Вы находитесь на странице: 1из 272

Advanced Soil Mechanics

Stefan Van Baars


© Edited and published by Stefan Van Baars. Edition May 2016

Key words: soil mechanics, foundation engineering, tunnelling

2
PREFACE
This book Advanced Soil Mechanics is part of the education of Civil Engineering at the
faculty of Science, Technology and Communication of the University of Luxembourg. It can
be seen as a continuation of the introductory courses of Soil Mechanics, as for example
written down in the book Soil Mechanics of A. Verruijt.
This book contains the major principles and design methods used in Geotechnical
Engineering, such as soil improvement, geotextiles, tunnelling, shallow and pile foundation,
sheet piles, anchors, struts and dewatering.
Probably a title for this book like “Geotechnical Engineering” would have been better, but
the current title “Advanced Soil Mechanics” expresses more the continuation of the book of
A. Verruijt and expresses more the mechanical focus of this book.

The chapters 14-22, 33-39, 45-46, 51-54, are copied from the book of A. Verruijt and
updated as a bridge to this book. Some other parts of this book have been collected from
other sources; books, internet, other lecture notes, including my own; and translated from
Dutch, German or French into English. Since this book contains pictures and parts of text
from many different authors, which are mixed, sorted and changed all over, it is impossible to
compile a complete track record of all references. In some cases the sources will remain
unknown unfortunately.

My special thanks go out to Prof. Dr. Ir. A. Verruijt, who taught me the fundamentals of Soil
Mechanics and to Prof. Drs. Ir. J. Vrijling who taught me the importance of simple hand
design engineering and the danger of experience, which is not fully understood. This book
tries to teach their visions in Advanced Soil Mechanics.

Luxembourg, May 2012 Stefan Van Baars

Cover Photo from Neidhardt Grundbau Gmbh, Germany

3
CONTENT

I SOIL IMPROVEMENT ....................................................... 11


1 Introduction ................................................................................................................. 12
2 Mechanical compaction ............................................................................................... 13
2.1 Surface compaction .............................................................................................. 13
2.2 In-depth compaction ............................................................................................. 15
3 Injection techniques..................................................................................................... 17
3.1 Permeation grouting.............................................................................................. 18
3.2 Jet-grouting .......................................................................................................... 19
3.3 Fracture grouting .................................................................................................. 20
3.4 Compaction grouting ............................................................................................ 21
4 Dewatering & drains ................................................................................................... 22
4.1 Introduction .......................................................................................................... 22
4.2 Vertical drainage .................................................................................................. 23
5 Ground freezing........................................................................................................... 25
5.1 Introduction .......................................................................................................... 25
5.2 Brine freezing (indirect cooling) ........................................................................... 26
5.3 Nitrogen freezing (direct cooling) ......................................................................... 26
5.4 Thermic Design .................................................................................................... 27
6 Geotextiles and reinforced soil .................................................................................... 29
6.1 Geotextiles............................................................................................................ 29
6.2 Reinforced soil / Terre armée ................................................................................ 31
7 Ground exchange ......................................................................................................... 34
7.1 Natural soils.......................................................................................................... 34
7.2 Artificial heavy or light products .......................................................................... 36

II TUNNELLING ...................................................................... 37
8 Tunnel types................................................................................................................. 38
9 Cut & cover tunnel ...................................................................................................... 40
9.1 Open building pit .................................................................................................. 40
9.2 Classical cut & cover ............................................................................................ 41
9.3 Aqueduct .............................................................................................................. 42
9.4 Wall-Roof method ................................................................................................ 43
10 Tunnel Boring Machine (TBM) .................................................................................. 46
10.1 Open face TBM .................................................................................................... 47
10.2 Hydro shield or slurry shield (SS) TBM ................................................................ 47
10.3 Earth Pressure Balanced shield (EPB) TBM ......................................................... 49
10.4 Hard rock TBM .................................................................................................... 49
10.5 Support pressure ................................................................................................... 50
10.6 Pneumatic caisson................................................................................................. 51
10.7 Settlements ........................................................................................................... 51
11 Drill & blast (rock) tunnelling .................................................................................... 54
11.1 German Method or Core Method .......................................................................... 55

4
11.2 Old Austrian Tunnelling Method or Old ÖTM ...................................................... 55
11.3 New Austrian Tunnelling Method or NÖTM ........................................................ 55
12 Immersed tunnel ......................................................................................................... 58
12.1 Construction method ............................................................................................ 58
12.2 Chin-nose and jack-leg support............................................................................. 60
12.3 Advantages and disadvantages.............................................................................. 61
13 Jacked box tunnel........................................................................................................ 63

III SHALLOW FOUNDATIONS .............................................. 65


14 Types of shallow foundations ...................................................................................... 66
15 Elastic stresses and deformations ............................................................................... 68
16 Boussinesq ................................................................................................................... 70
17 Flamant........................................................................................................................ 74
18 Deformation of layered soil ......................................................................................... 78
19 Bearing capacity of a strip footing.............................................................................. 81
19.1 Lower bound ........................................................................................................ 82
19.2 Upper bound......................................................................................................... 84
20 Prandtl & Nc ................................................................................................................ 86
21 Reissner & Nq .............................................................................................................. 89
22 Meyerhof & N ............................................................................................................ 90
23 Table of the bearing capacity factors ......................................................................... 91
24 Correction factors ....................................................................................................... 92
24.1 Inclination factors i............................................................................................... 92
24.2 Shape factors s ..................................................................................................... 93
25 CPT, undrained shear strength and Prandtl ............................................................. 95

IV PILE FOUNDATIONS........................................................ 97
26 Cone Penetration Test (CPT)...................................................................................... 98
27 Compression piles ..................................................................................................... 102
27.1 Pile types ............................................................................................................ 102
27.2 Bearing capacity ................................................................................................. 109
27.3 CPT, SPT and HDP ............................................................................................ 110
27.4 Tip resistance using Meyerhof (Prandtl, Terzaghi or Brinch Hansen) .................. 111
27.5 Tip resistance using a CPT method (Koppejan) .................................................. 112
27.6 Shaft resistance................................................................................................... 115
27.7 Settlement of piles .............................................................................................. 119
28 Tension piles .............................................................................................................. 120
28.1 Differences between tension and compression piles ............................................ 120
28.2 Reduction of the cone value qc due to excavation................................................ 120
28.3 Reduction of cone resistance due to tensile pile force ......................................... 121
28.4 Clump criterion .................................................................................................. 122
28.5 Edge Piles .......................................................................................................... 123

5
28.6 MV-pile or Jacket-grouted pile ........................................................................... 123

V UNDERGROUND MEGASTRUCTURES........................ 125


29 Underground megastructures ................................................................................... 126
30 Sustainable resources ................................................................................................ 130

VI BUILDING PIT ................................................................. 131


31 Building pit ................................................................................................................ 132
31.1 Introduction ........................................................................................................ 132
31.2 Pit collapse ......................................................................................................... 133
31.3 Pit design ............................................................................................................ 135

VII WALLS AND LATERAL STRESS ................................... 137


32 Walls .......................................................................................................................... 138
32.1 Wall types........................................................................................................... 138
32.2 Soldier pile wall (Berliner wall) .......................................................................... 139
32.3 Sheet pile wall .................................................................................................... 140
32.4 Combi wall ......................................................................................................... 141
32.5 Bored pile wall ................................................................................................... 142
32.6 Diaphragm wall .................................................................................................. 143
33 Lateral stresses in soils .............................................................................................. 147
33.1 Coefficient of lateral earth pressure ..................................................................... 147
33.2 Elastic material ................................................................................................... 149
33.3 Elastic material under water ................................................................................ 150
34 Rankine ...................................................................................................................... 151
34.1 Mohr-Coulomb ................................................................................................... 151
34.2 Active earth pressure .......................................................................................... 153
34.3 Passive earth pressure ......................................................................................... 154
34.4 Neutral earth pressure ......................................................................................... 156
34.5 Menard-penetrometer and CAMKO-pressuremeter ............................................. 156
34.6 Groundwater ....................................................................................................... 157
35 Coulomb ..................................................................................................................... 159
35.1 Active earth pressure .......................................................................................... 159
35.2 Passive earth pressure ......................................................................................... 161
36 Tables for lateral earth pressure ............................................................................... 164
36.1 The problem ....................................................................................................... 164
36.2 Example ............................................................................................................. 165
36.3 Tables ................................................................................................................. 166
37 Sheet pile walls........................................................................................................... 170
37.1 Homogeneous dry soil ........................................................................................ 170
37.2 Pore pressures ..................................................................................................... 174
38 Blum ........................................................................................................................... 177
39 Sheet pile wall in layered soil .................................................................................... 182

6
39.1 Computer program ............................................................................................. 182
39.2 Computation of anchor plate ............................................................................... 183
39.3 Horizontal bedding constants for subgrade reaction models ................................ 184
39.4 Finite Element Modelling ................................................................................... 186
39.5 Example quick design ......................................................................................... 186
40 Sheet pile profiles ...................................................................................................... 188

VIII ANCHORS, STRUTS AND WALES ............................... 191


41 Supports .................................................................................................................... 192
42 Anchors...................................................................................................................... 193
42.1 Anchoring .......................................................................................................... 193
42.2 Anchor installation ............................................................................................. 195
42.3 Soil nailing ......................................................................................................... 199
42.4 Holding capacity of anchor plate ........................................................................ 201
42.5 Holding capacity of a grouted anchor ................................................................. 203
42.6 Overall stability .................................................................................................. 206
43 Struts ......................................................................................................................... 208
44 Wales ......................................................................................................................... 210

IX FLOORS............................................................................. 213
45 Floatation & Archimedes .......................................................................................... 214
46 Natural floor & heave ............................................................................................... 215
47 Unsupported under water concrete floor ................................................................. 216
48 Supported under water concrete floor ..................................................................... 218
48.1 General............................................................................................................... 218
48.2 Floatation of the floor ......................................................................................... 219
48.3 Transfer of forces to piles & fracture of the pile joint .......................................... 220
48.4 Fracture of the floor ............................................................................................ 221

X GLOBAL STABILITY & FAILURE................................ 225


49 Failure modes ............................................................................................................ 226
50 Global translation (sliding) ....................................................................................... 227
51 Global rotation (circular sliding) .............................................................................. 229
51.1 Safety factor ....................................................................................................... 230
51.2 Fellenius ............................................................................................................. 231
51.3 Bishop ................................................................................................................ 233
51.4 Global failure ..................................................................................................... 234

XI DEWATERING ................................................................... 237


52 Groundwater flow ..................................................................................................... 238
52.1 Hydrostatics ....................................................................................................... 238

7
52.2 Groundwater head............................................................................................... 239
52.3 Darcy.................................................................................................................. 239
52.4 Permeability ....................................................................................................... 240
52.5 Flow in a vertical plane ....................................................................................... 241
53 Flow net...................................................................................................................... 243
53.1 Potential and stream function .............................................................................. 243
53.2 Flow under a structure ........................................................................................ 245
54 Flow towards well ...................................................................................................... 248
54.1 Confined aquifer ................................................................................................. 248
54.2 Unconfined aquifer ............................................................................................. 250
54.3 Semi-confined aquifer ......................................................................................... 252
54.4 Superposition ...................................................................................................... 253
54.5 Examples ............................................................................................................ 257
55 Dewatering ................................................................................................................. 259
55.1 Open dewatering and wellpoint drainage............................................................. 259
55.2 Dewatering problems .......................................................................................... 261
55.3 Hindrance for the surroundings ........................................................................... 261

XII SAFE DESIGN .................................................................... 263


56 Limit states and design rules ..................................................................................... 264
56.1 Limit states ......................................................................................................... 264
56.2 Design rules ........................................................................................................ 265
57 Material and load factors .......................................................................................... 266
57.1 Design theory ..................................................................................................... 266
57.2 Load factors ........................................................................................................ 268
57.3 Material factors................................................................................................... 269
57.4 Inadequate standards ........................................................................................... 270
58 Control sensors .......................................................................................................... 271

8
9
10
I Soil improvement

11
1 Introduction
Soil mechanical engineering is just like other fields of engineering the control of the strength,
stiffness or conductivity, which means the control of the stresses, strains and permeability.
Sometimes geotechnical tools like piles, anchors, walls and geotextiles can be optimised or
even avoided when the soil would has had a higher strength, stiffness, permeability or a
combination of these.
There are many different ways of improving the strength, stiffness or permeability of the
ground. The following chapters will discuss the following options:
1. Mechanical densification
2. Injection techniques
3. Dewatering & drains
4. Ground freezing
5. Geotextiles
6. Ground exchange

These techniques will be used when the extra cost of using them is less than the extra cost,
which would have to be accepted for extra use of geotechnical tools due to the bad soil
conditions.

12
2 Mechanical compaction
The goal of mechanical compaction is the reduction of the amount of pores. The less pores,
the higher the stiffness of the ground and also the higher the strength (angle of internal
friction).
For cohesive soils like clay and peat, this reduction can only be achieved by consolidation
and creep of the material, which means dewatering is required. For these soils with a low
permeability, dewatering is a very slow process. Since mechanical compaction is a quick load
technique, it will not work well for these soil types.
For non-cohesive soils like sand and gravel, or mixed soils, there are good possibilities for
this technique.
There are two types of mechanical densification techniques:
1. Surface compaction
2. In-Depth compaction

2.1 Surface compaction


There are different types of machines for surface densification:
 Surface vibrators,
 Vibrating rollers,
 Vibro rammers,
 Explosion rammers,
 Impact slabs

Figure ‎2-1. Surface vibrator.

Figure ‎2-2. Vibrating roller.

13
Figure ‎2-3. Vibro rammer.

Figure ‎2-4. Explosion rammer.

Figure ‎2-5. Impact slab.

14
2.2 In-depth compaction
At deep compactions, depths of more than about 2 m are compacted, which cannot be
reached with the surface compaction techniques mentioned before. For soft soils and deep
compaction still very heavy impact slabs can be used:
1) One can drop weights of 15 to 30 tons from 30 m high, with a fixed number of drops.
2) After that the crater is filled with appropriate ground (sandy, not clayey soil)
3) After that the steps 1) and 2) can be repeated.

Figure ‎2-6. In-depth impact compaction. (Picture: Brüchner Grundbau)

But for sandy soils another technique is used, based on long vibrating needles. These are put
in the ground with the help of their own weight, the vibrations, and sometimes also water
injection. Most needles are about 30 to 40 cm thick, 3 to 4,5 m long and weigh between 1,5
and 3 tons. The frequencies of the horizontal vibrations are mostly between 30 and 60 Hz.
Because of the vibrations, these techniques should not be used near existing buildings.

15
Figure ‎2-7. Vibration needles, with water injectors at the tip. (Photo: BIUG Geotechnik)

For sandy soils, this vibration technique, with or without water injection, works fine. For
clayey soils, one cannot decrease the amount of pores, but one can increase the strength,
stiffness and permeability (for drainage to speed up the consolidation), by vibrating gravel or
small stones into the ground.

Figure ‎2-8. Vibrator (Rüttler) in sandy soils and gravel (Schotter) injection in clayey soils.

16
3 Injection techniques
Injection techniques are techniques in which mixtures of materials are injected under pressure
in the pores and other open spaces of the underground in order to decrease the permeability
or to increase the strength and/or stiffness. One needs large injection pipes and large
machines for this.

Figure ‎3-1. Injection pipes and machinery.

There are several grouting techniques:


a) Permeation grouting: a very fluid liquid grout (water and cement) is injected in the
pores (of mostly sand) to make it more impermeable and stronger.
b) Jet-grouting: Soil is mixed with a powerful grout beam. The beam is rotated to make
columns or walls.
c) Fracture-grouting: a very fluid liquid grout (water and cement) is injected with high
pressure in order to create horizontal cracks (unfortunately mostly vertical cracks
appear). The horizontal cracks are made for compensating settlements..
d) Compaction Grouting: Very stiff grout is blown up like a balloon in order to
compensate settlements.

Figure ‎3-2. Grouting techniques.

17
3.1 Permeation grouting
Permeation grouting is often used to reduce the leakage under a dam, especially when the
difference in groundwater head has washed away the smaller sand particles in the ground
over the years. Hardening of the grout takes ½ or 1½ hours, in which there should not be any
groundwater flow, so the lake must be empty.

Figure ‎3-3. Permeation grouting under a dam.

The injected material depends on the size of the pores. In larger gravel: sand-cement mixture;
in fine gravel: cement; in coarse sand: cement-bentonite; in fine sand: micro-cement; in silts:
sodium silicates or acrylamide. When the injected materials are not water and cement but
other chemical products, we talk about chemical injections.
The permeation grouting technique is also conducted before an excavation in bedrock to
reduce the inflow of gas and water into the proposed excavation.

Figure ‎3-4. Permeation grouting before excavation.

18
3.2 Jet-grouting
The Jet-grouting technique can be used for different purposes. An example is the
underpinning of a shallow foundation that is settling too much. In this way, more or less a
deep foundation is made.

Figure ‎3-5. Foundation support.

By combining a long line of grouting columns even a retaining wall can be made.

Figure ‎3-6. Retaining wall from jet-grouted columns.

In fact, besides the walls, also the floor can be jet-grouted. In this way a complete swimming
pool can be made in a small and difficult to access back yard.

Figure ‎3-7. Jet-grouted swimming pool.

19
There is a risk in jet-grouting. The grouting tubes are never installed exactly at where the
design requires. Also these tubes are never installed fully perpendicular. Due to this, the
centres of the columns are never as designed. Also the radiuses of the columns are never
exactly as designed. This can cause small leaks in the grouted floor, which are very difficult
to locate and to repair. A famous case is the tunnel under the Big Market Street in Den Hague
in the Netherlands, where this problem caused major problems. The final, very time and
money consuming method, was to install air-locks and seal off the whole tunnel, pushing out
the ground water and repairing the leaks, all under air-pressure.

Figure ‎3-8. Designed (left) and constructed (right) grout floor with leak.

3.3 Fracture grouting


Fracture grouting is mainly used for lifting buildings, which have settled too much. This
mostly happens due to insufficient deep constructed shallow foundations and settlements
caused by excavations like building pits and tunnels. The grout pressure should be far more
than the vertical pressure at that particular level in the ground.

Figure ‎3-9. Fracture grouting (Pictures Hayward Baker, USA).

20
3.4 Compaction grouting
The technique of compaction grouting is, just as fracture grouting, mostly used for improving
shallow foundations.

Figure ‎3-10. Compaction grouting (Picture Hayward Baker, USA).

Both the fracture grouting and the compaction grouting can be seen as a way of
compensation grouting. This can be done vertically through the house floor, diagonally under
a house foundation or horizontally from a pit over a longer distance. This last option is very
expensive and is mostly used for the construction of bored tunnels in soft soils near
foundations of buildings, in order to compensate the settlements of several nearby houses.

Figure ‎3-11. Horizontal compensation grouting from pit (Photo Keller Grundbau).

21
4 Dewatering & drains
4.1 Introduction
The consolidation and dewatering reduces the amount of pores in cohesive and low-
permeable soils (clay and peat). In this way, it increases the stiffness and strength. This is the
same as the effect of the mechanical compaction for the non-cohesive and high-permeable
soils (sand and gravel). The low-permeable soils however need far more time to consolidate
and to creep. Installing vertical drains, improves the dewatering and therefore also the
consolidation process.
Vertical drains are often used to excel the consolidation process, because the excess
groundwater pressure can first flow horizontally to the drain and then vertically to the
surface. If the drainage distance reduces from 5 m to only 1 m, then the consolidation time
reduces 25 times; or sometimes from 6 years to only 3 months.

Figure ‎4-1. Plastic vertical drains.

The best way to speed up the consolidation process is to use a combination of vertical drains,
horizontal drains and a temporarily fill to increase the pressure temporarily.

Figure ‎4-2. Combination of vertical drains, horizontal drains and a temporarily fill.

22
4.2 Vertical drainage
The degree of consolidation U indicates how much water (pressure) has already dissipated.
By approximation the following applies for one-dimensional consolidation:
2 cvt
U (4.1)
 h2
in which: cv = coefficient of (vertical) consolidation (oedometer test)
h = drainage height. For double sided drainage: h = ½ H.

The above degree of consolidation is not valid for vertical drainage. For that, see the next
section.
The following is valid for the vertical consolidation "constant" cv:
k
cv  v
 w mv
No sensible estimates can be made of cv. This is because both the vertical permeability kv and
the vertical soil stiffness mv vary too much per soil layer. Furthermore, kv and mv decrease
during consolidation due to their dependence on the stresses! The only reasonable solution is
to do oedometer tests for the correct stress path.

The duration of one-dimensional consolidation:


1.78  h 2
t99% 
cv
At time t99%, the pressure has adjusted 99%. From then on one can assume: U  1 .

If the natural drainage is too slow, one can speed up the drainage with a vertical drainage
system. The best-known vertical drainage systems are sand drains and synthetic geo-drains.
Vertical drainage is applied to accelerate the consolidation process. The most important
reason for accelerating the consolidation process is the problem that one starts construction
too late. An additional matter is that the drainage costs can sometimes be earned back by
interest savings, by being able to purchase the land later. Drawbacks of vertical drainage are:
 the extra costs,
 the environmental drawbacks (plastic in the soil),
 the degradation of the seal against vertical water transport (“piercing the polder”).

One must realise that the vertical drainage does not accelerate the secondary settlement
(creep). If one doesn’t satisfy a certain residual settlement requirement, vertical drainage is
not a solution in all cases. Preferably one should consider working with a longer settlement
period or with temporary embankments.

In the case of vertical drainage, there is a constant horizontal drainage distance for all soil
layers. This makes calculating the degree of consolidation per layer easier. The following is
valid:
8Th

Uh  1 e (4.2)
With:
ch  t
Th 
D2

23
kh
ch 
mv   w
1  v
mv  
Eoed  v
n2  3 1 1 
  2 ln(n)   2 1  2  
n 1  4 n  4n  
D
n
d
2 b  t 
d

D  1.05  l (for triangular pattern of drainage)
 1.13  l (for square pattern of drainage)

in which:
Uh = horizontal degree of consolidation (for vertical drainage) [-]
ch = (horizontal) consolidation coefficient (oedometer test) [m2/s]
kh = (horizontal) permeability [m/s]
w = unit weight of water [kN/m3]
l = distance between two vertical drains [m]
D = equivalent drain distance [m]
d = equivalent drain thickness [m]

Note:
 The horizontal permeability of a soil layer can be greater (or smaller) than the vertical
permeability.
 The distance l between two drains is generally between 1 and 3 m.
 A triangular pattern is always more economical than a square pattern.
 The cross section of narrow synthetic drains is b  t = 100  4 mm2. The cross section of wide
synthetic drains is b  t = 300  4 mm2.

24
5 Ground freezing
5.1 Introduction
Ground freezing is a construction technique that has been used for over one hundred years to
provide temporary earth support and groundwater control. Applications are found in the
underground civil, mining and environmental remediation industries.
The ground freezing process uses a series of drilled freeze pipes and large refrigeration plants
to convert existing pore water in the soil to ice, creating a strong, water-tight frozen earth
material similar to rock or concrete. The frozen ground acts as an excavation support system
requiring no bracing, tiebacks or additional shoring. It has been used extensively for shafts
extending over several tens of metres deep. Its impermeable characteristic eliminates the
need for dewatering, making the technique practical for large groundwater barriers. These
barriers can be used to reduce or eliminate flow into excavations and contain contaminants or
restrict the flow of contaminated plumes.
Ground freezing may be used in any soil or rock formation regardless of structure, grain size
or permeability. However, it is best suited for soft ground rather than rock conditions.
Freezing may be used for any size, shape or depth of excavation and the same cooling plant
can be used from job to job. As the impervious frozen earth barrier is constructed prior to
excavation, it generally eliminates the need for compressed air, dewatering, or the concern
for ground collapse during dewatering or excavation.

Figure ‎5-1. Pit only supported by frozen ground, CERN, Switzerland.

Freezing is normally used to provide for a stronger foundation, a temporary support for an
excavation or to prevent ground water to flow into an excavated area. Successful freezing of
permeable water-bearing ground provides simultaneously for a seal against water and a
substantial strengthening of incoherent ground. No extraneous materials need to be injected
and apart from the contingency of frost heave, the ground normally reverts to its normal state.
It is applicable to a wide range of soils, but it takes considerable time to establish a
substantial ice wall and the freeze must be maintained by continued refrigeration as long as
required. This leads to the differences between the two different freezing techniques.

25
5.2 Brine freezing (indirect cooling)
In an indirect cooling system, a primary refrigeration plant is used to abstract the heat from a
secondary coolant circulating through pipes driven into the ground. The primary refrigerant
most commonly used will typically be some alternative to Freon, which, due to its ozone-
depleting characteristics, had to be phased before 1996. Other primary refrigerants are
ammonia, NH4 (–33.3°C) and carbon dioxide, CO2, which is not commonly used. The
secondary coolant, circulated through the network of tubes in the ground is usually a solution
of Calcium Chloride. With a concentration of 30% such as brine, this has a freezing point
well below that of the primary coolant.
The primary refrigeration process is basically the Carnot cycle of compression and expansion
reversed. The time required to freeze the ground will obviously depend on the capacity of the
freezing plant in relation to the volume of ground to be frozen and on the spacing and size of
freezing tubes and water content in the grounds.
The advantage of this calcium-chloride solution (brine) technique is that it is cheap to cool,
but the disadvantage is that one has to invest first in installing the whole equipment.

Figure ‎5-2. The brine freezing system (schematic diagram).

5.3 Nitrogen freezing (direct cooling)


With this method a large portable refrigeration plant is not necessary, and the temperature is
much lower, so this freezing method is therefore quicker in application. The nitrogen gas (N2)
boils at –196°C at normal pressure. It provides in this way the required cooling. It is brought
to, and stored on site under moderate pressure in insulated containers as a liquid. There is a
particular advantage for emergency use, i.e quick-freezing without elaborate fixed plant and
equipment. This may be doubly advantageous on sites remote from power supplies. In such
conditions the nitrogen can be discharged directly through tubes driven into the ground, and
allowed to escape to the atmosphere. Precautions for adequate ventilation must be observed.
When there is time for preparation, an array of freezing tubes is installed for the nitrogen
circulation, including return pipes exhausting to the atmosphere.
The speed of ground freezing with N2 is much quicker than with other methods, days rather
than weeks, but liquid nitrogen is costly. Therefore the method is particularly appropriate for
a short period of freezing of up to about 3 weeks. It may be used in conjunction with the
other processes with the same array of freezing tubes and network of insulated distribution
pipes, in which liquid nitrogen is first used to establish the freeze quickly and is followed by

26
ordinary refrigeration to maintain the condition while work is executed. This can be of
particular help when a natural flow of ground water makes initial freezing difficult.
The advantages are that it is rather cheap to install and the ground freezes quickly. The
disadvantage is that the nitrogen is expensive to use for a long time, so then the brine system
becomes cheaper.

Figure ‎5-3. two nitrogen tanks for freezing.

Figure ‎5-4. Examples of freezing: left complex building pit connection; right: complex safety
tunnel lateral to main tube of Western Scheld Tunnel.

5.4 Thermic Design


The targets for the design of a freeze are:
 Calculation of the location of the freezing tubes
 Calculation of the required energy
 Calculation of the required freezing time
 Selection of the freezing system

Sometimes the freeze should be retained within certain limits in order to prevent damage of
nearby structures (foundations of buildings, building pits, underground stations or tunnels).
For these cases, a combination of freezing tubes and heating tubes has to be used.

27
The boundary conditions for the thermic design depend on the strength and stiffness of the
frozen soil:
 The average temperature of the frozen wall should be lower than -10 ºC.
 The thickness of the wall should be at least 2.5 m within the isotherms of -2 ºC.

The sensitivity of the design should be checked on three important aspects:


 The influence of the groundwater flow
 The influence of the colder nitrogen tubes on the brain tubes (in case both are used)
 The influence of failure of one of the freezing tubes

For those calculations, complex Finite Element Models are used, specifically designed for
these types of thermic designs (see Figure below).

Figure ‎5-5. Examples of design (picture Crux Engineering, NL).

The strength of the frozen soils strongly depends on the temperature of the soil (see below).

Figure ‎5-6. Strength of sand and clay versus temperature.

28
6 Geotextiles and reinforced soil
6.1 Geotextiles
Geotextiles are permeable fabrics which, when used in association with soil, have the ability
to separate, filter, reinforce, protect or drain. As the use of geotextile fabrics has expanded
there has been the introduction of geotextile composites and the development of products
such as geogrids and meshes. Overall these materials are referred to as geotextiles and related
products. All have a wide range of applications and are currently used to benefit in many
civil engineering applications including roads, airfields, railroads, embankments, retaining
structures, reservoirs, canals, dams, bank protection and coastal engineering.
Geomembranes are impermeable membranes used widely as cut-offs and liners. Until recent
years, geomembranes were used mostly as canal and pond liners; however, one of the largest
current applications is the containment of hazardous or municipal wastes and their leachates.
In many of these applications geomembranes are employed with geotextile or mesh
underliners which reinforce or protect the more flexible geomembrane whilst also acting as
an escape route for gases and leachates generated in certain wastes.

Figure ‎6-1. Different types of geotextiles.

There are different ways to distinguish geotextiles, one way would be:
1. Erosion protecting
2. Sealing
3. Filtering
4. Reinforcing
5. Draining
6. Separating

Figure ‎6-2. Six different functions of geotextiles (G).

The function of draining is already discussed in Chapter 4 “Dewatering & drains”.

29
Figure ‎6-3. Erosion protection.

Figure ‎6-4. Erosion protection.

Figure ‎6-5. Erosion protection.

30
Figure ‎6-6. Sealing.

Figure ‎6-7. Combination of reinforcing & separating.

6.2 Reinforced soil / Terre armée


(Some part of the text in this section is from Terre Armée, France).
Soil reinforcement is a technique based on the formation of a composite material through the
association of compacted frictional fill and linear soil reinforcement. The soil reinforcement
may be either metallic (like the high adherence galvanised steel strips) or synthetic (like the
high tenacity polyester-based strips).
Reinforced earth provides:
 Strength: the resistance and stability of the composite structure provides significant
load-bearing capacity.
 Cost effectiveness: the ease and speed of construction are significant advantages in
reducing overall cost.
 Reliability: the durability of the materials used is well documented and the safety of
the structures is high.
 Adaptability: the technology provides solutions to complex cases and often proves to
be the best answer to circumstances such as restricted right-of-way, unstable natural
slopes, marginal foundation conditions and large settlements.
 Aesthetic appearance: the variety of facings can meet all architectural requirements.

31
Figure ‎6-8. Terre armée construction.
A big advantage is that the terre armée elements can be prefabricated and brought to places
where it is difficult and time consuming to produce and construct large concrete retaining
structures.

Figure ‎6-9. Terre armée at complex places.

One can combine the terre armée wall with the construction of a steep slope by using several
layers (sandwiches) of soil and geogrids, see Figure 6-10.

Figure ‎6-10. Cross section of combined terre armée wall with slope for Shinkan-sen train
yard at Biwajima, Nagoya (Tatsuoka et al.,1992).

32
One should think of making the original slope stepwise and not smooth, in order to prevent
shear failure along the old slope. By using gravel gabions at the toe of the slope, the drainage
is solved, for preventing the terre armée wall to build op excess pore pressure after rainfall.

Figure ‎6-11. Combined terre armée wall and slope.

Micro/local stability
A reinforced earth wall is in fact a wall consisting of prefab slabs piled on top of each other.
For horizontal micro stability, each slab is attached to two strips of geo-textile (or zinc-coated
steel), which disappear horizontally into the soil behind the wall. Due to the weight of the
soil, the strips can absorb a tensile force and hold the individual wall elements in place. The
force of this strip can be:
Fstrip  2bl 'v tan  (6.1)

Macro/global/total stability
The collapse mechanism of the entire wall can be solved with Fellenius and Bishop but that
requires calculating the extra tensile force of the strips correctly. For the macro stability, only
the part of the strip outside of the slip circle (see Figure 6-12) can absorb an additional force
in the order of:
Fstrip  2bleff  'v tan  (6.2)

Fstrip

Figure ‎6-12. Extra stability due to reinforced earth.

Note:
Not many computer programmes can calculate reinforced earth. Some programmes that make
calculations with geotextiles do not include the correct effective length in the calculations. Finite
Element Models (Plaxis), however, automatically calculate this length properly.

