

This equation has solutions of the form:

43

Geotechnical Engineering

w f1 ( z ct ) f 2 ( z ct )

(7.5)

Where f1 and f2 are arbitrary functions and

(7.6)

c is the propagation velocity (m/s) of the wave in the pile.


Each disturbance in the pile will be transported through the pile with the wave velocity given in equation
(7.6). This velocity c is approximately 4200 m/s for concrete piles and 5000 m/s for steel piles.
This means that when the block falls on the pile it will take L/c seconds before this is noticed at the tip of
the pile.

t
Table 7.1. Parameters used in numerical calculation

parameter
L
area
circumference
E

number of elements
force Newton
duration of force (tem)
fr

value
Dimension
20
m
0.16
m2
1.6
m
4.00E+10
kPa
2400
kg/m3
20
-4.0E+06
N
0.0041
s
3000
N.s/m3

Figure 7-3. Sketch pile with


friction.

In reality there is friction between pile and soil. A simple approximation of a pile with friction is shown in
Figure 7-3. The friction has a static component and a component that is proportional to the pile velocity
with respect to the ground. Neglecting the static component in the friction the differential equation
becomes:

2w
2 w f r S w

A t
z 2
t 2

(7.7)

Where fr is the friction in N/m2/(m/s) thus N.s/m3 and S is the circumference (m)
This equation can be solved numerically. For the situation of a constant force that is present during 4 ms
and the parameters mentioned in Table 7.1, the calculated displacement and forces in the pile are shown
in Figure 7-4 to Figure 7-7. Figure 7-4 and Figure 7-5 are for the situation of a pile where the

44

Pile driving
displacement of the pile tip is zero, thus when the pile tip has reached firm soil and is not displaced any
further and Figure 7-6 and Figure 7-7 for the situation of free displacement.

0.012
0.010
0.008

0.004

0.010
0.008
0.006

0.002

displacement (m)

displacement (m)

0.006

1
4
8
12
16
20

0.012

1
4
8
12
16
20

0.000
-0.002
-0.004

0.004
0.002
0.000
-0.002

-0.006

-0.004

-0.008

-0.006

-0.010

-0.008

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

time (s)

-0.010
0.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.10

time (s)

Figure 7-4. Movement of pile element 1(top), 4, 8, 12,


16 and 20 (tip) after impact. No movement at tip.

Figure 7-5. Movement of pile element 1(top), 4,8,12,16


and 20 (tip) after impact. No movement at tip, detail
0.1s.

45

0.18

0.18

0.16

0.16

0.14

0.14

0.12

0.12

displacement (m)

displacement (m)

Geotechnical Engineering

0.10
0.08
1
4
8
12
16
20

0.06
0.04
0.02

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0.08
1
4
8
12
16
20

0.06
0.04
0.02
0.00
0.00

0.00
0.0

0.10

1.0

0.01

0.02

0.03

0.04

8106

8106

6106

6106

4106

4106

2106

2106

0106
-2106

0
4
8
12
16
20

-410

-610

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.10

time (s)

Figure 7-8. Stresses in pile on element 0(top), 4,8,12,16


and 20 (tip) after impact. No movement at tip.

46

0.06

0.07

0.08

0.09

0.10

Figure 7-7. Movement of pile element 1(top), 4,8,12,16


and 20 (tip) after impact. Free tip, detail 0.1s

Force (N)

Force (N)

Figure 7-6. Movement of pile element 1(top), 4,8,12,16


and 20 (tip) after impact. free tip

-8106
0.00

0.05
time (s)

time (s)

0106
-2106
-410

0
4
8
12
16
20

-6106
-8106
0.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.10

time (s)

Figure 7-9. Stresses in pile on element 0(top), 4,8,12,16


and 20 (tip) after impact. Free tip.

Pile driving
The left hand side picture shows the complete movement; on the right hand side a detail is shown for only
the first 0.1 s. For no pile tip movement the residual pile movement after impact is of course also zero.
With a free pile tip a lasting movement occurs. The detail in Figure 7-7 shows that this movement is not
constant but wanders through the pile. Sometimes there is no movement and with the next shock wave
there is further movement.
It is interesting to see the pressures as calculated for the first 0.1 s in Figure 7-8 and Figure 7-9. When the
pressure wave is reflected at the tip of the pile, the force is for the lower end of the pile twice the original
loading force for the situation without pile movement shown in Figure 7-8. This is also the case for the
tension wave after reflection of the stress wave to the top part of the pile. But with a free tip the force
remains below the original loading force. This means that the loading increases if the pile resistance
increases, even at the same impact force. A consequence of this will be dealt with in the next section.

Figure 7-10. Contour lines force in pile versus position (y-axis) and time (x-axis).

Figure 7-10 shows the results in another way. It shows a contour plot of the stresses calculated in the pile
for the situation without tip movement as a function of the time step and the position in the pile. It shows
the triangles at the tip (the series1 to series 19 presents the position on the pile) with high and with low
stresses (blue and purple). The time step of 0.00245 s is chosen in a way that within one time step the
wave travels over 1 m (=one element). This time step gives the most accurate results in the numerical
solution procedure chosen.

7.3

Stresses in piles and impedance

The stresses in the pile can become critical. To check the stresses for various pile types the strain can be
written as a function of the velocity of the pile (v) and the wave velocity (c), see also eq. (7.2):

w w t v

z t z c

(7.8)

The stress in the pile is then:

v
c

E. E v E

(7.9)

47

Geotechnical Engineering
With the usual parameters for steel, concrete and timber this leads to:

40v steel
10v concrete
2 4v timber

(7.9)

Where the velocity in the pile v has to be in m/s and is in MPa or N/mm2.
The starting velocity v in the pile will be more or less the same as the end velocity of the block when it
hits the pile. The transfer of a certain driving energy is necessary. However, if the required velocity of the
block is too high, leading to damage to the pile, it is possible to increase the weight of the block and to
reduce its velocity or to install a cushion pad (Heimutsvulling) between the block and the pile. This
elastic material (normally wood) leads to a lower impact force and a longer duration of the impact, so that
more energy can be transferred to the pile without damaging the pile. A well designed cushion pad has an
impedance that matches the impedance of the pile.
The impedance of a pile (Zp) is defined as the relation between the velocity and the force on the pile:

F Z pv

(7.11)

In the same way the impedance of the cushion pad (Zb) defined:

F Zbv

(7.12)

From eq. (7.9) it follows directly:

Zp

EA p
c

(7.13)

With Ap the cross-sectional area of the pile and for Zb holds:

Z b m.k

(7.14)

With m the mass of the block and k the spring constant of the cushion pad.
The loading on the pile is a function of the total impedance:

1
1
1

Zt Z p Zb
as shown in Figure 7-11.

48

(7.15)

Pile driving

Figure 7-11. Pile loading as a function of the impedance of the pile and the cushion pad.

7.4

Impact force

Up to now a constant impact was assumed in the calculations. In reality this force is a bit more
complicated as can be seen from Figure 7-12. This figure is also still a simplification from reality but it
shows the various processes involved.

Figure 7-12. Force as a function of time for a diesel block.

49

Geotechnical Engineering

7.5

Testing piles

The same principles as described above can also be used to test whether or not piles have been damaged
during pile driving.

foto GME consultants

Figure 7-13. Pile testing.

Figure 7-14. Possible results PIT.


(http://www.tradeindia.com/fp1282666/Pile-Integrity-Test-html)

The method is called Hamertje Tik in Dutch and Pile Integrity Testing (PIT) in English. The top of
the pile is hit with a hammer. The waves travelling through the pile are measured and recorded. The
procedure is shown in Figure 7-13. Possible results are shown in Figure 7.14. In case there is something
wrong there will be extra reflections. This can be a crack in the pile or for cast-in-situ piles also a location
where the pile is thicker or thinner.
A real report, as shown in Figure 7-15, shows a bit more complicated behaviour, but also here the pile
with an anomaly can easily be distinguished.

50

Pile driving

Figure 7-15. Results of PIT (http://www.piletest.com/show.asp?page=PocketPETReport.gif&img=1)

Another way of pile testing is to measure the waves in the pile during pile driving. The advantage is that it
is immediately known when something goes wrong. However, the interpretation is a bit more difficult due
to the non-linear behaviour of the soil during driving. Strain and velocity are measured separately to be
able to divide the signal in the incoming and the reflecting wave.

7.6

Pile driving by vibration

Piles and sheet piles are regularly installed by vibration. The idea is that the vibrations decrease the soil
resistance so that a small static force is sufficient to overcome the total soil resistance. The vibrations are
made by rotating an out of balance mass at frequencies of 25 to 40 Hz. It is known from experience that
the amplitude of the vibrations has to be more than 5 mm, which leads to the following equation:

Me
d
m
With
m
Me
d

(7.16)

: the total mass of the (sheet) pile and vibrating block (and some soil during removal),
: the moving mass times the amplitude of the centre of gravity of the moving mass and
: the amplitude of the vibrations.

When a pile is removed from the soil using vibrations, the effective mass is higher (because in the
beginning there is also soil attached to the pile). This means that a vibration block with a higher eccentric
moment is needed to remove for example sheet piles.
The loading on a (sheet) pile during installation by vibration is the sum of the static loading (the weight of
the block) and the dynamic loading caused by the vibrations. This dynamic force (Fc) can be written as:

51

Geotechnical Engineering

Fc M e 2

(7.17)

With the angular velocity.


This dynamic force is much larger than the static force given by the weight of the pile and the vibrating
block.

7.7

Influence on surroundings

Pile driving by hammering or vibration will lead to vibrations in the surrounding soil and buildings and
may lead to damage to these buildings. In densely built areas, these vibrations are quite often the main
reason to choose another installation system, for example vibrating or certain not-driven in-situ cast pile
systems.
Vibrations spread through the soil in different types of waves, see Figure 7-16. The fastest wave is the
compression wave (a), followed by the shear wave (b) and the Rayleigh wave or surface wave (c) is the
slowest one. The deformation patterns for the various waves are shown in Figure 7-17. The compression
wave and shear wave have a geometrical damping that is inversely proportional to r, with r the distance
from a circular source. Therefore, only the Raleigh wave is significant at some distance, since this wave
has a geometrical damping that is inversely proportional to r. Apart from the geometrical damping there
is also the material damping in the soil that is proportional with e-ar. Where a is a soil dependent damping
factor. The wave amplitude as a function of the distance r from the source can therefore be written as:

u (r ) u 0

1 ar
e
r

(7.18)

In which u0 is the vibration velocity at the source and u(r) the amplitude at a distance r from the source.

Figure 7-16. Different wave types in the soil.

52

Pile driving

Figure 7-17. Various wave types, (a) compression waves, (b) shear waves and (c) Raleigh waves.

Calculation methods are available to estimate the velocity of vibrations, but it appeared that it is difficult
to quantify the source of the vibrations. The CUR in the Netherlands suggests an empirical approach
based on a large number (300) of field measurements. Vibrations can lead to damage, but can also be a
nuisance for the surroundings and the people living there. This can lead to a situation where pile driving is
forbidden, Figure 7-18.

Figure 7-18. Cartoon as illustration in a newspaper article describing how a judge forbids pile
driving in a residential area. (the Dutch text means: Pack up and go!)

Group effects during installation


Normally not one pile has to be installed but a number of piles. For large buildings or piled embankments,
this can be a considerable number, although two piles less than 3 times the diameter from each other can
already influence each other. In case of driven piles this leads to densification of the granular soil layers.
Due to this densification, the bearing capacity of the piles increases. This is an advantage, although this
advantage is generally not taken into account when designing a piled foundation. A disadvantage is that it
will take more effort to get the piles at the required depth. An example of the influence of densification
can be seen in the CPTs performed before and after the piling for the Millennium Tower in Rotterdam.
For the foundation of this building precast concrete square piles of 450x450 mm were driven into the
Holocene layers (with low CPT values) and 8 to 10 m into the dense Pleistocene sand with a spacing
1.5x2.0 m2. This means that, assuming only horizontal displacement of the sand, there will be a
densification of 6.75 %. The CPT value increases significantly due to this densification, see Figure 7-19.
The pile tip resistance and pile friction thus also increases significantly. Consequently it can become
difficult to reach the desired depth. However, it is also possible that other failure mechanisms become
relevant, see Section 8.3.2.

53

Geotechnical Engineering

Refusal of CPT
Figure 7-19. CPT before and after piling for the millennium tower in Rotterdam, shown right
(Lecture notes Van Tol).