33
7 Ground exchange
7.1 Natural soils
Ground exchange is a very simple and cheap per cubic metre solution for improving the
conditions (strength, stiffness or permeability) of the ground. This solution is often used to
avoid pile foundations and use shallow foundations instead.
The biggest advantage is found when:
1. A soft soil layer lies on top of a strong and stiff soil layer
2. The soft soil layer is thin
3. The groundwater level is low

The question whether the exchange is profitable can only be answered by calculating the
extra costs of the exchange and by comparing this with the savings of using a lighter structure
(for example, a shallow foundation instead of a pile foundation). Besides looking at the costs,
one also has to:
 Check the ground for its quality, both the old and the new ground,
 Look at the chemical situation and strange or large parts (> 5 cm stones, waist, etc.),
 Find a place where to leave the removed ground (other projects, soil bank?) and
where to get the better ground, and
 Get the necessary licenses for all this.

Figure ‎7-1. Transport of soil is simple and relatively cheap.

Transport of the ground is simple as long as the ground is strong enough to carry the trucks.
One has to be careful for soft grounds though.

Figure ‎7-2. Be careful with soft grounds.

34
Metal plates can help the transport on soft ground, if the ground is even softer; mostly the
ground water table is high and maybe dredging becomes a better solution.

Figure ‎7-3. Plates spread the load.

Dredging and shipping is by far the cheapest solution per cubic metre of ground exchange
and transport, of course only when there is water (sea, lake, river or canal) nearby.

Figure ‎7-4. Dredging techniques are the best solution for large sand supplying near water.

To optimise the costs, one has to find the right equipment for both the ground type and
project size.

Figure ‎7-5. Small tools for small project and large tools for large projects.

35
7.2 Artificial heavy or light products
Sometimes it is better not to use natural soils as a replacement. There are cases in which very
light or very heavy materials are much better.

Heavy products
Heavy materials are used in geotechnical engineering when there are buoyancy problems (see
Chapters 45 and 46), for example in the case of a bored tunnel which is coming too near the
surface at both ends and the amount of ground on the tunnel ends is too light. Another
solution to make the whole tunnel deeper, is especially expensive for both entrances, because
these will then become much longer and deeper. This can be avoided by exchanging the
existing soil above the tunnel ends by slag. Slag is a by-product of smelting ore in the
furnaces of a steel mill, to separate the metal fraction from the unwanted fraction. It can
usually be considered to be a mixture of metal oxides and silicon dioxide. This waste product
is very heavy (up to 40 kN/m3) and cheap. Slag is also often used for the foundation of roads,
becomes it becomes very strong and stiff after a while.

Light Materials
Light Materials are mostly used in combination with soft soils in order to reduce the
settlements. EPS (Expanded Polystyrene) foam is extremely light, only 0.15 kN/m3 to 0.35
kN/m3, but also strong enough to make roads and even complete bridge abutments (in

Figure ‎7-6. Left: Expanded Polystyrene Foam blocks on both sides of a train line at Alphen
at the Rhine, the Netherlands.‎Picture‎from‎M.‎Duškov,‎InfraDelft‎bv.‎Right:‎Argex‎
Norway) on it. In Japan it is often used, because it is also strong enough to resist earthquakes.
Another well-known light weight material is Argex, an expanded clay product with a weight
of only 3.7 kN/m3 to 6.5 kN/m3. It is also rather strong and fire resistant. It is often used to
reduce the settlements or in combination with sheet pile walls or terre armée walls in order to
reduce the active pressure on the wall.

36
II Tunnelling

37
8 Tunnel types
A tunnel is an underground passageway, completely enclosed except for some openings for
air, fire escape and of course the openings at each end. A tunnel may be for foot or vehicular
road traffic, for rail traffic, or even for water. Some tunnels are aqueducts to supply water for
consumption or for hydroelectric stations, and some are sewers. Other uses include electricity
or telecommunication cables. The most common types and methods of tunnel construction
that are in use are:
 Cut-and-cover tunnels are built by excavating a trench, constructing the concrete
structure in the trench, then covering it with soil. The tunnels may be constructed in
place or by using precast sections
 Bored or mined tunnels, built without excavating the ground surface. These tunnels
are usually labelled according to the type of material being excavated. Sometimes a
tunnel passes through the boundary between different types of material; this often
results in a difficult construction known as mixed face (Chapter 8).
 Rock tunnels are excavated through the rock by drilling and blasting, by mechanized
excavators in softer rock, or by using rock tunnel boring machines (TBM). In certain
conditions, Sequential Excavation Method (SEM) is used.
 Soft ground tunnels are excavated in soil using a shield or pressurized face TBM
(principally earth pressure balance or slurry types), or by mining methods, known as
either the sequential excavation method (SEM).
 Immersed tunnels are made from very large precast concrete or concrete-filled steel
elements that are fabricated in the dry, floated to the site, placed in a prepared trench
below water, and connected to the previous elements, and then covered up with
backfill.
 Jacked box tunnels are prefabricated box structures jacked horizontally through the
soil using methods to reduce surface friction; jacked tunnels are often used where
they are very shallow but the surface must not be disturbed, for example beneath
runways or railroads embankments.

These tunnels can be grouped in different ways. One way is shown in Figure 8-1.

Figure ‎8-1. Tunnel types.

38
Over the last century many rock tunnelling techniques have been developed. For example:
• The English method
• The German “core” construction method (sequential)
• The old Austrian method (sequential with bottom pilot opening)
• The Belgian method (sequential with top heading and immediate final support)
• The New Austrian Tunnelling Method (NATM) (full face or sequential)
• The Norwegian Method of Tunnelling (Nordic – NMT, usually hard rock full face)
• The old Italian construction method
• The Italian method (full face based on ground pre-improvement)
• The needle beam method (American beam method)

In fact many countries developed their own techniques, looking more at their own geology
than at the obtained experiences of other countries.

In this book the following tunnelling methods will be discussed:

Cut & Cover:


• Open pit
• Classical building pit (Bottom-up)
• Wall-roof method (Top-down)
TBM (Tunnel Boring Machine) methods:
• Open Face TBM
• Hydro shield (Slurry Shield) TBM
• EPB (Earth-Pressure Balance) TBM
• Hard rock TBM
Blast & drill (rock) tunnel
• German method
• Old ÖTM (Alte Östereichische Tunnel Methode)
• NÖTM (Neue Östereichische Tunnel Methode)
Special methods:
• Immersed tunnel
• Jacked box tunnel

39
9 Cut & cover tunnel
9.1 Open building pit
Cut-and-cover is a simple and relatively cheap method of construction for shallow tunnels
where a trench is excavated and roofed over with an overhead support system that is strong
enough to carry the load of what is to be built above the tunnel. By far the cheapest method to
make a tunnel is to make a simple building pit with slopes. This is also only possible when
there are no groundwater problems, or when for the duration of the tunnel construction,
dewatering (pumping) is allowed. In urban areas this is often not allowed (especially when
there are wooden pile foundations or shallow foundations on soft soils, subject to
settlements). When it is allowed, it is often only allowed for a year, which is mostly
insufficient.

Figure ‎9-1. Open pit tunnelling (Tunnel OA08 Esch-Belval, Luxembourg).

Figure ‎9-2. Open pit tunnelling (Tunnel Route du Nord, Walferdange, Luxembourg).

40
9.2 Classical cut & cover
Two basic forms of cut-and-cover tunnelling are available:
1. Classical cut & cover (Bottom-up method): A trench is excavated, with ground
support as necessary, and the tunnel is constructed in it. The tunnel may be of in situ
concrete, precast concrete, precast arches, or corrugated steel arches; in early days
brickwork was used. The trench is then carefully back-filled and the surface is
reinstated.
2. Wall-Roof method (Top-down method): First side support walls and capping beams
are constructed from ground level by such methods as diaphragm walling, slurry
walling, or contiguous bored piling. Then a shallow excavation will be made for the
making of the tunnel roof. This will be made of precast beams or in situ concrete. The
surface is then reinstated except for access openings. This allows early reinstatement
of roadways, services and other surface features. Excavation then takes place under
the permanent tunnel roof, and the base slab is constructed.

Shallow tunnels are often of the cut-and-cover type (if under water, of the immersed-tube
type), while deep tunnels are excavated, often using a tunnelling shield. For intermediate
levels, both methods are possible.

Large cut-and-cover boxes are often used for underground metro stations, such as Canary
Wharf tube station in London. This construction form generally has two levels, which allows
economical arrangements for ticket hall, station platforms, passenger access and emergency
egress, ventilation and smoke control, staff rooms, and equipment rooms. The interior of
Canary Wharf station has been likened to an underground cathedral, owing to the sheer size
of the excavation. This contrasts with most traditional stations of the London Underground,
where bored tunnels were used for stations and passenger access.

Figure ‎9-3. Schematisation of the Classical Cut & Cover method.

41
Figure ‎9-4. Cut & cover tunnelling for the entrances of a submersed tunnel.

9.3 Aqueduct
An aqueduct is a special type of tunnel which is built in a similar way to the classical cut &
cover method. In this method the canal is temporarily led aside, the building pit is made, the
aqueduct is made in the dry building pit and finally the water is led back over the aqueduct.
Sometimes an aqueduct is created by making one side first and then the other side, in order to
maintain the water connection, although temporarily smaller, in order not to block the
shipping.
Nice examples for an aqueduct are the Gouwe Aqueduct and the Aqueduct Ring-canal
Haarlemmer Lake in the Netherlands.

Figure ‎9-5. Gouwe Aqueduct, Gouda, the Netherlands.

42
Figure ‎9-6. Aqueduct Ring-canal Haarlemmer lake, the Netherlands.

9.4 Wall-Roof method


In the Wall-Roof method first the walls are made, then the roof and finally the excavation
and the rest of the tunnel. This is done to disturb the infrastructure (traffic) as little as
possible in urban areas.

Figure ‎9-7. Schematisation A of the Wall-Roof method.

There are different types of schematisations for this method. The last version (C) is the best
to show the complexity of making a tunnel under a busy street. This method is often used for
railway stations in big cities, like the Ceintuurbaan Railway Station of the North-South Line
in Amsterdam.

43
Figure ‎9-8. Schematisation B of the Wall-Roof method.

Figure ‎9-9. Schematisation C of the Wall-Roof method.

44
Figure ‎9-10. Station Ceintuurbaan in Amsterdam is made with the
Wall-Roof method.

Figure ‎9-11. Artist impression of the train station Ceintuurbaan.

45
10 Tunnel Boring Machine (TBM)
(Most text and some pictures in this section are from Trem de Alta Velocidade Brasil,
Misistério dos Transportes)
A tunnel boring machine (TBM) is a machine used to excavate tunnels with a circular cross
section through a variety of soil and rock strata. They can bore through anything, from hard
rock to sand and clay. Tunnel diameters can range from a metre (done with micro-TBMs) to
almost 20 metres. Tunnels of less than a metre or so in diameter are typically done using
trenchless construction methods or horizontal directional drilling rather than TBMs.
Tunnel boring machines are used as an alternative to drilling and blasting (D&B) methods in
rock and conventional "hand mining" in soil. TBMs have the advantages of limiting the
disturbance to the surrounding ground and producing a smooth tunnel wall. This significantly
reduces the cost of lining the tunnel, and makes them suitable to use in heavily urbanized
areas. The major disadvantage is the upfront cost. TBMs are expensive to construct, and can
be difficult to transport. However, as modern tunnels become longer, the cost of tunnel
boring machines versus drill and blast is actually less—this is because tunnelling with TBMs
is much more efficient and results in a shorter project.
The bore process contains two stages. The first stage contains the excavation with a cutting
wheel and at the same time the advance of the shield with hydraulic jacks and also the
grouting for the tail void. The second stage is the temporarily withdrawal of some jacks and
the erection of precast concrete lining segments, to build a new ring of segments. After this
the first stage is repeated.

Figure ‎10-1. Bore process with TBM.

Modern TBMs typically consist of the rotating cutting wheel, called a cutter head, followed
by a main bearing, a thrust system and trailing support mechanisms. The type of machine
used depends on the particular geology of the project, the amount of ground water present
and other factors.

46
Figure ‎10-2. Transport of the precast concrete lining.

In soft ground, there are three main types of TBMs: open-face TBM’s, Slurry Shield (SS) and
Earth Pressure Balance Machines (EPB).

10.1 Open face TBM


Open face TBMs in soft ground rely on the fact that the face of the ground being excavated
will stand up with no support for a short period of time - this makes them suitable for use in
rock types with a strength of up to 10MPa or so, and with low water inflows. Face sizes in
excess of 10 metres can be excavated in this manner. The face is excavated using a cutter
head. The shield is jacked forwards and cutters on the front of the shield cut the remaining
ground to the same circular shape. Ground support is provided by use of precast concrete, or
segments that are bolted or supported until a full ring of support has been erected. A final
segment, called the key, is wedge-shaped, and expands the ring until it is tight against the
circular cut of the ground left behind by cutters on the TBM shield. Many variations of this
type of TBM exist.

Figure ‎10-3. Open face TBM.

10.2 Hydro shield or slurry shield (SS) TBM


The basic principle of this TBM is to maintain the face pressure during the excavation phase
by filling the working chamber, located behind the cutter head, with (bentonite) slurry. In soft
ground with very high water pressure and large amounts of ground water, Slurry Shield

47
TBMs are needed. These machines offer a completely enclosed working environment. Soils
are mixed with bentonite slurry, which must be removed from the tunnel through a system of
slurry tubes that exit the tunnel. Large slurry separation plants are needed on the surface for
this process, which separate the dirt from the slurry so it can be recycled back into the tunnel

Advantages
 Allows soft, wet, or unstable ground to be tunnelled with a speed and safety not
previously possible
 Suitable for ground with high water pressures (below water table)
 Limits ground settlement and produces a smooth tunnel wall. This significantly
reduces the cost of lining the tunnel, and makes it suitable to use in heavily urbanized
areas

Disadvantages
 The major disadvantage is the upfront capital cost. TBMs are expensive to construct,
difficult to transport, and require significant backup systems and power.
 Machinery is needed to separate the excavated soil from the slurry.
 Drive can be hindered by large stones and boulders

Figure ‎10-4. Cutter head (shield) of SS TBM.

Figure ‎10-5. Working chamber with bentonite slurry.

48
10.3 Earth Pressure Balanced shield (EPB) TBM
This is a mechanised tunnelling method in which soil is admitted into the tunnel boring
machine (TBM) via a screw conveyor arrangement which allows the pressure at the face of
the TBM to remain balanced without the use of slurry.

Advantages
 Allows soft, wet, or unstable ground to be tunnelled with a speed and safety not
previously possible
 Limits ground settlement and produces a smooth tunnel wall. This significantly
reduces the cost of lining the tunnel, and makes it suitable to use in heavily urbanized
areas
 Needs no slurry and machinery to separate the soil and slurry.

Disadvantages
 The major disadvantage is the upfront capital cost. TBMs are expensive to construct
(also due to large ground forces on TBM), difficult to transport and, require
significant backup systems and power.

Figure ‎10-6. Cross-Section through Earth Pressure Balanced Shield TBM.

10.4 Hard rock TBM


This method involves the use of a Tunnelling machine with a shield and cutter head suitable
for hard rock.

Advantages
 They offer a continuous and controlled means of tunnelling capable of high rates of
advance under favourable conditions.

Disadvantages
 The major disadvantage is the upfront capital cost. TBMs are expensive to construct,
difficult to transport, require significant backup systems and power.
 Their applicability is limited to long tunnels where the high rates of advance and
tunnel quality can offset their high capital cost.

49
Figure ‎10-7. Hard rock TBM.

10.5 Support pressure


All TBMs have to create a horizontal support pressure with the shield or with the slurry in
order to stabilise the effective earth pressure and also the groundwater pressure.
The maximum pressure of the slurry or shield is the vertical total ground stress at the top of
the tunnel, because then a blowout can occur: the uplifting of the ground, creating a leak.
The minimum pressure is the groundwater pressure at the bottom of the tunnel, in order to
avoid the water to flow into the tunnel. In fact, even a little more pressure is needed to
support the 3D arch of ground around the bore front. Failure in this way is sometimes called
blowin.

Figure ‎10-8. Blowout.

Figure ‎10-9. Blowin.

50
10.6 Pneumatic caisson
Sometimes a pneumatic caisson is used to make a part of the tunnel, for example having a
starting point for the tunnel boring machine (TBM). The technique originates from the
making of a pillar of a bridge in a river with a caisson. In the example below, the caisson in
front of the central station of Amsterdam, there were even two functions; a starting point for
a TBM and a foundation for a future bridge. The caisson is sunken down by pneumatic
sinking with hydraulic ejectors and pumps, which pump out the sand/water mixture. This
technique is mostly used when there is a sandy sub soil (easy excavation) and little space
(otherwise a building pit would be cheaper) or a bridge pillar in a river.

Figure ‎10-10. Caisson before and after sinking.

Figure ‎10-11. Pneumatic sinking with hydraulic ejectors and pumps.

10.7 Settlements
The volume of ground loss experienced during tunnelling can be related to the volume of
settlement expected at the ground surface (Peck, 1969). For a single tunnel in soft ground
conditions, it is typically assumed the volume of surface settlement is equal to the volume of
lost ground. However, the relationship between the volume of lost ground and the volume of
surface settlement is complex. Volume change due to bulking or compression is typically not
estimated or included in the calculations. Ground loss will produce a settlement trough at the
ground surface where it can potentially impact the settlement behaviour of any overlying or
adjacent bridge foundations, building structures, or buried utilities transverse or parallel to
the alignment of the proposed tunnel excavation. Empirical data suggests the shape of the
settlement trough typically approximates the shape of an inverse Gaussian curve.

51
Figure ‎10-12. Overview reasons for settlement ground level.

Figure ‎10-13. Settlement trough.

The shape and magnitude of the settlement trough is a function of excavation techniques,
tunnel depth, tunnel diameter, and soil conditions. In the case of parallel adjacent tunnels,
surface settlement is generally assumed to be additive. The shape of the curve can be
expressed by the following mathematical relationships (Schmidt, 1974):
Settlement trough on ground level, Gauss curve S(x) and Decrease in volume V:
S  x   Smax e  x
2
/2 i

(10.1)
V  Smax i 2

in which:
S(x) settlement of the ground level;
Smax maximum ground level settlement;
x horizontal distance to the axis of the tunnel;
z depth of the tunnel axis.

52
i coefficient which determines the distance between the bending points; for this the
following empirical relations are used:
- cohesive i = 0.43 z + 1.1
- non-cohesive i = 0.28 z - 0.1

Figure ‎10-14. Settlement trough.

Example:

53
11 Drill & blast (rock) tunnelling
Before the advent of tunnel boring machines, the method of drilling and blasting was the only
economical way of excavating long tunnels through hard rock, where digging is not possible
The drill and blast method follows the following steps repeatedly:
1. A number of holes are drilled into the rock
2. These are then filled with explosives (about 0.3 to 4.5 kg/m3)
3. The explosives are connected to a detonator
4. Safety rules and checks have to be carried out constantly
5. Detonating the explosives (the blast) will cause the rock to collapse and thus lengthen
the tunnel.
6. The vibrations have to be controlled
7. The air has to be controlled, especially after the blast (CO 2, CO, Nitrogasses)
8. The rubbles are removed
9. The new tunnel surface must be reinforced (wood, supports, reinforcement bars,
bolds, shotcrete)
10. Repeating these steps will eventually result in a tunnel.

The positions and depths of the holes (and the amount of explosive each hole receives) are
determined by a carefully constructed pattern, which, together with the correct timing of the
individual explosions, will guarantee that the tunnel will have an approximately circular
cross-section. The tunnel lining is mostly a face with sprayed concrete (shotcrete), with or
without rock bolts, or when the rock is of good quality; no lining at all. The typical
performance could be low; about 0.1 m to 1 m per day.

Advantages
 Suitable for hard rock where digging is not possible
 Its flexibility, mobility, and low capital cost constitute real advantages in many
situations, such as those involving short lengths of tunnel or low rates of advance.

Disadvantages
 There is a high risk of over breaking the tunnel profile and damaging the surrounding
rock
 High levels of noise and vibration make this unsuitable for an urban area.

In general German terms are used, see the Figure below. On top is the crown (Kalotte), The
middle part is Bench (Strosse) and below the Base (Sohle).

Figure ‎11-1. Definitions of positions in the tunnel.

54
There are different ways to make drill & blast rock tunnels. The most well-known in Europe
are:
 The German method,
 The old ÖTM (Old Austrian Tunnelling Method) and
 The NÖTM (New Austrian Tunnelling Method).

11.1 German Method or Core Method


The Core Method (Kernbauweise), also called the German Method was first used in France.
First the area around the core (Kern) is broken away (Ulme and Kalotte) and are inspected.
Then the core is broken away. Finally the base (Sohle) is closed. This method limits the
settlements of the tunnel.

Figure ‎11-2. German Method or Core Method.

11.2 Old Austrian Tunnelling Method or Old ÖTM


The old ÖTM (Alte Östereichische Tunnel Methode) uses a method that can be seen in the
figure below. The first excavation is at the base. This is also used for inspection. It is possible
to dig at several places at the same time; therefore this is a very fast method.

Figure ‎11-3. The Old ÖTM (Austrian Tunnelling Method).

11.3 New Austrian Tunnelling Method or NÖTM


There is no exact definition of this method. It is an improved method of the Old ÖTM. There
is a patent of Rabcewicz from 1948 and his dissertation of 1950 that explain this method. The
first idea is that stresses and settlements are limited by using an optimum of shapes and

55
blasting steps. The second is that deformations are measured to control the whole process;
which decides when each following step will start.
First the left and right drift walls (left and right ellipses in Figure 11-4 and Figure 11-5) are
excavated or drilled, starting at the crown and the bench and base at the end.

Figure ‎11-4. The NÖTM (New Austrian Tunnelling Method).

Figure ‎11-5. Excavating the bench.

Figure ‎11-6. Drilling.

56
Figure ‎11-7. Explosives injection.

Figure ‎11-8. Reinforcement and shotcrete.

57
12 Immersed tunnel
12.1 Construction method
An immersed tunnel is a type of underwater tunnel composed of segments, constructed
elsewhere and floated to the tunnel site to be sunk into place and then linked together. They
are commonly used for road and rail crossings of rivers, estuaries, sea channels and harbours.
Immersed tunnel segments are often used in conjunction with a cut and cover tunnel, which is
usually needed to connect the tunnel from near the water's edge to the entrance (portal) at the
land surface.

Figure ‎12-1. Immersed tunnel segments in a dry dock.

One can distinguish several construction phases:


 A trench is dredged in the bed of the water channel.
 Tunnel elements are constructed in a dry area, for example in a dry dock, a fabrication
yard, on a ship-lift platform or in a factory unit.
 The ends of the element are then temporarily sealed with bulkheads.
 Each tunnel element is transported to the tunnel site - usually floating, occasionally on
a barge, or assisted by cranes.
 The tunnel element is lowered to its final place on the bottom of the dredged trench.
 The new element is placed against the previous element under water. Water is then
pumped out of the space between the bulkheads.
 Water pressure on the free end of the new element compresses the rubber seal (Gina
profile) between the two elements, closing the joint. On top of the joint there is an
additional omega profile for safety.
 Backfill material is placed beside and over the tunnel to fill the trench and
permanently bury the tunnel.
 Approach structures can be built on the banks before, after or concurrently with the
immersed tunnel, to suit local circumstances.

58
The segments of the tube may be constructed in one of two methods. In the United States, the
preferred method has been to construct steel or cast iron tubes, which are then lined with
concrete. In Europe there is no reasonable explanation for such a method. In Europe a simple
reinforced concrete box construction is the standard; the sections are cast in a basin
(construction dock) which is then flooded to allow their removal.

Figure ‎12-2. Floating and towing of tunnel elements.

Figure ‎12-3. Placing of element.

Figure ‎12-4. Ballast tanks for sinking the tunnel element.

59
Figure ‎12-5. Left: covered tunnel; right: entrance of tunnel.

12.2 Chin-nose and jack-leg support


Every time a new element is placed, the new element will hang with its “nose” on the “chin”
of the previous element.

Figure ‎12-6. Chin-nose construction between the two end sides (bulk heads).

At the end of each element (at the nose side) the segment will rest on two legs, controlled by
jacks to optimise the position of the tunnel. After installation, sand is injected under the
segment. From then on, the chin-nose support and the two jack leg supports are out of
function.

60
Figure ‎12-7. Jack leg.

12.3 Advantages and disadvantages


The main advantage of an immersed tunnel is that it can be considerably more cost effective
than the alternative option; a bored tunnel beneath the water. Other advantages relative to
these alternatives include:
 The speed of construction
 Minimal disruption to the river/channel, if crossing a shipping route (but boring is
even better)
 Resistance to seismic activity (earthquakes)
 Safety of construction (for example, work in a dry dock as opposed to boring beneath
a river)
 Flexibility of profile

Disadvantages include:
 The tunnel is rather shallow, so rather exposed (usually with some rock armour and
natural siltation) on the river/sea bed, risking a sunken ship/anchor strike
 Direct contact with water necessitates careful waterproofing design around the joints
(same for bored tunnel) and round the jack legs.

61
 The segmental approach requires careful design of the connections, where
longitudinal effects and forces must be transferred across (think of the chin nose
structure and the jack legs).

62
13 Jacked box tunnel
The method of the jacked box tunnel is a method of tunnel construction where cables are
used to pull (or hydraulic jacks are used to push) a box shaped tunnel through the ground.
This technique is commonly used to create tunnels under existing structures, such as roads or
railways. Jacked boxes can be wider than 20 m. A cutting head is normally used at the front
of the box being jacked and excavation is normally by excavator from within the box.
Mostly a platform is made to have a weight to connect the reaction force of the pulling or
pushing. The space between the tunnel and the platform is injected with bentonite slurry in
order to reduce the forces.
The figure below shows the construction phases. The fact that the second phase (moving the
tunnel and digging the tunnel face at the same time) goes so quickly, makes this method
perfect to install the tunnel over a long weekend. This keeps any hindrance to rail transport to
a minimum.

Figure ‎13-1. Jacked box tunnel construction phases.


(Figure from Allenby & Ropkins, IAEG2006)

63
Figure ‎13-2. Three jacked box tunnels ready to install.

Figure ‎13-3. Jacked box tunnel for railway line near A20 Lübeck-Moisling, Germany.

Figure ‎13-4. Jack pulling cable (left) Digging at the front (right).

64
III Shallow foundations

65
14 Types of shallow foundations
There are two types of foundations; shallow foundations and pile foundations. Shallow
foundations are used when the upper layer is both strong and stiff enough to carry the load of
the structure. If not, pile foundations have to be used in order to transfer the load of the
structure to deeper layers which are more strong and stiff.

There are four types of shallow foundations:


1. Plate foundation (or single foundation)
2. Strip foundation
3. Grid foundation
4. Slab foundation

The choice of the type of shallow foundation depends mostly on the structure and
corresponding load. A pillar is founded, for example, usually on a single plate foundation. A
wall is usually founded on strip foundation, a large floor usually on a grid foundation or slab
foundation.

Figure ‎14-1. Types of shallow foundations.

Over time the used construction materials for foundations have changed a lot. The oldest
foundations were simply made of well stacked dry stones. Later, dry stone-concrete mixtures
were used, brick structures and in more contemporary times, concrete without reinforcement.
In many countries nowadays only reinforced concrete is used. For renovations and for new
structures built near to older buildings, it is still important to understand the behaviour of the
old foundations.

66
Figure ‎14-2. Materials used in shallow foundations.

For the design of the settlement of a shallow foundation, see chapter 15-18.
For the design of the bearing capacity of a shallow foundation, see chapter 19-22.

67
15 Elastic stresses and deformations
The stresses and deformations are very important in the calculation of shallow foundations.
As long as we assume that we have non-failing soils, behaving in a rather linear elastic way,
we can do simple design calculations by hand.
Because of equilibrium of moments the stress tensor must be symmetric,
 xy   yx ,
 yz   zy , (15.1)
 zx   xz .
The equations of equilibrium constitute a set of six equations involving nine stress
components. In itself this can never be sufficient for a mathematical solution. The
deformations must also be considered before a solution can be contemplated.
For a linear elastic material, the relation between stresses and strains is given by Hooke’s
law,
1
 xx  [ xx  ( yy   zz )],
E
1
 yy  [ yy  ( zz   xx )], (15.2)
E
1
 zz  [ zz  ( xx   yy )],
E
1 
 xy   xy ,
E
1 
 yz   yz , (15.3)
E
1 
 zx   zx .
E
where E is the modulus of elasticity (Young’s modulus), and  is Poisson’s ratio. The
equations (15.2) and (15.3) add six equations to the system, at the same time introducing six
additional variables.
The six strains can be related to the three components of the displacement vector,
u x
 xx   ,
x
u
 yy   y , (15.4)
y
u
 zz   z ,
z
u x u y
 xy   12 (  ),
y x
u u
 yz   12 ( y  z ), (15.5)
z y
u u
 zx   12 ( z  x ).
x z

68
These are the compatibility equations. In total there are now just as many equations as there
are variables, so that the system may be solvable, at least if there are a sufficient number of
boundary conditions.
The minus sign in the equations has been introduced in geotechnical engineering because a
volume decrease leads to a positive strain and a volume increase to a negative strain.
A number of problems solutions of the system of equations can be found in literature on the
theory of elasticity. In soil mechanics the solutions for a half space or a half plane, with a
horizontal upper surface, are of special interest. In the next chapters some of the most
important solutions for soil mechanics are further discussed.
For the solved elastic problems the strains and effective stresses can be calculated for drained
or consolidated loads, with the help of two stiffness parameters (E, or K,G). If these
parameters are replaced by two undrained parameters ( Eu , u ) the strains and total stresses
can be calculated for undrained problems (short loads on clay or peat). When calculating
these two parameters, one should realize that an undrained load, due to the incompressibility
of water, causes no volume change, so
 u  0.5. (15.6)
The undrained compression modulus becomes infinite:
Eu
Ku   . (15.7)
3(1  2 u )
Besides this the water has no shear stresses, both in the drained as in the undrained state.
From this it follows that
E Eu
Gu  G   . (15.8)
2(1  ) 2(1  u )
From which can be derived that
(1  u )
Eu  E . (15.9)
(1  )
Replacing the drained stiffness parameters ( E, ) from the previous chapters by the
undrained stiffness parameters ( Eu , u ) and calculating these values in the way demonstrated
above, the solutions can be used not only for the drained, but also for the undrained case.

69
16 Boussinesq
In 1885 the French scientist Boussinesq obtained a solution for the stresses and strains in a
homogeneous isotropic linear elastic half space, loaded by a vertical point force on the
surface, see Figure 16-1.

Figure ‎16-1. Point load on half space.

A derivation of this solution is given in Appendix B, see also any textbook on the theory of
elasticity (for instance S.P. Timoshenko, Theory of Elasticity, paragraph 123). The stresses
are found to be
3P z 3
 zz  , (16.1)
2 R5
P 3r 2 z 1
 rr  [ 5  (1  2 ) ], (16.2)
2 R R( R  z )
P 1  2 R z
   (  ), (16.3)
2 R 2
Rz R
3P rz 2
 rz  . (16.4)
2 R5
In these equations r is the cylindrical coordinate,
r x2  y 2 , (16.5)
and R is the spherical coordinate,
R x2  y 2  z 2 . (16.6)
The solution for the displacements is
P(1  ) r 2 z z
ur  [ 3  (1  2 )(1  )], (16.7)
2 ER R R
u  0, (16.8)
P(1  ) z2
uz  [2(1  )  2 ]. (16.9)
2 ER R

70
The vertical displacement of the surface is particularly interesting. This is
P(1  2 )
z0 : uz  . (16.10)
 Er
For r  0 this tends to infinity. At the point of application of the point load, the
displacement is infinitely large. This singular behaviour is a consequence of the singularity in
the surface load, and as in the origin, the stress is infinitely large. The fact that the
displacement in that point is also infinitely large may not be so surprising.
Another interesting quantity is the distribution of the stresses as a function of depth, just
below the point load, i.e. for r  0 . This is found to be

Figure ‎16-2. Vertical normal stress  zz .

3P
r 0 :  zz  , (16.11)
2 z 2
P
r 0 :  rr     (1  2 ) . (16.12)
4 z 2
These stresses decrease with depth, of course. In engineering practice, it is sometimes
assumed, as a first approximation, that at a certain depth, the stresses are spread over an area
that can be found by drawing a line from the load under an angle of about 45º. That would
mean that the vertical normal stress at a depth z would be P  z 2 , homogeneously over a
circle of radius z. That appears to be incorrect (the error is 50 % if r  0 ), but the trend is
correct, as the stresses indeed decrease with 1 z 2 . In Figure 16-2 the distribution of the
vertical normal stress  zz is represented as a function of the cylindrical coordinate r, for two
values of the depth z.

71
Figure ‎16-3. Uniform load over circular area.
The assumption of linear elastic material behaviour means that the entire problem is linear, as
the equations of equilibrium and compatibility are also linear. This implies that the principle
of superposition of solutions can be applied. Boussinesq’s solution can be used as the starting
point of more general types of loading, such as a system of point loads, or a uniform load
over a certain given area.
As an example, consider the case of a uniform load of magnitude p over a circular area, of
radius a. The solution for this case can be found by integration over a circular area (S.P.
Timoshenko, Theory of Elasticity, paragraph 124), see Figure 16-3.
The stresses along the axis r  0 , i.e. just below the load, are found to be
z3
r 0 :  zz  p(1  3 ), (16.13)
b
z z3
r 0 :  rr  p[(1  )  12 (1  )], (16.14)
b b3

in which b  z 2  a 2 .
The displacement of the origin is
pa
r  0, z  0 : uz  2(1  2 ) . (16.15)
E
This solution will be used as the basis of a more general case in the next chapter.
Another important problem, which was already solved by Boussinesq (see also Timoshenko)
is the problem of a half space loaded by a vertical force on a rigid plate. The force is
represented by P   a 2 p , see Figure 16-4. The distribution of the normal stresses below the
plate is found to be
1
p
z  0, 0  r  a :  zz  2
. (16.16)
1  r 2 a2

72
Figure ‎16-4. Rigid plate on half space.
This stress distribution is shown in Figure 16-4. At the edge of the plate the stresses are
infinitely large, as a consequence of the constant displacement of the rigid plate. In reality the
material near the edge of the plate will probably deform plastically. It can be expected,
however, that the real distribution of the stresses below the plate will be of the
form shown in the figure, with the largest stresses near the edge. The centre of the plate will
subside without much load.
The displacement of the plate is
 pa
z  0, 0  r  a : uz  (1  2 ) . (16.17)
2 E
When this is compared with the displacement below a uniform load, see (16.15), it appears
that the displacement of the rigid plate is somewhat smaller, as could be expected.

73
17 Flamant
In 1892 Flamant obtained the solution for a vertical line load on a homogeneous isotropic
linear-elastic half space, see Figure 17-1. This is the two dimensional equivalent of
Boussinesq’s basic problem. It can be considered as the superposition of an infinite number

Figure ‎17-1.‎Flamant’s‎Problem.
of point loads, uniformly distributed along the y-axis. A derivation is given in Appendix B.
In this case the stresses in the x,z-plane, resulting from the line load, are
2F z 3 2F
 zz   cos3  , (17.1)
 r 4
r
2F x2 z 2F
 xx   sin 2  cos  , (17.2)
 r4  r
2 F xz 2 2 F
 xz   sin  cos 2  . (17.3)
 r 4
r
In these equations r  x 2  z 2 . The quantity F has the dimension of a force per unit length,
so that F r has the dimension of a stress.
Expressions for the displacements are also known, but these contain singular terms, with a
factor ln r . This factor is infinitely large in the origin and at infinity. Therefore these

Figure ‎17-2. Strip load.


expressions are not so useful.
On the basis of Flamant’s solution several other solutions may be obtained using the principle
of superposition. An example is the case of a uniform load of magnitude p on a strip of width
2a, see Figure 17-2. In this case the stresses are
p
 zz  [(1  2 )  sin 1 cos 1  sin 2 cos 2 ], (17.4)

74
p
 xx  [(1   2 )  sin 1 cos1  sin  2 cos 2 ], (17.5)

p
 xz  [cos2 2  cos2 1 ]. (17.6)

In the centre of the plane, for x  0 , 2  1 . Then the stresses are
2p
x0 :  zz  [1  sin 1 cos 1 ], (17.7)

Figure ‎17-3. Stresses for x=0.


2p
x0 :  xx  [1  sin 1 cos 1 ], (17.8)

x0 :  xz  0. (17.9)
That the shear stress  xz  0 for x  0 is a consequence of the symmetry of this case. The
stresses  xx and  zz are shown in Figure 17-3, as functions of the depth z. Both stresses tend
towards zero for z   , of course, but the horizontal normal stress appears to tend towards
zero much faster than the vertical normal stress. It also appears that at the surface the
horizontal stress is equal to the vertical stress. At the surface this vertical stress is equal to the
load p, of course, because that is a boundary condition of the problem. Actually, in every
point of the surface below the strip load the normal stresses are  xx   zz  p .

75
Figure ‎17-4. Strip load next to a smooth rigid wall.
It may be interesting to further explore the result that the shear stress  xz  0 along the axis
of symmetry x  0 in the case of a strip load, see Figure 17-2. It can be expected that this
symmetry also holds for the horizontal displacement, so that ux  0 along the axis x  0 .
This means that this solution can also be used as the solution of the problem that is obtained
by considering the right half of the strip problem only, see Figure 17-4. In this problem the
quarter plane x  0 , x  0 is supposed to be loaded by a strip load of width a on the surface
z  0 , and the boundary conditions on the boundary x  0 are that the displacement ux  0
and the shear stress  xz  0 , representing a perfectly smooth and rigid vertical wall. The wall
is supposed to extend to an infinite depth, which is impractical. For a smooth rigid wall of
finite depth the solution may be considered as a first approximation.
The formulas (17.7) and (17.8) can also be written as
2p a az
x0 :  zz  [arctan( )  2 2 ], (17.10)
 z a z
2p a az
x0 :  xx  [arctan( )  2 2 ]. (17.11)
 z a z
Integration of the horizontal stress  xx from z  0 to z  h gives the total force on a wall of
height h,
2 a
Q ph arctan( ). (17.12)
 h
For a very deep wall ( h a ) this becomes, because arctan(a h)  a h ,

2
h : Q pa  0.637 pa. (17.13)

The quantity pa is the total vertical load F (per unit length perpendicular to the plane of the
drawing). It appears that the horizontal reaction in an elastic material is 0.637F.
For a very shallow wall ( h a ) the total lateral force will be, because then
arctan(a h)   2 ,
h0 : Q  ph. (17.14)

76
Figure ‎17-5. Line load next to a smooth rigid wall.
This is in agreement with the observation made earlier that the value of the horizontal stress
at the surface, just below the load, is  xx  p . For a very short wall the horizontal force will
be that horizontal stress, multiplied by the length of the wall.
Another interesting application of Flamant’s solution is shown in Figure 17-5. In the case of a
surface load by two parallel line loads it can be expected that at the axis of symmetry ( x  0 )
the horizontal displacement and the shear stress will be zero, because of symmetry. This
means that the solution of that problem can also be used as the solution of the problem of a
line load at a certain distance from a smooth rigid wall, because the boundary conditions
along the wall are that ux  0 and  zx  0 for x  0 . By the symmetry of the problem shown
in the left half of Figure 17-5, these conditions are satisfied by the solution of that problem.
The horizontal stresses against the wall are given by equation (17.2) for Flamant’s basic
problem, multiplied by 2, because there are two line loads and each gives the same stress. In
this formula the value of x should be taken as x  a , where a is the distance of the force to
the wall. The horizontal stress against the wall in this case is
4 Fa 2 z
 xx  . (17.15)
 (a 2  z 2 ) 2
The distribution of the horizontal stresses against the wall is also shown in Figure 17-5. The
maximum value occurs for z  0.577a , and that maximum stress is
F
 xxmax  0.4135 .
a
The total force on a wall of depth h can be found again by integration of the horizontal stress
over the depth of the wall. This gives
2 F
Q . (17.16)
 1  a 2 h2
If a  0 this is Q  0.637 F . If a increases the value of Q will gradually become smaller. A
force at a larger distance from the wall will give smaller stresses against the wall.

It should be kept in mind that only the extra forces and stresses resulting from the loads are
mentioned. The weight of the soil itself also causes stresses, both in vertical and horizontal
direction.

77
18 Deformation of layered soil
An important problem of soil mechanics practice is the prediction of the settlements of a
structure built on the soil. For a homogeneous isotropic linear elastic material the
deformations could be calculated using the theory of elasticity. That is a completely
consistent theory, leading to expressions for the stresses and the displacements. However,
solutions are available only for a half space and a half plane, not for a layered material (at
least not in closed form). Moreover, soils exhibit non-linear properties (such as a stiffness
increasing with the actual stress), often have anisotropic properties, and in many cases the
soil consists of layers of different properties. For such materials the description of the
material properties is already a complex problem, let alone the analysis of stresses and
deformations.

Figure ‎18-1. Load on layered soil.

For these reasons an approximation is often used, based on a semi-elastic analysis. In this
approximation it is first postulated that the vertical stresses in the soil, whatever it true
properties are, can be approximated by the stresses that can be calculated from linear elastic
theory. On the basis of these stresses the deformations are then determined, using the best
available description of the relation between stress and strain, which may be non-linear. If the
soil is layered, the deformations of each layer are calculated using its own properties, and
then the surface displacements are determined by a summation of the deformations of all
layers. In this way the different properties of the layers can be taken into account, including a
possible increase of the stiffness with depth.
The procedure is not completely consistent, because in a soil consisting of layers of different
stiffness, the stress distribution will not be the same as in a homogeneous linear elastic
material. A partial justification may be that the stresses following from an elastic
computation at least satisfy the equilibrium conditions. Also it has been found, by
comparison of solutions for layered materials with the solution for a homogeneous material,
that the distribution of the vertical stress  zz is not very sensitive to the material properties,
provided that the differences in material properties are not very large, i.e. excepting extreme
cases such as a very stiff layer on a very soft subsoil.

Example
The computation method can best be illustrated by considering an example. This example
concerns a circular fluid reservoir, having a diameter of 20 metres, which is being
constructed on a foundation plate on a layer of fairly soft soil, of 20 metres thickness, see
Figure 18-2.

78
Figure ‎18-2. Reservoir on soft soil.

Below the soft soil the soil is a hard layer of sand or rock. The compressibility of the soft soil
is about C10  50 . The pressure of the foundation plate on the soil is 20 kPa, and the
additional load by the fluid in the reservoir is 100 kPa. The problem is to determine the
settlement caused by the load, in the centre of the reservoir.

depth weight found. 1 load  log   1  h


(m) (kPa) (kPa) (kPa) (kPa) kPa (m)
1 10.00 19.98 29.99 99.90 129.89 0.6366 0.025
3 30.00 19.52 49.52 97.63 147.15 0.4730 0.019
5 50.00 18.21 68.21 91.06 159.27 0.3683 0.015
7 70.00 16.23 86.23 81.14 167.37 0.2880 0.011
9 90.00 14.01 104.04 70.06 174.07 0.2236 0.009
11 110.00 11.90 121.90 59.49 181.39 0.1726 0.007
13 130.00 10.04 140.04 50.20 190.24 0.1330 0.005
15 150.00 8.48 158.48 42.40 200.88 0.1030 0.004
17 170.00 7.19 177.19 35.96 213.15 0.0802 0.003
19 190.00 6.14 196.14 30.70 226.84 0.0631 0.002

Table ‎18-1: Computation of settlement.

The example has been elaborated in Table 18-1. The soil has been subdivided into 10 layers,
of 2 metres thickness each. The first column of the table gives the average depth of each
layer. The second column gives the effective stress due to the weight of the soil, assuming
that the effective unit weight of the soil is 10 kN/m3, so that for each metre depth the stress
increases by 10 kPa. The third column gives the additional stress due to the weight of the
foundation plate. These stresses have been calculated using the formula for the stresses below
a uniform circular load, equation (16.13),
z3
 zz  p(1  ), (18.1)
b3

in which b  z 2  a 2 , and a is the radius of the circular area. The fourth column is the sum
of the second and third columns. This is considered as the initial stress, before the application
of the load, but after the construction of the foundation plate. The fifth column gives the
stresses caused by the load, the weight of the fluid in the reservoir. These stresses have also
been calculated by equation (18.1). The sixth column is the sum of the fourth and the fifth

79
column. These are the final effective stresses. The seventh and eight columns contain the
actual computation of the deformations of each layer, using Terzaghi’s logarithmic formula,
and the value C10  50 . By adding the deformations of the layers the total settlement of the
reservoir is obtained, which is found to be 0.10 m. That is quite large, and it may mean that
the construction of the reservoir on such a soft soil is not feasible.
The procedure described above can easily be extended. It is, for instance, simple to account
for different properties in each layer, by using a variable compressibility. The method is also
not restricted to circular loads. The method can easily be combined with Newmark’s method
to calculate the stresses below a load of arbitrary magnitude on an area of arbitrary shape. It
is also possible to incorporate creep by adding the formula of Koppejan in the calculation.
The method can also be elaborated with little difficulty to a computer program. Such a
program may use a numerical form of Newmark’s method to determine the stresses, and then
calculate the settlements of the loaded area by the method illustrated above. The formula to
compute the deformation of each layer may be Terzaghi’s formula, but it may also include a
time dependent term, to account for creep and consolidation.

80
19 Bearing capacity of a strip footing
The biggest problem for a shallow foundation, just as any other type of foundation, is a
failure due to an overestimation of the bearing capacity. This means the correct prediction of
the bearing capacity of the shallow foundation is often the most important part of the design
of a civil structure.

Figure ‎19-1. Overloaded shallow foundation of a row of silos I.

Figure ‎19-2. Overloaded shallow foundation of a row of silos II.

One of the simplest problems for which lower limits and upper limits can be determined is
the case of an infinitely long strip load on a layer of homogeneous cohesive material (   0 ),
see Figure 19-3.

81
Figure ‎19-3. Strip footing.
The weight of the material will be disregarded, at least in this chapter. This means that it is
assumed that   0 . The problem is a first schematisation of the shallow foundation of a
structure, using a long strip foundation, made of concrete, for instance.
It will first be attempted to obtain a lower bound for the failure load, using an equilibrium
system. Such a system should consist of a field of stresses that satisfies the conditions
of equilibrium in all points of the field, that agrees with the given stress distribution on the
soil surface, and that does not violate the yield condition in any point.

19.1 Lower bound


An elementary solution of the conditions of equilibrium in a certain region is that the stresses
in that region are constant, because then all
conditions are indeed satisfied. In a two-dimensional field these equilibrium conditions are,
in the absence of gravity,
 xx  zx
  0, (19.1)
x z
 xz  zz
  0, (19.2)
x z
 xz   zx . (19.3)
The main difficulty is to satisfy the boundary condition, because the normal stress  zz is
discontinuous along the surface, see Figure 19-3. This difficulty can be surmounted by noting
that in a statically admissible field of stresses (an equilibrium system), not all stresses need be
continuous.
Formally this can be recognized by inspection of the equations of equilibrium, eqs. (19.1) –
(19.3). All partial derivatives in these equations must exist, which means that the stresses
must at least be continuous in the directions in which they have to be differentiated. It
follows that the shear stress  xz must be continuous in both directions, that the normal stress
 xx must be continuous in x-direction, and the normal stress  zz must be continuous in z-
direction. However, two of the partial derivatives,  xx z and  zz x , do not appear in
the equations of equilibrium, and therefore no conditions have to be imposed on the
continuity of these two normal stresses in these directions. This means that  xx may be
discontinuous in z-direction, and that  zz may be discontinuous in x-direction. Such a
discontinuity is shown, for the vertical direction, in Figure 19-4.

82
Figure ‎19-4. Stress discontinuity.
This figure shows a small element, with all the stresses acting upon its boundaries. The
normal stress  xx must be continuous in x-direction, because of equilibrium, as can most
easily be seen by letting the width of the element approach zero. Then the continuity of the
stress  xx can be seen as a consequence of Newton’s principle that the reaction must be equal
to the action. The normal stress  zz , however, may jump across the vertical line, without
disturbing equilibrium.
In Figure 19-4 the stress  zz is discontinuous in x-direction. The partial derivative  zz x
is infinitely large at the location of the vertical axis, but the element and all of its parts, are
perfectly well in equilibrium.
This property of equilibrium systems has been applied by Drucker, one of the originators of
the theory of plasticity, to construct equilibrium fields for practical problems. In this method
the field is subdivided into regions of simple form, in each of which the stress is constant, so
that the equations of equilibrium are automatically satisfied. The various subregions then are
connected by requiring that all the stresses transferred on the boundary surfaces are
continuous, allowing the normal stresses in the direction of these boundaries to be
discontinuous.

Figure ‎19-5.‎Drucker’s‎equilibrium‎system.

An example is shown in Figure 19-5, for the case of a strip footing. In a vertical strip below
the load the stresses are supposed to be  xx  2c ,  zz  4c , and  xz  0 . In the two regions
to the left and right of this strip the stresses are  xx  2c ,  zz  0 , and  xz  0 . On the two
vertical discontinuity lines only the vertical normal stress  zz is discontinuous. The other
stresses are continuous, as required by equilibrium. This field of stresses satisfies all the
conditions of equilibrium, and satisfies the boundary conditions on the upper surface. The

83
shear stress  zx  0 , and the normal stress  zz  0 if | x | a , and  zz  p if | x | a , where
2a is the width of the loaded strip. The stress distribution should also satisfy the condition
that the yield condition is never violated. This can be checked most conveniently by
considering the Mohr circles for this case, as shown in Figure 19-6.

Figure ‎19-6.‎Circles‎of‎Morh‎belonging‎to‎Drucker’s‎equilibrium‎system.

In order that all circles remain within the yield envelope the value of the load p should be
such that p  4c . The stress distribution satisfies all the conditions for a statically admissible
stress field, and it can be concluded that p  4c is a lower bound for the failure load. If the
true failure load is denoted by pc, it now has been shown that
pc  4c. (19.4)
It is possible that by considering more than two discontinuity lines slightly higher lower
bounds can be found. This will not be investigated here, however.

19.2 Upper bound


An upper bound for the failure load can be obtained by considering the mechanism shown in
Figure 19-7.

Figure ‎19-7. Mechanism 1.

This mechanism consists of a displacement field in which half a circle, of radius a, rotates
over a small angle, without internal deformations. This half circle slides along the remaining
part of the body. The displacement field is compatible, and satisfies the boundary conditions
on the displacements (that is very simple: these are only present on the circular slip surface).
The load corresponding to this deformation can be determined by examining the moment

84
equilibrium. If the circle rotates, a shear stress occurs at the periphery. If the shear stress is
assumed to be maximal, so   c , the moment with respect to the axis of rotation of the
internal friction stresses at the periphery of the circle equals
 ca 2 ,
because the length of the circular arc is  a . The eccentricity of the external load is 1
2 a , so
the exerted moment becomes
1
2 pa 2 .
Equating these two moments gives
p  2 c.
This is an upper bound for the failure load pc,
pc  6.28c. (19.5)
A somewhat lower upper bound can be found by choosing the centre of the circle somewhat
higher, see Figure 19-8.

Figure ‎19-8. Mechanism 2.

If the angle at the top is 2 , it follows


2cR2  12 pa 2 ,
and because a  R sin  , in which R is the radius of the circle and a the width of the load,
4c
p .
sin 2 
For   12  the previous upper bound is recovered. The smallest value is obtained for  =
1.165562, or  = 66.78. The centre of the circle then is located at a height 0.429a. The
corresponding value of p is 5.52c. This is an upper bound, hence
pc  5.52c. (19.6)
It can be concluded at this stage that it has been shown that
4c  pc  5.52c. (19.7)
In the next chapter the failure load will be approximated even closer.

85
20 Prandtl & Nc
The first step in solving the bearing capacity of a strip of infinite length, on weightless soil,
was made by Ludwig Prandtl. He published in 1920 an analytical solution for the problem of
a strip load on a half plane, see Figure 20-1 and Figure 20-2, on the basis of the assumption
that in a certain region at the soil surface, the stresses satisfy the equilibrium conditions and
the Mohr-Coulomb failure criterion. In this entire region the soil then is on the verge of
yielding.

Figure ‎20-1.‎Prandtl’s‎schematisation (original drawing).

This solution is both statically admissible and kinematically admissible, and must therefore
give the true failure load. The lower bound part of Prandtl’s solution, with an equilibrium
system of stresses, will be presented in this chapter. The proof that this solution is also
kinematically admissible, which is much more difficult, will be omitted here. A complete
proof can be found in textbooks on the theory of plasticity.
The analytical solution gives the maximum foundation pressure or bearing capacity of the
soil under a limit pressure, p, causing kinematic failure of the weightless infinite half-space
underneath. The strength of the half-space is given by the angle of internal friction,  and the
cohesion, c. The sliding soil part is subdivided into three zones:

1. Zone 1: A triangular zone below the strip load with a width B  2  b1 . Since there is
no friction on the ground surface, the directions of the principal stresses are horizontal
and vertical; the largest principal stress is in the vertical direction.
2. Zone 2: A wedge (Prandtl’s wedge) with the shape of a logarithmic spiral, where the
principal stresses rotate through 90 from Zone 1 to Zone 3. The pitch of the sliding
surface equals the angle of internal friction  , creating a smooth transition between
Zone 1 and Zone 3 and also creating a zero frictional moment on this wedge (see
Equation 13).
3. Zone 3: A triangular zone adjacent to the strip load. Since there is no friction on the
surface of the ground, the directions of principal stress are horizontal and vertical with
the vertical component having the smallest amplitude.

86
Figure ‎20-2. Strip foundation with failure mechanism.
The interesting part of the solution is that all three zones are fully failing internally according
to the Mohr-Coulomb failure criterion, while the outer surfaces are simultaneously fully
sliding according to the Coulomb failure criterion. Only the latter criterion exists in the case
of a Bishop slope stability calculation.
The Prandtl-wedge failure mechanism can be confirmed by laboratory tests, see Figure 20-3.

Figure ‎20-3. Prandtl-wedge failure planes in sand.

Prandtl published an analytical solution for the bearing capacity of this three-zone problem
considering the soil to be weightless (   0 ) and without a surcharge (q = 0 ):
p  cNc , (20.1)
in which:
1  sin 
Nc  ( Nq  1) cot  , with Nq  K p exp( tan  ) and K p  . (20.2)
1  sin 

In the case of a frictionless soil (   0 ), this becomes:


p  (  2)c  5.14c. (20.3)
If we check the solution of Prandtl of the cohesion bearing capacity factor Nc with finite
element calculations, then we see that for lower friction angles (   25 ) this solution is
rather accurate, see Figure 20-4, but for higher friction angles, another failure mechanism
start to appear, see Figure 20-5. Therefore the solution of Prandtl is a bit unsafe. FEM
calculations made by the author, both displacement and stress controlled, show it would be
more correct to apply a cohesion bearing capacity factor of, see Figure 20-6:
Nc  ( Nq  1) cot  , with Nq  cos   K p  exp( tan  ) . (20.4)

87
Figure ‎20-4. FEM calculation of failure mechanism; left: low friction angle; right: high.

Figure ‎20-5. FEM calculation of failure mechanism; left: low friction angle; right: high.

Figure ‎20-6. Cohesion bearing capacity factor: Analytical versus FEM.

88
21 Reissner & Nq
An important problem in foundation engineering is the computation of the maximum or
ultimate load (the bearing capacity) of a very long foundation, at a certain depth below the
soil surface. The influence of the depth of the foundation is accounted for by considering a
surcharge at the foundation level, to the left and the right of the applied load.
For this reason, the solution of Prandtl was extended by Reissner in 1924 with a surrounding
surcharge, q. The analytical solution for the bearing capacity of this three-zone problem by
Prandtl with the extension of Reissner can be written as:
p  cNc  qNq , (21.1)
where the coefficients Nc en Nq are dimensionless constants, for which Prandtl obtained the
following expression,
Nc  ( Nq  1) cot  , (21.2)
and Reissner:
1  sin 
Nq  K p exp( tan  ), with K p  . (21.3)
1  sin 
If we check the solution of Reissner of the surcharge bearing capacity factor Nq with Finite
Element Model calculations, then we see that also in this case, for lower friction angles (
  25 ) this solution is rather accurate, but for higher friction angles other failure
mechanisms start to appear. Therefore the solution of Reissner is a bit unsafe. Finite element
calculations made by the author, both displacement and stress controlled, show it would be
more correct to apply a cohesion bearing capacity factor of:
1  sin 
Nq  cos2 K p exp( tan  ), with K p  . (21.4)
1  sin 

Figure ‎21-1. Surcharge bearing capacity factor: Analytical versus FEM.

89
22 Meyerhof & N
In this chapter the case of a strip footing on cohesive material, considered in chapters 19
and 20, is extended to a general type of shallow foundation, on a soil characterised by its
cohesion c, friction angle  and volumetric weight  . The soil is assumed to be completely
homogeneous. Although the formulas were originally intended to be applied to foundation
strips of buildings, at a shallow depth below the soil surface, they are also applied to large
caisson foundations used in offshore engineering for the foundation of huge oil production
platforms.

In 1940, equation (21.1) of Prandtl and Reissner has been extended by Keverling Buisman
for the soil weight,  . It was Von Terzaghi (1943) who first wrote this extension as:
p  cNc  qNq  12  BN , (22.1)
where B is the total width of the loaded strip, and  is the volumetric weight of the soil. For
the coefficient N various suggestions have been made by Keverling Buisman (1940), Von
Terzaghi (1943), Meyerhof (1951; 1953; 1963), Caquot and Kérisel (1953), Brinch Hansen
(1970), Vesic (1973) and Chen (1975), on the basis of theoretical analysis or experimental
evidence.
The equation by Brinch Hansen, for the soil weight bearing capacity coefficient, was based
on calculations of Lundgren-Mortensen and also of Odgaard and Christensen. The Chen
equation for the soil weight-bearing capacity coefficient became the currently used equation:
N  2( Nq  1) tan . (22.2)
In 2015, Van Baars showed, based on finite element calculations, that also this solution is
unsafe and proposed a more accurate and safer solution:
N   tan   e tan  (22.3)
Later the formula (22.1) has been further extended with various correction coefficients, in
order to take into account the shape of the loaded area, the inclination of the load, a possible
inclined soil surface, and a possible inclined loading area. Most of these effects were
assembled into a single formula for the vertical bearing capacity by Meyerhof (1963) and
later Jørgen Brinch Hansen (1970),
pv  ic sc cNc  iq sq qNq  i s 12  BN . (22.4)
In this equation the coefficients ic and iq are correction factors for a possible inclination of the
load (inclination factors), and sc and sq are correction factors for the shape of the loaded area
(shape factors). Some other factors may be used (for a sloping soil surface, or a sloping
foundation foot), but these are not considered in this book.

90
23 Table of the bearing capacity factors
In Table 23-1 the values of Nc, Nq and N are given, based on the finite element calculations,
(eqs: 20.4, 21.4 and 22.3), as a function of the friction angle  . In the limiting case   0 the
value of Nc  2   , as found in Chapter 20. If c = 0 and   0 the bearing capacity must be
equal to the surcharge, i.e. p  q . Even a layer of mud can support a certain load, provided
that it is the same all over its surface. This is expressed by the value N q  1 for   0 .

 Nc Nq N  Nc Nq N
0 5.142 1.000 0.000 20 12.778 5.651 3.588
1 5.360 1.094 0.058 21 13.449 6.163 4.028
2 5.590 1.195 0.122 22 14.166 6.724 4.517
3 5.831 1.306 0.194 23 14.933 7.339 5.060
4 6.085 1.426 0.274 24 15.755 8.015 5.665
5 6.353 1.556 0.362 25 16.637 8.758 6.339
6 6.634 1.697 0.459 26 17.584 9.576 7.092
7 6.931 1.851 0.567 27 18.603 10.479 7.934
8 7.244 2.018 0.687 28 19.702 11.476 8.877
9 7.574 2.200 0.818 29 20.888 12.578 9.935
10 7.922 2.397 0.964 30 22.172 13.801 11.125
11 8.291 2.612 1.125 31 23.563 15.158 12.466
12 8.680 2.845 1.302 32 25.075 16.668 13.980
13 9.092 3.099 1.498 33 26.720 18.352 15.693
14 9.528 3.376 1.714 34 28.516 20.234 17.637
15 9.991 3.677 1.953 35 30.480 22.342 19.848
16 10.482 4.006 2.218 36 32.633 24.709 22.371
17 11.004 4.364 2.510 37 35.001 27.375 25.258
18 11.558 4.756 2.833 38 37.612 30.386 28.571
19 12.149 5.183 3.191 39 40.499 33.796 32.387
20 12.778 5.651 3.588 40 43.703 37.671 36.797

Table ‎23-1: Commonly used bearing capacity factors (Prandtl, Reissner and Chen).

The Eurocode does not recommend which equations have to be used for the calculation of the
bearing capacity; in fact this can be found in the country annex, which means that every
country selects its own equations. The problem of these annexes is that no explanation is
given as to why a certain equation has been selected; even references to authors or
publications are never given. Besides this, equations other than the equations of Meyerhof
and Brinch Hansen are used and sometimes the inclination and shape factors of Brinch
Hansen are often used, although they are incorrect.

91
24 Correction factors

24.1 Inclination factors i


In the case of an inclined load, i.e. loading by a vertical force and a horizontal load at the
same time, the horizontal component of the load is limited at the foundation surface, due to
the Coulomb shear failure,
ph  c  pv tan . (24.1)
But also the vertical bearing capacity is considerably reduced by the additional horizontal
load.
In 1953 Meyerhof published his results of laboratory experiments on inclined loading on
“purely cohesion materials “ and “cohesionless materials”, for cases in which the horizontal
component of the load is smaller than its maximum possible value. The correction factors for
the inclination of the load were in 1963 expressed by him as:
2
     
2

iq  ic  1   , i  1   , for:   . (24.2)
 90    
In 1970 Brinch Hansen proposed other inclination factors:
phor
ic  1  , iq  ic2 , i  ic3 . (24.3)
c  pvert tan 
Although these inclination factors are often used, they are however a disallowed mixture of
the Coulomb failure criterion, which should only be applied at the interface, and the Mohr-
Coulomb bearing capacity failure of the half-space below the interface. A clear indication of
the incorrectness of his solution is the fact that the surcharge inclination factor, iq , depends
here on the cohesion, c , while the factor N q for any inclination, and therefore also iq ,
should not depend on the cohesion, c . The same even applies for the cohesion inclination
factor, ic . This indicates that these inclination factors should not be used.

Based on the Prandtl wedge, one can assume that there is a failure mechanism for inclined
loads as in Figure 24-1.

p
q
B
b1 b1 1 b3 b3
3 3
1
 1  1
h1 rR h3
rL
1 1
r3
2

Figure ‎24-1. Failure mechanism for an inclined load.

92
By using this failure mechanism, Van Baars (2014, 2015) found analytical solutions for the
inclination factors. By combining these equations with the equation of the Coulomb shear
failure ( ph  c  pv tan  , Eq. 22.8), an analytical solution is found for ic and iq , which is
rather close to the empirical solution of Meyerhof based on his laboratory tests, but also
rather close to the results of finite element calculations. These calculations also indicate that
the soil-weight inclination factor i of Meyerhof must be corrected. This finally results in:

     
2 5

iq  ic  1   , i  1   , for:   . (24.4)
 90   90 

24.2 Shape factors s


If the shape of the foundation area is not an infinitely long strip, but a rectangular area, of
width B and length L (where it is assumed, for definiteness, that the width is the shortest
dimension, i,e, L  B ), correction factors for the shape are used. Meyerhof (1963) proposed
the following shape factors:
B B 1  sin 
sq  s  1  0.1K p sin  , sc  1  0.2 K p with: K p  (24.5)
L L 1  sin 
Later Brinch Hansen (1970) proposed other shape factors:
B B B
sc  1  0.2 , sq  1  sin  , s  1  0.3 . (24.6)
L L L
In 2015 Van Baars proved that the solutions of both Meyerhof and Brinch Hansen are unsafe
and proposed the following solutions based on finite element calculations:

B B B
sc  1  0.12 , sq  1  0.55 , s  1  0.55 . (24.7)
L L L

There is still no international agreement on the precise values of these correction factors
either. It may be noted that for B L  0 , the formulas all give a factor 1, in agreement with
the basic results for an infinite strip. It should also be remembered that B L  1 , by
definition.

Figure ‎24-2. Rectangular area.

Some justification for values of the coefficients sc and sq larger than 1 is that when loading a
rectangular plate, deformations and failure will also take place in the third dimension, so that
a larger slip surface contributes to the bearing capacity of the plate.