It is possible to estimate the influence of densification of the sand on the CPT value and thus the pile tip
resistance using one of the empirical relations between the CPT value and the relative density of the sand.
An example of such a relation is the relation given by Schmertmann (1976):

Dr 0.34. ln

qc
61. '0v .71

(7.19)

Where:
Dr
: the relative density calculated using the void ratio (-)
qc
: the cone resistance (kPa)
v
: the vertical effective stress (kPa)
With this formula it can be derived that the ratio (f) between the cone resistance before and after pile
driving is a relation of the relative density before (Drb) and after (Dra) pile driving:

D Drb
f exp ra

0.34

(7.20)

Starting with a relative density of 62% this leads to an increase in tip resistance as shown in Figure 7.20.
The empirical relations between cone resistance and relative density are not very accurate, but they give
an indication of what can be expected. It shows that a significant increase in pile tip resistance is possible.
Starting in rather dense sand this increase can be larger, because dilatancy in the sand leads to an
additional increase in horizontal stress.

54

Pile driving
3.00
2.75

1.7x1.7

2.50

1.5x2.5

2.25
2.00

1.75

2.5x2.5

1.50

5x5

1.25

3.5x3.5

1.00

2x2.5

3x2.5

4x4

0.75
0.50
0.25

square piles 0.45*0.45 m

0.00
0.00

0.02

0.04

0.06

0.08

pile area / total area (-)

Figure 7-20. Theoretical increase in cone resistance f when driving a pile group in sand with a relative density of
62% by pile groups.

The increase in pile resistance will hamper pile driving and can lead to pile damage. Higher tensile forces
will occur in the piles that can lead to cracks. To minimize the risk of damage in heavy pile driving
works, the following measures are recommended:
The use of a heavy hammer
A heavier hammer has a higher energy at the same velocity.
The use of an appropriate cushion pad.
Use of high quality concrete for the piles (B65).
Piles used have to be older than 28 days, so that the concrete in the piles has the maximum
strength.
Use of a higher pre-tension in the reinforcement (up to 7 N/mm2).
Start with the pile in the center and work from there outwards.
Take into account the better soil conditions in the design (only possible for tension piles, since it
is not known how far the soil underneath the pile tips is densified).
Decrease the soil stress by excavation, if possible in the design.
Use jets or make a boring for the first meters (be careful not to disturb the soil too much).
Table 7.2. Pile characteristics Caland Tunnel (The Netherlands)
Pile characteristics

East bank

West bank

Dimensions
[mm]
Penetration level
[NAP m]
Concrete type

450 x 550

450 x 550

35

32 (pile length 15 to
27 m)
C55/67 (B65)

Pre-tension
[N/mm2]
Hammer
Damage %

C45/55 (Dutch
code B55)
4.5
Hydraulic IHC SC
110
13.5

6.1
Hydraulic IHC
SC 110
negligible

These measures really help as shown in table 7.2. Tension piles were used to keep the ramps of the
Caland Tunnel in the Netherlands in position. During the pile driving on the East bank 13.5% of the piles
were damaged (tension cracks but in most cases damage of pile head). For the West bank the piles were
of a better concrete and had a higher pre-tension and the penetration level was 3 m reduced. These
changes resulted in negligible damage on the West Bank.

55

Geotechnical Engineering

Pile settlement and group effects after installation

8.1

Negative skin friction

Paragraph 3.3.5 showed that negative skin friction can lead to excessive settlement of piles and the
buildings on top of it. Negative skin friction can be expected when ground surface settles more than 20
mm.
Negative skin friction will not necessarily lead to complete failure since if the pile settles somewhat, the
negative skin friction changes partly to positive friction. This reduces the force on the pile tip. This finally
ends in a stable situation, although with some deformation.
The development of negative skin friction is a slow process. Therefore drained parameters have to be
used. The following relations are used for a single pile:

K 0 'v tan

(8.1)

Fn Os h avg

(8.2)

With

skin friction
ratio between horizontal and vertical effective stress
vertical effective stress and
friction angle between the soil and the pile.

K0
v

And in the second formula


Fn

negative skin friction


average skin friction over the pile,
the circumference
the height over which negative skin friction is expected.

avg
Os
h

According to the Dutch Standard NEN6743, K0.tanin Holocene layers


As shown in the equation above, the negative skin friction is proportional to the circumference of the pile
cross-section. The pile tip force, however, is proportional to the area of the pile cross-section. This means
that negative skin friction has more impact on long slender piles.
Negative skin friction will most probably occur when a loading is applied on top of soft soil layers. This
is often the case in new housing areas where a few metres of sand are placed on top of the existing soil.
Negative skin friction occurs because soil settles and hangs on the piles. Therefore the influence of
negative skin friction will be relatively large in a closely spaced pile group, with only a limited amount of
soil between the piles.
The influence of a pile group can be calculated using the method developed by Zeevaert-De Beer, De
Beer (1966): In this method a cylindrical volume of soft settling soil is assumed around each pile. This is
an approximation since not all soil between piles can be captured with cylinders, see Figure 8-1.

56

Pile settlement and group effects after installation

pile

'v

soil around
pile

'v'vzz

Figure 8-1. Top view, soil around piles.

Figure 8-2. Cross section pile and soil slice.

Following this approximation the following derivation is possible, see also Figure 8-2:

'v
O
s
z
A

(8.3)

With (8.1) and:

m K 0 tan

Os
A

(8.4)

(8.3) can be written as:

'v
m 'v 0
z

(8.5)

With the boundary condition v = 0 for x = 0, the differential equation can be solved leading to:

'v

(1 e zm )

(8.6)

The formula resembles the formula of a silo. The negative friction leads to unloading of the deeper soil
layers.
The negative skin friction at a pile position z is:

Fs ,nk A( '0,v 'v )

(8.7)

Where:

Os
A
z
Fs,nk

o,v
v

: the density of the soil


: the circumference of the pile
: the area of the soil around the pile (top view, see Figure 8-1)
: the position along the pile
: the negative skin friction at a pile position z
: the vertical stress in the soil at depth z without pile (.z)
: the vertical stress in the soil at depth z with the pile

Combining (8.6) and (8.7) leads to :

57

Geotechnical Engineering

Fs; nk Az '[1

1
(1 e zm )]
zm

(8.8)

If the surcharge (P0) is applied this becomes:

Fs; nk Az '[1 (

p
1
0 )(1 e zm )]
zm ' z

(8.9)

For large pile spacings, the calculated negative skin friction becomes the same as for a single pile as could
be expected, see Figure 8-3.

negative skin friction (kN)

600

single pile limit


500
400

400 mm

Zeevaert
De Beer

300

16 m clay
' = 5 kN/m3
P0 = 40 kPa
K0 = 0.5
tan = 0.5

200
100
0
0

10

distance between piles (m)

Figure 8-3. Comparison single pile calculation and Zeevaert-De Beer for a pile group.

The forces caused by negative skin friction can be considerable, see also Figure 8-3.

8.2

Settlement

Loading a pile with a load significantly less than the failure load will result in some elastic settlement.
This settlement has 3 components:
1. The pile will shorten due to the loading
2. Some movement of the pile tip is necessary to mobilize the pile tip resistance
3. Some movement of the shaft is necessary to mobilize the shaft resistance

8.2.1

Pile shortening due to loading

The pile behaves linear elastic during loading. This means that if the average loading on the pile is
known, the shortening can be calculated. The average loading depends not only on the total load but also
on how the load is distributed over the pile. If negative skin friction occurs additionally, the average load
will be higher and thus the pile shortening will be greater, see Figure 8-4.

58

Pile settlement and group effects after installation

force in pile

Bepaling elastische
verkorting Wel

Fs,gem, nf
Fs,gem

Bepaal gemiddelde
kracht Fs;gem in
paalschacht
Dan

soil settles
more than
pile

pile settles
more than
soil

Wel= Fs;gem L/ Ep Ap
forces in pile with and without
negative fricion

Figure 8-4. Average force in pile with (Fs,gem,nf) and without negative skin friction (Fs,gem).

The shortening of the pile due to its loading can be written as:

Wel

Fs , gem L
(8.10)

E p Ap

This settlement is usually not more than a few millimetres.

8.2.2

Settlement of tip and shaft

The settlement that can be expected from the tip and the shaft during loading is determined empirically.
The load settlement diagram of the tip has been determined, using a specially equipped pile that allows
determining the load settlement diagram of the tip separately. The load settlement curve for the shaft has
been determined in the same way. The resulting curves are shown in Figure 8-5 and
Figure 8-6 respectively. In these curves the deformation of the pile itself is not taken into account.

59

Geotechnical Engineering
Fs , shaft

tip
tip

Displacement mm

Fs , shaft, max

elastic
area

Piles in clay and bored piles


In sand
driven piles
CFA piles
bored piles

Figure 8-5. Empirical load settlement curve of pile


tip.

Soil replacing piles in


dense sand

Figure 8-6. Empirical load settlement curve of pile shaft.

The load settlement curves are different for different piles. The difference between the two figures is that
the tip settlement is presented as a function of the equivalent pile diameter while the shaft settlement is
presented as an absolute value and thus independent on the pile diameter. For a given pile diameter,
Figure 8-5 can be used to construct a load settlement plot in absolute values. The total load-settlement
relation for the pile can then be constructed, see Figure 8-7.

Figure 8-7. Graph to determine the total load settlement curve of a pile.

60

Pile settlement and group effects after installation

8.3

Group effects after installation

8.3.1

Elastic deformation

A pile group will settle more than an individual pile. The pile group creates an area of high pressure
below the piles. This high pressure can create settlement of deeper layers, see Figure 8-8. In this figure a
soft soil layer is overlain with a stiff soil layer, the bearing layer for the piles. For one pile the loading of
the soft layer will hardly result in any settlement, since the increase in soil stress, caused by the loading of
one pile will only be small. For a group of piles, the zone with an increased stress is larger and therefore
also extends to a greater depth. This will create an additional settlement in the area W2.

z;4D

Figure 8-8. Sketch of pile group and the corresponding CPT


(green line represents the CPT tip resistance).

The settlement (W2) can be written as:

W2

m ' z , 4 D 0.9 A4 d

Where
m
z,4D
A4D
Eg,gem
W2

(8.11)

E g , gem
factor depending on the geometry,
stress at 4Deq below the pile tip,
loaded area below the group
average Youngs modulus in the soil below the pile group up to a depth of W2 below the
pile group.
shortest distance of the loaded area at 4Deq below the pile tips of the piles in the group.

Youngs modulus, Eg,gem, is determined assuming that it is 5 times the value of qc, the measured cone
resistance. As shown in Figure 8-8, the loaded area A4D is determined by assuming a stress distribution
that spreads under and angle of 45 degrees with depth. This method is only a rough first estimation of the
additional settlement caused by the pile group. The soil layers close to the pile tip have a larger influence
than the layers at larger depth. Furthermore the determination of Eg,gem by using qc, is just an
approximation. If it turns out that the group effect on the pile settlement is critical, the results should be
verified with numerical calculations.
All piles get the same displacement if they are loaded by a stiff cap. Numerical calculations showed that
the outer piles of such a pile group get more loading than the others. So, the piles do not all have the same
load. Plastic soil behaviour will therefore start with the outer piles at somewhat larger loads. This will
then equalize the loads on the different piles.

61

Geotechnical Engineering

8.3.2

Plastic deformation

It was shown in Section 0 that the tip resistance of piles in a pile group increases due to densification. For
friction piles, the pile resistance of a group can be lower than the resistance of the sum of the single piles.
It is possible that the friction that can be mobilized by the total group or by some piles in the group is less
than the sum of the friction of the single piles, see Figure 8-9.

Figure 8-9. Possible failure modes for friction piles in a group.

8.4

Horizontal loading due to soil movement

So far, we have considered axially loaded piles. Piles can however also be loaded horizontally. A classic
example is an abutment. If the abutment (bruggenhoofd) is made on soft soil, the soil underneath the
abutment will settle. This causes horizontal soil movements. Piles used for the foundation of a bridge or a
viaduct will get a horizontal loading Figure 8-10 that creates moments (torques) in the pile. These
moments will be even larger when the top of the pile is fixed by the structure on top of it and cannot
move. The result can be cracks in the pile. This was shown by the following example. A test pile was
placed before the construction of an embankment with a height of 6-7m. The embankment was
constructed on 15 m soft soil layers of clay and peat. The pile was placed in the toe of the embankment.
The pile was pulled out of the soil after the embankment had been placed.