93
Figure ‎24-3. Eccentric load.

In case of an eccentric resultant force of the load, the width B and the length L may be
reduced such that the resulting force is applied in the centre of the reduced area, see
Figure 24-3. Part of the foundation plate then does not contribute to the bearing capacity, at
least for this loading case. It may, of course, give a contribution to the bearing capacity of
other loading cases.
As mentioned before, there is no general agreement about the values of many of the
correction factors, because the results obtained by researchers in different countries, from
theoretical or experimental studies, appear to give different results. Great care is needed
when using data from literature. When a certain value has been obtained by one single
researcher, and deviates from the results of many others, that value may well be in error.
It is also very inconvenient that there is sometimes no agreement about the basic formula
(22.1). In some older (German) publications the factor 12 is omitted. Then the values of N
are (approximately) half as large, so that the final result is the same, but it may give rise to
some confusion when using a formula from one publication, and taking the coefficients from
another publication. In this book Von Terzaghi’s original formula has been used, as is
common practice internationally.
The formulas presented in this chapter have originally been derived for foundations on land,
with relatively modest dimensions, say a few square metres. The third term in Meyerhof’s
formula (22.4) then is rather small, because of the small value of B, and it is often omitted. In
offshore engineering the development of gravity foundations has meant that production
platforms may be founded on huge concrete caissons that are placed on the sea bed, in deep
water. The surface area may be up to 80 m  80 m . For the design of such structures the
bearing capacity of the foundation is of great importance, and then the third term in the
formulas (22.1) or (22.4), which describes the influence of the unit weight of the soil (i.e. the
gravity term), is the most important term of all, giving the major contribution to the bearing
capacity, especially in sand. This is why considerable attention has been paid to a more
accurate definition of this term.
It must be emphasised that all the considerations of this chapter are restricted to dry soils, in
which there is no difference between effective stresses and total stresses. For saturated soils
the formulas should be expressed in terms of effective stresses. Usually this can be
accomplished simply by replacing the volumetric weight  by the effective volumetric
weight  s   w . This is a simple, but very fundamental adjustment.

94
25 CPT, undrained shear strength and Prandtl
In the Netherlands the cone penetration test is mainly used as a model test for pile
foundations. In the Western parts of the Netherlands the soil usually consists of 10 – 20
metres of very soft soil layers (clay and peat), over a rather stiff sand layer. This soil structure
is very well suited for wooden or concrete piles of about 20 cm – 45 cm diameter, reaching
just into the sand. The weight of the soft soil acts as a surcharge on the sand, which has a
considerable cone resistance. The allowable stress on the sand depends upon its friction angle
 , its cohesion c (usually very small, or zero), and the surcharge q. The dimensions of the
foundation pile have very little influence, because this parameter appears only in the third
term of Meyerhof’s formula, which is a small term if the width is less than, say, 1 metre. This
means that the maximum pressure for a large pile and the thin pile of a cone penetrometer
will be practically the same, so that the allowable pressure on a pile can be determined by
simply measuring the cone resistance. This will be elaborated in Chapter 26.
The cone penetration test can also be used to determine physical parameters of the soil,
especially the shear strength. It can be postulated, for instance, that in clays the cone
resistance will be determined mainly by the undrained shear strength of the soil (su). In
agreement with the analysis of Meyerhof the relation will be of the form
qc   v  Nc su , (25.1)
where  v is the local vertical stress caused by the surcharge, and Nc is a dimensionless
factor. For a circular cone in a cohesive material a cone factor Nc of the order of magnitude
15 – 18 is usually assumed, on the basis of plasticity calculations for the insertion of a cone
into a cohesive material of infinite extent. By measuring the cone resistance qc the undrained
shear strength su can be determined. The results are not very accurate, because of theoretical
shortcomings and practical difficulties, but the measurement has the great advantage of being
done in situ, in the least disturbed soil. The alternative would be taking a sample, bringing it
to a laboratory, and then doing a laboratory test. This process includes many possible sources
of disturbance, which can be avoided by doing a test in situ.

95
96
IV Pile foundations

97
26 Cone Penetration Test (CPT)
The cone penetration test (CPT) is a testing method used to determine the geotechnical
engineering properties of soils and the soil stratigraphy. It was initially developed in the
1950s at the Dutch Laboratory for Soil Mechanics in Delft (now Deltares) to investigate soft
soils. Based on this history it has also been called the "Dutch cone test". Today, the CPT is
one of the most used and accepted in soil methods for soil investigation worldwide.

The test is used to determine:


 soil types,
 (stratigraphic) soil layering,
 soil stiffness and strength (empirical)
 bearing capacity of pile foundations
 permeability and consolidation parameters,
 groundwater pressure
 environmental data (pollution)

The test method consists of pushing an instrumented cone, with the tip facing down, into the
ground at a controlled rate (between 1.5 - 2.5 cm/s). The resolution of the CPT in delineating
stratigraphic layers is related to the size of the cone tip, with typical cone tips having a cross-
sectional area of either 10 or 15 cm², corresponding to diameters of 3.6 and 4.4 cm.
At least two things are constantly measured:
1. the tip resistance /cone resistance /cone pressure qc [MPa], and
2. the shaft friction or sleeve resistance s [MPa]

Also other items are measured, for example the water pressure, the inclination of the conus,
or chemical products like hydrocarbons or aromatics. One parameter is calculated from the
first two, which is the friction ratio Rf:
s
Rf  100% (26.1)
qc

Figure ‎26-1. Left:CPT truck (Fugro). Right: CPT cone (Geomil).

98
The friction number is quite helpful in determining the soil type, see Table 26-1, but a check
with a few additional borings remains very important.

Soil Type Friction Rf cone pressure qc


Sand, coarse 0.4%
Sand, medium 0.6% 5-30 MPa
Sand, fine 0.8%
Sand, silty 1.1%
Sand, clayey 1.4% 5-10 MPa
Sandy clay or loam 1.8%
Silt 2.2%
Clay, silty 2.5%
Clay 3.3% 0.5-2 MPa
Clay, organic (peaty) 5.0%
Peat 8.0% 0-1 MPa

Table ‎26-1: Friction number and cone resistance for different soil types.

For example in the sounding (CPT) of Rotterdam in Figure 26-3, the soil types and layering
are very clear. There are very soft layers until MSL -17.5 m: a clear clay layer until -5.5 m;
peat until -11.5 m; clay until -16.0 m; clayey peat until -17.5 and first then soft sand.

Figure 26-2 shows the soil layering of the Hubertus Dune in Den Hague, the Netherlands.
This is used to plan a bore tunnel (see black lines under dune).

Figure ‎26-2. Soil layering of the Hubertus Dune in Den Hague.

99
Figure ‎26-3. CPT of Rotterdam, with a clear layering starting with clay, peat, clay, mixed
clay and peat and finally deep sand (Fugro).

100
One always has to be careful with the interpretation of a CPT. The following CPT for
example is made in Belgium and shows a CPT of sand containing glauconite. The sand of the
formation of Berchem (between -17 m and -25 m) in the CPT does not follow the general
idea that sands have a friction number of 0.5 to 1.0 %. The qc value is typical around 15 MPa
and the friction number Rf = 4 to 6%.

Glauconite is an iron potassium phyllosilicate (mica group) mineral with a characteristic


green colour and has a very low weathering resistance and is very weak. It crystallises with a
monoclinic geometry. Its name is derived from the Greek glaucos (γλαυκος) meaning
'gleaming' or 'silvery', to describe the appearance of the blue-green colour. It is normally
found in dark green rounded pellets with the dimension of a sand grain size.

Figure ‎26-4. CPT of sand containing glauconite.


Picture from Geotechniek April 2012

101
27 Compression piles
In many river delta areas, for example the Western part of the Netherlands, the soil consists
of layers of soft soil (clay and peat), of 10  20 metre thickness, on a stiff sand layer, of
pleistocene origin. The bearing capacity of this sand layer is derived for a large part from its
deep location, with the soft layers acting as a surcharge. The properties of the sand itself, a
relatively high density, and a high friction angle, also help to give this sand layer a good
bearing capacity, of course. The system of soft soils and a deeper stiff sand layer is very
suitable for a pile foundation. In this chapter a number of important soil mechanics aspects of
such pile foundations is briefly discussed.

27.1 Pile types


Piles can be made from wood, steel, (partially) reinforced concrete or prestressed concrete.
The installation can be done by driving, vibrating, boring or pressing, but not all installation
techniques can be applied to the different pile types.

The following distinction can be made between the piles, but it is a strong simplification:
 Wooden piles (driven or pressed)
 Steel piles (driven, vibrated or bored)
 Precast concrete piles (driven)
 In situ concrete, piles (driven with steel pipe)
 In situ concrete, piles (bored with steel pipe support)
 In situ concrete, piles (bored with slurry support)

Wooden piles
Wooden piles have existed for a long time, probably more than thousand years. In the 15th
century piles were used a lot in Amsterdam and also in Venice there is similar old tradition of
using wooden piles. The palace on the dam in Amsterdam was built in 1665 on 13.659
wooden piles. In the Netherlands there are still 25 million wooden piles in use.

Figure ‎27-1. Left: Wooden pile driving in Amsterdam in the 15th century.
Right: Palace on the dam (Amsterdam).

Wooden piles are cheap, light, easy to transport, easy to cut off, but have two disadvantages:
they have a limited bearing capacity and they will rot above the groundwater table and

102
therefore need a concrete extension, or something similar and should never face a sinking
groundwater (not even a leak draining sewage system nearby).

Figure ‎27-2. Wooden piles with concrete extension against rotting.

Steel piles
Steel piles come in many different profiles: H-profiles, pipes/tubes or other profiles. They
exist in all lengths, widths and wall thicknesses. If there is no closed foot of a tube, and the
soil conditions are correct, the piles can be vibrated into the ground, otherwise they will be
driven. For such an open foot, one has to check if there is soil plugging, if not, tubes can be
filled afterwards with concrete to make them stronger and to increase tip resistance.

Figure ‎27-3. Installation of steel piles.

Pile driving with a vibratory pile hammer uses a system of counter-rotating eccentric weights,
powered by hydraulic motors, and designed in such a way that horizontal vibrations cancel
out, while vertical vibrations are transmitted into the pile. The pile driving machine is lifted
and positioned over the pile by means of an excavator or crane, and is fastened to the pile by
a clamp and/or bolts. Vibratory hammers can either drive in or extract a pile; extraction is
commonly used to recover steel "H" piles used in temporary foundation shoring.
Even very large piles can be vibrated into the ground, see the following pictures.

103
Figure ‎27-4. Vibration of steel piles.

Pile driving with a pile driver includes a heavy weight placed between guides so that it is able
to freely slide up and down in a single line. It is placed upon a pile. The weight is raised,
which may involve the use of hydraulics, steam, diesel, or in the past even manual labour.
When the weight reaches its highest point it is then released and smashes on to the pile in
order to drive it into the ground. Steel piles are often driven from the inside, directly on the
foot, in order to protect the shaft of the steel pile.

Figure ‎27-5. Internal driving of steel piles


(Picture‎from‎P.‎van‎‘t‎Wout, Geotechniek, April 2012).

The advantages of steel piles are: they are very strong and therefore perfect for high
horizontal and vertical loads and can be used in heavy ground for pile driving. The
disadvantages are: expensive and rusting (corrosion). Therefore these piles are used when it
is not possible for other pile types to be used.

104
Precast concrete piles
Precast concrete piles are the most commonly used. They can be bought in many sizes up to
0,5 m wide and 40 m long. They are mostly driven into the ground, even for rather stiff sands
(cone resistance up to qc = 65 MPa). Precast piles are better in quality than in-situ piles. Other
advantages are that they are relatively cheap, strong and available in all sizes. There are few
disadvantages. Only for very high loads, steel piles have to be used and in case of problems
with driving vibrations or hard soil layers, in-situ (bored) concrete piles should be used.

Figure ‎27-6. Precast concrete piles.

 [mm] Max. L [m]  [mm] Max. L [m]


180 16 380 31
220 20 400 32
250 22.5 420 33
290 26 450 34
320 28 500 35
350 30

Table ‎27-1: Normal sizes of pre-tensioned precast concrete piles.

In situ concrete driven piles

One can drive a steel tube/casing with a loose steel plate at the tip, into the ground, fill it with
concrete, lift the tube and lose the plate. This is the way a concrete in-situ pile is made. The
advantage is that the pile is still driven, so the soil is compacted and there is a higher bearing
capacity than bored piles. Also no transport of long piles into urban areas are needed,

105
Disadvantages are: the pile is of a poorer quality than precast piles and the driving still causes
vibrations.
The sizes of these piles depend on the company but are mostly between 300 and 700 mm.
The maximum length can be more than 40 m.

Figure ‎27-7. Driven in-situ concrete piles.

Figure ‎27-8. Making of driven in-situ concrete piles.

In situations where one wants to combine the advantage of a precast concrete pile core with
the high bearing capacity of the in-situ driven piles, the so-called vibro-combination-piles are
used. The installation and construction of these piles are identical to the in-situ driven piles,
however before the lifting of the steel tube/casing, a precast pile core is hung into the tube.
The space between core and tube is filled with liquid concrete. This method is often chosen
to avoid any risk on corrosion of the steel reinforcement bars.

In situ concrete bored piles in steel casing

Another way of in-situ pile making is boring. A steel casing supports the hole. The
advantages are: no vibrations and also possible in complex grounds with stones are hard
soils. The disadvantage is the lower bearing capacity due to the looser ground by boring.

106
Figure ‎27-9. Making of bored in-situ concrete piles.

In situ concrete, piles, with slurry support


There are also bore techniques without a steel casing supporting the ground. Mostly a
bentonite slurry is used to create support pressure.

Figure ‎27-10. Bore for making bored in-situ concrete piles.

A big disadvantage of bored piles is the relaxation of the ground due to boring of the holes
for the in-situ bored piles. This reduces the bearing capacity of these piles.

Figure ‎27-11. Relaxation of the ground due to bored piles.

107
Other piles
There are many other pile types, for example the Franki pile, in which, after the installation
of the pile, the plastic concrete plug at the tip is driven out of the pile to create a much large
pile tip.

Figure ‎27-12. Franki pile.

Or the fundex pile, a drilled pile, which can be installed in small and low height areas, perfect
for improving a foundation afterwards, but not as good as a driven pile and above all very
expensive because it is complex and one loses an expensive steel drill as pile tip.

Figure ‎27-13. Fundex pile.

108
27.2 Bearing capacity
A pile may also have a bearing capacity due to friction along the length of the pile. This is
very important for piles in sand layers. In applications in very soft soil (clay layers), the
contribution of friction is generally very unreliable, because the soil may be subject to
settlements, whereas the pile may be rigid, if it has been installed into a deep sand layer. It
may even happen that the subsiding soil exerts a downward friction force on the pile,
negative skin friction, which reduces the effective bearing capacity of the pile. Friction is of
course very important for tension piles, for which it is the only contributing mechanism.
The maximum value of the skin friction can be determined very well using a friction cone,
which is a penetration test in which the sleeve friction is also measured. The local values are
often very small, however, so that the measured data are not very accurate. For sandy soils
the friction therefore is often correlated to the cone resistance.

The ultimate bearing capacity of a compression pile is determined by both the tip bearing
capacity and the shaft bearing capacity:
Fr ;max  Fr ;max:tip  Fr ;max:schaft (27.1)
and:
Fr ;max:tip  Atip pr ;max;tip (27.2)
L
Fr ;max:shaft  Op;avg  p (27.3)
o r ;max; shaft

in which:
Fr;max = maximum bearing force
Fr;max;tip = maximum tip resistance force
Fr;max;shaft = maximum shaft friction force
Atip = surface area of the tip of the pile
pr;max;tip = maximum tip resistance according to the sounding ( ≈  p qc )
Op;avg = circumference of the pile shaft
L = length of the pile
pr;max;shaft = maximum pile shaft friction according to the sounding ( ≈  s qc )

Notes
 In case prefab piles with enlarged feet are used, L may not be larger than the height of the
enlarged foot H (see Figure 27-18).
 For soil (non-rock) the value of the tip resistance is maximised to 15 MPa (see Figure 27-19).
 For open steel piles, which are not filled with concrete afterwards, first check if the pile tips are


plugging with the soil inside: Fr ;max:tip  min Atip pr ;max;tip ; Op;avg 
o
Lplug
p
r ;max; shaft

 One must not forget to reduce the net bearing capacity of the pile by the value of negative shaft
friction.
 For  s see Table 27-3 in Chapter 27.6. It should be roughly the same as the friction ratio Rf:

109
27.3 CPT, SPT and HDP
In areas with soft soils, mostly qc-methods based on the Cone Penetration Tests (CPT) are
used (for example method Koppejan) for predicting the maximum tip resistance, because this
method measures the tip resistance and shaft friction continuously, and is therefore the most
accurate method. If there are stones in the ground or the ground is too hard, the CPT cannot
be used. In these cases the Menard test can be used to measure the maximum ground
resistance (France) (See chapter 34.5). Sometimes the Heavy Dynamic Probing is used
(Germany) or the Standard Penetration Test (USA and Japan), but these are less accurate.
Another way is to calculate the bearing capacity of the pile with the Meyerhof equation. This
is done when there are no CPT’s, SPT’s or HDP-tests available, but this equation leads to a
maximum tip resistance which is far lower (but therefore safe), since this equation neglects
the strength of the ground above the pile tip.

In Japan the norm for the tip resistance is based on the blow count N from the SPT, and not
on the cone resistances qc of a CPT. For comparison between Japan, Europe and the USA for
cases without a CPT, see the table below.

Ground Resistance Japan Europe USA


(JPTS) (EC7) (USACE)
Sandy Toe [kPa] 300N Nq ’ v0 Nq ’ v0
Shaft [kPa] 2N Ks ’ v0 tan Ks ’ v0 tan
Clayey Toe [kPa] 6 cu 9 cu + ’ v0 9 cu
Shaft [kPa] cu  cu  cu

Table ‎27-1 Comparison of static bearing capacity, Kikuchi 2012

In Japan the friction angle for calculating the coefficient of bearing capacity Nq (see
chapter 21) may be determined from the blow count N:
100 N
  25  3.2 (27.4)
p 'v 0  70

in which p 'v 0 is the effective overburden pressure at sounding.

In Germany the cone resistance qc may be estimated from the blow count of the heavy
dynamic penetration test (Heavy Dynamic Probing) ( qc  n10 cm ), but this is rather inaccurate.

110
27.4 Tip resistance using Meyerhof (Prandtl, Terzaghi or Brinch Hansen)
For the determination of the bearing capacity of a foundation pile it is possible to use a
theoretical analysis, on the basis of Meyerhof’s or Brinch Hansen’s general bearing capacity
formula. In this analysis the basic parameters are the shear strength of the sand layer
(characterized by its cohesion c and its friction angle  ), and the weight of the soft layers,
which can be taken into account as a surcharge q.
The maximum tip resistance can be determined analogously to the bearing capacity of a
shallow foundation, which is based on a slide plain according to Prandtl. This entails simply
using the 2-dimensional solution for 3-dimensional collapse (conservative) and simply
disregarding the shear strength, not the weight, of the soil above the foundation plane
(conservative).

 Bef N s is negligible relative to  v;z;o;d


The following example shows that the term 0.5 e;d  Nq sq .
Therefore the tip resistance can be calculated most simply according to:
 Nq sq
pr;max;tip   v;z;o;d

Figure ‎27-14. Slide plains under a pile according to Prandtl.

Example
Given is a pile that is driven 11 metres into the ground according to Figure 27-14. The pile is in
approximately 1 metre of sand; above the sand layer is 10 metres of clay. Groundwater level is 1
metre below ground level. The following data is given:
Sand:  = 40, c = 0, w,d = 20 kN/m3
Clay: w,d = 15 kN/m3
Pile dimensions: 0.4 x 0.4 m2
According to Prandtl, because the cohesion is zero the maximum pressure below the tip of the pile
 Nq sq iq  0.5 e;d
will be: pr;max;tip   v;z;o;d  Bef N s i
According to Table 23-1 in Chapter 22: Nq = 64 and N = 106
The pile is only subjected to axial loads, so: iq = i =1
The shape factors can be derived from (Bef = Lef):
sq = 1+sin 40 = 1.64 s = 1-0.3 = 0.7
The initial effective ground pressure is:
 v;z;o;d
  10 15  1 20  10 10  70 kN/m2
The maximum acceptable tip stress is:
pr;max;tip  70  64 1.64  0.5 10  0.4 106  0.7
 7.35 103  0.15 103  7.5 103 kN/m 2
 Bef N s <<  v;z;o;d
So: 0.5 e;d  N q sq

111
27.5 Tip resistance using a CPT method (Koppejan)
In engineering practice, a simpler, more practical and more reliable method has been
developed, on the basis of a cone penetration test, considering this as a model test. It would
be even better to perform a pile loading test, in which the pile is loaded, for instance by
concrete blocks on a steel frame, with a test load approaching its maximum bearing capacity.
This is very expensive, however, and the CPT is usually considered reliable enough.

Figure ‎27-15. CPT and pile.

In a homogeneous soil it can be assumed that under static conditions the failure load of a long
pile is independent, or practically independent of the diameter of the pile. This means that the
cone resistance measured in a CPT can be considered to be equal to the bearing capacity of
the pile tip (or pile base or pile foot). A possible theoretical foundation behind this statement
is that the failure is produced by shear deformations in a zone around the pile, the dimensions
of which are determined by the only dimension in the problem, the diameter of the pile. If the
pile diameter is taken twice as large, the dimensions of the failure zone around the pile will
also be twice as large. The total force (stress times area) is then four times as large, see
Figure 27-15. This is also in agreement with the theory behind Meyerhof’s formula, provided
that the third term (representing the weight of the soil below the foundation level, and the
width of the foundation) is small. This will be the case if the pile diameter is small compared
to its length.
In reality the soil around the pile tip usually is not perfectly homogeneous. Very often the soil
consists of layers having different properties. For this case, practical design formulas have
been developed, which take into account the different cone resistance below and above the
level of the pile tip. Moreover, in these design formulas the possibility that the failure mode
will prefer the weakest soil can be accounted for. In engineering practice of some countries
the Koppejan formula is often used. In this formula the resistance of the pile is assumed to
consist of three contributions,
p  12 [ 12 (qc;l  qc;2 )  qc;3 ]. (27.5)
In this equation is qc;l the smallest value of the cone resistance qc below the pile tip, up to a
depth of 4d, where d is the diameter of the pile, qc;2 is the average cone resistance in that
zone, and qc;3 is a representative low value of the cone resistance above the pile tip, in a zone
up to 8d above the pile tip. In this way a representative average value of the cone resistance
around the pile tip is obtained, in which engineering judgement is combined with experience.

In the Netherlands, this equation for the maximum tip resistance is written as:
pr;max;tip   s  p 1
2  q
1
2 c;l;avg  qc;ll;avg   qc;lll;avg  (27.6)

112
in which:

qc;l;avg = the average value of the cone resistances qc;z;corr’ along section I,
from the level of the tip of the pile to a level at least 0.7 times the equivalent
centre line (Deq) and at most 4 times Deq deeper. The bottom of section I must
be selected between these boundaries so that pr;max;tip is minimal;
qc;ll;avg = the average value of the cone resistances qc;z;corr’ along section II,
from the bottom of section I to the level of the tip of the pile; the value of the
cone resistance to be used in calculations may never exceed that of a lower
level;
qc;lll;avg = the average value of the cone resistances qc;z;corr’ along section III,
which runs up from the pile tip level to a level 8 times the equivalent centre
line (Deq) above; as for section II the value used for the cone resistance may
never exceed that of a lower level, starting with the value of cone resistance at
the end of section II.
In the case of auger piles, this section must start with a cone resistance equal
to or less than 2 NM/m2, unless the sounding took place within 1 metre of the
side of the pile after the pile was placed in the ground;
Deq = equivalent pile tip diameter:
Round pile: Deq  D
4
Square/rectangular pile: Deq  a b/a

a = width of pile, shortest side
b = width of pile, longest side (with a  b  1.5a )
p = the pile class factor according to Table 27-2
(these factors are according to the latest research all 30% too high!)
 = the factor that takes into account the influence of the shape of the foot
of the pile, determined using Figure 27-16 and Figure 27-17;
s = the factor that takes into account the influence of the shape of the
cross section of the foot of the pile, determined using Figure 27-18.

Note:
The equation above is based on a 3D extended
Prandtl-like failure mechanism as shown on
the right. This mechanism is for area III
(above the pile tip) impossible since during
failure the soil cannot reduce in width and
disappear near the pile. Therefore this failure
mechanism and corresponding calculation
method is incorrect. More research is needed
here, see for example:
White, D.J. and Bolton, M.D. (2005)
Comparing CPT and pile base resistance in
sand. Proc. Institution of Civil Engineers,
Geotechnical Engineering 158 (GE1), 3-14

113
pile class / type p
 soil displacing placement methods
- driven piles 1.0
- driven piles, formed in the soil 1.0
- screwed piles, formed in the ground 0.9
- prefabricated screwed piles 0.8
 piles with little soil displacement, such as 1.0
steel profiles and open steel tubes
 piles made with soil removal
- auger piles 0.8
- drilled piles 0.5
- pulsated piles 0.5

Table ‎27-2 Values of the pile class factor p

pile tip

Figure ‎27-17. Factor for the shape of the foot of the pile 
Figure ‎27-16. Shape of the
foot of the pile

Limit value is 15 MN/m2 for


all non-cohesive soil types

1. sand and gravel OCR ≤ 2


2. sand and gravel 2 < OCR ≤ 4
3. sand and gravel OCR > 4
Figure ‎27-18. Values of s
OCR = degree of overconsolidation

Figure ‎27-19. Value of the maximum tip


resistance in sand and gravel

114
27.6 Shaft resistance
The maximum pile shaft friction derived from the sounding is:
pr;max;shaft;z  s qc;z;a (27.7)
in which:
pr;max;shaft;z = maximum pile shaft friction at depth z;
s = factor according to Table 27-3 and Table 27-4 that takes into account
the influence of the method of realisation. Table 27-3 applies for fine
to coarse sand. An additional reduction factor of 0.75 applies for
extremely coarse sand.
For gravel the reduction factor is 0.5.
qc;z;a = cone resistance, whereby the peaks in the sounding diagram with
values equal to or greater than 15 MN/m2 are removed if these values
occur over sections of at least 1 metre and otherwise at values of 12
MN/m2.

Pile class / type s


 ground displacing placement methods:
- driven smooth prefab concrete pile and steel tube pile with closed tip 0.010
- pile made in the soil, whereby the concrete column presses directly 0.014
onto the ground and the tube was driven back out of the ground
- idem, but with a tube removed by vibration 0.012
- tapered wooden pile 0.012
- screwed piles
with grout injection or grout mix 0.009
without grout 0.006
 piles with little ground displacement
- steel profiles 0.0075
 piles made with ground removal
- auger piles 0.006
- bored piles 0.006
- pulsated piles 0.005

Table ‎27-3 Maximum values of s in sand and in sand containing gravel

ground type relative depth z/D s


clay/silt : qc  1MPa 5 < z/D < 20 0.025
clay /silt : qc  1MPa z/D  20 0.055
clay /silt : qc > 1MPa not applicable 0.035
peat not applicable 0

Table ‎27-4 Values of s for clay, silt or peat

Notes:
 The shaft friction factor is in fact the same as the friction value of a CPT-sounding test as
mentioned before. The only difference is that for s one does not use an average, but for safety’s
sake, a lower limit value.

115
 In the case of peat, one assumes no frictional strength because it is not certain that the shear stress
doesn’t relax to zero due to the creep of the peat.
 For open steel piles (tubes), one should use in sand layers a shaft friction factor on the inside of
the pile of s = 0.004 for vibrated piles and s = 0.006 for driven piles, BUT this should never
lead to a higher pile tip resistance than for a closed pile, because in this case the soil in the pile
will simply plug!

Maximum negative shaft friction


The effective bearing capacity of the pile is not only determined by the shaft bearing capacity
and the tip bearing capacity, but also by the negative shaft friction, which influences the
capacity unfavourably. The ground (of course only in the case of ground settlements) can
hang onto the pile. This causes a downward shaft friction force at the top of the pile. For the
calculation of the negative shaft friction, refer to the example calculation of a prefab concrete
pile on the following page.

Note:
For groups of piles the following applies: The negative shaft friction per pile can never be greater than
the total weight of the settling ground divided by the number of piles. The weight of the weak ground
per pile is therefore an upper limit for the calculation of the negative shaft friction ( m;g = 1.0 !!!).

Simple method
A simple method to estimate the shaft bearing capacity (in case there is no CPT), is to find
the cumulative shear stress of all ground layers along the pile:
Fshaft  Os  h   'v  K  tan  (27.8)
n

In the case of ground displacing piles (prefab driven piles) the ground is tensed horizontally,
which means: K  K0 . In the case of drilled piles, one should assume K  K0 . One can
simplify the calculation by assuming K  tan   0.25 (Methode Zeevaert).

Example
To show how to calculate the bearing capacity of a pile, the following example is of a prefab concrete
pile at the point of sounding test S12396. The pile’s dimensions are 160 x 160 mm2. The following
table describes the soil layer composition. The italic values can be calculated from the previous
(given) layer thicknesses and ground weights.

Layer Top Thickness (h) sat ’v,rep ’v,avg h*’v,avg


3
[m NAP] [m] [kN/m ] [kPa] [kPa] [m kN/m]
Peat (MV) -4.61 0.6 10.3 0.0 0.1 0.1
Clay, sandy -5.2 1.6 17.0 0.2 5.8 9.3
Peat deep -6.8 1.5 11.1 11.4 12.2 18.3
Clay -8.3 1.5 14.7 13.0 16.6 24.8
Sand loose -9.8 7.4 18.9 20.1 - -
Sand deep -17.2 3.8 20.0 - - -

Table ‎27-5 Layer thickness and soil weight

116
The negative shaft friction is calculated as follows:
Fs , nk , rep  Os  h   'v , avg  K0, rep  tan  rep
n

in which:
 h  '
n
v , avg  0.1  9.3  18.3  24.8  52.5 kN/m

K0,rep  tan  rep  0.25 (Somewhat conservative, but saves a lot of calculation)
Substitution gives:
Fs ,nk ,rep   4  0.16  52.5  0.25  8.4 kN
The normally-shaped (s = 1.0), square ( = 1) and driven (p = 1.0) piles are driven to NAP –16.0 m.
One then finds for Koppejan-s method (see Figure 27-20 Sounding):
qI  10.0 kN and qII  9.7 kN and qIII  8.2 kN
hence:
qI  qII
 qIII
pr ;max;tip  s p  2  9.0 MPA
2
The tip bearing capacity is:
Fr ;max;tip  Atip pr ;max;tip  0.162  9.0  231 kN

The shaft friction develops in "Sand, loose". Between NAP –10 m and NAP – 16 m one finds per
metre an average qc of respectively 4 MPa, 3 MPa, 4.5 MPa, 4 MPa, 8.5 MPa and 9.5 MPa (see
Figure 27-20). This leads to an average qc of 5.6 MPa.
L
Fr ;max; shaft  Otip   s qc  dz  4  0.16  0.010  5.6  16.0  9.8  222 kN
0

The load bearing capacity of the pile consists of the bearing capacities of the tip and of the shaft:
Fr ;max;d  Fr ;max;tip  Fr ;max;shaft  231  222  453 kN
In the case of a relatively flexible foundation plate (M=1) and with 4 sounding tests per area (N=4)
one applies a sounding factor of:
 = 0.80.
The material factor (method factor) is:
 m;b  1.25
With these factors the maximum load bearing capacity of the pile becomes:

Fr ; found ;max;d  Fr ;max;d  290 kN
 m;b
After the subtraction of the negative shaft friction, the pile can carry the following load:
Fr ; found ;all ;d  Fr ; found ;max;d  1.0  Fs;nk  282 kN

117
S12396

0
0 5 10 15 20 25 30

-2.000

-4.000
MV -4.61 m NAP

-6.000

-8.000
Depth [m NAP]

-10.000

-12.000
q III = 8.2 MPa

-14.000

-16.000

qI = 10.0 MPa
-18.000

q II = 9.7 MPa

-20.000
Qc [MPa]

Figure ‎27-20 Sounding (Cone Penetration Test) with pile

118
27.7 Settlement of piles
Not only the ultimate bearing capacity of piles is important, but also the settlement. The
settlement of a pile consists of the elastic deformation of the pile and the soil displacement of
the pile tip. This soil displacement can be estimated with the figure below.
For example a bored pile which will be loaded up to 50% of its ultimate bearing capacity,
will have a 3% deformation, relative to the equivalent diameter of the pile Deq, which is the
same as the diameter for a round pile or Deq  d 4 /  for square piles.

Figure ‎27-21. Load displacement diagram for several pile types


(picture from Dutch Norm NEN 6740).

Constructers can base their “spring-stiffness” of the soil below the pile in their computer
calculation on this figure.

119
28 Tension piles
28.1 Differences between tension and compression piles
The calculation of tension piles is theoretically the same as for compression piles, but there
are a number of differences:
1. Tension piles are only subject to shaft friction, not to tip resistance. Therefore the pile
tip factor p is not important for tension piles.
2. The tensile shaft factor t for tension piles is normally taken a little bit lower than the
compression factor s.
3. Tension piles are most commonly used in excavated building sites. Due to the
excavation of the building site, the ground pressure and thus the value of qc and the
friction force on the pile are reduced.
4. Due to the tensile force of the tension piles the vertical effective ground pressure, and
with it the value of qc, are reduced.
5. The single pile should not only be tested for possible slip, the clump weight should
also be verified. The pile could emerge with the soil around it. See chapter 50.
6. The safety factors for tension piles are greater than for compression piles, firstly
because inaccuracies are greater for shaft friction than for tip resistance and secondly
because this is increased further by statement 4.
7. The self-weight of the pile (usually under water) affects the system positively.

Due to statements 1-6, the tensile capacity of a pile is many times smaller than the
compression capacity of a pile.
There are no norms for the calculation of tension piles. Usually the rules according to the
CUR-report 98-9 are assumed, but they are disputable.
Particularly for very large tensile forces, grout anchors are used instead of tension piles.

Calculation of the bearing capacity of a tension pile is not easy. One must take a lot of factors
into account. For instance, the tensile force in a pile reduces the surrounding effective ground
pressure and thereby reduces the pile’s tensile force. This leads to a complex method of
calculation. Hence, in general, tension piles are calculated using a computer programme. For
the sake of completeness, the theory of such computer programmes is explained below.

28.2 Reduction of the cone value qc due to excavation


Most calculations assume piles are driven into the ground after the building sites have been
excavated. As a consequence of the reduced effective ground pressure, the cone resistance
has to be reduced. The following relationship can be used for this reduction:
0.71
 ' 
qc ,corr  qc  v;new 
 '  (28.1)
 v;old 
in which:
qc ,corr = corrected cone resistance at depth z [MPa]
 'v;new = vertical effective ground pressure after excavation at depth z [kPa]
 'v;old = vertical effective ground pressure before excavation at depth z [kPa]

120
The reduction of the effective ground pressure due to excavation (  'v;old -  'v;new ) is
calculated as described in appendix B2 of CUR-report 162 (according to Flamant). This
calculation takes the limited width of the excavated building site into account.
q
 'zz       sin 1 cos 1  sin  2 cos  2 
 1 2 (28.2)

in which:
q   hi i
n

a a
q
x

1 2

r1 r2

z
Figure ‎28-1 Strip load (Flamant)

The reduction of the effective ground pressure due to excavation can be calculated for a pile
positioned in the middle of the excavated building site. For piles nearer the edge of the
excavated building site, the reduction of the effective ground pressure will be less. One can
study this effect more closely using Flamant.

28.3 Reduction of cone resistance due to tensile pile force


The load on the pile group reduces the effective ground pressure in the soil layers from which
the pile takes its tensile strength. This reduction is taken into account by applying a reduction
factor f2 to the cone resistance. In order to do this, the soil along the length of the pile is
divided into layers of, for example, 0.5 metres thick. The factor f2 is subsequently calculated
per layer according to the following equation:

 i 1

 M i  M i2   2 'v; z ;0; j   ' di    2 'v; z ;0; j   ' di  2 Tn 
 n 0 
f 2; z ;i  (28.3)
2 'v; z ;0; j   ' di
with:
Op;i  t q c;z;i di 
Mi 
A  m,ten  var
in which:
Mi = reduction of the vertical effective ground pressure in layer i [kPa]
’v;z;0;j = vertical effective stress after excavation on top of layer i [kPa]
di = thickness of layer i [m]
Ti = maximum possible friction between the pile and the soil [kPa]

121
Op = average circumference of the pile in layer i [m]
t = friction factor as indicated in table 24 of CUR-report 98-9
qc;z;i = average cone resistance in layer i after reduction due to
excavation [kPa]
A = surface area taken up by the pile [m2]
m,ten = material factor for piles with tensile loads (=1.4)
var = material factor for variable loads (=1.0)
 = factor, dependent on the number of soundings. See table 1
of NEN 6743 or Table 57-4 (  0.8 )

Notes:
1. In many cases the load on the tension piles is normative during implementation. In this phase the
variation of the load is negligible. Consequently one uses var = 1.0 in this phase.
2. An underwater concrete floor is considered a flexible structure according to the definition in NEN
6743, hence, when determining the factor  according to table 1 of NEN 6743, the number of
piles is set as 1.

28.4 Clump criterion


An important form of global translation is the vertical uplift (heave) of a building pit
including the tension piles and ground in between, due to the groundwater pressure.
Therefore, not only should one check whether the pile will fail due to sliding and whether it
can be pulled “clean” out of the soil, one should also verify whether all piles, including the
soil, can be pulled out as a “clump”. The weight of the clump Fclump per pile is often
calculated as follows:

Fclump  A· ' L  23 A  12 3  (28.4)
 A· '  L  A/3 
in which:
A = surface area per tension element [m2]
’ = effective volumetric weight of the soil [kN/m3]
L = distance between the bottom of the underwater
concrete floor and the bottom of the tension element [m]

This formula shows that the length L is reduced because a cone-shaped fracture surface at 45
degrees is taken into account at the centre point of the tension element (pile).
The maximum acceptable tensile force is then:
Fa  Fclump  m,clump (28.5)
in which:
m,clump = material factor for the weight of the clump (=1.1)

Of course, the weight of the underwater concrete and of the tension element (pile) may also
be included (m = 1.1).

The problem with the method above is that it is not correct to take the pore water pressure
only at a constant level of the pile tip (pile base), while the bottom of the “soil clump” has a
far more complex shape. Therefore an easier way for a quick design check of the clump
criterion can be found in Chapter 50 “Global translation”.

122
28.5 Edge Piles
Particularly around an open access road to a tunnel, edge piles sometimes have greater tensile
loads than the other piles. In this case, to determine the tensile load capacity, one must take
into account:
 limited (Flamant-) reduction of the sounding resistance along the edge of the excavated
building site due to the excavations.
 group action based on the average pile tensile forces in a cross section and thus not based
on the tension in the edge piles themselves.
 possible horizontal forces in sheet piling next to the pile.

28.6 MV-pile or Jacket-grouted pile


Any normal compression pile can act as a tension pile. Of course the bearing capacity for
tension is much less than for compression because for tension there is no pile tip resistance
and the shaft friction is reduced due to several effects written before.
If the tension capacity of a normal compression pile is insufficient, one can think of special
tension piles like the MV pile. The original name of this pile is the “Müller-Verpress Pfahl”.
This a jacked-grouted pile: A steel profile (H, I, ][, or tube) is driven in the ground. This can
be done skew up to 45 degrees.

Figure ‎28-2 Skew MV pile installation (Photo: Voorbij Foundation Eng. NL)

At the tip there is a tray in which, during pile driving, grout is injected at high pressure. This
creates a higher shaft friction. The tension bearing capacity of this pile is therefore calculated
with a tensile shaft factor t = 1.2%. The compression bearing capacity of this pile is
calculated with a compression shaft factor of s = 1.4% and a pile tip factor of p = 1.0%
with shape factor  = 1.0.

123
Figure ‎28-3 Tray at tip of MV pile
If these MV piles have insufficient tensile bearing capacity, mostly grout anchors are used,
similar as for retaining walls.
If the jacking of a MV pile makes too much noise or vibration, one can also vibrate the pile
into the ground; these piles are called RI piles (Rüttel Injektion Pfahl).

124
V Underground Megastructures

125
29 Underground megastructures
No people on earth make larger structures than civil engineers. The art of civil engineering is
changing and shaping the world. The largest, or at least, the longest structures on earth are
civil infrastructures like roads, railways and water ways such as canals and man shaped
rivers, but also parts of these structures like bridges, aqueducts, underground railway and
metro stations and tunnels for either roads, railways or water ways.

Figure ‎29-1. Road, railway and water tunnels.

Since these megastructures are always constructed on or in the ground, geotechnical


engineering always deals with those structures. There are also non-infrastructural
geotechnical megastructures, such as underground shopping centres or underground
museums.

Figure ‎29-2. Underground shopping centres and museums.

126
Also underground living is more and more in focus because underground houses can be
better insulated and need less space, since the garden is on the roof.

Figure ‎29-3. Living underground.

Architects have also designed structures like the futuristic “ground scraper” or an
underground mega sport centre as shown in the pictures below.

Figure ‎29-4. Futuristic‎“ground‎scraper”‎and‎mega‎sport‎centre.

127
For military purposes, bunkers have already been made for a long time and these also are a
type of geotechnical megastructures.

Figure ‎29-5. Bunkers.

128
But not even the structure-to-be-built itself is a geotechnical megastructure. The creation of
the building pit is sometimes also a huge undertaking, especially when there is a high ground
water level. The loads of the ground and groundwater increase roughly linearly with depth,
the resultant forces, already quadratic and the bending moments in the retaining walls
increase even faster with depth than that. So, it is complex to construct deep.

Figure ‎29-6. Huge building pit (European Court, Kirchberg Luxembourg).

For the design and planning of geotechnical megastructures one needs a very good
geotechnical survey first of all, with enough wide spread and sufficiently deep cone
penetration tests and borings. Laboratory tests are also required in the survey for determining
the necessary parameters for soil weight, strength, stiffness and sometimes also permeability.
The design will have to go in several steps; first a rough design with rules of thumb, then a
quick design with general design rules (Meyerhof/Terzaghi or Boussinesq, etc.), then a
thorough design with simple computer programs and finally a sophisticated design with 2-D
or even 3-D finite element models such as Plaxis. The time, transport and financial planning
must be part of the construction design because it is decisive for the costs, especially for
megastructures.

129
30 Sustainable resources
The natural resource which is mostly used is water, mostly due to agriculture. Due to
evaporation and rainfall, this water comes back to earth, although not always at the same
location. The second most used resource is sand, mainly for construction.

One of the human activities which consumes most natural resources, especially sand,
concrete, steel, energy and water, is the construction of infrastructure. The world has only a
certain amount of resources, so human mankind must think wisely when to use what
materials and how to reuse materials as much as possible. Also the durability of a material is
very important.

Luckily the most important material in geotechnical engineering, the soil itself, is mostly
already available and is a material that does not corrode or rust, does not burn and cannot be
broken, since it is already broken (sand and clay particles are left-over from rock due to
weathering and erosion).

Since the other materials used are mostly concrete and steel, the same applies here as for the
use of these materials in concrete and steel structures. Concrete and steel can only partly be
reused but can fulfil their task for a very long time underground, where there is no UV-light,
and especially below the groundwater table, little oxygen.

Other materials used in geotechnical engineering are the geotextiles. As long as no UV-light
or chemicals can harm the geotextiles, they can be used for a long time. They do wear out
unfortunately, because the plasticiser in plastics will be washed out, which is not good for the
environment and our drinking water, and the plastic will become more brittle.

Some techniques can cost a lot of energy (like ground freezing) or a lot of chemicals
(chemical injections for stopping leaking) and should therefore be reduced as much as
possible, but are nevertheless perfect solutions for unexpected and local problems, where
other solutions are not possible or at least very complex and expensive.

130
VI Building pit

131
31 Building pit
31.1 Introduction
(Most text in this section is from COWI, Denmark. Pictures are copied from internet)
Most building pits are made for the construction of underground infrastructure (tunnels for
roads and railway lines; railway stations) and for underground parking garages.
In urban areas underground structures often have to be constructed under space constraints.
Structures may have to be constructed with limited space between and close to existing
buildings and in some situations structures may even have to be constructed below existing
buildings and structures. Construction under space constraints requires special attention to
neighbouring buildings and structures, which in some areas may require temporary support
during construction or a permanent reconstruction of the affected part of the building and
foundation.

There are two different ways for making an excavation possible: using slopes and using
retaining walls. The biggest advantage of using a slope is that it is cheaper. The biggest two
advantages for a retaining structure is that it uses less space (important in urban areas) and
that it is more impermeable, so only limited dewatering is needed.

Figure ‎31-1. Building pit with slopes and dewatering

The problem of groundwater can often be solved with a dewatering system, see Chapter 54.
There are side effects with dewatering, such as settlements. Dewatering, therefore, is often
only allowed for 1 year in urban areas.

132
Retaining wall structures are primarily used for retaining construction pits. Walls are either
temporary or part of the permanent underground structure. Normally retaining walls are
exposed to external ground and water pressure. During the construction period and in the
permanent situation they are subject to a number of different load combinations depending on
the design of permanent works, temporary works and the construction method. Retaining
wall structures include embedded retaining walls such as sheet pile walls, bored pile walls
and diaphragm walls.

Figure ‎31-2. Building pit with retaining wall

Control of the groundwater is an important consideration in the design of underground works,


and it is often one of the critical factors when choosing a retaining wall. The impact of
construction on the existing groundwater regime may have to be minimised, particularly in
urban areas where there are buildings with foundations that are often sensitive to the
settlements that may be induced by groundwater lowering. The amount of water that has to
be abstracted during construction may also have to be minimised for environmental reasons.
The deflection of retaining walls in urban areas may have to be limited in order to protect
nearby buildings and structures from damage. Monitoring of the performance of the retaining
wall structure and the ground movements are often necessary, particularly in urban areas.

31.2 Pit collapse


The worst case scenario is when a building pit collapses. A terrible example is the collapse of
the Nicoll Highway in Singapore in April 2004, which left 4 people dead. Mistakes in the
design led to this, along with the problem of a manager who switched off the building pit
alarm system instead of investigating, when it sounded after detecting large deformations.

133
Figure ‎31-3. Building pit collapse Metro near Nicoll Highway, Singapore, 2004

Another terrible example is the collapse of the building pit for a station in the North-South
train Line in Cologne in 2009. An employee had stolen a large number of steel supports and
had sold them for old steel prices. The missing supports were not noticed by any workers, nor
any inspection. Two people died in the collapse and many ancient documents from the city
archive were lost.

Figure ‎31-4. Building pit collapse, North-South Line station, Cologne, Germany, 2009

The collapse in Singapore is a clear example for the importance of good engineering and
design. Both the collapses in Singapore and Cologne, Germany, show clearly the importance
of thorough inspection and control of a project.

134
31.3 Pit design
The design of the pit depends on many aspects. Most important is the functionality, as that is
the reason that the money will be spent. The ground conditions are also very important.
These are leading in the choice of the type of pit and substructures.
One of the first big choices is the decision between a temporarily building pit structure and a
permanent structure. The latter means that the walls and floors of the building pit will be
incorporated into the final structure.
Another choice to be made is the type of vertical building pit boundaries. This depends on the
previous decision, the ground condition due to dewatering and of course the amount of space
available. In case of a slope, see Chapter 51 on Slope stability and the Chapters 52 and
further for the dewatering.
In case of a retaining wall, the decision has to be made as to what type of wall. See
Chapters 32 about this. For the design of the retaining walls, see the proceeding chapters.
Usually walls for deeper pits will have to be supported. For the supports, there are two types
of options; anchors or struts. For this see the Chapters 41, 42 and 43.
For the design of the wales see Chapter 44.
For the design of the floors, see the Chapters 45 to 48.
Finally, the overall stability has to be checked. For this see the Chapters 50 to 51.3.

135
136
VII Walls and lateral stress

137
32 Walls
32.1 Wall types
There are different types of classical retaining walls, for example the:
a) gravity wall,
b) cantilever wall (or L wall)
c) crib wall
d) supported (thin) wall
e) gabion wall (not in the figure below)

Figure ‎32-1. Common types of retaining walls. (from McGraw-Hill Encyclopaedia)

Other types are the internally stabilised earth retaining structures:


f) reinforced soil
g) soil nailing

Gravity walls depend on the weight of their mass (stone, concrete or another heavy material)
to resist pressures from behind and will often have a slight 'batter' setback, to improve
stability by leaning back into the retained soil. Short landscaping walls are often made from
mortarless stone or segmental concrete units (masonry units). Dry-stacked gravity walls are
somewhat flexible and do not require a rigid footing in frost areas. Earlier in the 20th century,
taller retaining walls were often gravity walls made from large masses of concrete or stone.

Cantilevered retaining walls are made from an internal stem of steel-reinforced, cast-in-place
concrete or mortared masonry (often in the shape of an L or inverted T). These walls
cantilever loads (like a beam) to a large, structural footing, converting horizontal pressures
from behind the wall to vertical pressures on the ground below. This type of wall uses much
less material than traditional gravity walls.

138
Today, taller retaining walls are increasingly built as composite gravity walls such as:
geosynthetic or with precast facing; gabions (stacked steel wire baskets filled with rocks);
crib walls (cells built up log cabin style from precast concrete or timber footing, beams and
ties, filled with soil); or soil-nailed walls (soil reinforced in place with steel and concrete
rods).

The thin walls are usually sheet pile walls. Sheet pile retaining walls are typically used in soft
soils and tight spaces. Sheet pile walls are made out of steel, vinyl or wood planks which are
driven into the ground. For a quick estimate the material is usually driven 1/3 above ground,
2/3 below ground, but this may be altered depending on the environment. Taller sheet pile
walls will need a tie-back anchor, or "dead-man" placed in the soil a distance behind the face
of the wall, which is tied to the wall, usually by a cable or a rod. Anchors are placed behind
the potential failure plane in the soil. Struts are also possible if there is an opposite wall
nearby.

These walls all have different advantages and disadvantages, but a big difference between the
thin wall and the others types, is that a thin wall can be installed before a lot of ground has
been removed. Therefore these are commonly used for building pits.

There are several types of thin walls:


 Soldier pile wall (Berliner wall)
 Sheet pile wall
 Combi wall
 Bored pile wall
 Diaphragm wall

The order of the list above is more or less from the cheapest to most expensive, but of course
not all options are always possible due to the soil conditions and the size of the excavation.

32.2 Soldier pile wall (Berliner wall)

Figure ‎32-2. Berliner wall

139
The soldier pile wall is the easiest way to make a wall. At a few metres distance from each
other, a hole is bored in the ground. Steel profiles are installed and mostly grouted at the
footing of the profiles to decrease deformations later. During the excavations, wooden beams
are installed, starting at the top. Anchors can also be installed and connected to the vertical
profiles. This method only works when the soil is rather dry, which means the pit should be
above the groundwater table, with or without dewatering by pumping. A little cohesive soil is
very useful because the installation of the beams is after the step by step excavation.

32.3 Sheet pile wall


The sheet pile wall is frequently used. It is a bent sheet of steel (can also sometimes be
concrete) which is designed to hold the bending moments and to be waterproof. Installation is
mostly through impact driving or vibratory driving. Both the vibratory driving and more so
the impact driving cause vibrations. If this creates too much damage or nuisance, the sheet
pile can be pressed (however only in very soft soils) or cannot be used at all. Also if the
ground resistance is too high, it is better to use a bored pile wall or diaphragm wall.

Figure ‎32-3. Sheet pile wall (left failure, right installation)

140
Figure ‎32-4. Left: Vibratory driving near house. Right: Pressing of sheet pile
(Photos AJ Snethlage)

32.4 Combi wall


The combi wall is similar to the sheet pile wall, it is composed of one or two sheet piles and
then one round pile (pipe) in a constant pattern. The benefit is that this sheet-pipe
combination is far stiffer and stronger than only a sheet pile. Therefore, this is used for deep
excavations or large differences in ground level, for example at quays in harbours.

Figure ‎32-5. Combi wall (left principle, right installation)

141
Figure ‎32-6. Reinforcement with reinforced concrete in pipe and with anchoring

32.5 Bored pile wall

Figure ‎32-7. Principle of a bored pile wall

Large and small diameter bored cast-in-place piles are often used to construct efficient and
economic temporary or permanent retaining walls. They are especially useful at those places
where the ground is too hard for sheet pile driving or where vibration nuisance should be
avoided. The bored hole is prevented from collapsing by either using a steel tube or a
bentonite suspension. The wall consists of interlocking bored piles. Primary piles are
constructed first using a 'soft' cement-bentonite mix (commonly 1 N/mm2) or 'firm' concrete
(commonly 10 N/mm2). Secondary piles, formed in structural reinforced concrete, are then
installed between the primary piles with a typical interlock of 150 mm. These walls may need
a reinforced concrete lining for permanent works applications, depending on the expected
bending moments.

142
Figure ‎32-8. Bored pile wall supported by struts (Photo: Skanska UK)

Figure ‎32-9. Bored pile wall, left: unsupported, right: supported by anchors

32.6 Diaphragm wall


(Parts of this section are from www.diaphragmwall.com and Bilfinger Berger Spezialtiefbau
GmbH)

Diaphragm walls are concrete or reinforced concrete walls constructed in slurry-supported,


open trenches in the ground. The walls are usually excavated under bentonite slurry.
There are two types of diaphragm walls:

143
 cast-in-place diaphragm walls; these are made from liquid concrete placed by using
the tremie installation method, and the
 pre-cast diaphragm walls; which are made by installing pre-cast concrete panels.

Diaphragm walls can be constructed to depths of up to 100 metres and to widths of 0.45 to
1.50 metres. Diaphragm wall construction methods are relatively quiet and cause little or no
vibration. Therefore, they are especially suitable for civil engineering projects in densely-
populated inner city areas. Due to their ability to keep deformations low and provide low
water permeability, diaphragm walls are also used to retain excavation pits in the direct
vicinity of existing structures. If there is a deep excavation pit at the edge of an existing
structure and groundwater is present, diaphragm walls are often used as the most technically
and economically favourable option. Diaphragm walls can be combined with any type of
support (anchor or strut) system.

Figure ‎32-10. Diaphragm wall support by struts

Diaphragm wall construction begins with the trench being excavated in discontinuous
sections or "panels". Stop-end pipes are placed vertically at each end of the primary panel to
form joints for adjacent secondary panels.
Once the excavation of a panel is complete, a steel reinforcement cage (sometimes only a
steel sheet pile) is placed in the centre of the panel. Concrete is poured in one continuous
operation through one or more tremie pipes that extend to the bottom of the trench. The
tremie pipes are extracted as the concrete rises; however, the discharge end of the tremie pipe
always remains embedded in the fresh concrete.

144
Figure ‎32-11. Phases of the installation of a diaphragm wall

The bentonite slurry that is displaced by the concrete is saved and reused for subsequent
panel excavations. As the concrete sets, the end pipes are withdrawn. Similarly, secondary
panels are constructed between the primary panels to create a continuous wall. The finished
wall may be cantilever or require anchors or props (struts) for lateral support.
Diaphragm walls are commonly used in congested areas. They can be installed in close
proximity to existing structures with minimal loss of support to existing foundations. In
addition, construction dewatering is not required, so there is no associated subsidence.

The following group of photos shows the installation of a diaphragm wall: a) installation
support, b) digging of tranche, c) installation reinforcement, d) concrete pouring.

145
Figure ‎32-12. Installation of a diaphragm wall (Photos Spoorzone Delft, NL)

The design of a diaphragm wall is not too complex. There are a limited amount of sizes
(mostly 0,7 or 1,0 m thick). The depth and length can be anything that is required. The
maximum moment found in the calculation leads to the amount of reinforcement chosen. In
order to avoid mistakes during construction, the reinforcement is made symmetric, so both
sides have the same amount of reinforcement. If not, a simple mistake of installing the
reinforcement will not be notified, until the moment of collapse of the diaphragm wall. If one
wants to reduce the maximum moment, one can install extra layers of supports (anchors or
struts).

146
33 Lateral stresses in soils
33.1 Coefficient of lateral earth pressure
In this chapter, and the next chapters, the analysis of the horizontal stresses will be discussed.
This is of particular interest for the forces on a retaining structure, such as a diaphragm wall
or a sheet pile wall.

Figure ‎33-1. Half space.

Even in the simplest case of a semi-infinite soil body, without surface loading, see
Figure 33-1, it is impossible to determine all stresses caused by the weight of the soil. It
seems reasonable to assume that in a homogeneous soil body with a horizontal top surface
the shear stresses  zx ,  zy and  xy are zero, and it also seems natural to assume that the
vertical normal stress  zz increases linearly with depth,  zx   z . These assumptions ensure
that the condition of vertical equilibrium is satisfied. The horizontal stresses  xx and  yy ,
however, cannot be determined unequivocally from the equilibrium conditions.
Actually, it can be stated that the stresses must satisfy the equations of equilibrium,
 xx  yx  zx
   0, (33.1)
x y z
 xy  yy  zy
   0, (33.2)
x y z
 xz  yz  zz
     0, (33.3)
x y z
 yz   zy , (33.4)
 zx   xz , (33.5)
 xy   yx . (33.6)
These equations constitute a set of six conditions for the nine stress components, at every
point of the soil body. It seems probable that many solutions of these equations are possible,
and it cannot be decided, without further analysis, what the best solution is. It seems natural
to assume, at least for a homogenous material, or a material consisting of horizontal layers,
that the stress state may be such that vertical normal stress increases linearly with depth, in
proportion to the unit weight of the soil. More precisely, it is assumed that the stresses can be
written as
 zz   z, (33.7)

147
 xx   yy  f ( z), (33.8)
 yz   zy  0, (33.9)
 zx   xz  0, (33.10)
 xy   yx  0. (33.11)
This field of stresses satisfies all the equilibrium conditions, and the boundary conditions on
the upper surface of the soil body, i.e. for z  0 the stresses on a horizontal plane are zero,
 zz  0 ,  zx  0 , and  zy  0 . To assume that all shear stresses in the soil body seems a
realistic assumption if all vertical columns of the soil have the same properties. There will
probably be no shear stress transfer between these columns.
The function f ( z ) in equation (33.8) remains arbitrary, and in principle the stresses  xx and
 yy need not be equal. It has been assumed that the horizontal stress in any horizontal plane
is the same in all directions, so there are no preferential directions in the horizontal plane.
Theoretically speaking, the function f ( z ) , which describes the horizontal stresses, need not
be continuous. Discontinuities in this function are allowed, and may occur especially if there
are discontinuities in the soil properties. It may be remarked that even the expressions for the
vertical normal stress  zz and for the shear stresses do not follow necessarily from the
equilibrium conditions. It may well be that these stresses depend upon x and y, if the soil
stiffness is not constant in horizontal planes. In case of a very soft inclusion in a rather stiff
soil body, the stresses may be concentrated in a region around the soft inclusion. This is
called arching, as the stiffer soil may form a certain arch to transmit the load from upper
layers to the subsoil. In homogeneous soil, however, or in soils without large differences in
stiffness, the stress distribution given above can be considered as a reasonable approximation.
Such a soil body has often been created, in its geological history, by gradual sedimentation,
often under water. In such conditions the gradual increase of the thickness of the soil body
will normally lead to a stress state of the form given above.
The stress state described by equations (33.7) – (33.11) can be made somewhat more
practical by writing f ( z )  K zz , where K is an unknown coefficient, that may depend upon
the vertical coordinate z. The horizontal stresses then are
 xx   yy  K zz  K z, (33.12)
where K is the coefficient of lateral earth pressure. It gives the ratio of the lateral normal
(effective) stress to the vertical (effective) stress. Theoretically speaking, the problem has not
been cleared, because the value of K is still unknown, but it seems safe to assume that the
horizontal stresses will also increase with depth, if the vertical stresses do so. Thus, it can be
assumed that the coefficient K will not vary too much, at least compared to the original
function f ( z ) .
It may be mentioned that for historical reasons the coefficient K is denoted as a coefficient of
earth pressure, in agreement with most soil mechanics literature. This is one of the few
instances where the word earth is used in soil mechanics, rather than the word soil, or ground.
No special meaning should be attached to this terminology. In this book the coefficient will
sometimes also be denoted as the horizontal (or lateral) stress coefficient.
The value of the lateral earth pressure coefficient K depends upon the material, and also on
the geological history of the soil. In this chapter some examples of possible values, or the
possible range of values, will be given, for certain simple materials. It will appear to be
illustrative to describe the relations between vertical and horizontal stresses in a stress path.

148
Often the horizontal stress will indeed be smaller than the vertical stress, so that K  1 , but
this is not absolutely necessary.
Stresses in a fluid are different from a solid material. In a fluid the shear stresses can be
neglected, compared to the pressure. This means that the normal stress is equal in all
directions. This means that
K  1. (33.13)
When K  1 the horizontal stress is equal to the vertical stress.
Soil is not a fluid, but certain very soft soils come close: the mud collected by dredging often
is similar to a thick fluid. Very soft clay, with a high water content, also behaves similar to a
fluid. When spread out it will flow until an almost horizontal surface has been formed. For
such soils, the value of K will be close to 1.

33.2 Elastic material


A possible approach to the behaviour of soils is to consider the soil as an elastic material. In
such a material the stresses and strains satisfy Hooke’s law. In a situation in which there can
be no lateral deformation, the stresses must satisfy the condition
1
 xx  [ xx  ( yy   zz )]  0,
E
1
 yy  [ yy  ( zz   xx )]  0,
E
if the z-direction is vertical. In a medium of large horizontal extent it can be expected that
 xx   yy . Then
 xx   xx  o :  xx   yy  K zz ,
with

K . (33.14)
1 
If Poisson’s ratio varies between 0 and 0.5, the value of K varies from 0 to 1.
If   12 the horizontal stresses are equal to the vertical stresses. In that case there are no
volume changes, just as in a fluid. The stress path then is equal to the stress path in a fluid.
If the horizontal strains are not zero, but it is still assumed that the two horizontal stresses,
 xx and  yy , are equal, these stresses are
 E
 xx   yy   zz   xx . (33.15)
1  1 
In case of a negative horizontal strain, the horizontal stress decreases, and K is getting
smaller. A positive horizontal strain, for instance due to some lateral compression, will result
in a larger horizontal stress. The value of K then will seem to increase. These are general
tendencies, with a validity beyond elasticity.
In some older publications equation (33.14) has been proposed as a generally applicable
relation for soil and rock. This is untrue. An elastic analysis supposes that the stresses are
being developed gradually, by gravity being applied gradually, on an existing soil in an
unstressed state. And during this entire process the relation between stress and strain should
be linear, and no horizontal deformations should occur. Geological history is usually much

149
more complex, and the material behaviour is non-linear. This means that the value of the
lateral stress coefficient K in general cannot be predicted with any accuracy. It can be
expected that in a region between two deep rivers the value of K will be relatively small,
whereas in a valley between two mountain ridges that are moving towards each other due to
tectonic motion, the stress coefficient K will be relatively large.

33.3 Elastic material under water


In order to take groundwater into account, the soil may be schematised as a linear elastic
material, which is being deposited under water.
If the weight of the material is again carried by the vertical stresses, the vertical total stress
will increase linearly with depth,
 zz   s z, (33.16)
in which  s is the total volumetric weight of the soil, including the water in the pores. The
pore pressures are assumed to be hydrostatic,
p   w z, (33.17)
so that the vertical effective stresses are
 zz   zz  p  ( s   w ) z. (33.18)
It is now postulated that in the process of the development of these stresses no horizontal
deformations of the soil skeleton can occur. The deformation of this soil skeleton is
determined by the effective stresses, and in this case, for a linear elastic material, it follows
that
 
 xx   yy   zz  ( s   w ) z. (33.19)
1  1 
This means that

K . (33.20)
1 
The coefficient of horizontal earth pressure K is in general the ratio between the horizontal
and vertical effective stresses
 xx  K zz . (33.21)
Therefore, it can be deduced that the horizontal total stress becomes
 xx   xx  p  K ( s   w ) z   w z. (33.22)
The lateral stress coefficient K should be used for the effective stresses only. The horizontal
total stresses should be determined by adding the pore pressure to the horizontal effective
stress.

150
34 Rankine
The possible stresses in a soil are limited by the Mohr-Coulomb failure criterion. Following
Rankine (1857) this condition will be used in this chapter to determine limiting values for the
horizontal stresses, and for the lateral stress coefficient K.
For reasons of simplicity the considerations will be restricted to dry soils at first. The
influence of pore water will be investigated later.

34.1 Mohr-Coulomb
The stress states in a soil can be limited, with a good approximation by the Mohr-Coulomb
failure criterion. This criterion is that the shear stresses on any plane are limited by the
condition
   f  c   tan  , (34.1)
where c is the cohesion, and  is the angle of internal friction. The criterion can be illustrated
using Mohr’s circle, see Figure 34-1.

Figure ‎34-1. Mohr-Coulomb.