Figure 8-10. Stress distribution around a pile before (left) and after (right) horizontal movement to the right.

62

Pile settlement and group effects after installation

Figure 8-11. Test location pile loaded horizontally by soil movement (lecture notes Frits van Tol).

The construction of the embankment led to a large horizontal deformation, see Figure 8-12.

Figure 8-12. Measured horizontal deformation of the concrete pile (lecture notes Frits van Tol).

This resulted in cracks as shown in Figure 8-13.

63

Geotechnical Engineering

Figure 8-13. Measured cracks in the pile due to horizontal bending (lecture notes Frits van Tol).

The following procedure is used to estimate the horizontal loading on the pile. It should be noted this is
an estimation, not an accurate calculation. The relation between the pressure on the pile and the
movement of the soil with respect to the pile is assumed to be described with the P-Y curve, shown in
Figure 8-14. In this figure, p is the horizontal pressure exerted by the soil on the pile, pu the ultimate
horizontal pressure exerted by the soil on the pile, Y the lateral deformation and Yu the lateral deformation
at the ultimate horizontal pressure. The other parameters can be described as:
For clay:

z
) su 'v 9.su
D

pu (3 0.5
Y50 2.5 50D

(8.12)
1/ 3

Y
p
0.5
pu
Y50

for

p pu

else

p pu

In this formulas Y is the lateral deformation, Y50 the lateral deformation at 50% of pu, 50 the strain at 50%
of the deviatoric stress in a triaxial test and D the pile diameter. 50 decreases for stiffer clay and is
approximately 1-2% between 50-100 kPa. Furthermore, it is assumed that there is no axial loading on the
pile.

Figure 8-14. P-Y curve for calculation of horizontal loading on pile.

64

Pile settlement and group effects after installation


The shape of the assumed curve results in a stiff response of the soil for the first horizontal deformation,
as shown in Figure 8-14, but with a decreasing stiffness as the loading reaches pu.
The API (American Petroleum Institute) and other organisations have also defined these P-Y curves for
sand. However, it appears that for sand a lateral displacement Y/D=1% is already sufficient for full
mobilisation of the maximum lateral pressure on the pile and therefore such a P-Y curve is only useful
when the deformations are very small.

65

Geotechnical Engineering

Pile load tests

9.1

Type of tests

There are 3 types of pile load tests. They differ in the rate at which the load is applied. The reason why
dynamic loading is used is quite simple. The bearing capacity of most piles is quite significant and that
means that testing the piles needs quite some weight to get the required settlement, see Figure 9-1. It is
possible to reduce the amount of weight needed by installing anchors in the surrounding soil and mobilize
the reaction force from there, but that leaves the problem of possible influence and is also not very simple
(depending on the soil conditions).

Figure 9-1. Load necessary for a pile load test.

A lower load can be used when the acceleration or deceleration of the load is more than 1 g. The most
well-known example is pile driving. In the numerical example discussed in Chapter 7, the block has a
velocity of around 4 m/s and stops within 4 ms. This means that the deceleration of the block during
impact is 1000 m/s2, thus around 100 times the acceleration of gravity (9.81 m/s 2). This means that only
1/100 of the load of a static pile loading test is necessary to move the pile. The disadvantage is also clear
from the same chapter: a dynamic load introduces waves in the pile and, different from static loading, the
loading differs from time to time. Thus, the dynamic test is the cheapest test (up to 10 tests a day), but
also the least accurate one. The static test is the most expensive test (one test a week), but also results in
the highest accuracy and the, what is called statnamic tests, is situated in between them (4 tests a day).

9.2

The dynamic loading test

A dynamic test is comparable to pile driving of an instrumented pile. The waves in the pile are analysed
with a numerical programme as discussed in Chapter 7 but more complicated. The bearing capacity and
friction are back calculated from the measured waves in the piles. Opinions differ about the quality of
such a back calculation. In some papers it is mentioned that this is quite good, some people doubt the
quality and dont trust the results of a dynamic loading test. The fact that the other, more expensive
methods still exist implies that the dynamic test cannot give all the answers for varying circumstances.
Sometimes the dynamic test is calibrated with a static loading test performed on one pile. This increases
the reliability of the dynamic tests in the direct neighbourhood.

66

Pile load tests

9.3

The static loading test

The static test is the most simple when it comes to interpretation of the test. More and more load is added
to the pile, normally by means of a hydraulic jack, until the pile fails (that means: until the displacement
is 10% of the pile diameter, see also Figure 8-5). Force and displacement are recorded. A simple version
of the test is shown in Figure 9-2.

Figure 9-2. Example of a pile load test

More information can be obtained by introducing unloading cycles before pile failure to see where plastic
deformation starts and by measuring the friction force and the tip force separately.

9.4

The statnamic test

The statnamic test is performed with a load that is accelerated to 25 g. This means that there is a
considerable mass reduction possible, compared to a static test. The idea is that the loading is still much
longer than the time that it takes for a wave to travel through the pile, approximately 0.1 s. So at a certain
moment the whole pile is loaded, as is the case in a static test, see the sketch of the load distribution in
Figure 9-3.

67

Geotechnical Engineering
STN

DLT

LOAD

TIME

DEPTH
H

DLT

STN

SLT

z
z
Figure 9-3. loading for a dynamic loading test (DLT), a statnamic test (STN) and a static loading test (SLT).
Loading in DLT and STN as a function of time and stress ( ), velocity (v) and displacement (u) as a function of pile
depth (z).

The total loading on the pile (Ftot) is a function of the static force that only depends on the
displacement/penetration of the pile, a force that depends on the strain rate (thus on the penetration
velocity) and a force that depends on the acceleration. In formula:

d 2u du
Ftot m 2 c f u
dt
dt

(9.1)

Where
u
is the pile movement,
m
the mass of the pile
c and f functions of the penetration velocity and the penetration respectively.
We need to know the last term of eq.(9.1).to compare the result of a statnamic test with the result of a
static test. The loading is measured and the first term on the right hand side of the equation is rather easy
to determine. However, the second term is difficult to determine. Therefore, the point in the loading curve
where du/dt=0 is used, the point Umax in Figure 9-4. The acceleration of the pile is determined from the
displacement as a function of time, see Figure 9-5.
This makes it possible to make a reasonable accurate prediction of the loading capacity. The accuracy of a
static test is 10%, of a dynamic test 30% and of a statnamic test it is said to be 20%. It is well possible
that there is quite some engineering judgement in these numbers since not many piles have been tested
with more than one method.

68

Pile load tests


Bochum, Pfahl 57
Statnamic signals, Cycle 3
Force [MN]

Force

LOAD [MN]
LOAD [MN]

2
1
0

Fstn(max)
Fstn(max)

Displacement [mm]

TIME [ms]
TIME [ms]

Displacement

0
5
10
15

Velocity [m/s]

Umax
Umax
DISPLACEMENT
DISPLACEMENT
[mm]
[mm]

LOAD [MN]
LOAD [MN]

Velocity

0.4
0.2
0.0
-0.2

Fstn(max)
Fstn(max)

Umax
Umax
DISPLACEMENT
DISPLACEMENT
[mm]
[mm]

Figure 9-4. Example loading curve in a statnamic test.

Acceleration

Acceleration [m/s2]
20
0
-20
-40
0

50

100

150

200

250

Time

Figure 9-5. Result of statnamic test and determination of


the acceleration.

The influence of pore water pressure generation in sand and creep effects in clay are well-known sources
of inaccuracy for the statnamic test.

69

Geotechnical Engineering

10

Physical modelling in geotechnics

10.1

Why physical modelling

A physical model in geotechnics is a (scaled) simplified model of the real world. Normally it contains
soil, but not the complex layering that can occur in the field.
Physical modelling is regularly used in geotechnics, especially to test new concepts or to investigate
mechanisms. In quite some engineering areas, physical models have been replaced by numerical models.
In geotechnics (as in coastal engineering) this is not often the case, because the soil behaviour is not
completely understood in case of large deformations, complicated soil structure interaction, cyclic and
dynamic loading. Therefore a description in a numerical model has limitations.
Physical models are quite often used in combination with numerical models. In that case, the physical
model is a simplified version of reality, for example a pile in homogeneous soil. The results of the
physical model are compared with the results of a numerical model. If there is sufficient agreement, the
numerical model can be used to simulate the real situation with more complicated loading conditions or
soil layering etc.
Most applications of physical modelling in geotechnics are for the foundation of offshore structures,
underground structures or fundamental research on various topics as piles, soil-structure interaction or 2phase flow.

10.2

1-g and N-g models

Typical for geo-engineering is that not only 1-g models (models under the normal acceleration of gravity)
are used, but also, what is called N-g models, models within an increased acceleration field. This
acceleration field is generated with a geotechnical centrifuge. The reason for the N-g tests is that the soil
behaviour is non-linear and stress dependent. Therefore it is preferred that the scale model has the same
stresses as the field situation. In a 1-g surrounding, a scale model has lower stresses because volumes and
mass are scaled with N3 and areas with N2. The stress underneath for example a cube of 1x1x1 m will be
10 times higher than the stress underneath a cube of 0.1x0.1x0.1 m. In an N-g model the acceleration is
increased leading to the same stress distribution in model and prototype, see Figure 10-1.

G / n3

n
2
A/ n

Figure 10-1. Stress underneath a cube, reality and model in an n-g situation.

Scaling rules follow from a dimensional analysis, see Table 10-1.


The permeability is an exception. Strict application of the scaling rules would demand that the grain size
is reduced with N (that would lead to a reduction of the permeability with N2) and that the viscosity of
the pore water is also reduced by a factor of N (increasing the permeability with N). Both demands
together would lead to a permeability that is N times lower in the model than in the prototype and since
the pressure gradient in the model is N times higher this would lead to the same water velocity. However

70

Physical modelling in geotechnics


both demands are not very practical for a sand or clay model. Therefore the usual procedure is that the
same sand is used in the model as is used in reality and the viscosity is also not scaled. This leads to a
permeability that is N times too high. Consequently, the time for consolidation scales with N2 and not
with N (as follows from Newtons law for a model that has N times smaller dimensions, but an N times
higher gravity). If both dynamic forces and consolidation effects are important, it is necessary to increase
the viscosity by a factor of N. In case the structure in prototype consists of large stones (for example a
breakwater) the turbulent permeability is independent of the viscosity of the water and scales with the
grain size diameter. Thus for such a structure a geometrical downscaling with a factor of N is required.
This is advantageous: the large unscaled stones would not fit in a model.
Table 10-1. Scaling rules.
Parameter

Acceleration
Length
Velocity
Mass
Force
Stress
Permeability
Time
(consolidation)
Time dynamic)

Value centrifuge
model assuming
1 in prototype
N
1/N
1
1/N3
1/N2
1
N
1/N2
1/N

Value in model
assuming 1 in
prototype
1
1/N
1/N
1/N3
1/N3
1/N
1
1/N2
1/N

Dimension

m/s2
m
m/s
kg
N
N/m2
m/s
s
s

71

Geotechnical Engineering
Table 10-2. Table 3.1 from Mayne et al. (2009): Physical modelling options.

72

Slope Stability

Slope Stability

11

(Until Section 11.3 taken from Soil Mechanics, Verruijt and Broere, 2011)
For the analysis of the stability of slopes of arbitrary shape and composition various approximate methods
have been developed. Most of these assume a circular slip surface. Using a number of simplifying
assumptions a value for the safety factor F, the ratio of strength and load, is determined. The circle giving
the smallest of F is considered to be critical. The multitude of methods (developed by Fellenius, Taylor,
Bishop, Morgenstern-Price, Spencer, Janbu, among others) in itself illustrates that none of them is exact.
The results should always be handled with care. A value F = 1.05 gives no absolute certainty that the
slope will stand. In this chapter two of the simplest methods are presented.

11.1

Circular slip surface

Most methods assume that the soil fails along a circular slip surface, see Figure 11-1.

Figure 11-1: Circular slip surface.

The soil above the slip surface is subdivided into a number of slices, bounded by vertical interfaces. At
the slip surface the shear stress is , which is assumed to be a factor F smaller than the maximum possible
shear stress, i.e.
1

= ( + tan )

(11.1)

The factor F is assumed to be the same for all slices, an assumption that is common to all methods.
The equilibrium equation to be used in conjunction with a circular slip surface is the equation of
equilibrium of moments with respect to the center of the circle. This equation gives
sin =

cos

(11.2)

Where

73

Geotechnical Engineering
h
b

average height of a slice,


width of a slice,
volumetric weight of the soil in the slice
radius of the circle.