If it is assumed that  zz and  xx are principal stresses, and that  zz is known (by the weight
of the load and the soil), it follows that the value of the horizontal stress  xx cannot be
smaller than indicated by the small circle, and not larger than defined by the large circle. The
ratio between the minor and the major principal stress can be determined by noting, see
Figure 34-2, that the radius of Mohr’s circle is 12 (1   3 ) , and that the location of the centre
is at a distance 1
2 (1   3 ) from the origin. It follows that for a circle touching the envelope,
1
( 1   3 )
sin   2
.
2 ( 1   3 )  c cot 
1

So that

151
1  sin  cos 
3   1  2c . (34.2)
1  sin  1  sin 

Figure ‎34-2. Ratio of principle stresses.

The two coefficients in this equation can be related by noting that

cos  1  sin 2  (1  sin  )(1  sin  ) 1  sin 


   .
1  sin  1  sin  1  sin  1  sin 
This means that equation (34.2) can be written as
 3  Ka1  2c Ka , (34.3)
Apart from the constant term 2c K a there appears to be a given ratio of the minor and the
major principal stress.
Formula (34.3) can be written in inverse form as
1  K p 3  2c K p , (34.4)
where
1  sin 
Kp  . (34.5)
1  sin 
The coefficients Ka and Kp, which give the smallest and the largest ratio of the two principal
stresses (apart from a constant term), are denoted as the coefficients of active earth pressure
(Ka) and passive earth pressure (Kp), respectively.
If the cohesion is zero ( c  0 ) it can be seen that
c0 : Ka  K  K p . (34.6)
If   30 (this is a common value for sand, on the small side), it follows that
c  0 and   30 : 1
3  K  3. (34.7)

152
1
The lateral stress coefficient K appears to be limited by values about 3 , and about 3. The
precise limits depend upon the angle of internal friction  .
As seen in the previous chapter for the elastic case, lateral extension of the soil leads to a
smaller value of the lateral stress coefficient K, whereas lateral compression leads to a larger
value of the coefficient K. The extreme situations are denoted as active earth pressure and
passive earth pressure, respectively. The case of active earth pressure is supposed to occur
when a retaining structure is being pushed away by the soil stresses. Passive earth pressure
denotes that the structure is being pushed into the ground, in which a reaction is being
developed.
The large difference between the minimum and the maximum lateral stress is characteristic
for frictional materials such as soils.

34.2 Active earth pressure


It can be expected that the smallest value of the horizontal stress occurs in the case of a
retaining wall that is moving away from the soil, see Figure 34-3. The Mohr circle for that
case is also shown in the figure. The pole for the vectors normal to the planes is the rightmost
point of the circle. This means that the critical shear stress acts on planes making an angle
4   2  with the horizontal direction, that is an angle of 4   2  with the vertical
1 1 1 1

direction. These planes have been indicated in the left half of Figure 34-3. It is sometimes
imagined that the soil indeed slides along these planes in case of failure.

Figure ‎34-3. Active earth pressure.

The vertical stresses along the wall are


 zz   z, (34.8)
in which  is the volumetric weight of the soil, and z is the depth below soil surface. The
horizontal stresses now are, with (34.3),
 xx  Ka z  2c Ka . (34.9)
The total horizontal force on a wall of height h is obtained by integration from z  0 to z  h
. This gives
Q  12 Ka h2  2ch K a . (34.10)

The distribution of the horizontal normal stress  xx against the wall is shown in Figure 34-4.

153
Figure ‎34-4. Horizontal stress in case of active earth pressure.

It appears that at the top of the wall tensile stresses are generated, over a depth of 2c  K a .
This may be possible in the soil for a short while, in undrained conditions, with negative
stresses in the water. In fully drained conditions this is not possible, because then there
should be tensile stresses between the particles, or between the particles and the wall.
Therefore, it is usually assumed that in a top layer of the soil, of height 2c  K a , cracks
will appear in the soil, and between the soil and the wall. For this case, the stress distribution
is shown in Figure 34-5. For the vertical stresses, the cracked zone acts as a surcharge.

Figure ‎34-5. Horizontal stresses, with cracks in the soil.

The total horizontal force now is


Q  12 Ka hr 2 , (34.11)
in which hr is the reduced height of the wall,
hr  h  2c  Ka . (34.12)

34.3 Passive earth pressure


The case of passive earth pressure, in which the horizontal soil stress has its maximum value,
can be considered to correspond to a smooth vertical wall that is being pushed in horizontal
direction into the soil, see Figure 34-6. The Mohr circle has been shown in the figure as well,
with the pole in this case being located in the leftmost point of the circle. The critical shear
stress    f  c   tan  occurs on planes making an angle 14   12  with the horizontal

154
direction. These planes have also been indicated in the left half of the figure. In this case the
potential sliding planes are shallower than 45º.

Figure ‎34-6. Passive earth pressure.

The horizontal stresses against the wall in this case are


 xx  K p z  2c K p . (34.13)
They are shown in Figure 34-7.

Figure ‎34-7. Horizontal stresses in case of passive earth pressure.

The total horizontal force on a wall of height h is obtained by integration of the horizontal
stresses from z  0 to z  h . This gives
Q  12 K p h2  2ch K p . (34.14)
In the passive case the cohesion c appears to lead to a constant factor in the expression for the
horizontal stresses. There is no reason for cracks to appear, as there are no tensile stresses in
this case.
The two extreme states of stress considered here are often denoted as the Rankine states, after
the English scientist Rankine (1857), who indicated that these stress states are the limiting
conditions. In the case of a solid retaining wall, on a good foundation, the actual horizontal
stresses will be somewhere between these two extremes. As the limits are so far apart (there
may be a factor of about 9 between them), this leaves the horizontal stress  xx undetermined
to a high degree.

155
34.4 Neutral earth pressure
It has been found that the possible states of stress in a soil may vary between fairly wide
limits, especially if the friction angle is large. For a normal sand, with   30 , the smallest
value of the horizontal stress is 13 of the vertical stress (which is usually known from the
surcharge and the weight of the overlying soil), and the largest value is 3 times the vertical
stress. The lateral stress in a non-preloaded horizontal terrain, which is a totally different case
than a moving retaining wall, is unknown, at least from a strictly scientific viewpoint. In
reality there may be some additional information that may help to determine the probable
range of the horizontal stress. If the horizontal displacements of the wall are practically zero,
the ratio of the horizontal stress to the vertical stress is denoted as the neutral earth pressure
coefficient K0. In an elastic material this value would be K0   (1  ) , but this is not a very
good estimate, as soil is not an elastic material, and the history of the development of stresses
in the soil may be more important than the condition of zero lateral deformation.
Nevertheless, in many cases it is unlikely that the coefficient K0 would be larger than 1, as
this would require some form of motion of the wall towards the soil mass. Also, the active
state of stress (say 13 ) may also be unlikely, if the wall is rather stiff and strong. All this
suggests that the neutral stress coefficient may perhaps vary in the range from 0.5 to 1.0. For
soft clay the value may be close to 1, and for sands, values of about 0.6 or 0.7 have been
found to give reasonable results. Lacking any better information, the value may be estimated
by the formula proposed by Jaky,
K0  1  sin  , (34.15)
but there is no real basis for this formula, except that it gives values between 0 and 1, with
the value being close to 1 if the friction angle  is very small (as it is for soft clays).

34.5 Menard-penetrometer and CAMKO-pressuremeter


The best procedure is to try to measure the value of K0, using an instrument of which
determines a response from the horizontal stress.

Figure ‎34-8. Menard-penetrometer. (Picture: Ph. Reiffsteck, LCPC Paris)

This can be done with the Menard-penetrometer which is more or less the same as the
CAMKO-pressuremeter, which was developed in Cambridge (UK). These pressuremeters

156
also measure the stiffness and the undrained strength of the ground. A rubber membrane
around a pipe is pressurised, and the resulting deformation is measured. The idea is that the
soil response will be different for lateral pressures above and below the original neutral
stress. The membrane consists of three parts, with the central part being the measuring cell.
A similar instrument is Marchetti’s dilatometer, which consists of a hollow circular plate that
is pushed into the soil, in a vertical position. By expanding the plate the lateral response is
measured, and from this response the lateral stiffness and the neutral stress coefficient may
be estimated. Another method is to inject water into the soil from a tubular instrument. The
idea is that a vertical crack may be produced in the soil if the water pressure exceeds the
horizontal total stress, because the soil skeleton cannot transfer tensile stresses. In petroleum
engineering this process is called hydraulic fracturing, and it is used to improve the
permeability of porous rock.

Figure ‎34-9. Horizontal stress as a function of the displacement.

A possible relation between the horizontal stress against a retaining structure and its
horizontal displacement is shown in Figure 34-9. If the displacement is zero, the lateral stress
coefficient will be K0. If the structure is being pressed towards the soil, the lateral stress will
gradually increase, until finally the passive coefficient Kp is reached. On the other hand, if the
structure moves away from the soil, the lateral stress will decrease, until its lowest value is
reached, as defined by the active coefficient Ka. In advanced computations a relation as
shown in Figure 34-9 may be used to determine the displacement of the structure and the
stresses against it. In many cases it can be argued that one of the two limiting values can be
considered as appropriate, see the next chapter.
The horizontal stress at failure ('v*Kp) is also used in some countries (France) as the tip
resistance of a foundation pile.

34.6 Groundwater
In the case of a soil saturated with water it should be noted that the Mohr-Coulomb criterion
describes the limiting states of effective stress in the soil. The correct procedure should be
that the smallest and the largest horizontal stress can be deduced from the vertical effective
stress, using the active or passive stress coefficient. The horizontal total stress can be
obtained by adding the pore water pressure to the horizontal effective stress.

157
Figure ‎34-10. Groundwater in the soil.

As an example a retaining wall is shown in Figure 34-10. The wall is assumed to be 8 metre
high, with the groundwater level at 2 metre below the soil surface. The question is to
determine the horizontal total stress at a depth of 8 m, assuming that the soil is sand, with c =
0 kPa and   30 , for the case of active earth pressure. The volumetric weight of the soil is
16 kN/m3 when dry, and 20 kN/m3 when saturated with water. It is assumed that there is no
capillary rise in the sand. The vertical total stress at a depth of 8 m is the weight of 2 m dry
soil and 6 m of saturated soil, which gives  zz = 152 kPa. Because the pore pressure at that
depth is 60 kPa is, the vertical effective stress is  zz = 92 kPa. The active stress coefficient is
K a  13 , so that the horizontal effective stress is  xx = 31 kPa. The total stress is found by
adding the pore pressure, i.e.  xx = 91 kPa. It is interesting to note that this consists of 66 %
water pressure, and only 34 % of effective stress. This illustrates that the contribution of the
pore water pressure may be very large. This is especially true in the case of active earth
pressure.

158
35 Coulomb
Long before the analysis of Rankine the French scientist Coulomb presented a theory on
limiting states of stress in soils (in 1776), which is still of great value. The theory enables the
determination of the stresses on a retaining structure for the cases of active and passive earth
pressure. The method is based upon the assumption that the soil fails along straight slip
planes.

Figure ‎35-1. Active earth pressure.

35.1 Active earth pressure


For the active case (a retreating wall) the procedure is illustrated in Figure 35-1. It is assumed
that in case of displacement of the wall towards the left, a triangular wedge of soil will slide
down, along a straight slip plane. The angle of the slope with the vertical direction is denoted
by  . It is also assumed that at the moment of sliding, the weight of the soil wedge is just in
equilibrium with the forces on the slip surface and the forces on the wall. For reasons of
simplicity, it is assumed, at least initially, that the force between the soil and the wall (Q) is
directed normal to the surface of the wall, i.e. shear stresses along the wall are initially
neglected. In later chapters such shear stresses will be taken into account as well. The
purpose of the analysis is to determine the magnitude of the force Q. The principle of
Coulomb’s method is that it is stated that the wall must be capable of withstanding the force
Q for all possible slip planes. Therefore, the slip plane that leads to the largest value of Q is
to be determined. The various slip planes are characterized by the angle  , and this angle
will be determined such that the maximum value of Q is obtained. The starting point of the
analysis is the weight of the soil wedge (W), per unit width perpendicular to the plane shown
in the figure,
W  12  h2 tan  . (35.1)
This weight must be balanced by the horizontal force Q (horizontal because the wall has been
assumed to be perfectly smooth), and the forces N and T on the slip plane. The direction of
the shear force T is determined by the assumed sliding direction, with the soil body moving
down, in order to follow the motion of the wall to the left. Furthermore, because the length of
the slip plane is h cos  ,
ch
T  N tan . (35.2)
cos 
The equations of equilibrium of the soil body, in horizontal and vertical direction, are
Q  T sin   N cos  0, (35.3)

159
W  N sin   T cos  0. (35.4)
With eq. (35.2) the shear force T can be eliminated. This gives
N
Q cos(   )  ch tan  , (35.5)
cos 
N
W sin(   )  ch. (35.6)
cos 
From these two equations the normal force N can be eliminated,
cos(   ) cos 
Q W  ch . (35.7)
sin(   ) cos  sin(   )
With eq. (35.1) this gives
sin  cos(   ) cos 
Q  12  h2  ch . (35.8)
cos  sin(   ) cos  sin(   )
This equation expresses the force Q as a function of the angle θ. The relation is rather
complex (the angle  appears in 6 places), so it does not seem to be particularly easy to
determine the maximum value. However, the expression can be simplified by using various
trigonometric relations, such as sin  cos(   )  cos sin(   )  sin  . This gives
1
 h2 sin   ch cos 
Q  12  h2  2
. (35.9)
cos  sin(   )
Now the angle  appears in 2 places only, in the denominator of the second term. The
maximum value of Q can be determined by the maximum value of the function
f ( )  cos sin(   ).
The maximum of this function occurs when its derivative with respect to  is zero.
Differentiation gives
df
 cos(2   ),
d
and a second differentiation gives
d2 f
 2sin(2   ).
d 2
It now follows that df d  0 if 2    12  or

df
0:   14   12 . (35.10)
d
Then d 2 f d  2 , so that the function f indeed has a maximum for this value of  . This
2

means that the horizontal force Q also has a maximum for   14   12  This maximum value
is, after some elaboration,
  14   12  : Q  12  h2 Ka  2ch Ka , (35.11)
in which Ka is the coefficient of active earth pressure, defined before,

160
1  sin 
Ka  . (35.12)
1  sin 
These results are in full agreement with the results obtained in the previous chapter on active
earth pressure, see equation (34.11). The value for the horizontal force Q corresponds to the
sliding of a wedge of soil along a slip plane making an angle 14   12  with the vertical
direction. These are just the planes shown in Figure 34-3. In the previous chapter it was
found that along these planes the stresses first reach the Mohr-Coulomb envelope. It might be
noted that in this analysis possible tension cracks in the soil have been ignored.
Coulomb’s method contains a possibly confusing step, in the procedure of maximizing the
force Q to determine the appropriate value of the angle  . This might suggest that the
procedure gives a high value for Q, whereas in reality the value of Q indicates the smallest
possible value of the horizontal force against a retaining wall, as can be seen from Rankine’s
analysis. The confusion is caused by the assumption in Coulomb’s analysis that the soil slides
along a slope defined by an angle  with the vertical, and not along any other plane. For a
value of  other than the critical value   14   12  , the force Q may be smaller, but in that
case there will be other planes on which the maximum shear stress exceeds Coulomb’s
maximum  f  c   n tan  . In the analysis it ought to be investigated whether the assumed
slip plane, at an angle  , is indeed the most critical plane. This is the case only if
  14   12  , as can be seen from Rankine’s analysis. In that analysis the stresses on all
planes are considered, by using Mohr’s circle. In Coulomb’s analysis the stresses on planes
other than the assumed sliding plane are not considered at all.
In engineering practice, the horizontal stress against a retaining wall or a sheet pile wall is
often calculated using the active stress coefficient Ka. This may seem on the unsafe side,
because it gives the smallest possible value of the horizontal stress, and will occur only in
case of failure of the soil. The application is based upon the following argumentation. It is
admitted that the analysis following Rankine or Coulomb, for the active stress state, yields
the smallest possible value for the lateral force. In reality the lateral force may be higher,
especially if the foundation of the retaining wall is stiff and strong. If the lateral force is so
large that the wall’s foundation cannot withstand the force, it will deform, away from the soil.
During that deformation the lateral force will decrease. Eventually this deformation may be
so large that the active state of stress is attained. If the foundation and the structure are strong
enough to withstand the active state of stress, the deformations will stop as soon as this active
state is reached. These deformations may be large, but the structure will not fail. Thus, the
structure will be safe if it can withstand active earth pressure, provided that there is no
objection to a considerable deformation. For instance, the pile foundation of a quay wall in a
harbour can be designed on the basis of active earth pressure against the quay wall, if it is
accepted that considerable lateral deformations (say 1 % or 2 % of the height of the wall) of
the quay wall may occur. If this is undesirable, for aesthetical reasons or because other
structures (the cranes) might be damaged by such large deformations, the foundation must be
designed for larger lateral forces. This will mean that many more piles are required.

35.2 Passive earth pressure


For the case of passive earth pressure (i.e. the case of a wall that is being pushed towards the
soil mass, by some external cause) Coulomb’s procedure is as follows, see Figure 35-2.
Because the wedge of soil in this case is being pushed upwards, the shear force T will be
acting in a downward direction. The weight of the wedge is, as in the active case,
W  12  h2 tan  . (35.13)

161
Figure ‎35-2. Passive earth pressure.

The equations of equilibrium in x- and z-direction now are

Q  T sin   N cos  0, (35.14)


W  N sin   T cos  0. (35.15)
After elimination of T and N from the equations, and some trigonometric manipulations, the
force Q is found to be
 h2 sin   ch cos 
1
Q  h 
1 2 2
. (35.16)
2
cos  sin(   )
Again this force appears to be dependent on the angle  . The minimum value of Q occurs if
the function
f ( )  cos sin(   ),
has its largest value. Differentiation gives
df
 cos(2   ),
d
It now follows that df d  0 if 2    12  , or

df
0:   14   12 . (35.17)
d
Then d 2 f d  2 , and the function f indeed has a maximum for that value of  . That
2

means that the horizontal force Q has a minimum.


This minimum is
  14   12  : Q  12  h2 K p  2ch K p , (35.18)
where Kp is the passive earth pressure coefficient, defined before,
1  sin 
Kp  . (35.19)
1  sin 
Again, the result is in complete agreement with the value obtained in Rankine’s analysis.
Coulomb’s procedure appears to lead to the maximum (passive) earth pressure.

The effect of active and passive behaviour and the different slopes of the failure planes can
easily be shown in a small test on a clay sample. By making a small excavation in a piece of

162
clay, supported by a simple sheet pile, and loading this in an ordinary centrifuge of a washing
machine, both planes become visual, see Figure 35-3.

Figure ‎35-3. Active (left) and passive (right) failure plane in a clay sample.

Coulomb’s method can easily be extended to more general cases. It is possible, for instance,
that the surface of the wall is inclined with respect to the vertical direction, and the soil
surface may also be sloping. Also, the soil may carry a given surface load. For all these cases
the method can easily be extended. The general procedure is to assume a straight slip plane,
consider equilibrium of the sliding wedge, and then maximizing or minimizing the force
against the wall. Analytical, graphical and numerical methods have been developed. In the
next chapter a few tables are presented.

163
36 Tables for lateral earth pressure
The computation of lateral earth pressure against retaining walls is such an important
problem of soil mechanics that tables have been produced for its solution, all on the basis of
Coulomb’s method. These tables can be found in many handbooks, such as the German
“Grundbau Taschenbuch”. Following Coulomb these tables apply to soils without cohesion (
c  0 ), that is, for sand or gravel. In this chapter some tables are given for the active and the
passive earth pressure against a retaining wall, with a surface that is practically vertical, and a
sloping soil surface.

36.1 The problem


The general problem considered in this chapter concerns a retaining wall, having a surface
inclined at an angle  with the horizontal direction. The soil surface is horizontal, or it may
be sloping at an angle  with the horizontal direction, see Figure 36-1. The wall may be
perfectly smooth, or it may have a certain friction, so that the direction of the force Q is at an
angle  with the direction normal to the wall. The friction angle  is supposed to be given.
Because the wall often is rather smooth, its value is often taken somewhat smaller than the
friction angle of the soil itself, say   23  . The angle  is considered positive in the active
case, illustrated in Figure 36-1, in which the sliding soil wedge is expected to slide in
downward direction, along the surface of the wall. In the case of passive earth pressure it can
be expected that the soil will move in upward direction along the surface of the wall. The
angle  should then be given a negative value.

Figure ‎36-1. Horizontal earth pressure.

The tables record the values of the coefficient K in the formula


Q  12 K h2 . (36.1)
This coefficient would be equal to 1 in the case of a fluid against a vertical wall.
It should be noted that Q is the total force. The angle of this force with the vertical direction
is    . The horizontal component of this force is
Qh  Q sin(   ). (36.2)
If the tables are used to determine the horizontal force, the multiplication by the factor
sin(   ) should be performed by the user.

164
The values of the active coefficient Ka were already calculated by Coulomb. He obtained
sin 2 (   )
Ka  2
. (36.3)
sin  sin(   ) 1  {sin(   )sin(   )} {sin(   )sin(   )} 
2

For the passive case the formula is


sin 2 (   )
Kp  2
. (36.4)
sin 2  sin(   ) 1  {sin(   )sin(   )} {sin(   )sin(   )} 

It may be mentioned that the active coefficients in the tables may be somewhat too small, and
that the passive coefficients may be too large. This may be because in reality the soil has not
yet reached a critical state, but also because in Coulomb’s method only straight slip surfaces
are considered. In reality, a curved slip surface, for instance a circular slip surface, may give
a higher active earth pressure or a lower passive pressure. This last possibility can easily be
imagined: if the soil can fail along a circular slip surface for a force that is smaller than the
critical straight sliding plane, there is no reason why the soil would not fail along the circular
slip surface. A chain breaks if the weakest link fails.
It has been found that using circular slip surfaces leads to a very small increase of the active
coefficients. The passive coefficients, however, may become considerably lower when
circular slip surfaces are also taken into account. In particular, all values larger than 10 in the
tables are unreliable. This can be very dangerous, for instance when calculating the
maximum holding force of an anchor. This may be severely overestimated by using tables
based upon straight slip planes only (as in this chapter). More reliable values are given in the
tables in “Grundbau Taschenbuch, Teil 1”.
It should be noted that in some tables the definitions (and the notations) of the angles , 
and  differ from the definitions used here. Great care should be used when taking values
from an unfamiliar table.

36.2 Example
As an example, the case of a wall at an inclination of 80º is considered. The slope of the soil
is 10º, see Figure 36-2. The soil is sand, with  = 30º, and the friction angle between the wall
and the soil is  = 20º. The problem is to determine the horizontal component of the force
against the wall, in the case of active earth pressure.

Figure ‎36-2. Example: Active earth pressure.

165
In this case Table 36-2 gives Ka  0.438 , so that the force on the wall is Q  0.219 h2 . Its
horizontal component is, with (36.2), Qh  0.190 h2 .

Figure ‎36-3. Example: Passive earth pressure.

In the case of passive earth pressure, when the wall is moving to the right, it will push the soil
wedge up. It can be expected that the wall will then exert a shear force on the wall in a
downward direction, with the value of  being negative,  = -20º, see Figure 36-3. In this case
Table 36-3 gives K p  7.162 . The force on the wall then is Qh  3.527 h2 , because in this
case  = 100º.

36.3 Tables
On the following pages some values of Ka and Kp are given in tabular form.

166
  90,   0 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 0.704 0.589 0.49 0.406 0.333 0.271 0.217 0.172
5° 0.662 0.556 0.465 0.387 0.319 0.26 0.21 0.166
10° 0.635 0.533 0.447 0.373 0.308 0.253 0.204 0.163
15° 0.617 0.518 0.434 0.363 0.301 0.248 0.201 0.160
20° 0.607 0.508 0.427 0.357 0.297 0.245 0.199 0.160
25° 0.604 0.505 0.424 0.355 0.296 0.244 0.199 0.160
30° 0.606 0.506 0.424 0.356 0.297 0.246 0.201 0.162
  90,   10 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 0.970 0.704 0.569 0.462 0.374 0.300 0.238 0.186
5° 0.974 0.679 0.547 0.444 0.359 0.289 0.230 0.180
10° 0.985 0.664 0.531 0.431 0.350 0.282 0.225 0.177
15° 1.004 0.655 0.522 0.423 0.343 0.277 0.221 0.174
20° 1.032 0.654 0.518 0.419 0.340 0.275 0.220 0.174
25° 1.070 0.658 0.518 0.419 0.340 0.275 0.221 0.175
30° 1.120 0.669 0.524 0.422 0.343 0.278 0.223 0.177
  90,   20 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 0.883 0.572 0.441 0.344 0.267 0.204
5° 0.886 0.558 0.428 0.333 0.259 0.199
10° 0.897 0.549 0.42 0.326 0.254 0.195
15° 0.914 0.546 0.415 0.323 0.251 0.194
20° 0.940 0.547 0.414 0.322 0.250 0.193
25° 0.974 0.553 0.417 0.323 0.252 0.195
30° 1.020 0.565 0.424 0.328 0.256 0.198
  90,   30 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 0.750 0.436 0.318 0.235
5° 0.753 0.428 0.311 0.229
10° 0.762 0.423 0.306 0.226
15° 0.776 0.422 0.305 0.225
20° 0.798 0.425 0.305 0.225
25° 0.828 0.431 0.309 0.228
30° 0.866 0.442 0.315 0.232

Table ‎36-1: Active earth pressure coefficient, Ka.

167
  80,   0 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 0.757 0.652 0.559 0.478 0.407 0.343 0.287 0.238
5° 0.720 0.622 0.536 0.460 0.393 0.333 0.280 0.233
10° 0.699 0.603 0.520 0.448 0.384 0.326 0.275 0.229
15° 0.687 0.592 0.511 0.441 0.378 0.323 0.273 0.228
20° 0.684 0.588 0.508 0.438 0.377 0.322 0.273 0.229
25° 0.689 0.591 0.510 0.440 0.379 0.325 0.276 0.232
30° 0.702 0.600 0.517 0.446 0.385 0.330 0.281 0.237
  80,   10 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 1.047 0.784 0.654 0.550 0.461 0.384 0.318 0.261
5° 1.067 0.766 0.636 0.534 0.448 0.374 0.311 0.255
10° 1.097 0.759 0.626 0.524 0.440 0.368 0.307 0.253
15° 1.138 0.759 0.622 0.520 0.437 0.366 0.305 0.252
20° 1.191 0.768 0.625 0.521 0.438 0.367 0.306 0.254
25° 1.259 0.785 0.634 0.528 0.443 0.371 0.310 0.257
30° 1.346 0.811 0.650 0.539 0.452 0.379 0.317 0.264
  80,   20 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 1.015 0.684 0.548 0.444 0.360 0.291
5° 1.035 0.676 0.538 0.436 0.354 0.286
10° 1.064 0.674 0.534 0.432 0.351 0.283
15° 1.103 0.679 0.535 0.432 0.350 0.284
20° 1.155 0.690 0.540 0.435 0.354 0.286
25° 1.221 0.708 0.551 0.443 0.360 0.292
30° 1.305 0.734 0.568 0.456 0.370 0.300
  80,   30 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 0.925 0.566 0.433 0.337
5° 0.943 0.563 0.428 0.333
10° 0.969 0.564 0.427 0.332
15° 1.005 0.570 0.430 0.333
20° 1.051 0.582 0.437 0.338
25° 1.111 0.600 0.448 0.346
30° 1.189 0.624 0.463 0.358

Table ‎36-2: Active earth pressure coefficient, Ka.

168
  90,   0 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 1.420 1.698 2.040 2.464 3.000 3.690 4.599 5.828
-5° 1.569 1.901 2.313 2.833 3.505 4.391 5.593 7.278
-10° 1.730 2.131 2.635 3.285 4.143 5.309 6.946 9.345
-15° 1.914 2.403 3.029 3.855 4.976 6.555 8.872 12.466
-20° 2.130 2.735 3.525 4.597 6.105 8.324 11.771 17.539
-25° 2.395 3.151 4.169 5.599 7.704 10.980 16.473 26.696
-30° 2.726 3.691 5.036 7.013 10.095 15.273 24.933 46.087
  90,   10 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 2.099 2.595 3.235 4.080 5.228 6.841 9.204
-5° 2.467 3.086 3.908 5.028 6.605 8.923 12.518
-10° 2.907 3.700 4.783 6.314 8.569 12.076 17.944
-15° 3.456 4.496 5.969 8.145 11.536 17.225 27.812
-20° 4.166 5.572 7.652 10.903 16.370 26.569 48.891
-25° 5.122 7.093 10.181 15.384 25.117 46.474 108.431
-30° 6.470 9.371 14.274 23.468 43.697 102.545 426.159
  80,   0 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 1.363 1.582 1.843 2.156 2.535 3.002 3.587 4.332
-5° 1.480 1.737 2.045 2.418 2.879 3.456 4.193 5.158
-10° 1.600 1.905 2.273 2.725 3.292 4.017 4.966 6.244
-15° 1.732 2.096 2.540 3.094 3.802 4.730 5.981 7.726
-20° 1.883 2.321 2.861 3.549 4.450 5.666 7.363 9.838
-25° 2.060 2.590 3.257 4.127 5.299 6.937 9.329 13.021
-30° 2.274 2.923 3.759 4.881 6.450 8.742 12.286 18.184
  80,   10 :
 \ 10° 15° 20° 25° 30° 35° 40° 45°
0° 1.935 2.308 2.767 3.343 4.079 5.043 6.340
-5° 2.218 2.668 3.233 3.96 4.914 6.201 7.998
-10° 2.541 3.093 3.805 4.742 6.010 7.783 10.372
-15° 2.922 3.614 4.528 5.767 7.504 10.045 13.969
-20° 3.387 4.272 5.474 7.162 9.636 13.465 19.844
-25° 3.975 5.131 6.759 9.148 12.854 19.039 30.500
-30° 4.740 6.295 8.583 12.137 18.084 29.127 53.188

Table ‎36-3: Passive earth pressure coefficient, Kp.

169
37 Sheet pile walls
An effective way to retain a soil mass is by installing a vertical wall consisting of long thin
elements (steel, concrete or wood), that are driven into the ground. The elements are usually
connected by joints, consisting of special forms of the element at the two ends. Compared to
a massive wall (of concrete or stone), a sheet pile wall is a flexible structure, in which
bending moments will be developed by the lateral load. They should be designed so that they
can withstand the largest bending moments. Several methods of analysis have been
developed, of different levels of complexity. The simplest methods, which will be discussed
in this chapter, are based on convenient assumptions regarding the stress distribution against
the sheet pile wall. These methods have been found very useful in engineering practice, even
though they contain some rather drastic approximations.

Figure ‎37-1. Anchored sheet pile wall.

37.1 Homogeneous dry soil


A standard type of sheet pile wall is shown in Figure 37-1. The basic idea is that the pressure
of the soil will lead to a tendency of the flexible wall to displace towards the left. By this
mode of deformation, the soil pressures on the right side of the wall will become close to the
active state. This soil pressure must be equilibrated by forces acting towards the right. A
large horizontal force may be developed at the lower end of the wall, embedded into the soil
on the left side, by the displacement. In this part, passive earth pressure may develop if the
displacements are sufficiently large. The usual schematisation is to assume that on the right
side of the wall active stresses will be acting, and below the excavated soil level at the left
side of the wall passive stresses will develop. As the resulting force of the passive stresses is
below the resulting force of the active stresses, complete equilibrium is not possible by these
stresses alone. This is due to the fact the condition of the equilibrium of moments is not
satisfied. Equilibrium can be ensured by adding an anchor at the top of the wall, on the right
side. This anchor can provide an additional force to the right. Without such an anchor the
sheet pile wall would rotate, until at the extreme lower end of the wall passive earth pressures
would be developed on the right side. With an anchor equilibrium can be achieved, without
the need for very large deformations. It may be noted that the anchoring force can also be
provided by a strut between two parallel walls. This is especially practical in the case of a
narrow excavation trench.
For the sheet pile wall to be in equilibrium the depth of embedment should be sufficiently
large, so a passive zone of sufficient length can be developed. In case of a very small depth,
with a thin passive zone at the toe, the lower end of the wall might be pushed through the
soil, with the structure rotating around the anchor point. The determination of the minimum
depth of the embedment of the sheet pile wall is an important part of the analysis, which will

170
be considered first. For reasons of simplicity it is assumed that the soil is homogeneous, dry
sand. The assumed stress distribution is shown in Figure 37-1.
If the retaining height (the difference of the soil levels at the right and left sides of the wall) is
h, the length of the toe is d, and the depth of the anchor rod is a, then the condition of
equilibrium of moments around the anchor point gives
1
2 Ka (h  d )2 ( 32 h  32 d  a)  12 K p d 2 (h  32 d  a)  0.

It follows that
Kp
(h  d )2 ( 23 h  23 d  a)  d 2 (h  23 d  a). (37.1)
Ka
This is an equation of the third degree equation in the variable d. It can be solved iteratively
by writing
d 2Ka d 1  (d h)  32 (a h)
( )2  (1  )2 . (37.2)
h 3K p h 1  23 (d h)  (a h)

Starting from an initial estimate, for example d h  0 , ever better estimates for d h can be
obtained by substituting the estimated value into the right side of eq. (37.2). This process has
been found to iterate fairly rapidly. About 10 iterations may be needed to obtain a relative
accuracy of 10-6. The results for a series of values of K p K a and a h are recorded in
Table 37-1.

K p Ka
ah 4 6 8 9 10 12 14 16
0.00 0.793 0.550 0.438 0.401 0.371 0.326 0.294 0.269
0.05 0.785 0.545 0.433 0.396 0.367 0.323 0.290 0.265
0.10 0.777 0.539 0.428 0.392 0.363 0.319 0.287 0.262
0.15 0.768 0.532 0.422 0.386 0.385 0.314 0.282 0.258
0.20 0.759 0.524 0.416 0.380 0.352 0.309 0.278 0.254
0.25 0.749 0.516 0.409 0.374 0.346 0.303 0.273 0.249
0.30 0.737 0.507 0.401 0.366 0.339 0.297 0.267 0.243
0.35 0.724 0.496 0.392 0.358 0.330 0.289 0.260 0.237
0.40 0.710 0.484 0.381 0.348 0.321 0.281 0.252 0.229
0.45 0.693 0.470 0.369 0.336 0.310 0.270 0.242 0.220
0.50 0.674 0.454 0.354 0.322 0.296 0.258 0.230 0.209

Table ‎37-1: Depth of sheet pile wall  d h  .

The magnitude of the anchor force can be determined from the condition of horizontal
equilibrium,
T  12 Ka (h  d )2  12 K p d 2 . (37.3)
The values of T Fa are given in Table 37-2. The quantity Fa is the total active force,

171
Fa  12 Ka (h  d )2 . (37.4)
It appears that the anchor carries a substantial part of the total active load, varying from 20%
to more than 50%. The remaining part is carried by the passive earth pressure.
If the length of the sheet pile wall (h + d) and the anchor force are known, the shear force Q
and the bending moment M can easily be calculated, in any section of the wall. For the case
K a  13 , K p  3 and a h  0.2 the results are given in Table 37-3. At the location of the
anchor, the shear force instantly increases by the magnitude of the anchor force. At the top
and at the toe of the wall, both the shear force and the bending moment are zero.

K p Ka
ah
4 6 8 9 10 12 14 16
0.00 0.218 0.244 0.258 0.263 0.267 0.274 0.279 0.283
0.05 0.226 0.254 0.269 0.275 0.279 0.286 0.292 0.296
0.10 0.235 0.265 0.281 0.287 0.292 0.300 0.306 0.310
0.15 0.245 0.277 0.295 0.301 0.306 0.315 0.321 0.326
0.20 0.255 0.290 0.309 0.316 0.322 0.331 0.338 0.344
0.25 0.267 0.305 0.326 0.334 0.340 0.350 0.358 0.364
0.30 0.280 0.321 0.345 0.353 0.360 0.371 0.380 0.387
0.35 0.294 0.340 0.366 0.375 0.383 0.395 0.405 0.413
0.40 0.311 0.361 0.390 0.401 0.409 0.423 0.434 0.443
0.45 0.329 0.386 0.419 0.431 0.441 0.456 0.469 0.478
0.50 0.351 0.415 0.453 0.466 0.478 0.496 0.510 0.521

Table ‎37-2: Anchor force T Fa  .

The largest bending moment, which determines the profile of the pile sheets, is 0.032 h3 .
The results of this example are shown in graphical form in Figure 37-2.
A simple verification of the order of magnitude of the results can be made by considering the
sheet pile wall as a beam on two supports, say at z h  0.2 and at z h  1.2 . The length of
the beam then is h, and the average load is Ka (0.7h) . If this load is thought to be distributed
homogeneously along the beam, the maximum bending moment would be
M  qh2 8  0.029 h3 , which is reasonably close to the true value given above.

172
Figure ‎37-2. Shear force and Bending moment.

If the sheet pile wall is designed on the basis of the maximum bending moment, there is no
safety against failure. In order to increase the safety of the structure the passive earth pressure
is often reduced, by using a conservative value for Kp. The tables then remain valid, but the
result will be a somewhat greater length, as can be seen from Table 37-1.
If K p K a is taken smaller, the needed value of d h will be larger. In the next chapter a more
advanced method to reduce the risk of failure will be presented.

z h f h Q  h2 M  h3
0.00000 0.00000 0.00000 0.00000
0.10000 -0.03333 -0.00167 -0.00006
0.19999 0.06666 -0.00667 -0.00044
0.20001 0.06667 0.09381 -0.00044
0.30000 0.10000 0.08548 0.00855
0.40000 0.13333 0.07381 0.01654
0.50000 0.16667 0.05881 0.02320
0.60000 0.20000 0.04048 0.02819
0.70000 0.23333 0.01881 0.03119
0.80000 0.26667 -0.00619 0.03184
0.90000 0.30000 -0.03452 0.02984
1.00000 0.33333 -0.06619 0.02483
1.10000 0.06667 -0.08619 0.01699
1.20000 -0.20000 -0.07952 0.00848
1.30000 -0.46667 -0.04619 0.00197
1.38047 -0.68125 0.00000 0.00000

Table ‎37-3: Sheet pile wall.

Computer programs for the calculation of the minimum length of the sheet pile wall, the
corresponding anchor force, and the distribution of shear forces and bending moments, can
be based on the previous equations. The output will depend on the values of the input data,

173
like Ka, Kp and a h . The program first calculates the values of the depth of embedment d h
and the anchor force T, and then gives, for an arbitrary value of z h , to be given by the user,
the shear force Q and the bending moment M. The program can be improved in many ways,
especially by adding more advanced forms of input and output, such as graphs of the shear
force and the bending moment, to be shown on the screen or on a printer.

37.2 Pore pressures


In the previous sections the soil was assumed to be dry, for simplicity. In general, the soil
may consist of soil and water however, and the excavation may even contain free water. Thus
the general problem of a sheet pile wall should take into account the presence of groundwater
in the soil. The failure of soils, as described by the Mohr-Coulomb criterion, for instance,
refers to effective stresses, so the relations formulated above for the earth pressure
coefficients Ka and Kp, should be applied to effective stresses only. This means that the
vertical effective stresses should be calculated first, before the horizontal effective stresses
can be determined. The horizontal total stresses can then be determined in the next step by
adding the pore pressure.

The general procedure for the determination of the horizontal stresses is as follows.

1. Determine the total vertical stresses, from the surcharge and the weight of the
overlying soil layers.

2. Determine the pore water pressures, on the basis of the location of the phreatic
surface. If the pore pressures can be assumed to be hydrostatic (if there is no vertical
groundwater flow) these can be determined from the depth below the phreatic surface.
Above the phreatic surface the pore pressures may be negative in case of a soil with a
capillary rise.

3. Determine the value of the vertical effective stress, as the difference of the vertical
total stress and the pore pressure. If the result of this computation is negative, it may
be assumed that a crack will develop, as tension between the soil particles usually is
impossible. The vertical effective stress then is zero.

4. Determine the horizontal effective stress, using the appropriate value of Ka or Kp at


the depth considered, and, if applicable, the local value of the cohesion c.

5. Determine the horizontal total stress by adding the pore pressure to the horizontal
effective stress.

The algorithm for this procedure can be summarised as:


 zz  qz    dz,
p   w ( z  zw ), if z  zw  hc then p  0,
 zz   zz  p, if  zz  0 then  zz  0,

 xx  K zz  2c K ,
 xx   xx  p.

174
In these equations it has been assumed that the phreatic level is located at a depth z  zw , and
that in a zone of thickness hc above that level capillary water is present in the pores. Above
the level z  zw  hc there is no water in the pores, which can be expressed by p  0 . It has
also been assumed, in this procedure, that the particles cannot transmit tensile forces. It may
also be noted that in computations such as these, open water above the soil, may also have to
be considered as soil, having a volumetric weight  w . The pore pressures in such a water
layer will be found as zero, and the horizontal total stress will automatically be found equal
to the vertical total stress. For the analysis of the forces on a wall, these forces are essential
parts of the analysis.
For the analysis of a sheet pile wall, the stress calculation must be performed for both sides of
the wall separately, as on the two sides, the soil levels and the groundwater levels may be
different.
An example is shown in Figure 37-3. In this case an excavation of 6 m depth is made into a
homogeneous soil. On the right side the groundwater level is located at a depth of 1 m below
the soil surface, and on the left side the groundwater level coincides with the bottom of the
excavation. For simplicity it is assumed that on both sides of the sheet pile wall the
groundwater pressures are hydrostatic. This might be possible if the toe of the wall reaches
into a clay layer of low permeability. Otherwise, the groundwater pressures should include
the effect of a groundwater movement from the right side to the left side. That complication
is omitted here. An anchor has been installed at a depth of 0.50 m, at the right side. The
length of the wall is initially unknown, but is assumed to be 9 m, for the representation of the
horizontal stresses. The soil is homogeneous sand, having a dry volumetric weight of 16
kN/m3, a saturated volumetric weight of 20 kN/m3. It is assumed that for this sand
Ka  0.3333 , K p  3.0 , c = 0 and hc  0 .

Figure ‎37-3. Example: The influence of groundwater.

In order to present the stresses against the wand, the simplest procedure is to calculate these
stresses in a number of characteristic points. At a depth of 1 m, for instance, at the right side,
the vertical total stress is  zz =16 kPa . Because the pore pressure is zero at that depth the
horizontal effective stress is  xx  5.3 kPa , and the horizontal total stress is equal to that
value, because p  0 . At a depth of 9 m, the total stress is larger by the weight of 8 m
saturated soil, so that  zz  176 kPa . At that depth the pore pressure is p = 80 kPa, and the
horizontal effective stress is now  zz  96 kPa . Because Ka  0.3333 the horizontal effective
  32 kPa . Finally, the horizontal total stress is  xx  112 kPa .
stress is  xx

175
At the left side of the wall all stresses are zero down to the level of the bottom of the
excavation, at 6 m depth. At a depth of 9 m:  zz  60 kPa and p = 30 kPa. This gives
 zz  30 kPa and, because K p  3 ,  xx  90 kPa . The horizontal stress is obtained by
adding the pore pressure, i.e.  xx  120 kPa .
Even in this simple case, of a homogeneous soil, the determination of the horizontal loads on
the wall is not a trivial problem. In many problems of engineering practice, the analysis may
be much more complicated, as the soil may consist of layers of different volumetric weight
and composition, with variable values of the coefficients Ka and Kp. This may lead to
discontinuities in the distribution of the horizontal stress. The groundwater pressures also
need not be hydrostatic. In the case of a permeable soil the determination of the groundwater
pressures may be a separate problem.

Another difficulty can be seen in Figure 37-3. Because the water level left of the sheet pile
wall is different from the water level on the right, a flow under the sheet pile wall will occur.
Due to this, the pore pressures around the sheet pile tip will be levelled.
The length of the sheet pile wall is initially unknown. It can be determined by requiring that
equilibrium is possible with the toe of the wall being a free end, with Q = 0 and M = 0. As in
the simple case considered before, see Figure 37-1, the length can be determined from the
condition of equilibrium of moments with respect to the anchor point. The simplest procedure
is to first assume a certain very short depth of the embedment, with full passive pressures at
the left side, then calculating the bending moment at the toe, and then gradually reducing the
embedment depth until this bending moment is zero.

176
38 Blum
In the previous chapter a procedure has been presented for the determination of the minimum
length of a sheet pile wall, required to ensure equilibrium. This method is such that whenever
the wall is shorter than that minimum length, no equilibrium is possible, and the wall will
certainly fail. This suggests that it is advisable to choose the length of the wall somewhat
larger than the minimum length, as a total failure of the wall would be disastrous. If the
length is taken somewhat larger than required, the bending moments may be reduced. A
method of analysing the deformation and bending of the wall has been developed by Blum.
This method is presented in this chapter.

If the length of the sheet pile wall is somewhat larger than strictly necessary to ensure
equilibrium, the passive earth pressure need not be developed over the entire length of the
embedded part of the wall. It may be expected that the pressures against the wall will be of
the form shown in Figure 38-1. Due to the extra length of the sheet pile wall the toe will act
as a clamped edge, in which the lowest part may have a tendency to move to the right,
building up a pressure towards the left. Together with the incomplete passive pressure
towards the right this will constitute the clamping moment. Blum suggested to schematise the
loads on the wall as shown in the right half of the figure. The resulting force R is equivalent
to the pressure to the right at the extreme lower part of the wall. Its precise distribution is left
undetermined. The toe of the sheet pile wall is now assumed to be a clamped edge, and it is
also assumed that at the toe the bending moment is zero, but a shear force (of magnitude R) is
allowed. In order that this force may indeed develop, and that there is enough material to
form a clamped boundary, the actual length should be somewhat larger than assumed in the
schematisation: usually the embedment depth is taken 20 % larger than calculated.

Figure ‎38-1.‎Blum’s‎schematisation.

One of the ideas behind Blum’s schematisation is that the clamping moment will probably
lead to a reduction of the bending moments in the sheet pile wall, so that a lighter profile may
be used. Thus the additional costs involved by taking a longer sheet pile wall are somewhat
balanced by a lighter profile. That this is acceptable can be argued by noting that a failure by
a wall that is too short is indeed disastrous, but in case of failure by exceeding the maximum
bending moment, some additional strength is available beyond the onset of plastic
deformation of the steel. If a plastic deformation in bending is developed, the bending
moment will at least be constant, and may even increase somewhat. Also, the soil pressures
may be redistributed by the large deformations.

The basic principle of Blum’s method of analysis is that the sheet pile wall is considered as
fully clamped at its toe, with the additional condition that the bending moment at the toe is

177
zero. The shear force, however, in general will be unequal to zero. This shear force is
supposed to be the resultant force of the stresses in the vicinity of the toe, including some
length below the toe. The clamping of the edge is supposed to be so strong that the
displacement and the rotation (which is the first derivative of the displacement) are zero, and
even the second derivative is zero, so that the bending moment is zero. There are three
equations (horizontal equilibrium, displacement and moment equilibrium) with three
unknowns (d, R and T). The length of the wall will be determined by the conditions of
equilibrium, with active soil stresses on the high side and full passive stresses on the low
side, and the condition that the horizontal displacement is zero at the level of the anchor. The
procedure can best be illustrated by means of an elementary example.

Figure ‎38-2. Example.

The example refers to a sheet pile wall retaining a height h of homogeneous saturated soil,
see Figure 38-2. To enable an analytical solution it is assumed that on the two sides of the
wall the groundwater table coincides with the soil surface. To further simplify the problem
the anchor is supposed to be acting at the top of the wall. The embedment depth d is
unknown. This is one of the parameters that have to be determined by the analysis.
At the active side of the wall the vertical total stress is
 zz   z, (38.1)
in which  is the volumetric weight of the saturated soil. The pore pressures are
p   w z, (38.2)
so that the effective stresses are
 zz  (   w ) z. (38.3)
The horizontal effective stresses now are, for a cohesionless soil with c  0 ,
 xx  Ka (   w ) z. (38.4)
The horizontal total stresses are obtained by adding the pore pressures,
 xx  [ Ka (   w )   w ]z. (38.5)
This can also be written as
 xx  Ka z, (38.6)
where

178
Ka  Ka (1   w  )   w  . (38.7)
If Ka  0.3333 and  w   0.5 , then Ka  0.6667 . It should be noted that the simple
expression (38.6), linear in z, is valid only if the soil is homogeneous, with c  0 , and if the
groundwater table coincides with the soil surface. In a more general case the computation of
the horizontal total stresses proceeds in exactly the same way, but the result cannot be
expressed in the simple form of eq. (38.6).
In the same way the horizontal stresses at the passive side, for z  h , can be determined. The
result is
 xx  K p ( z  h), (38.8)
where
K p  K p (1   w  )   w  . (38.9)
If K p  3.0 and  w   0.5 , then K p  2.0 .
The resulting active and passive forces are
Fa  12 Ka (h  d )2 ,

Fp  12 K p d 2 .

The condition that the bending moment at the toe of the sheet pile wall must be zero, at the
depth of the clamped edge, i.e. the point of application of the force R, give
T (h  d )  16 Ka (h  d )3  16 K p d 3 . (38.10)
For the computation of the horizontal displacement of the top of the sheet pile wall (which
must be zero), the contribution of the three forms of loading can best be considered
separately, see Figure 38-3.
The first loading case is the anchor force T, acting at the top of the sheet pile wall. This force
leads to a displacement of the top of magnitude
T ( h  d )3
u1  . (38.11)
3EI
This is a well-known basic problem from applied mechanics.
For the case of a triangular load f  az on a clamped beam of length l, the loading case in
the central part of Figure 38-3, the displacements can be found using the classical theory of
bending of beams, from applied mechanics. By integrating the differential equation
EId 4u dz 4  f , with the boundary conditions that at the top the bending moment and the

Figure ‎38-3.‎Loads‎on‎the‎clamped‎wall‎in‎Blum’s‎schematisation.

179
shear force are zero, whereas at the toe the horizontal displacement u and its first derivative
(the rotation) are zero, the displacement of the top can be obtained as
al 5
u0  . (38.12)
30 EI
The rotation of the top is found to be
al 4
0  . (38.13)
24 EI
Using these formulas the horizontal displacement of the top of the sheet pile wall caused by
the active soil pressure on the right side is, with eq. (38.6) and (38.12),
K a (h  d )5
u2   . (38.14)
30 EI
The minus sign indicates that this displacement is directed towards the left.
The displacement caused by the passive soil pressures at the left side of the sheet pile wall, as
described by eq.(38.8), are found to be
K p d 5 K p d 4 h
u3   . (38.15)
30 EI 24 EI
The first term in this expression is the displacement at the top of the load, the second term is
the additional displacement due to the rotation at the top of the load. Together these two
quantities constitute the displacement at the top of the sheet pile wall. The upper, unloaded
part of the wall, does not deform in this loading case.
The sum of the three displacements (38.11), (38.14) and (38.15) must be zero. This gives,
with (38.10), and after multiplication by EI K p ,

K a (h  d )5 d 3 (h  d )2 K a (h  d )5 d 5 d 4 h
      0,
K p 18 18 K p 30 30 24
or, after some rearranging of terms,
d 8( K a K p )(1  d h)5
( )3  . (38.16)
h 20(1  d h)2  15 d h  12(d h)2

From this equation the value of d h can be solved iteratively, using an initial estimate,
possibly simply d h  0.0 .
The computations can be made by a computer program based on the previous equations, only
requesting the input of the volumetric weights of water and (saturated) soil, and the values of
the active and passive pressure coefficients, and then computes the values of d h and T  h2
, using the equations (38.16) and (38.10). For the case that  w  10 kN/m3 ,   20 kN/m3 ,
K a  1 3 and K p  3 , the results will be: d h  1.534 and T  h2  0.239 . It appears that in
this case the sheet pile wall needs a rather long embedment depth (more than 1.5 times the
retaining height). This is the price that has to be paid for a more favourable distribution of the
bending moments. The profile of the steel elements can be somewhat lighter, but the length is
considerably larger than in the simple method of the previous chapter.

180
Figure ‎38-4. Shear force and bending moment.
The distribution of the shear force and the bending moment is shown in Figure 38-4. The
shear force at the top is the anchor force. The value at the toe is Blum’s concentrated force R.
It appears that this force results in a reduction of the bending moments in the sheet pile wall,
as mentioned before. For the determination of the profile of the wall it is favourable that the
positive and negative bending moments are of the same order of magnitude.
The results of the computations for a number of values of the earth pressure coefficients Ka
and Kp are given in Table 38-1. It has been assumed that the volumetric weight of the water is
 w  10 kN m3 , and that the volumetric weight of the saturated soil is   20 kN m3 , a
common value.

 Ka Kp d h T  h2
10º 0.7041 1.4203 5.228 0.881
15º 0.5888 1.6984 3.406 0.554
20º 0.4903 2.0396 2.481 0.394
25º 0.4059 2.4639 1.917 0.300
30º 0.3333 3.0000 1.534 0.239
35º 0.2710 3.6902 1.255 0.196
40º 0.2174 4.5989 1.040 0.165
45º 0.1716 5.8284 0.868 0.141

Table ‎38-1:‎Blum’s‎method‎for‎homogeneous‎soil.

The concentrated force R is an essential element in Blum’s method. It should be remembered


that this force actually represents the distributed load at the extreme toe of the sheet pile wall,
which is produced by the deformation of the sheet pile wall. For the generation of this
concentrated force the wall should be given some additional length, by choosing the length of
the wall somewhat larger than the theoretical value computed in the analysis. It is often
assumed that the length of the embedment depth (the distance d in the example) should be
taken 10 % or 20 % larger than computed. All this leads to a wall of considerable length. This
is the price that has to be paid for the advantages of Blum’s analysis: a lighter profile, and
small displacements.
It may be noted that the example considered in this chapter is perhaps a very unfavourable
case: the level of groundwater at the right side is very high, and on the left side it is very low.
In the next chapter a more general method will be described. In more general cases it is
observed that Blum’s method leads to long sheet pile walls. The safety is large, but at a price.

181
39 Sheet pile wall in layered soil
For a sheet pile wall in a layered soil, the method of analysis is the same as for a wall in
homogeneous soil, as considered in the previous chapter. The main difference is that the
computation of the horizontal stresses against the wall is more complicated. The computation
can best be performed using a computer program. In this chapter a simple program is
discussed, using Blum’s method.

Figure ‎39-1. Layered soil.

The complications are that the weight and the properties of the various layers may be
different, and the zero level of the groundwater may also be different for each layer. The
simplest approach is to consider the determination of the horizontal stresses against the wall
as a separate problem, which precedes the analysis of the sheet pile wall. In principle, these
stresses can easily be determined by analysing the stresses from the top of the soil downward,
in each step adding the weight, and using the appropriate values of the lateral stress
coefficients. The horizontal effective stress follows from  zz   zz  p , and the horizontal
 follows from the formula for active or passive earth pressure. The
effective stress  xx
  p . At the
horizontal total stress finally is obtained by adding the pore pressure,  xx   xx
interfaces between succeeding layers the horizontal total stress may be discontinuous,
because the stress coefficients may be discontinuous.

39.1 Computer program


The computations start always with the input data. A program needs as well as general data
like the anchor depth, the excavation depth and the number of layers. It also requires data for
each layer, such as the thickness of the layer, the cohesion, the coefficient of active earth
pressure, the coefficient of passive earth pressure, the volumetric weight of the soil in dry
condition, the volumetric weight of the soil in saturated condition, and the zero level of the
groundwater, considered with respect to the top of the wall. This must be given for both the
left and the right sides of the wall.
The computer program can start with an initial assumption of the length of the sheet pile
wall, namely the sum of all layer thicknesses. The distribution of the loads on the two sides
of the wall, active on the right side and passive on the left side, is then calculated for a large

182
number of points, from the data defining the layer thicknesses, their weight, the stress
coefficients and the applicable depth of the groundwater. The stresses can be cut off for
tension. The program will find a resulting load. This will be the difference of the soil
pressures at the right and the left sides. Next the shear force and the building moment are
calculated numerically, ignoring the anchor force, which is unknown at that stage.
The anchor force T is calculated from the condition that the bending moment is zero at the
toe of the wall. Next the rotation  and the horizontal displacement u are computed by
integrating the bending moment twice, using the boundary condition at the clamped toe of the
wall, where   0 and u = 0. Actually, this calculation should involve the bending stiffness
EI, because the equations are d dz  M EI and du dz   , but as the conditions are that
these variables must be zero, the value of the constant factor EI is irrelevant.
If the displacement at the anchor point is found to be positive, the length of the wall is
shortened by a small amount, and the computation is repeated until this displacement of the
anchor point turns negative. The procedure assumes that the initial estimate of the length of
the wall is sufficiently large. Therefore the thickness of the deepest layer must be sufficiently
large.
The output of such a program can consists of the computed sheet pile length, anchor force,
shear force and bending moment.

The simple example of the previous chapter, for the case of a cohesionless soil, as shown in
Figure 38-2, can be determined using the following data.
The depth of the excavation (the thickness of the first layer) has been taken as 1 m. The
thickness of the second layer is 2 m. The initial estimate for the length is then 3 m. This will
probably be sufficiently long. To simulate the excavation the dry volumetric weight of the
soil at the left side is assumed to be 0.0, and its saturated weight is assumed to be 10.0, the
volumetric weight of water. Although there is no water there, changing the groundwater level
might change that.
Running a program as described before, shows that the length of the sheet pile wall must be
2.532 m, and the anchor force T = 4.751 kN/m. This means that T  h2 = 0.238. These
results are in agreement with the results of the analytical solution of the previous chapter.

39.2 Computation of anchor plate


The anchor is supposed to consist of a steel rod, connected to a plate. This plate should be
capable of resisting the anchor force T. The maximum force of the anchor plate can be
determined by an application of Coulomb’s theory, see Figure 39-2. At the left side of the
plate the stresses are supposed to be the passive earth pressure, and at the right side the active
earth pressure is assumed to act. The maximum force is then, assuming that the plate is
continuous,
Tmax  12 ( K p  Ka ) b2 . (39.1)
This value must be larger than the anchor force required for equilibrium of the sheet pile
wall. A sufficiently large safety factor must be taken into account (for instance 1.5). The
anchor rod must also be strong enough to transfer this force. In engineering practice the
anchor usually consists of a series of plates, at certain distances. The value of equation (39.1)
gives the force per unit length. This must be multiplied by the distance of the anchors to
obtain the force in a single anchor.
The anchor plate does not need to run from its largest depth b to the soil surface. This might
be an obstacle for other use of the soil. However, even when the anchor reaches only to a

183
certain small depth, the sliding surface might be the same as for an anchor up to the surface.
In practice, an anchor is considered as fully embedded if its height is at least 12 b .

Figure ‎39-2. Anchor.

The distance of the anchor plate from the sheet pile wall should be sufficiently large to enable
the necessary passive earth pressure to be developed. In principle, the distance must be so
large that the active region of the sheet pile wall and the passive region of the anchor plate do
not overlap, see Figure 39-2. The complete system, of sheet pile wall, anchor and soil body,
forms a retaining structure, which must be stable as a whole. In order to verify this condition
a stability analysis of the total system must also be made, for instance using a circular slip
plane passing below the sheet pile wall and the anchor plate.

39.3 Horizontal bedding constants for subgrade reaction models


The biggest problem of the previous method is that the lateral stresses are either active or
passive. There is nothing in between. Therefore, in engineering practice, some more
advanced methods are often used for the analysis of a sheet pile wall. A familiar method is to
consider the wall as a beam supported by a large number of elasto-plastic springs. The
characteristics of these springs are defined such that the maximum soil stress is the passive
earth pressure, and the minimum soil stress is the active earth pressure. The actual soil stress
is supposed to depend upon the displacement, with the passive stress being developed if the
displacement is towards the soil, and sufficiently large. The active state of stress is developed
when the wall displaces away from the soil, and the displacement is sufficiently large. The
actual displacements are calculated by considering the differential equations of equilibrium
and deformation of the sheet pile wall. In such an analysis, the excavation process can be
modelled in a number of loading and unloading stages, with anchors being installed in
different stages. The main advantages of this method are that the analysis includes a
reasonably accurate computation of the deformations, and that it enables analysis of walls
with multiple anchors.
The following figure and table explain which bedding constants should be used according to
CUR 166 to calculate a sheet pile wall with a spring model.
The following figure is used to calculate the horizontal stress-displacement relationship. In
the figure the following applies:
 h' ,min  K a  v'  2 c K a
 h' ,max  K p  v'  2 c K p (39.2)
 '
h ,n  K0  '
v

184
’ h
1.0 ’ h,max
k3
0.8 ’ h,max 1

k2
0.5 ’ h,max 1

k1
1
’ h,n
u
’ h,min

Figure ‎39-3. Horizontal stress-displacement relationship.

The coefficients of Ka and Kp are the active and passive coefficients of horizontal ground
pressure and the K0 is the neutral coefficients of horizontal ground pressure; mostly
calculated with K0  1  sin  .

k1 [kN/m3] k2 [kN/m3] k3 [kN/m3]


klow khigh klow khigh klow khigh
sand qc [MPa]
loose 5 12000 27000 3270 7360 1000 2250
moderate 15 20000 45000 5460 12270 1670 3750
firm 25 40000 90000 10900 24550 3330 7500
clay fundr [kPa]
weak 25 2000 4500 400 900 200 450
moderate 50 4000 9000 1090 2460 240 530
firm 200 6000 13500 2570 5790 670 1500
peat fundr [kPa]
weak 10 1000 2250 275 615 85 185
moderate 30 2000 4500 400 900 200 450

Table ‎39-1: Horizontal bedding constants (CUR 166).

The stress-displacement relationship is split into three branches. The bedding constants for
these three branches are given in Table 39-1. They depend on the type of soil. The bedding
constants are determined according to CUR 166.

Note:
 To calculate the displacement normative k-values are used, i.e.: klow.
 To calculate the forces and moments, two calculations are carried out. One with klow./1.3 and a
second with khigh. It is impossible to tell which of the two is normative in advance, although in
most cases the calculation with the low bedding constants is normative.

185
39.4 Finite Element Modelling
The subgrade reaction models (beam supported by a large number of elasto-plastic springs)
already show good results, but some rather large errors can also be found. The reason for this
is that regarding the soil as a row of independent springs, is incorrect. Every part of the
ground is connected through pressure and friction with ground around. This means the
calculated forces are incorrect. Moreover the calculation of the bedding constants is
impossible because of this. It even depends on the size of the sheet pile wall. The only
solution is to describe the soil correctly in a two dimensional – plain strain way, or even three
dimensional. This can be done with finite element models.
The finite element method (FEM) is a numerical technique for finding approximate solutions
of partial differential equations. These equations could be, for example Hook’s law equations
for the ground, and bending moment equations for the sheet pile wall. The solution approach
is based on rendering these equations into an approximating system of ordinary differential
equations, which are then numerically solved using techniques such as the Conjugated
Gradient Method.
Modern geotechnical applications require advanced constitutive models for the simulation of
the non-linear and time-dependent behaviour of soils. In addition, since soil is a multi-phase
material, special procedures are required to deal with hydrostatic and non-hydrostatic pore
pressures in the soil. Although the modelling of the soil itself is an important issue, many
geotechnical engineering projects involve the modelling of structures and the interaction
between the structures and the soil. An example is such a finite element model is the Plaxis
software, which is equipped with special features to deal with the numerous aspects of
complex geotechnical structures.

Figure ‎39-4. Exaggerated deformations of a building pit in Plaxis 2D

39.5 Example quick design


For an acceptable design nowadays, one needs computer models. However, a simple hand
calculation to check the order of magnitude is still always welcome:

Question:
A designer tries to guess quickly whether the chosen sheet pile wall is correct. The sheet
pile wall is 10 m long. Both non-cohesive soil layers have an angle of internal friction of
35 degrees. The friction angle between sand and sheet pile wall is 25 degree. It is
estimated that the support point in the deeper sand layer is about 1 m deep. The unit
weight of the dry sand is 17 kN/m3. The sheet pile is a type AZ18 with a section modulus
of W = 1800 cm3/m. The failure stress of the steel is 480 N/mm2.

186
Figure ‎39-5. Sheet pile dimensions
M
Hint:   and M  18 ql 2
W

Answer:
Given:   25   35  Ka  0, 244 (see tables Chapter 36.3)
l = 6 m. Average depth between 1 m depth and 1+6 = 7 m depth is 4 m depth:
 v  4 17  68 kPa
 h  Ka v  0, 244  68  16,6 kPa at 4 m depth.
Bending Moment: M  18 ql 2  18  hl 2  18 16,6  62  74,7 kNm/m
M
 with   480.000 kN/m 2
W
 W  1,56 104 m3 /m
 W  156 cm3 /m
So, the sheet pile wall is strong enough (1800 > 156).

Given:   25   35  K p  10,98 (see tables Chapter 36.3)


Rotation around anchor point:
Fa  12  l  d   d  Ka  207, 4 kN / m ea  23 10  1  5, 67m
2

M a  Faea  1175,3 kNm/m


Fp  12  d   d  K p  1493,3 kN / m e p  23  4  5  7, 67m
2

M p  Fpe p  11448,3 kNm/m


M a  M p So, the sheet pile wall is deep enough.