More generally it can be defined that bh is the weight of the slice, possibly consisting of a sum of parts
with different unit weight. If all slices have the same width, it now follows from (11.1a) and (11.1b) that
F=

(+ tan )/ cos
sin

(11.3)

This is the basic formula for many computation methods. The various methods usually differ in the
method of calculating the normal effective stress

11.2

Fellenius

In Fellenius method, the oldest method for the analysis of slope stability, it is assumed that there are no
forces between the slices. The only remaining forces acting on a slice, see Figure 11.1b, then are the
weight bh, a normal stress and a shear stress at the bottom of the slice.

Figure 11-2: Fellenius

The normal stress can most conveniently be expressed into the known weight by considering the
equilibrium of the slice in the direction perpendicular to the slip surface. This gives
= h cos

(11.4)

And, because = + p, with p the pore water pressure:


= h cos - p

(11.5)

Substitution into (11.3) gives:


F=

{+( ) tan / cos }


sin

This is the Fellenius formula.

74

(11.6)

Slope Stability
For a slope in homogeneous soil the computation can be executed by assuming a certain location of the
circle and subdividing the sliding soil wedge into 10 or 20 slices. By measuring the values of and h for
each slice the value of the stability factor F can be determined. This must be repeated for a large number
of circles, to determine the smallest value of F. In non-homogeneous soil the computation is somewhat
more complicated because for each slice the value of h must be determined as the sum of the
contributions of a number of layers in the slice.
Several objections can be made against this method. To begin with, a sound theoretical base lacks for all
slip surface methods for materials with internal friction. It is thus not know how the resulting stability
factor F corresponds with the real safety of the slope.
But there are other objections as well. Disregarding the forces transmitted between the slices is a severe
approximation, and vertical equilibrium is violated. Furthermore, there is an internal inconsistency in
stating on one hand that sliding occurs along the circle, and on the other hand stating that the horizontal
and vertical directions are the directions of principal stress (as it is assumed that there are no shear
stresses on vertical planes). This inconsistency can best be seen by considering the slice for which = 0.
At that slice = h, and it is assumed that there is a shear stress ( p)/F on that slice. This violates the
assumption that the vertical direction is a direction of principal stress. Horizontal equilibrium of that slice
is also clearly violated. For other slices vertical equilibrium is violated, as only the condition of
equilibrium perpendicular to the slip surface is taken into account.
Fellenius method has the property that in a number of special cases it confirms certain limiting values.
For instance, for an infinite slope in a dry frictional material without cohesion, one obtains from (11.6),
assuming a straight slip surface at a depth d below the slope, and taking p = c = 0,

F=

cos tan
sin

tan

= tan

(11.7)

(From here some changes were made in the text of Verruijt and Broere, 2011)
This is in perfect agreement with the stability formula for an infinite slope that can be derived without
using slices.
In the case of a slope under water, in the absence of groundwater flow, see Figure 45.2, the limiting value
(11.7) is not immediately recovered, because the formula for the effective stress as used by Fellenius is
only valid when the water table is lower than the soil surface and thus not for under water slope.
Furthermore, the pore water in the Fellenius formula leads to a strength reduction because the normal
stress is reduced but the loading reduction due to the pore water is not taken into account. The Fellenius
formula for a slope of frictional material without cohesion (sand) filled with water results in the same
stability formula as what can be calculated for an infinite slope with a horizontal water flow with the
same gradient as the slope angle (see also section 11.4.5):

w / cos 2 tan

tan

(11.8)

With w =
For a problem with a horizontal phreatic surface, the Fellenius formula might be modified by using the
volumetric weight under water, ( - w)h rather than h, and using the excess water pressure with respect
to the hydrostatic water pressure for p.

75

Geotechnical Engineering

11.3

Bishop

A method that is frequently used in engineering practice is Bishops method. In this method, the forces
between the slices are not neglected, but it is assumed that the resultant force is horizontal, seeFigure
11-3. By considering the vertical equilibrium of each slice only, the horizontal forces do not enter into the
computations, however.
The basic equation is again the equation of moment equilibrium, eq. (46.3). Vertical equilibrium of a slice
now requires that
sin

sin

h = + cos = + p + cos

(11.9)

Figure 11-3: Bishop

If in this equation the value of is written, as = (c + )/F, the result is


tan tan
)

(1+

= h p tan

(11.10)

Substitution of into (11.3) now leads to the final equation for Bishops method,
+()
cos (1+tan tan /)

F=

sin

(11.11)

Because the stability factor F also appears in the right hand side, it must be determined iteratively, by
starting from an initial estimate (for instance F = 1.0), and then calculating an updated value using
(11.11). This must be repeated until the value of F no longer changes. In general the procedure converges
rather fast. As the computations must be executed by a computer program anyhow (many circles have to
be investigated) the iterations can easily be incorporated into the program.
If = 0 the Bishop and Fellenius methods are identical. For > 0, Bishops method usually gives
somewhat higher values for F. Because Bishops method is more consistent (vertical equilibrium is
satisfied), and it confirms known results for special cases, it is often used in geotechnical engineering.

76

Slope Stability
Various other methods have been developed, but the results usually differ only slightly from those
obtained by Bishops method. That may explain its popularity.
Taking into account the theoretical weakness of Fellinius method, it is remarkable that for an infinite
slope of constant soil material with a failure surface at a constant depth d, both methods result in exactly
the same value of F. This is not straightforward since the result of Bishops method has to be determined
iteratively. However, assuming that both methods lead to the same value of F, it is possible to insert the F
obtained by Fellenius method in the F at the right hand side of the formula of Bishop. For a constant
depth d, this leads to the same result for both formulas.

11.4

Other mechanism that lead to slope instability

Slip circle instability is an important failure mechanism for slope stability but not the only one.
This section summarizes other possible failure mechanisms.

11.4.1

Squeezing of soft layers underneath a slope

In case a sandy slope is constructed on top of a weak soil layer, squeezing of the weak soil layer may
occur. The weak soil is squeezed from underneath the sand, see Figure 11-4.

Figure 11-4. Squeezing of peat from underneath a sand layer during road construction.

A simple and save approach to calculate the stability against squeezing is presented below. In this
approach it can be taken into account that there is a geotextile between the dam and the soft layer. This
can take the shear stresses that occur when the soft soil is loaded with the dam, see below, and therefore
increase the stability against squeezing.

depth (m)

5
Y dam
y mud

sand

-5
-10

Pp
-20

-10

Pa
0

10

20

geotextile

Pa
30

40

50

layer of soft soil


60

70

X (m)

Figure 11-5. Sketch to calculate squeezing.

77

80

Geotechnical Engineering

The maximum passive resistance that can be mobilized at X=0 m in Figure 11-5 can be written as:

Pp

1
y ' h 2 2 su h
2

Where:
Pp

h
su

(11.12)

= the total passive resistance over the clay layer [kN/m]


= the effective unit weight of the clay [kN/m3]
= the thickness of the clay layer [m]
= the undrained shear strength [kN/m2]

The passive pressure, see also Figure 11-5, can be written as:

Pp

1
y' h 2 2su h qh
2

(11.13)

Where:
q
= the overburden pressure [kN/m2]
The passive pressure in the layer of soft soil is higher under the top of the dam than it is under the toe of
the dam, due to the weight of the sand. However, squeezing the clay from underneath the middle of the
dam is only possible when apart from the passive pressure at x=0 also the shear stress between the
geotextile and the clay and the shear stress between the clay and the sand is exceeded. This results in the
formula:

1
1
y' h 2 2su h x y' h 2 2su h qh
2
2

(11.14)

Which can be simplified to:

q 4 su

x
h

(11.15)

In this equation is the shear resistance at the boundary between the top and bottom of the squeezed mud
layer and is nearly equal to su. = 1 when there is no geotextile between the sand and the mud that can
take the tensile forces from the squeezing mud. For a geotextile of sufficient strength = 2. The
minimum strength of the geotextile (without safety factors) = su.L kN/m where L is the length of the
geotextile between the toe of the dam and its centre (30 m in the example shown in Figure 11-5).

11.4.2

Liquefaction

In slopes of loose sand under water another dangerous failure mechanism may be present. The
slope may lose its stability by excess pore water pressures. The mechanism is simple to describe although
difficult to model. In the loose sand the grains form a grain skeleton, see
Figure 11-6. Sketch changes in grain skeleton and pore water pressures during liquefaction.

. If there is some disturbance, the grains move with respect to each other and the grain skeleton breaks
down, see Figure 11-6b and settles down in a lower porosity, Figure 11-6c. The disturbance can be
erosion of the sand that steepens the slope, some minor seismic activity, low tide that leads to an outward

78

Slope Stability
directed hydraulic gradient along the slope, an anchor dropped on the slope etc.. Without a grain skeleton
to support the grains they will fall down. This will take some time, because they can only fall down if the
water is expelled from between the grains. In this situation the sand body behaves as a sand-water mixture
and flows along the slope. This will go on until the sand has settled at a lower porosity. This process can
repeat several times in very loose sand, see Figure 11-7. This figure shows the results of a test where a
sand body was made using hydraulic fill below the water line. The very loose sand liquefied several times
leading to a jump wise increase of the pore pressure and a slow decrease afterwards, where every increase
in pore pressure also led to a decrease of porosity that was measured in the liquefied sand with an
electrical resistivity probe.
A slope failure caused by liquefaction in sand leads generally to very gentle slopes (1:20 to 1:40
typically). Since liquefaction can hardly be avoided when making a sand dam below the water line by
hydraulic fill this dam will have very gentle slopes. This can only be avoided by making a gravel or mine
stone berm or use geosynthetic tubes or containers.
grain skeleton

grain skeleton partly destroyed

grain skeleton higher density


decrease
in height

pore press.
hydrostatic

excess pore
press. due to liquefaction

pore press.
hydrostatic

Figure 11-6. Sketch changes in grain skeleton and pore water pressures during liquefaction.

Figure 11-7. Porosity and excess pore pressure both measured at 1.4 m from the bottom in a 2 m high sand body that
liquefied during a hydraulic fill experiment below the water line.

11.4.3

Retrogressive flow slide or Breaching

Another possibility to get an instable slope is a retrogressive flow slide caused by breaching of the sand.
The mechanism was described for the first time by the US-Army Corps of Engineers in the 80-ties. It was
noticed along the Mississippi that underwater slopes failed and that gentle slopes were formed,

79

Geotechnical Engineering
comparable to those found in case of liquefaction, although there was no loose sand in the slope. The
following process occurs: a local instabilityon the slope may cause a sand-water mixture running down
along the slope. If the slope is long enough, then this sand water mixture erodes the slope. This erosion
goes on until equilibrium is reached between the erosion caused by the sand-water mixture and
sedimentation of sand from the mixture. This equilibrium is reached at a very gentle slope. The process is
sketched in Figure 11-8 for the situation where a local instability does not lead to a large change in the
underwater slope, because the sand-water mixture flow is too small to create a lot of erosion on the slope;
and for the situation where an instable slope occurs. Gentle slopes are only possible for rather high slopes,
because sufficient sand-water mixture has to be present to keep the erosion going. For this reason this
mechanism was never found in model tests. This process is much slower than the liquefaction process. It
is reported that the breaching may occur for more than half an hour. The breach velocity can be calculated
from:

v
cot
1 S
cot
vw

(11.15)

Where:

vw

k
1 n0 cot
n

(11.16)

Where:
vS
: is the breach velocity

: slope angle
vw
: active wall velocity (in Dutch the walletjessnelheid)
k
: permeability of the sand
n
: change in porosity between the packed sand of the slope and the maximum porosity

: relative density
n0
: porosity of the sand

: friction angle of the sand.


The active wall velocity vw is low for sands with a low permeability and a high relative density. For very
loose sands with a porosity close to the maximum porosity, vw and thus vs will be very high, especially
when > .

Figure 11-8. Retrogressive flow slide in an underwater


slope of sand by breaching. (Top: more or less stable
slope, Bottom: unstable slope).

80

Figure 11-9. Stable end slopes as a function of the


active wall height.

Slope Stability

11.4.4

Breaching or liquefaction

As described in the sections above, the mechanisms for breaching or liquefaction are quite different and
easy to distinguish when you are present when the failure occurs. However, after a failure occurred, it is
in most cases not obvious what mechanism has caused the failure. The end result is more or less the same:
a slope with a very flat angle. Since above the water line the capillary forces result in some apparent
cohesion in the slope, the slope above the water line is mostly steep, see Figure 11-10. Furthermore, sand
layers often consist of layers with a lower and higher density. It is therefore possible that a local
liquefaction in a sand layer of loosely packed sand triggers the breaching of a larger part of the slope.
Consequently, it is not known what mechanism causes most of the underwater slope failures and this is
still a research area. What is known until now, is that under water steep slopes over a large slope height
(see Figure 11-9) are likely to be unstable and that the risk of instability increases when the slope (partly)
consists of loose sand.

Figure 11-10. Slope failure in sand. Liquefaction or breaching.

11.4.5

Groundwater flow within a slope

Groundwater flow can stabilize or destabilize a slope. An inward directed flow will increase its stability,
but an outward directed flow or a flow parallel to the slope will destabilize that slope. The destabilization
is most important. Experiments have shown that an inward directed flow perpendicular to a sandy slope
results in a stable slope for higher slope angles than the friction angle of the sand. But an outward directed
flow leads to a 2 times as high decrease of the critical slope angle (Van Rhee & Bezuijen, 1992).
The following formulas apply for the safety factor Fs of slopes of granular material (sand, gravel) with a
constant friction angle:
For a dry slope or an underwater slope without flow:

Fs

tan
tan

(11.17)

For a slope with a downward directed groundwater flow parallel to the slope, thus with a gradient of
sinand the same pressure on both sides perpendicular to the slope, see Figure 11-11:

81

Geotechnical Engineering

Fs

w tan
tan

(11.18)

For a slope with a horizontal outward directed groundwater flow, which must have a gradient of tan,
see Figure 11-12:

Fs

w / cos 2 tan

tan

(11.19)

In these formulas is:


Fs
safety factor,

density of the soil,


w
density of the water,

slope angle and

friction angle.

w.d.l

w.d.l.tan/sin

i=sin
d

d/sin

.d.l.sin

i=tan

.d.l.sin

.d.l.cos
.d.l

Figure 11-11. Sketch stability slope granular material,


flow parallel to slope.

.d.l.cos
.d.l

Figure 11-12. Sketch stability slope granular


material horizontal flow.

In both cases the safety factor is much less than in case of a dry slope. For example, for the situation with
flow parallel to the slope the safety factor is approximately only half the safety factor of a dry slope (for
=20 kN/m3 and w=10 kN/m3). A downward water flow along a slope therefore quite often leads to a
dangerous situation.

11.4.6

Monitoring pore water pressures

Construction of a sand body, for example for a road on soft soil, may lead to unstable situations, see
Figure 11-13. , The weight of this sand body causes shear stresses in the soft soil underneath at the edges
of the sand body, but the soft soil has to consolidate before the effective stress increases. Before
consolidation of these soft layers, there will be an increase in pore water pressure, while only limited
shear stress can be taken by the soft soil. The common procedure is that the sand is applied in layers and
that slip circle stability calculations are made (see Section 11.1 until 11.3). From these calculations it is
determined what increase in pore pressure can be allowed to keep the slope stable. The settlement and
pore pressure is then monitored in the field, see Figure 11-14 and Figure 11-15. As long as the pore
pressure remains lower than the calculated values, additional sand layers can be installed. When the pore
pressures reach the critical values, the construction must be stopped until the pore pressures are low
again.

82

Slope Stability
Pore pressure gauge
settlement plate

Figure 11-14. Instruments to control the construction of


an embankment.

Exces pore pressure

Figure 11-13. Failure during construction of an


embankment.

Construction stages (m above start)


maximum expected (green line)
maximum allowed (red line)
measurements

Figure 11-15. Sketch of graph to control excess pore water pressures during embankment construction on soft soil
layers.

83

Geotechnical Engineering

12

Trenchless Technologies

12.1

Introduction

Some structures cannot, or only with great difficulty, be placed into the soil from the soil surface. For
example a pipeline that crosses a busy road or a tunnel that is made underneath a mountain or underpasses
some buildings or a river. Numerous techniques have been developed to be able to construct such
structures. In general these are called trenchless technologies. Three examples of such technologies are
described in this chapter: horizontal directional drilling, pipe jacking and mechanised tunnelling. These
examples are important construction methods, but not necessarily the most important. Other examples are
tunnels made by drilling and blasting, submerged tunnels and NATM (New Austrian Tunnel Method)
tunnels. The methods chosen are just presented to show what aspects are important for this kind of
underground work.
The three methods described have a different history. Horizontal directional drilling is developed in the
thirties of the last century in the oil industry, but it only became used in the seventies for civil engineering
works for the undercrossing of rivers, canals and roads. It is now widely used for relatively big projects as
a pipeline crossing of more than several kilometres but also for small projects as a telephone cable that
has to cross a ditch.
Mechanised tunnelling is an older technique developed in the beginning of the 19 th century. Major
improvements have been made after that. Nowadays the mechanised tunnelling technique is used rather
often in cities because of its low risk profile, or in soft soil conditions.
Pipe jacking can be seen as a simplified form of mechanised tunnelling.

12.2
12.2.1

Horizontal Directional Drilling.


What is HDD?

Horizontal directional drilling (HDD) is used for pipelines with a diameter up to 1500 mm. The bending
of the pipe, which is necessary for HDD, becomes too complicated for larger diameters.
Directional drilling starts with a pilot hole, see Figure 12-1 top figure. The pilot hole is created by
pushing a hollow pipe with an asymmetrical leading edge into to the soil. During pushing bentonite slurry
is pumped into the pipe and released as a slurry jet through a nozzle in front of the leading edge. The
direction of the leading edge is controlled by rotation of the pipe. The soil is taken away with the
bentonite slurry. The position of the leading edge can be determined from the surface. There are several
methods available to determine the position. In some cases there is a battery operated signaller built just
behind the drill head, the position of the drill head is determined from this signal. An alternative detection
system is the electromagnetic down-hole navigational system. This is used in conjunction with a series of
four electrical cables positioned directly above the desired path and secured in place. The cables can be
used directly on top of the street or highway and do not interfere with the traffic flow. The cables transmit
an electromagnetic signal that is picked up with the navigational instruments in the drill head. The
operator sees the exact position continuously on a computer screen. Both methods are less useful for river
crossings. For these a probe close to the drill head measures the inclination and azimuth continuously and
from that the position is determined. A quite accurate determination of the position appears to be possible.
When the pilot drilling is finished the diameter of the annulus created is enlarged by one or several
reamers that are pulled through the annulus, see Figure 12-1, middle. During these stages again bentonite
is added to keep the annulus open.
Finally the pipe is pulled in and the HDD is finished, Figure 12-1 bottom.

84

Trenchless Technologies

Figure 12-1. Principle of horizontal directional drilling. (J.D. Hair and associates, website, 2015).
A more geotechnical aspect of this process is how to create a stable borehole before the pipeline is
inserted. Only for stiff clay this borehole will remain open without any special measures, in saturated
sand it will certainly collapse. To avoid this collapse bentonite slurry is used as a drilling fluid. This is
water with around 4% of bentonite clay. This creates a slurry with a density that is a bit higher than the
density of water, furthermore the bentonite will create an internal and external filter cake that prevents a
water flow from the borehole to the pore water, see Figure 12-3. This results in a small effective stress on
the grains that is sufficient to create a stable borehole provided that the following functions are fulfilled:
1. The level of the slurry is higher than the piezometeric head during the drilling. A decrease of
water level over only a short period is sufficient to create instability.
2. The density of the slurry is sufficient that at every depth the pressure exerted by the slurry is
higher than the necessary pressure to have a stable borehole.
3. The slurry must have enough particles or enough shear strength to limit the penetration depth of
the slurry into the soil; otherwise the soil will collapse before the cake is formed.
4. The slurry has to be able to transport the soil particle out of the bore hole to the entry point.
Therefore it must have enough carrying capacity.
5. The pressure drop in the bore hole must be limited to prevent a blow-out.

85

Geotechnical Engineering
To fulfil these items it is necessary to pay some attention to the mud or slurry properties.

12.2.2

Mud properties

The penetration depth of slurry into the soil, called the mud spurt, can be calculated using an equation of
Mller-Kirchenbauer (1977):

sd

d10
p
3.5 f

(12.1)

Where:
sd
: the distance over which the slurry penetrated into the soil (m)
f
: the yield stress of the slurry (Pa)
P
: the pressure difference (pressure bore hole pore water pressure) (Pa)
Also other formulas are used. Some take into account the influence of the void ratio as well. The formula
presented above is based on a simplification of the real process assuming that the slurry can be seen as a
pure Bingham liquid. In reality it is a two phase material consisting of grains and liquid. The consequence
of that will be discussed below.
Furthermore, the bentonite slurry itself has to be stable. When the permeability of the bentonite is too
high, it works not as a plastering layer, because there will still be a flow of water from the bore hole to the
soil. Therefore the permeability is tested with a filtration test. The API (American Petroleum Institute)
describes a standardized test of such a filtration test. The standard prescribes the piece of equipment to be
used with a diameter of 78 mm and a pressure loading of 7 bar. The amount of water lost in half an hour,
filtering from the bentonite through a filter layer should be less than 15 gr in a test with this filtration
apparatus, see Figure 12-2.

Figure 12-2. Filtration test according to API.

The real behaviour of the slurry is a combination of mud spurt and filtration. The slurry consists of
small particles (the bentonite particles) in water. When this slurry penetrates into the sand, it will first
penetrate as if it is one homogeneous liquid, water and particles go with the same velocity. However, after

86

Trenchless Technologies
some time the slurry has penetrated partially into the liquid and due to the yield stress the driving force
becomes less. Now the particles are hampered and the water can still flow.
As a result the water in the slurry flows faster than the particles and the particles start to consolidate in the
grain matrix of the sand, see also Figure 12-3.

finer particles
penetrate into
sand

slurry
Figure 12-3. Sketch external and internal cake formation bentonite slurry and sand.

This figure also shows particles smaller than the sand particles and larger than the slurry particles (the
black points). These particles are sometimes added to speed up cake formation.
This consolidation of the fine particles in the sand matrix leads to a further blocking of the slurry
particles, which is called an internal cake formation. The in between particles lead to an external cake,
since they cannot penetrate into the matrix. Both processes will lead to a decrease in velocity between the
water and the particles. This can be seen clearly in a filtration test where bentonite slurry is pressed into
sand, see Figure 12-4.

Figure 12-4. Setup and result of a filtration test where bentonite and slurry is pressed into sand (Talmon et al., 2013).

The graph presents the volume of displaced fluid as a function of the square root of time. The difference
between the mud spurt part of the plot and the consolidation part is clearly visible. The transition between
the two parts is given by the Peclet number:

Pe

ud
cv

(12.2)

Where:

87

Geotechnical Engineering
Pe
u
d
cv

: Peclet number (-)


: the pore velocity of the slurry (m/s)
: the average diameter of the pores (-)
: the consolidation coefficient of the slurry (m2/s)

For a Peclet number smaller than 1 the velocities of the pore water and the slurry particles are different
(so there is consolidation), for a Peclet number larger than 1 the slurry behaves as one liquid.
The fourth function (see Section 12.2.1) is described as carrying capacity for the soil grains since the
slurry acts as a hydraulic conveying system for the spoil transport from the drilling location to the entry
point. The indicator for the carrying capacity is the gel strength of the suspension. It may be expressed
according to Triantafyllidis (2004) by means of the yield point of the suspension. The equation was
originally developed to describe a sinking process of a ball into bentonite suspension at rest; however, it
may also be applied for a description of soil grains behaviour within bentonite suspension [Schoesser et
al., 2011]:

f 0.7

2
d particle particle suspension
3

(12.3)

Where:

dparticle

particle
fluid

: the shear strength between the slurry and de ball [kN/m2]


(more or less equal to the shear stress in the slurry)
: the diameter of the particle [m]
: the volumetric weight of the particle [kN/m3]
: the volumetric weight of the slurry [kN/m3]

Higher values of f are used in practice to calculate the carrying capacity of the drilling fluid. The factor
0.7*2/(3in equation (11.22) is replaced by 0.5 or 1/. The reason is that the slurry is not only loaded
with the weight of the particle, but when the slurry flows through an annulus there is also a shear force
exerted on the slurry. This shear stress decreases the particle carrying capacity of the slurry.
The pressure drop caused by the slurry flow through the bore hole can be calculated assuming that the
yield stress is dominant and thus neglecting the viscous forces. Equilibrium of forces leads then to the
following equation:

ptip f

l.S

(12.4)

Aborehole

Where:
Ptip
: is the pressure at the tip above the atmospheric pressures assuming that at the exit point the
pressure is equal to the atmospheric pressure [kPa]
L
: the length of the borehole [m]
S
: the circumference of the borehole outside (the soil) and inside (the pipe) [m]
Aborehole : the open area of the borehole (not occupied by the drilling rod) [m2]
If this pressure is higher than the pressure corresponding to the vertical weight of the soil, it is assumed
that there is a risk for a blowout. This is a rather conservative approach. In reality the pressure can be
higher depending on the circumstances (type of soil, quality filter cake). However, in quite some projects
there is only limited knowledge on the circumstances and a safe approach has to be used.
The equations mentioned above can now be used to design the drilling fluid. The yield strength has to
be limited to prevent a blowout. However, if the yield strength is too low, the sand will settle from the
drilling fluid. If the viscosity is too high, the pressure drop in the bore hole will be too high and it is not
possible anymore to use equation (12.4). Furthermore, filtration losses have to be limited and thus the

88

Trenchless Technologies
permeability of the filter cake has to be rather low. Bentonite quality may vary and clay from the subsoil
can influence the hydraulic properties. Therefore it is necessary to check viscosity and yield stress
regularly during a project. The instrumentation used for these tests is described below.

12.2.3

Measuring viscosity and yield strength of drilling fluid

Marsh funnel
Two devices are used to measure the viscosity and yield strength of a drilling fluid.
The simplest one is the Marsh Funnel. This is a funnel of prescribed dimensions.

= 1 quart

Figure 12-5. Marsh funnel and measuring cup.

The funnel is filled with 946 cc (1 quart) of drilling fluid while the opening below is closed. Then the
closing is removed and the time it takes to empty the funnel is measured. From this time an apparent
viscosity () is calculated with the formula:

f t f 25

(12.5)

Where

: the apparent viscosity in centipoise [1 centipoise = 10-3kg/ms]


f : the density of the fluid [kg/m3]

tf : the time it takes to empty the funnel [s]


It should be noticed that this is an empirical formula and that dimensions on both sides of the formula are
not the same, so it can only be used in the given dimensions.
The shear rate ( in the outflow opening of a Marsh funnel is around 2000 1/s.
The Marsh funnel test is very robust and does not need electricity, so it can be used in the field.
Fann viscometer
The Fann viscometer is a laboratory equipment, see Figure 12-6. A varying shear rate is applied by
changing the number of revolutions with which two concentric cylinders rotate with respect to each other
through the drilling fluid. The dimensions of the cylinders are fixed and thus the applied shear rate can be
calculated as a function of the angular velocity. The yield strength is measured at 600, 300, 200, 100, 6
and 3 revolutions per minute (RPM). These values of RPM correspond with the shear rates of the
measurement points inFigure 12-7. The measurement is repeated for 3 RPM after 10 seconds of rest and

89

Geotechnical Engineering
after 10 minutes of rest to measure what is called the gel strength, the shear stress when the drilling fluid
has come to rest. Due to the thixotropic behaviour of the liquid, this gel strength is higher than the yield
strength. If the gel strength is too high, it becomes difficult to start drilling again after a stop. Figure 12-7
is a typical result of such a test. The red dots represent the yield strength when slowing down the
revolutions per minute. When the point with 3 RPM is measured, the rotations stop for 10 s and the gel
strength at 10 s is measured at 3 RPM (the blue dot). The same is done after 10 minutes.
120

100

=0.0417*+28.5 (Pa)

(Pa)

80
60
tau (Pa)
10''
10'

40
20
100

0
0

Figure 12-6. Fann viscometer.

200 300

500

600 RPM

1,000

. /s

1,500

2,000

Figure 12-7. Result of a Fann viscometer test.

Table 12-1. Example of results of a Fann viscometer (also mentioned: Fann VG viscometer, VG because it
measured viscosity and gel strength)

Table 12-1 shows a result of the measurements with a Fann viscometer. The shear force created by the
viscous force and yield force causes rotation of the inner cylinder, see Figure 12-6, that is connected with
a calibrated spring. So the rotation is linearly related to the shear force and the measured rotation leads
with a calibration factor to the shear force.
The viscosity and yield strength are determined using the data points measured at 600 and 300 RPM.
Fitting a line as shown in results Figure 12-7 in the viscosity that linearly depends on the shear rate and
the yield stress where the line crosses the vertical axis. This result shows clearly that describing a

90

Trenchless Technologies
bentonite slurry as a Bingham liquid is only an approximation. At low shear rates the shear strength will
be less than according to the blue line that describes a Bingham liquid.

12.2.4

Stability of bore hole

Davis et al.(1980) has derived the minimum pressure to keep a borehole in clay stable. They defined:
c
(12.6)
N v
su
With:
N
: the stability number (-)
v
: the vertical total pressure (kPa)
c
: the pressure in the cavity (kPa)
su
: the undrained shear strength (kPa)
The cylindrical cavity is stable when:
C
N 2. ln( 2. 1)
D
With C the cover and D the diameter of the cavity.

(12.7)

The effective stress is of importance for granular materials. Usually Terzaghis theory is used to calculate
the minimum pressure necessary in the bore hole. This theory has a general applicability when in the soil
there are stiff parts that do not settle and parts that do settle. The stiff parts will attract the loading and the
settling parts will be stable with only a limited vertical pressure as will be explained below.
Based on previous work of Janssen (1895), Terzaghi (1943) presented the well-known equation for the
vertical pressure on a trapdoor as given in Figure 12-8:

B( c / B)
K
1 e
K tan

tan z / B

qe

K tan z / B

(12.8)

In which v (kPa) is the vertical stress on the trapdoor, B (m) and z (m) are explained in Figure 12-8,
(kN/m3) the density of the sand, (deg) the soil friction angle, c (kPa) the soil cohesion and K (-) the ratio
between horizontal and vertical pressure. The equation results in a maximum possible vertical stress on
the trapdoor for large values of z (large overburden):

v ,max

B( c / B)
K tan

(12.9)

91

Geotechnical Engineering
q
z

slip planes

sand

trapdoor
2B

Figure 12-8. Sketch of trapdoor experiment.


A modified plot of the original experiments is shown in Figure 12-9.

Figure 12-9. Original data from Terzaghi: Vertical force on trapdoor as a function of the displacement
(Terzaghi 1936, Revised by Evans, 1984, described by Tien, 1996).
The plot shows the load on the trapdoor divided by the average load that is composed out of the weight of
the soil and a possible surcharge load as a function of the movement of the trapdoor. It can be seen that a
minimum displacement of the trapdoor is sufficient to have a large reduction in the load.
A hole made by directional drilling can be seen as a trapdoor, see Figure 12-10

92
Figure 12-10. Bore hole sketched as trapdoor.

Trenchless Technologies
q
z

slip planes

sand

bore hole
2B

Without any cohesion it can be derived from Equation (12.9) that the maximum loading exerted by the
soil and a possible surcharge equals roughly the weight of the soil times the diameter of the pipe (for
example if K=0.7 and tan=35 degrees then the product of K*tan0.49). The necessary borehole
pressure is created by the density of the slurry, which is larger than the density of water, and the
difference between the slurry level and the phreatic surface. During drilling there will be an extra pressure
in the bore hole because the yield stress between the slurry and the soil has to be overcome to maintain a
slurry flow, see above. However this pressure will not or only partly be present during standstill.

12.3

Pipe jacking

Pipe jacking is quite comparable with mechanised tunnelling: also here a machine that resemblances a
TBM is used and sometimes slurry is used at the front face of the pipe to keep the pressures at the soil
face high enough to prevent a collapse of the front face (see further in the TBM section) as well.
Differences are that the diameter of pipe jacking pipes is in most cases limited to less than 2 m and that
the machine is pressed forward by pressing the pipe from the location where drilling begins, see Figure
12-11. This means that the length over which a pipe jacking can be performed is limited, or an
intermediate jacking station has to installed, see Figure 12-12. A TBM pushes itself further against the
just placed lining. This always has the same effectivity regardless the length of the tunnel.

Figure 12-11. Pipe jacking (Herrenknecht).

93

Geotechnical Engineering

Figure 12-12. Intermediate jacking station.

The pipe needs to be lubricated to reduce the friction between the soil and the pipe. Also bentonite
slurry is used for the lubrication. In long pipe jacking projects, the pipe is not only lubricated from the
pipe front, but sometimes holes are drilled in the pipe to be able to inject extra bentonite slurry on the pipe
between the pipe front and the starting point of the pipe jacking.
Some rotation between the various pipe elements is possible due to flexible (often wooden) load transfer
rings. This allows steering during pipe jacking. However this is also a risk, especially at larger
deformations and high installation pressures. Rotation between the various pipes will cause that the
largest forces are transmitted at the outside of the pipe, which can cause damage, see Figure 12-13. Since
the damage is outside of the pipe, it is not found during an inspection of the inside of the pipe.

Figure 12-13. Damage on pipes during pipe jacking due to a combination of high installation pressures and rotation
between the pipes.

12.4

TBM tunnelling, geotechnical aspects

In TBM (Tunnel Boring Machine) tunnelling, also mentioned mechanised tunnelling, collapse of the
tunnel face is prevented by pressurized slurry or muck that is pumped between the tunnel face and the
TBM. The TBM is a complex machine that performs the cutting of the soil by the cutting wheel, as well

94

Trenchless Technologies
as the transport of the soil out of the tunnel and the installation of the concrete lining elements that will
form the tunnel, see Figure 12-14. The TBM pushes against the just placed lining elements to push itself
further into the soil.
In a structural engineering point of view to TBM tunnelling the soil is in most cases simulated as springs
around the tunnel lining, or as a soil continuum around the lining, see Figure 12-16. In this way the soil
reaction is treaded in a lot of design calculations. Unfortunately the lining-soil interaction is more
complicated. The lining has no direct contact with the soil, but there is a grout layer of the tail void grout
in between.

plungers
cutting wheel

screw conveyor

mixing chamber

erector

press. bulkhead

lining

Figure 12-14. Sketch EPB TBM.

plungers

rings

concrete

Figure 12-15. Detail TBM, the construction of the lining.

95

Geotechnical Engineering

Figure 12-16. Simplified calculation model for a tunnel in soft soil.

It is not possible to deal with all aspects of TBM tunnelling,that would need several separate courses.
Here it is focussed on two aspects: the face stability and the lining-grout-soil interaction. However, first
some definitions are dealt with that are useful in the remaining part of this chapter.

12.4.1

Definitions

Volume loss
The volume loss is a very important parameter to determine the influence of tunnel construction on the
surroundings. Generally, more soil will be removed by the construction of a tunnel than there is tunnel
installed. Three items contribute to the volume loss. Firstly, the pressure at the tunnel face is normally a
bit lower than the original soil pressure. This will cause some stress relieve and the soil will move in the
direction of the tunnel face. Secondly, the TBM is slightly tapered; the diameter at the front is in general a
few (3-5) centimetres larger than at the tail. Thirdly the tail void grouting fills a smaller volume than the
original volume leading to soil settlement. The relative volume loss is defined as the volume loss,
measured as a settlement trough on the surface divided by the tunnel volume, see Figure 12-17. The
actual volume loss at the tunnel can be more or less than what is measured at the surface due to dilatancy
or contraction of the soil due to the deformation. However, this is normally not taken into account. It
should be noted that although areas are compared, it is always mentioned volume loss. For different
situations, the Gaussian curve as first suggested by Peck (1969) appeared to be a good approximation of
the settlement trough.

96

Trenchless Technologies

In modern tunnelling operation the volume loss is between 0.25 and 1%.
0
-10

settlement (mm)

-20
-30

Settlement
trough

-40

time (s)

-50

282
294
312
328
339

x
s s 0 exp( 0.5( ) 2 )
i

-60
-70
-80
-6

-4

-2

0
x (m)

tunnel

As 2 .i.s0

AT

D2

Volume loss

As
AT

Figure 12-17. Measured settlement troughs at different times in a model test, Gaussian curve fits and definition of
volume loss. (Bezuijen & Van Seters, 2002).

Drilling and ring building phase


TBM tunnelling is an alternating process. The TBM drills during the drilling phase over the length of one
segment pushing with the plungers (4 in Figure 12-18) against the concrete lining elements. Then it stops
to construct a new ring using the lining elements (5) that are brought into the newly constructed tunnel.
Some of the plungers are retracted to create the space to install a new lining element. This is called the
ring building phase and normally this takes more time than the drilling phase. If the ring of elements is
completed, drilling starts again. In the past it has been tried to combine these phases to have simultaneous
drilling and ring building but up to now that has not been successful.

TBM shield
Cutting wheel
Pressure chamber
Plungers

Lining elements
Gantry
Tunnel from lining elements

Figure 12-18. Sketch TBM and tunnel. (Herrenknecht)

97

Geotechnical Engineering

12.5

Stability of the tunnel face

12.5.1

Introduction stability of tunnel face

The sections on the stability of the tunnel face deal with the formulas that are used in the current practice
to calculate the minimum and maximum pressure that is allowed before instability of the tunnel face
occurs. The formulas are based on limit equilibrium calculations. What is the minimum pressure to have a
just stable tunnel face and what is the maximum pressure above which a blowout can be expected. As
always with limit equilibrium calculations the results present the pressure range in which no failure has to
be expected. However, it is certainly possible that within this range the deformations are larger than
allowed. The volume losses that are realised nowadays, between 0.25 and 1%, indicate that the applied
pressures are still well in the elastic region, thus in daily practice the minimum face pressure will be
higher than follows form the calculation methods below and the maximum face pressure will be lower.

surface settlement prototype (mm)

Centrifuge tests show that surface settlement and thus also the volume loss can be significantly reduced
when pressures higher than the minimum pressure required for stability of the tunnel face (the vertical
lines in the figure) are applied. In these tests a model tunnel was tested with a flexible membrane at the
tunnel face and a liquid container behind the membrane. The test started with a relatively large pressure.
After that the deformation was increased by pumping away the water from behind the membrane, see
Figure 12-20. Consequently the pressure reduced until a minimum was reached. Note that the vertical axis
has a logarithmic scale.
displ. transducer

C=8 m
D=10 m

100.0

clay

10.0

clay

sand
1.0

model tunnel

membrane

control

0.1

minimum pressure minimum pressure


in sand
in clay

to pump
displ. transducer

140

145

150

155

160

165

170

face pressure (kPa)

Figure 12-19. Centrifuge tests that show the relation between


face pressure and surface settlement for a tunnel with 10 m
diameter (D) and 8 m cover (C), the thickness of the soil layer
above the tunnel. (modified after Bezuijen & Van Seters, 2002).

Figure 12-20. Sketch of model tunnel in


centrifuge.

The deformations, as measured in centrifuge tests, cannot be calculated with the methods described
below, numerical methods have to be used.
In the last section of the face stability it will be explained that also the limit equilibrium calculations have
some complications in saturated soil due to ground water flow, which is normally not taken into account
in the text book calculations.

12.5.2

Stability tunnel face in clay

For tunnels bored in clay the stability can be calculated according to theory developed by Davis et al
(1980), mentioned already above in the section Directional drilling, stability of borehole. Using the same
definition:

98

Trenchless Technologies

v c

(12.10)

su

With:
N
: the stability number (-)
v
: the vertical total pressure (kPa)
c
: the pressure in the cavity (kPa)
su
: the undrained shear strength (kPa)
For small values of C/D (Cover/Diameter), N can be written as:

N 4 ln 2 1
D

(12.11)

For C/D values higher than 0.86:

N 2 2 ln 2 1
D

(12.12)

It is recommended by the authors to use in this formula C/D=3 for values of C/D higher than 3. This
means that N is limited to 5.89, close to the empirical value of 6 that was found by Broms and Bennemark
(1967) based on laboratory experiments.

12.5.3

Stability tunnel face in sand

The stability of the tunnel face is calculated using a 3-D version of the silo theory described already for
directional drilling (Jancsecz & Steiner, 1994). It is valid for tunnels in sand, or sand with some small
layers of clay in between. It is assumed that in front of the tunnel face there is some movement of the soil
and arching will occur in the block of soil just in front and above the tunnel, see Figure 12-21. Due to this
arching there will be a loading on the wedge ABCDEF below the rectangle CDEF but this loading is less
than the normal vertical soil pressure. The weight of this triangle and the loading from above has to be
compensated by the friction forces from the sides ADE and BCF, the friction on ABFE and the pressure
at the tunnel face.

99

Geotechnical Engineering
J

failure surface according to Jancsecz

G
I
H

F
C

D
force on wedge
by tunnel face
pressure
tunnel
face
B

Figure 12-21. Possible failure surface in front of a tunnel face.

In homogeneous soils the arching can be calculated according to the formula:

F
U
c
. z . K tan
A
v A
1

K tan

(12.13)

where U is the circumference and A the area of the horizontal plane CDEF. Assuming a low the
arching over the small rectangle CDEF will be very efficient and the loading on the wedge will be small
and consequently also the necessary pressure over the tunnel face will be small. For a high the friction
over AEFCB will become very efficient and again only a limited pressure over the tunnel face will be
necessary to keep the wedge stable.
The critical angle is the angle that needs the highest face pressure to get a stable solution. It is usually
close to = 450-/2.
As in the section on directional drilling, also here a maximum pressure can be defined that will lead to
stable arching for all depths:

v ,max

F
c
A
K tan

(12.14)

And since the weight of the wedge remains the same at all depths (assuming constant soil conditions), this
means that there is also a tunnel pressure that will lead to a stable tunnel at all depths. This pressure only
depends on the friction angle, the cohesion, the diameter and K. However, it should be realized that this
pressure is the effective pressure. For a stable tunnel face it is also necessary to counterbalance the water
pressure and therefore the pressures can be considerable when tunnelling under a river or estuary.
The maximum possible drilling pressure is, as in horizontal directional drilling, taken equal to the
pressure that corresponds with the weight of the soil layers and water above the tunnel. This is a safe
assumption, because the real maximum drilling pressure will be higher in most cases.

100

Trenchless Technologies

12.5.4

Influence of pore pressures

What is described up to now is traditional theory. It assumes implicitly that there is a perfect plastering
of the tunnel face, because it is assumed that the pressure minus the pore water pressure is directly exerted
on the grains in front of the tunnel. This plastering is caused either by the filter cake that is formed by the
bentonite in the bentonite slurry for a slurry shield TBM, or by the air bubbles that create a nearly
impermeable layer at the boundary between the muck and the soil for an EPB shield TBM.
However, measurements have shown that there is no perfect plastering when drilling in saturated fine to
medium course sand. In this situation a filter cake in case of a slurry shield or a low permeable layer of air
bubbles in case of an EPB shield can only form if there is a certain flow from the mixing chamber of the
TBM to the soil just in front of the tunnel face. In saturated conditions this flow is only possible in
combination with groundwater flow in front of the TBM. In fine and medium course sand the
groundwater flow is generally slower than the progress of the tunnel during drilling. Consequently the
filter cake or the air bubbles are taken away before they can seal the front face completely, leading to
excess pore water pressures in the sand. Broere (2001) and Bezuijen et al. (2001) have shown that this
causes a lower stability of the tunnel face and a higher pressure has to be applied to have a stable tunnel
face.
The measurements at the 2nd Heinenoord tunnel in the Netherlands clearly show this phenomenon. Pore
pressures were measured in the future alignment of the tunnel, thus the pore pressure gauge was eaten
by the TBM.
200
measurement PPT 5
calculation

190

pressure (kPa)

180
170

tunnel

distance PPT

160
150
140

during drilling

130

0 ( 1 ( x / R) 2 x / R)

120
during stand still

110
100
0

10
15
20
25
distance from tunnel face (m)

30

35

Figure 12-22. Measured excess pore pressure in front of a slurry shield and approximation. (Bezuijen et
al. (2001).
It can be clearly seen that during drilling the induced groundwater flow creates an excess pore pressure in
the soil in front of the TBM. During stand still for ring building, a filter cake can form and the excess pore
pressures in the soil decrease because now there is a perfect plastering.
The influence of the excess pore water pressure on the stability of the tunnel face is up to now not
frequently used in the design of tunnels. Yet accidents with face instability are rather rare. This means
that the safety factors used in the traditional calculation methods also incorporate implicitly the influence
of this phenomenon. In critical situations it is necessary to take the influence of excess pore water
pressures in front of the TBM into account. This was the case when the GHT (Green Heart Tunnel) was
made in the Netherlands. When the TBM arrived in a polder there was a risk that the excess pore water
pressure may have lifted the low permeable top layers of this polder. The soil was ballasted over some
distance to prevent uplifting. Also for a critical bridge passage in the North South line in Amsterdam it
was necessary to take the influence of excess pore water pressure into account.

101

Geotechnical Engineering

12.6

Tail void grouting

As can be seen in Figure 12-18, the concrete lining is constructed from inside the TBM. Furthermore, the
TBM is slightly tapered to allow some manoeuvring of the TBM, as sketched in Figure 12-23.

AB

C D

excavated boundary
front shield

back shield
lining ext

Figure 12-23. Basic structure of a TBM with highlight to the progressive diameter reduction from the cutterhead
(A), to the front (B) and back (C) of the shield and the lining extrados (D). (Dias and Bezuijen, 2015).

For clarity the tapering is exaggerated in the lowest part of the figure. In reality the diameter of a 10 m
diameter tunnel is approximately 5 cm less at the back of the shield in comparison with the front
diameter. However, the combination of the influence of the tapering, the smaller diameter of the lining
and the possibility to increase the diameter of the cutting wheel by using overcutters in case some more
space is needed for the TBM (for example when a curvature has to be created in the tunnel) leads to a
significant diameter reduction. Such a diameter reduction without any counter measures would lead to an
unacceptable volume loss. Therefore grout at the back of the TBM is injected to prevent this volume loss.
This process is called tail void grouting. The principle is sketched in Figure 12-24.

Figure 12-24. Sketch tail void grouting at the back of a TBM.

The opening between the tail of the TBM and the lining segments is closed by brushes and grease. Grout
is injected through a tube in the TBM tail. The grout fills the gap between the lining segments and the soil
and, if the grout pressure that is applied is high enough, it will also fill the gap between the tapered TBM
tail and the soil.

102

Trenchless Technologies
Three types of grout are used:
- Cement based grout has a hardening time of several days.
- 2 component grout that has a hardening time of only minutes.
- Non-cement grout, which is just a granular mixture with quite a wide grading (fines up to fine
gravel) and water. This grout has no chemical hardening. It is just a slurry when injected and will
lose water because the injection pressure of each type of grout will be larger than the pore water
pressure. When sufficient water is lost, the grains will contact and in this way fix the lining.
The 2 component grout will fix the tunnel lining immediately in position after the lining leaves the TBM.
Quite often this is not the first choice because of costs.
For the other two grouts (cement based and non-cement grout), it will take some time before a sufficient
amount of water is removed and the grout is behaving as a solid. Also for cement based grouts, the
removal of water is the dominating process and therefore the hardening of the cement takes more time. In
the meantime the lining is more or less floating in the liquid grout and fixed in position by the TBM at the
front and the lining that is fixed into solid grout. This can cause some lifting of the tunnel just when it
leaves the TBM, see Figure 12-25 for an example.

vertical movement (mm)

22.5
Y4
Y3
Y2
Y1

20.0
17.5
15.0
12.5

10.0
7.5

2
1

5.0
2.5
0.0
-2.5
-5

10

15

20

25

distance from TBM (m)

Figure 12-25. Example of measured heave of the lining elements measured at 4 locations Y1 to Y4 after leaving the
TBM. Likely caused by buoyancy forces in the liquid grout just after the TBM. (Bezuijen and Talmon, 2005,
modified).

How long it will take before sufficient water is removed from the tail void grout in order to behave like a
solid depends on the permeability and stiffness of the soil around the tunnel. In clay the water flow from
the grout into the soil is limited and it can take quite some time before the water content in the grout is
sufficiently reduced. Although some tunnels are constructed successfully in clay using non-cementing
grout, there is also a recent example (London) where the drive with cement based grout in clay resulted in
slow and difficult tunnelling and the replacement of this grout by 2 component grout resolved the
problem.
For tunnelling in sand, the permeability of the grout determines the water flow from the grout. Here the
stiffness of the sand is of importance. When grout is injected at a constant pressure, a layer of solid grout
will form, as explained in Section 6.2. However, when drilling stops and the ring building starts, also the
grouting stops. If the water flow from the grout continues, then the soil will unload and the pressure will
decrease, see Figure 12-26. This means that also the driving force for further dewatering of the grout
decreases. For stiff sand the driving force will stop completely before all grout is dewatered. The next
drilling face (with again tail void grouting) will increase the pressure in the grout again and the pressure
will decrease when this phase stops. This goes on until all grout is dewatered.

103

Geotechnical Engineering

soil
(sand)

x
dewatered grout
rt

grout

q=0

lining
grout

lining
flow by expelled
pore water from
the grout

Figure 12-26. Sketch dewatering of grout around a tunnel.

The dewatering during ring building is described with a comparable differential equation as used to
describe the dewatering in Section 6.2, but with one extra term to describe the elastic unloading of the
soil:

x dx 2 G
1 ni

x
0
k dt g r
ni ne

(12.15)

Where:
k
: the permeability of the grout cake of consolidated grout (m/s)
x
: coordinate, see Figure 12-26. (m)
t
: time (s)

: density of water (kg/m3)


g
: acceleration of gravity (m/s2)
G
: the shear modulus (kPa)
r
: radius of the tunnel (m)
ni
: initial porosity of the grout (-)
ne
: porosity of the grout after consolidation
0
: applied difference in piezometric head between grout and pore water during drilling (m)
Due to the extra term describing the elastic unloading (the term with G), it is a bit more complicated to
solve this non-linear differential equation. However, it is quite easy to calculate the final thickness (x) of
the grout cake. In that case there will be no further consolidation, thus dx/dt=0, leading to:

x final 0.5.g

1 ni 0 r
ni ne G

(12.16)

When this value is smaller than the thickness of the original grout layer, only a partly dewatering can be
expected.

104

drilling velocity (mm/min)

Trenchless Technologies

60
40
20
0
03:00
4.0

06:00

09:00

12:00

15:00

Bb
Ka

Ba
A6

A4b
A5

A1

Kb

A4a

Ca

A3b

Cb
A4

A2

A3a
A3

A1b

A2b
A2a

A1a

grout pressure (bar)

3.5
3.0
A2a
A1b A1a

2.5

Cb
Ca
Kb
Ka
Bb

2.0
1.5
1.0

Bb
Ka
Kb
Ca

0.5
0.0
03:00

06:00

Cb
A1a
A1b
A2a

09:00

12:00

15:00

time

Figure 12-27. Measured grout pressures at a fixed position on the lining during drilling and ring building. The
position of the instruments is shown on the circle at the left (black dots) as well as the position of the injection lines
(small circles). (Bezuijen et al. 2003)

Figure 12-27 shows an example of measurements from the outside of the lining of a tunnel constructed in
sand. Grout pressures were measured with sensors installed on the lining. This means that during drilling
the sensors come into the grout at a certain moment (around 6:00 in the plot), resulting in a pressure jump.
As the TBM advances, the distance between the TBM and the lining elements with the sensors increases.
As drilling stops the grout pressure decreases. In this graph the pressure is still decreasing when the next
drilling session starts. The grout pressure increases again because again fresh grout is pumped into the tail
void. After the drilling of some rings (the drilling of each ring results in a drilling velocity of around 40
mm/min as can be seen in the top plot) most of the grout is dewatered at the position of the sensors and
the variation in pore pressure is less explicitly related to drilling or ring building. Also the quality of the
measurements will decrease by then, because the pressure sensors will be less accurate when the grout
becomes stiff.
It is clear that the measured pressures cannot be described with springs as shown in Figure 12-16. It is not
the soil but the grout liquid that determines the pressure distribution around the tunnel. After
consolidation of the grout the elasticity of the soil may become important, but by that time the volume
loss due to grout bleeding has reduced the pressures on the tunnel.
In theory, cemented grout is an extra lining around the tunnel. However, it is thinner and in all cases the
grout after hardening is not as strong as the reinforced concrete of the segments. Therefore the strength of
the grout is never taken into account when designing the thickness and reinforcement of the tunnel lining.

12.7

Buoyancy forces

The average density of the tunnel lining and the air in the tunnel is less than the density of water. This
means that the tunnel has the tendency to float in the pore water. Soil pressures and friction should be
sufficient to keep the tunnel in place. For the largest tunnels this is also done by ballasting the tunnel as
soon as possible.,just after construction of the lining. Normally a TBM is followed by two gantries with

105

Geotechnical Engineering
equipment and an overhead crane for the transport of the lining segments, see Figure 12-28. In the Groene
Hart Tunnel in the Netherlands (which was at the time of construction the largest TBM tunnel in the
world), ballasting the tunnel started already before the last gantry had passed, as can be seen in Figure
12-29.

Figure 12-28. Sketch of tunnel boring machine, back-up train (gantries), and tunnel lining at the Groene Hart
Tunnel. A lining segment is being erected inside the TBM. (Talmon & Bezuijen, 2011).

Figure 12-29. Ballasting next to the technical gallery of the Groene Hart Tunnel already before the last gantry has
passed.

After this ballasting, the finishing of the tunnel is no task any more for geotechnical engineers, but for
other disciplines.

106

Trenchless Technologies
References:
Begemann, H. K. S., 1965. The friction jacket cone as an aid in determining the soil profile, Proceedings
of the 6th International Conference on Soil Mechanics and Foundation Engineering, ICSMFE,
Montreal, September 8 - 15, Vol. 2, pp. 17 - 20.
Bezuijen A. Pruiksma J.P., Van Meerten, H.H. (2001). Pore pressures in front of tunnel, measurements,
calculations and consequences for stability of tunnel face. Tunnelling a decade of progress - GeoDelft
1995-2005. Taylor & Francis Group plc. London.
Bezuijen A., Talmon A.M. (2005) Grout properties and their influence on back fill grouting, Proc.
Geotechnical Aspects of Underground Construction in Soft Ground - 5th International Symposium
(IS-Amsterdam). CRC Press/Balkema, Leiden, Amsterdam, the Netherlands.
Broere, W. (2001). Tunnel Face Stability & New CPT Applications, PhD Thesis. DUP Science Delft
University of Technology.
Broms B., Bennermark H. (1967): Stability of clay in vertical openings. Journal of the Geotechnical
Engineering Division, ASCE 1967; 193: pp. 7194.
Budhu M.(2007), Soil Mechanics and Foundations (2nd Edition). John Wiley and Sons.
Cone (2013) http://en.wikipedia.org/wiki/Cone_penetration_test#History_and_development.
Davis E.H., Gunn M.J, Mair, R.J., Seneviratne H.N. (1980): The stability of shallow tunnels and
underground openings in cohesive material. Gotechnique 30 (4), pp.397416
De Beer, E. (1966), Berekening van de negatieve wrijving op palen, Tijdschrift der Openbare Werken van
Belgi, Nr. 6, December
De Beer, E. (1971-1972), Mthodes de dduction de la capacit portante dun pieu partir des rsultats
des essais de pntration, Annales des Travaux Publics de Belgique, Nr. 4,5,6, Brussels
De Beer, E. (1988), Different behaviour of bored and driven piles, Proc. Deep Foundation on Bored and
Auger Piles, Van Impe, ed. Balkema, Rotterdam, 47-82.
Dias T.G.S. and Bezuijen A. (2015) TBM Pressure Models Observations, Theory and Practice. Proc.
15th Pan-American Conference on Soil Mechanics and Geotechnical Engineering
Evans C.H. 1983. An examination of arching in granular soils, M.S. thesis MIT, Boston, USA.
Fellenius, B. H., and Eslami, A., (2000_. Soil profile interpreted from CPTu data. Year 2000
Geotechnics Geotechnical Engineering Conference, Asian Institute of Technology, Bangkok,
Thailand, November 27 - 30, 2000, 18 p.
Hai-Sui, Yu, (2000), Cavity Expansion Methods in Geomechanics, Springer.
Janbu, N. (1976), Static bearing capacity of friction piles, Proc. Of the 6th European Conference on Soil
Mechanics and Foundation Engineering, Vol. 1,2, 479-488.
Jancsecz S., Steiner W. (1994): Face support for a large mix-shield in heterogeneous ground conditions,
in Proc Tunnelling 94, pp. 531-550, Chapman and Hall, London
Janssen, H.A., 1895. Versuche ber Getreidedruck in Silozellen Zeitschrift Verein Deutscher Ingenieure,
BD XXXIX, pp. 1045-1049.
Mayne, P.W., Coop, M.R., Springman, S., Huang, A-B., and Zornberg, J. (2009). State-of-the-Art Paper
(SOA-1): GeoMaterial Behavior and Testing. Proc. 17th Intl. Conf. Soil Mechanics & Geotechnical
Engineering, Vol. 4 (ICSMGE, Alexandria, Egypt), Millpress/IOS Press Rotterdam: 2777-2872.

107

Geotechnical Engineering
Mueller-Kirchenbauer, H. (1977): Stability of slurry trenches in inhomogeneous subsoil. Proceedings of
9th International Conference on Soil Mechanics and Foundation Engineering, Vol. 2, Tokyo.
Robertson, P. K., 1990. Soil classification using the cone penetration test. Canadian Geotechnical
Journal, Vol. 27, No. 1, pp. 151 - 158.
Schmertmann, J.H. (1976). An updated correlation between relative density, II, and Fugro-type electric
cone bearing qc. Waterways Experiment Station, Vicksburg, Miss. Contract report, DACW 38-76-M
6646. 145p
Schoesser B. et al. (2011): Practice-Oriented Guideline for the Choice of an Adequate Bentonite
Suspension for Lubrication in Pipe Jacking, Proceeding of 29th international NO DIG conference,
Berlin, Paper 3A-3, pp. 1-12
Stoevelaar R., Bezuijen A., Lottum H.M., Van Tol A.M. (2011) Effects of crushing on pile point bearing
capacity in sand tested in a geotechnical centrifuge. Proc. of the 15th European Conference on Soil
Mechanics and Geotechnical Engineering A. Anagnostopoulos et al. (Eds.). pp 393-944
Talmon, A.M., Mastbergen D.R., Huisman, M, (2013) Invasion of Pressurized Clay Suspensions into
Granular Soil, Journal of Porous Media, 16 (4): 351365 (2013)
Terzaghi, K. and Peck, R.B., (1948), Soil Mechanics in Engineering Practice, Wiley, New York.
Terzaghi, K., 1943. Theoretical Soil Mechanics, John Wiley and Sons, New York, pp. 66-76.
Terzaghi, K, 1936. Stress distribution in dry and wet sand above a yielding trapdoor. Proc. First int. conf,
on soil mechanics and foundation engineering, Cambridge Massachusetts, pp 307-311.
Tien, Hsien-Jen, 1996. A literature study of the arching effect, MSc thesis, MIT, Boston, USA.
Triatafylidis T. (2004): Planung und Bauausfhrung im Spetialtiefbau, Teil 1: Schlitzwand- und
Dichtwandtechnik, Ernst & Sohn Verlag, Berlin (in German)
Tritech, (2016): http://www.aeyates.co.uk/tritech-piling-and-foundations-ltd/about/history
Tomlinson, M. J., (2001), Foundation design and construction, 7th edn. Harlow: Prentice Hall.
Verruijt A. and Broere W. (2011) Grondmechanica, VSSD, Delft
Vesic, A.S., (1975), Bearing capacity of shallow foundations, Foundation Engineering Handbook, H.F.
Winkerton and H.Y. Fang, eds. Van Nostrand-Reinhold, New York, p. 121
Van Impe, W.F., De Beer E., Lousberg, E., (1988), Prediction of the single pile bearing capacity in
granular soils out of CPT results, Proceeding ISOPT I, First International Symposium on penetration
testing, Specialty Session, pp. 1-34.
Van Impe, W.F., (1993), Evaluation of the influence of pile execution parameters on the soil condition
around the pile shaft of a PCS-pile, Proceeding BAP II, Deep Foundations on Bored and Auger Piles,
pp. 217-220.
Van Rhee C. and Bezuijen A. (1992) Influence of seepage on stability of sand slope, Journal of
Geotechnical Engineering, Vol. 118, paper 26655. pp 1236-1240
Van Mierlo, W.C. and Koppejan A.W. (1952). Lengte en draagvermogen van heipalen; vaststelling
hiervan en enige daarbij verkregen ervaringen.Bouw nr.3 1952 (19 januari).
WTCB, (2009), Richtlijnen voor de toepassing van de Eurocode 7 in Belgi, Uitgave van het
wetenschappelijk en technisch centrum voor het bouwbedrijf, Rapport nr. 12, 2009
Xu, X., Lehane, B., (2005), Evaluation of end-bearing capacity of closed-ended piles in sand from CPT
data, Frontiers in Offshore Geotechnics: ISFOG 2005 Gourvenec & Cassidy (eds).

108

S-ar putea să vă placă și