The anchor plate, both local (micro) and global (macro) stability, and cable should be
checked as well.

187
40 Sheet pile profiles
The cheapest wall is usually made of steel sheet piling, particularly if it is only needed
temporarily (excavated building site) and can be reclaimed. Usually two piles are joined by
welding and are vibrated into the ground together. There are two main types of sheet piles:
U-profiles and Z-profiles.

Figure ‎40-1. U-profile (left) and Z-profile (right)

Two U-profiles together are not symmetrical which leads to oblique deflection when
subjected to load. This reduces the strength and stiffness of these profiles by tenths of per
cent.
Z-profiles don’t have this problem. The table below contains a number of types of Z-profiles.
The AZ13 is designed to have a section Modulus of W = 1300 cm3/m, and so on for the other
sheet piles.
As a rule, the AZ13 is rarely used. AZ18 can be used for small excavations. AZ26 and AZ36
can be used for larger excavated building sites, especially if the deflection is normative.
AZ48 is the heaviest of the profiles. If this is insufficient regarding strength or stiffness, one
must switch to a combi-wall or even a diaphragm wall.

Type B H EI EA Mpl (S235) Mpl (S355) Weight


[mm] [mm] [kNm2/m] [kN/m] [kNm/m] [kNm/m] [kg/m2]
AZ13 670 303 41370 ~2.8106 306 462 107
AZ18 630 380 71820 ~2.8106 423 639 118
AZ26 630 427 116570 ~3.7106 611 923 155
AZ36 630 460 173880 ~4.7106 846 1278 194
AZ48 580 482 242910 - 1128 1704 241

Table ‎40-1 Stiffnesses and strengths of Z-sheet piles

It is cheapest to use steel quality S235. If this meets the demands for stiffness but not for
strength, it is advisable to opt for better quality steel rather than for heavier sheet piling.

Combi-walls are walls that consist of a combination of sheet piling and steel tube piles,
which have a far greater flexural rigidity. The disadvantages of combi-walls are that the tube
piles make the wall a lot more expensive and removing the tube pile leaves a large hole,
because the ground usually stays stuck in the tube (especially in cohesive ground layers).
Combi-walls have a far greater flexural rigidity. A combi-wall 1620 mm  19 mm + 2 PU20
has an EI equal to 2.28106 kNm2/m. That is almost 10 times as stiff as an AZ48!
In this case the flexural rigidity of the PU20 sheet piling is almost negligible compared with
the tube pile. In quay walls the sheet piling is often placed less deep than the tube piles.

188
It is simple to calculate the flexural rigidity of tube piles. After all, for tube profiles, the
moment of inertia is:
  D4  d 4    4  D3t
I y  Iz   (40.1)
64 64
in which:
D = outer diameter tube pile [m]
d = inner diameter tube pile = D - t [m]
t = thickness of tube wall = D - d [m]

The maximum stress can be calculated with


M I
 max  with as section modulus: W 1 (40.2)
W 2
D
For the modulus of elasticity one can use:
Esteel  2.1108 kN/m2
The flexural rigidity EI per metre of combi-wall therefore mainly depends on the type of tube
pile and on the collective width of sheet piling between these tube piles.

Figure ‎40-2. Connecting lock types

There are many different types of connecting locks, and systems to control the locks, but
these will not be discussed here.

189
190
VIII Anchors, struts and wales

191
41 Supports
In order to reduce the bending moments an displacements of retaining structures, supports
can be used. There are two types of supports:
1. Anchors
2. Struts
The advantages and disadvantages of these supports are written in the table below:

Supports Advantages Disadvantages


Anchor Not blocking entrance pit Expensive
No buckling Most cannot be reused
Outside pit under other buildings
Difficult for grouting in soft soils (clay/peat)
Strut Cheap Blocking entrance pit
Can be reused Buckling problem for large/wide pits
Not outside pit
No problems at soft soils (clay/peat)

Table ‎41-1: Advantages and disadvantages supports.

Many excavations are made in several stages. The next strut or anchor can only be installed
after another excavation step.

Figure ‎41-1. Various uses of anchors

Each step has to be checked with a calculation of the maximum moment in the retaining wall
and the maximum forces in the supports (struts or anchors).

192
42 Anchors
42.1 Anchoring
Some text and pictures are of Bilfinger Berger Spezialtiefbau GmbH, Frankfurt.

Figure ‎42-1. Various uses of anchors

Ground anchors are elements used to transfer the compressive strength load from structures
or other units into the ground around or below that which is being built. Some of the various
types of anchors are:
• standard grouted anchors;
• rock and soil nails;
• driven steel piles;
• driven reinforced concrete piles;
• injected grout piles;
• drilled piles.

Anchors can be used in numerous ways:


• to support retaining walls for excavation pits;
• to secure shorelines and river banks;
• to prevent buoyancy;
• to secure underpinning;
• for tensioning purposes;
• to secure embankments and rock walls;
• to prevent tipping and sliding of structures;
• to secure and excavate cave-like structures.

193
Standard Grouted Anchors
The standard ground anchors consist of tension elements made up of 2 to 9 strands that are
installed in cased or uncased drilled boreholes. The individual strands are wedged into the
anchor head and are, therefore, firmly connected to the element that is to be anchored. The
anchor transfers its load through adhesive stress to the pressure-grouted body. This, in turn,
transfers the load into the ground through suspended friction. The length of the pressure-
grouted body, known as the bonded length, varies depending on the type of soil, the diameter
of the pressure-grouted body and the amount of tension. The free anchor length is the area
between the anchor head and the pressure-grouted body. It extends from the anchor head to
the beginning of the anchorage length. The length of the grouted anchor is a result of the
static computation against total failure (evidence of the sliding joint depth, weight of the
ground unit). Depending on the service life of the anchor, there is a distinction between
temporary anchors (can be removed) and permanent anchors (better corrosion protection).

Figure ‎42-2. Anchors; Above: temporary; Below: permanent


(Picture: Dywidag, Unterschleissheim, Germany)

Temporary anchors
Temporary anchors, according to European Standard EN 1537, can be used for 2 years and
are provided with a single corrosion protection. This protection is made up of cement grout in
the anchorage area and, in the exposed anchoring area, of plastic pipes.

194
Permanent anchors
Permanent anchors have to carry their load over the lifetime of the structure. For this reason,
corrosion protection plays a very important role. The double corrosion protection system
consists in the anchorage area of a corrugated tube grouted with cement mortar. In the
exposed area, the smooth tube is filled with grease.

Figure ‎42-3. Construction pit with temporary supporting anchors and anchored
waterproof concrete invert

Figure ‎42-4. Construction pit with multiple anchoring supports

42.2 Anchor installation


Boreholes for the anchor installation are drilled as cased, uncased or partially cased borings.
Uncased or partially cased borings can only be done in solid ground or sound rock. Examples
of relevant drilling methods are: Auger drill that mechanically conveys the excavated ground
over the auger's spiral or a rotary drill that uses an air flushing system. A cased boring has to
be used in soft ground. Depending on the type of ground, the pipes can either be driven or
drilled. In the driving method, the ground is displaced to the side; and in the drilling method,

195
it is conveyed to the surface by compressed air or with the help of a flushing stream of
compressed air and/or water. Drilling equipment mounted on caterpillar tracks is primarily
used for driving and drilling works.
After the hole has been bored, the anchor tension elements are fed into the anchor and then,
in the case of cased drillings, the casing is withdrawn and a cement suspension is
simultaneously injected under high pressure. This causes the suspension in the area of the
pressure-grouted body to be pressed against the ground and mesh with it. In the case of
uncased drilling, this effect is attained only through post grouting. The strands can be
prestressed and wedged after the pressure-grouted body has hardened and the anchor head
has been installed. When permanent anchors are used, in order to protect against corrosion,
the anchor head is then covered with a protective cap and sealed with a corrosion preventing
paste.

Figure ‎42-5. Anchor installation

196
Figure ‎42-6. Anchor drilling

Figure ‎42-7. Anchor prestressing

Figure ‎42-8. Anchor production against pressurised water

197
Anchoring to Protect Against Pressurised Water
When retaining walls are being built to retain pressurised water, no water or soil is allowed to
flow through the anchor borehole into the construction pit during the anchor installation. A
special anchoring system has been developed for this purpose. Anchor production can be
done in all phases within a watertight head construction.

Increasing the Anchor's Load Bearing Strength with Postgrouting


The maximum bearing strength of grouted anchors can be increased by improving the
suspended friction between the ground and the pressure-grouted body. This improvement of
the suspended friction can be attained by post-grouting. For this purpose, and depending on
the post-grouting system, one or more post-grouting lances will be installed with the anchor
during the production phase. These lances serve the purpose of feeding water into the
pressure-grouted body so that it bursts and then a cement suspension can be injected into it.
This cement suspension increases the size of the pressure-grouted body as well as the load
bearing capacity. The fissures that result in the pressure-grout body are filled in by the
cement grout.

Figure ‎42-9. Anchored pile wall

Removable Pressure-Grout Anchor System


When construction pit retaining walls are anchored in inner urban areas, anchors very often
protrude into the ground of the neighbouring property, if the neighbour allows it, and, after
the construction project is completed, these anchors remain in the ground. This can entail
high costs per anchor.

198
Figure ‎42-10. Anchoring of underpinning and bulkhead

42.3 Soil nailing


Soil nailing turns naturally occurring soil into an artificial gravity wall. This is achieved with
the aid of soil nails which are anchored into a reinforced shotcrete shell. Tension and
shearing strength of the soil is increased by soil nailing so that the nailed precipitate can be
regarded and certified as a monolithic block. This is the reason that the nailed soil can also be
seen as reinforced soil (see chapter 6.2). The maximum distance between nails is 1.5 metres
both horizontally and vertically and can only be exceeded if the space has been proven to be
stable. The bore holes are to be drilled with a minimum horizontal pitch of 10°.
Producing the bore holes is the same as for grouted anchors. In compliance with generally
approved building regulations "Soil Nailing System B+B", the soil nails are made up of
regulated ribbed concrete steel BSt 500 S-GEWI (IV S GEWI) with thread grooves of the
following diameters: 18 mm, 20 mm, 25 mm, 28 mm, 32 mm, 40 mm and 50 mm. The soil
nails will be anchored on the exposed side by GEWI-anchors that meet with building
regulations approval.
Just like with standard grouted anchors, there is a difference between temporary soil nails (to
be used for less than 2 years) and permanent soil nails. Corrosion protection for permanent
soil nails is to be applied during the manufacturing phase.
The main applications for soil nailing are for temporary or permanent securing of fissured
ground such as in construction pit retaining walls and embankments; the securing of existing
sloped banks; and the stabilization of the ground during underpinning work. In the process,
the pitch (angle) of the wall can be chosen freely.
The main advantages of soil nail walls are:
 Soil nail walls are relatively cheap.
 Soil nail walls are relatively flexible and can handle larger settlements.
 This wall requires smaller right than ground anchors as soil nails are typically shorter.

The main disadvantages are:


 The system requires some soil deformation in order to mobilize resistance.
 Soil nail walls are not well-suited where large amounts of groundwater seep into the
excavation because of the temporary unsupported excavation face.
 Existing utilities can restrict the location, inclination, and length of the soil nails.

199
Figure ‎42-11. Typical soil nail wall (Picture: Deep Excavation, Astoria, NY, USA)

Figure ‎42-12. Soil nailing process (Picture: Deep Excavation, Astoria, NY, USA)

200
Figure ‎42-13. Installing soil nail

Figure ‎42-14. Securing an embankment with soil nailing

42.4 Holding capacity of anchor plate


(The following is from the book: Soil Mechanics, Basic Concepts and Engineering
Applications, from A. Aysen).
The anchor system consists of three component parts:
1. The anchor rod
2. The anchor tendon (cable)
3. The anchor head (bolt)
The anchor element can be made of plates, blocks or anchor grout. The plates can be made
from steel or precast or in-situ concrete. The influence zone of the plate must be outside the
influence zone of the retaining structure (wall). A block concrete (or dead man concrete)
anchor becomes most of its resistance from the passive resistance at the front of the block.

201
Figure ‎42-15. Common types of sheet pile anchorages

There is also the weight of the block, the shear resistance on the upper surface and the force
of the piles, if these are used. The bearing capacity of the ground anchor, or anchor rod
grouted in a drilled hole, depends on the shear resistance over its grouted length and also the
ground (in case of no rock) in front of the grout, around the tendon (like a plate anchor).
Anchor rods are often inclined at 15 to 25 degrees below the horizontal to increase the
effective stresses and thereby the corresponding shear resistance.

If the depth ration h/B > 2 to 3 , the ultimate holding capacity in plane strain conditions is
obtained from equilibrium of the horizontal forces:

For safety reasons, the effect of any surface loading is usually ignored in the calculation of
the passive thrust. For h/B > 2, it is assumed the plate will move all the ground above, so B =
h.

Figure ‎42-16. Ultimate holding capacity of a vertical plate anchor

For plates of finite width, the shear resistance developed at the vertical faces of the failure
surface is taken into account:

Where k0 is the coefficient of earth pressure (at least 0,4) and L is the width of the plate.
There are more extended solutions because here the vertical equilibrium is not always
satisfied.

202
An anchor tendon usually consists of a bar, wire or strand, used singurlarly or in groups. Bars
provide the simplest type of tendon. They can be more readily protected against corrosion
and for shallow or low capacity installations, they cost less. Bar anchors are used more in
North America and Germany. Wire and strand offer distinct advantages with respect to
tensile strength, easiness of storage, fabrication and transport. Tendon wires vary in diameter,
mostly from 5 to 8 mm and have a usual ultimate strength of 1670 N/mm2. In general, it is
used in groups of 10 to 100, but there is no limit to this number. A strand consists of a group
of wires, usually from 4 to 20, arranged in a helical form around a common axis of a straight
wire. Forces up to 7000 kN are possible.

The anchor head includes a stressing head, wedges and a distributing bearing plate used to
transfer the load to the structure. A protective cap may also be fitted over the anchorage to
accommodate accessibility and surveillance. The head must be set concentrically with the
tendon, and is normally fitted with minimum tolerance, which should not exceed 5 mm. An
angular deviation between tendon and anchor head in the axial position of more than 3
degrees will affect the load transfer efficiency. The assembly should allow access for the
grout injection tub, preferably in a central axis position.

42.5 Holding capacity of a grouted anchor


Some parts in this chapter are from‎the‎publication‎“Design‎and‎construction‎of‎an‎anchored‎
soldier‎pile‎wall‎for‎a‎large‎underground‎car‎park”‎from‎Villalobos,‎F.A.‎&‎Oróstegui,‎P.L.‎

The most important single factor influencing the choice of drilling method is the type of
ground, namely, rock or soil. For each type there are different drilling methods available
(rotary, percussive or rotary- percussive technique or vibratory driving technique).

If the ground is sand, the cone resistance qc from a Cone Penetration Test (CPT) is decisive.
For sand, no higher values than 20 MPa are allowed to be used for bored anchors and 15 MPa
for other type of anchors. The bearing capacity of the anchor can be calculated with:
Fanchor   Dgrout Lgrout grout (42.1)
in which:
Dgrout = mean diameter of grout (from 0.180 to 0.350 m) [m]
Lgrout = length of the grouting section [m]
 grout = qc friction = shear stress along grouting section [kPa]
 qc = cone resistance from CPT (max. 20 MPa) [kPa]
friction = shear number (  shaft / qc ) (take  = 0.8% to 1.1% for sand) [-]

If there is no CPT, the only option is to calculate the shear stress along the grouting section
with:
 grout  c   v tan  (42.2)
In which it is assumed that the shear failure will not be in or around the relatively strong and
rough grout section but will be in the soil, so c is the cohesion of the soil,  is the friction
angle of the soil and  v is the average effective stress along the grouting section, under
condition the isotropic pressure of the liquid grout was almost enough to lift the ground.

203
For a grout anchor of 5 m length, 300 mm diameter, and a cone resistance qc of 20 MPa in
sand, we can expect a bearing capacity of Fanchor  1000 kN (exclusive safety).
In rock it can be much higher.
There are two types of anchors: anchor piles (for example GEWI) and grout anchors. Piles
use steel tubes and grout anchors use steel cables.

Of course the steel of the anchor (cable) should be checked as well.


The anchor allowable load Tcable can be determined using the following expression:
Tcable  nAcable f yield / FS (42.3)
in which:
n = number of cables [-]
Acable = area of each cable [m2]
 fyield = cable yield stress [kPa]
FS = factor of safety (FS = 1.5) [-]

Table 42-1 resumes the cable technical characteristics for the post-stressed anchors installed
in a project.

Table ‎42-1: Anchor cable properties (ASTM 416, GRADE 270)

The resulting anchor allowable load as a function of the number of cables is shown in
Table 42-2. Table 42-2 and the values of Tcable in Table 42-1 were used to determine the
necessary number of cables for each anchor.

Table ‎42-2: Allowable load versus the number of cables

To verify the design loads taken by the anchors, loading tests can be carried out for all the
anchor rows. The anchors in the project that had three steel cables are shown in Figure 42-17,
and the properties shown in Table 42-1. The maximum capacity was defined as 90% of the

204
steel yielding load, resulting in 635 kN. Figure 42-18 shows the results of a test in the second
row for an anchor with a grouting length of 2.5 m. Initially increments were applied until half
of the maximum capacity (first loading stage). A linear response is clearly observed and
during unloading there is an important recovery of the displacements.

Figure ‎42-17. Loading test set up in the second row of anchors

Figure ‎42-18. Load displacement curve determined in an anchor loading test.

205
Figure ‎42-19. Installation of anchors (Photo: Neidhardt Grundbau Gmbh, Germany)

42.6 Overall stability


The calculation of the overall (or global, or macro) stability and the extra stability due to the
anchors is very important.
The part of the anchors inside the slip plane creates no extra support in the overall stability.
There are two ways to look at this. One can regard the force of the internal grout element as
an internal force which gives no support in the overall stability. The other way is that one can
regard the two forces on the anchor tendon (cable) of identical size, but with opposite
direction: at the anchor head and at the grout element; erasing each other’s effect on the
overall stability.
Unfortunately mistakes in this calculation method can be found in many books and manuals,
so be careful, and see Chapter 51.4.

Figure ‎42-20. Anchors inside and outside the slip plane

206
With more than one row of anchors, more than one calculation has to be made to check this
overall stability, because more than one overall failure mechanisms are possible. Remind: the
grout inside the slip plane is useless for the overall stability.

The length of the anchor tendon (cable) has to be chosen such that it is long enough to create
an overall stability. The capacity of the anchor should be at least such that smaller potential
slip planes are supported by this anchor.
For the use of soil nailing and geotextiles the same theory has to be applied.

Figure ‎42-21. Section through a sheet pile wall with a grouted soil anchor in place

207
43 Struts
Struts are nothing other than steel profiles that can absorb axial compression force (strut
force).

Figure ‎43-1. Struts

Figure ‎43-2. Struts (Picture Volker Steel and Foundation, Netherlands)

The problem is not so much the stress in the steel due to this force, as the possibility of
collapse due to buckling. As buckling is possible both in the vertical and horizontal planes,
round profiles are often used for large spans and loads. For smaller spans and loads H-
profiles are usually applied because these are cheaper and are still sufficient against buckling.
The general formulas for the buckling force, displacement and moment are:

208
 2 EI
Fbuckling 
l2
5 ql 4
u0 
384 EI
Fbuckling
n (43.1)
Fstrut
n
u u0
n 1
1
M  ql 2  Fstrut u
8

Take at least n > 2 in order to avoid buckling or large deformations. With this, it is simple to
calculate the flexural rigidity of tube piles. After all, for tube profiles:
  D4  d 4    4  D3t
I y  Iz   , (43.2)
64 64
in which:
D = outer diameter tube pile [m]
d = inner diameter of tube pile = D  2  t [m]
t = thickness tube wall =  D  d  / 2 [m]

For the modulus of elasticity one can use:


Esteel  2.1108 kN/m2
The maximum stress will be:
M I
 max  with: W 1 (43.3)
W 2
D

Notes:
 Struts have a self- weight q, which causes increases of the deflections and moments.
 Struts should also be tested for an additional uniformly distributed top load ( q  1 kN/m : diggers
drop a lot of soil which remains on the struts).
 Struts should not only be tested for buckling but also for shock loads ( F  10 kN : diggers can hit
a strut or cranes can drop things onto the strut).
 The strut should be tested for fluctuating temperature loads during day and night.
 The total building site should be tested for strut failure resulting from a strut being knocked out of
position by a digger or crane (calamity: so lower safety factors for the strength. A displacement
test is not important then).

209
44 Wales
The horizontal reaction from an anchored sheet pile wall is transferred to the tie rods by a
type of horizontal beam or flexural member known as a wale. It normally consists of two
spaced structural steel channels (U-profiles) placed with their webs back to back in the
horizontal position. The following figure shows common arrangements of wales and tie rods
located on both the inside and outside of a sheet pile wall. The channels are spaced with a
sufficient distance between their webs to clear the upset end of the tie rods. Pipe segments or
other types of separators are used to maintain the required spacing when the channels are
connected together. If wales are constructed on the inside face of the sheet piling, every
section of sheet piling is bolted to the wale to transfer the reaction of the piling. While the
best location for the wales is on the outside face of the wall (like most in Europe), where the
piling will bear against the wales, they are generally placed inside the wall to provide a clear
outside face (like most in USA).

Figure ‎44-1. Typical wale and anchor rod details (United States Steel design manual)

210
1
For a beam on two supports, the maximum moment in a bending beam is: M  ql and for
2

8
1 2
a beam on multiple supports is: M  ql . The maximum moment at the end of a beam on
12
1 2
multiple supports is set between these two values: M  ql . This is a good estimate for the
10
wales in the corners of an excavated building site.
Wales are designed using this moment. As it is assumed that the anchor force Fanchor is
gradually spread across the sheet piling via the wales, the following applies:
Fanchor
q (44.1)
l
in which l is the distance between the anchors.
The maximum moment in the wale is therefore:
1
M Fanchor l (44.2)
10
If the wall is supported by struts, the wale is usually made of I-profiles. These are placed
against the wall like an H, with the strut in between.
If the wall is supported by anchors, one sometimes uses two U profiles positioned opposite
each other, with the end of the anchor in between.

211
212
IX Floors

213
45 Floatation & Archimedes
The basic principle of the uplift force on a body submerged in a fluid is due to Archimedes.
This principle can best be explained by considering a small rectangular element, at rest in a
fluid. The material of the block is irrelevant, but it must be given to be at rest. The pressure in
the fluid is a function of depth only, and in a homogeneous fluid the pressure distribution is
p   z   gz, (45.1)
where  is the density of the fluid, g the acceleration of gravity, and z the depth below the
fluid surface.
The pressures on the left right hand sides are equal, but act in opposite directions, and are
therefore in equilibrium. The pressure below the element is greater than the pressure above it.
The resultant force is equal to the difference in pressure, multiplied by the area of the upper
and lower surfaces. Because the pressure difference is just  gh , where h is the height of the
element, the upward force equals  g times the volume of the element. This is just the
volumetric weight of the water multiplied by the volume of the element. Due to the fact that
each body can be constructed from a number of such elementary blocks, the general
applicability of Archimedes’ principle follows.
According Archimedes, the upward force of the water around a body under water must be
equal to the weight of the water what could have fit in the volume of the body. This upward
force is often denoted as the buoyant force, and the effect is denoted as buoyancy.
The buoyancy force on a body in a fluid may have a result of a body floating on the water, if
the weight of the body is smaller than the upward force. Floatation will happen if the body on
average is lighter than water. More generally, floatation may occur if the buoyancy force is
larger than the sum of all downward forces together. This may happen in the case of
basements, tunnels, pipelines or building pits. In principle floatation can easily be prevented:
the body must be heavy enough, and may have to be ballasted.
The problem of possible floatation of a foundation is that care must be taken to ensure the
effective stresses are always positive (pressure), taking into account a certain margin of
safety. In practice this may be more difficult than imagined, because perhaps not all
conditions have been foreseen. A clear example is the floating Schuman’s building along the
Rhine in Bonn which was under construction for the members of the federal parliament (this
parliament was in Bonn until a few years after the unification of Germany), see Figure 45-1.

Figure ‎45-1. Floated‎Schumann’s‎building‎in‎Bonn‎during‎the Rhine high water of 1993.

214
46 Natural floor & heave
The biggest problem of a building pit is often the way to keep the groundwater out. For the
sealing of the bottom of the pit there are several options. The easiest and cheapest method is
using natural impermeable soil layers (clay or peat) at a deeper level than the bottom of the
building pit. One has to check the heave (buoyancy) of the impermeable layer, so one has to
check whether the groundwater pressure below this impermeable layer is not higher than the
weight of the ground between this level and the level of the bottom of the building pit. As
long as the water pressure is less, the vertical effective stresses remain positive and there will
be no uplift.

Figure ‎46-1. Natural floor.

This means:
p  v , (46.1)
with:
p  hw w ,
 v  hsand  sand  hclay clay . (46.2)

If there is no natural impermeable layer, one can make one with injection, grouting or
freezing techniques. For these techniques, see the ground improvement techniques.

Another way is the use of underwater concrete floor, with or without tension piles.
For these floors, see the following chapters.

215
47 Unsupported under water concrete floor
An unsupported concrete floor of an excavation is considered here. Such structures are often
used as foundations of basements, or as the pavement of the access road for a tunnel. One of
the functions of the concrete plate is to give additional weight to the soil, so that it will not
float. Care must be taken that the water table can only be lowered when the concrete plate is
already present. Therefore, a convenient procedure is to build the concrete plate under water,
before the lowering of the water table, see Figure 47-1. After excavation of the pit, under
water, perhaps using dredging equipment, the concrete floor must be constructed, taking
great care of the continuity of the floor and the vertical walls of the excavation. When the
concrete structure has been finished, the water level can be lowered. In this stage the weight
of the concrete is needed to prevent floatation.

Figure ‎47-1. Excavation with concrete floor under water.

There are two possible methods to perform the stability analysis. The best method is to
determine the effective stresses just below the concrete floor. If these are always positive, in
every stage of the building process, a compressive stress is being transferred in all stages, and
the structure is safe. Whenever tensile stresses are obtained, even in a situation that is only
temporary, the design must be modified. The structure will not always be in equilibrium, and
will float or break. It is assumed that in the case shown in Figure 47-1 the groundwater level
is at a depth d = 1 m below the soil surface, and that the depth of the top of the concrete floor
should be located at a depth h = 5 m below the soil surface. Furthermore, the thickness of the
concrete layer (which is to be determined) is denoted as D. The total stress just below the
concrete floor now is
   c D, (47.1)
where  c is the volumetric weight of the concrete, say  c  25 kN m3 . The pore pressure
just below the concrete floor is
p  (h  d  D) w , (47.2)
so that the effective stress is
 zz   zz  p
  c D   w (h  d  D) (47.3)
 ( c   w ) D   w (h  d ).
The requirement that this must be positive gives
w
D  (h  d ) . (47.4)
c w
The effective stress will be positive if the thickness of the concrete floor is larger than the
critical value. In the example, with h – d = 4 m and the concrete 2.5 times heavier than water,

216
it follows that the thickness of the floor must be at least 2.67 m. It may be noted that the
required thickness of the concrete floor should be somewhat larger, namely 3.33 m if the
groundwater level would coincide with the soil surface. One must be very certain that this
condition cannot occur if the concrete plate is taken thinner as 3.33 m. It may also be noted
that in time of danger, perhaps when the groundwater pressures rises due to an emergency,
the foundation can be saved by submerging with water.
The analysis can be done somewhat faster by directly requiring that the weight of the
concrete must be sufficient to balance the upward force acting upon it from below. This leads
to the same result. The analysis using the somewhat elaborate process of calculating the
effective stresses may take some more time, but it can easier be generalised, for instance, in
case of a groundwater flow, when the groundwater pressures are not hydrostatic.
The concrete floor in a structure as shown in Figure 47-1 may have to be rather thick, which
requires a deep excavation and large amounts of concrete. In engineering practice more
advanced solutions have been developed, such as a thin concrete floor, supported by tension
piles. It should be noted that this requires a careful (and safe) determination of the tensile
capacity of the piles. Though a heavy concrete floor may be expensive, its weight is always
acting.

217
48 Supported under water concrete floor
Parts of this chapter are written by H.K.T Kuijper.

48.1 General
Underwater concrete can be both reinforced and non-reinforced. Reinforced underwater
concrete can be considered structural concrete, for which one needs to take greater care
concerning the shape, the dimensions and the quality of the concrete into account. Loads on
reinforced underwater concrete are usually transmitted by bending and shear forces.
Non-reinforced concrete transmits its loads differently to reinforced concrete, as non-
reinforced concrete cannot take tensile forces. Hence, as a rule, more concrete is required in
case of non-reinforced concrete. On the other hand, one does not need to place reinforcement
cages under water.

Figure ‎48-1. Concrete floor with tension pile heads, photo: HJ Foundation

The forces in a non-reinforced underwater concrete floor can be compared to those in


masonry arches. Between the floor supports (piles and sheet piling), pressure arches are
created, which transmit the loads to the supports.

sheet piling tensile piles sheet piling


underwater
concrete
pressure arch

Figure ‎48-2. Pressure arches in underwater concrete

218
An arched structure owes its strength to so-called arch thrust. This arch thrust is a horizontal
reaction force in the support (see Figure 48-2 and Figure 48-6). The arch thrust makes
equilibrium with the strutting force in the floor, which originates from the sheet piling.

The non-reinforced concrete floor can also be compared with a pre-stressed floor. In this case
the strutting forces can be considered pre-stressing forces in the floor. The floor is subjected
to a compression load, so the non-reinforced concrete is able to absorb moments without the
presence of tensile stresses.

The non-reinforced concrete floor can therefore be schematised as a number of arches


between the supports and as a plate that is pre-stressed in two directions with supports in a
number of points and lines.

At the supports, however, the reaction force has to be transmitted by a shear force between
the concrete and the pile. If steel piles are used, the concrete is usually well attached to the
pile, however, if the shape of the pile encourages the creation of gravel pockets or soil
enclosures, one cannot simply depend on the attachment. Clean contact areas are of great
importance for a good attachment. If prefab concrete piles are used, the top metres of the
sides of those piles are often given ribs to improve the join.

There are three ultimate limit states that apply for an underwater concrete floor:
1. Floatation of the floor
2. Fracture of the joint between the floor and the tension piles
3. Fracture of the floor

48.2 Floatation of the floor


In the case of floatation of the floor as an ultimate limit state; one must consider the
equilibrium of the floor.
The upwards forces are:
• the upwards groundwater pressure under the floor
• the effective ground pressure under the floor (loads caused by swell of lower clay and
peat layers)

The downward forces are:


• the weight of the floor
• the maximum friction along the pile shaft (see Figure 48-4 on the left) or the weight of
the piles and anchors, plus the attached clump weight (weight of the soil under water that
is pulled up by the piles, see Figure 48-4 on the right)

Equilibrium of the friction along the pile shaft gives:


Fwater  Fpile   w  hw  hsand  BW  qct 4bpile hsand (48.1)
Equilibrium of the attached clump weight gives:
p   v   whw   sand hsand   c hc (48.2)

219
Figure ‎48-3. Single pile failure (left) and global failure / piles with soil clump (right)

48.3 Transfer of forces to piles & fracture of the pile joint


The reaction force that the piles have to provide is transferred from the floor to the pile
through the contact area between the floor and pile.
For a clean steel pile, the steel and concrete attach like reinforcement steel attaches to
concrete. Next to the pile, the load is transferred to the underwater concrete floor by means of
shear stress (see Figure 48-4).
For a smooth prefab concrete pile, the attachment between the underwater concrete and the
pile is inferior. For this reason, prefab piles usually have ribs along the top few metres, to
ensure a certain amount of catch resistance (see Figure 48-5).
The underwater concrete floor is loaded by the upward water pressure less its own weight.
The floor transfers the load to the piles. A design calculation will have to decide whether the
thickness determined practically is sufficient. An example of such a calculation follows at the
end of this chapter.

Steel vibro combination pile prefab concrete pile

ribs

shear stress shear stress

water pressure water pressure

pile, shaped in the ground

Figure ‎48-4. Shear stress along a pile shaft

220
Figure ‎48-5. Ribs at the top of prefab piles

A fracture of the join between the floor and the pile is possible if the maximum acceptable
shear stress along the piles is exceeded.
To describe this limit state one cannot simply use the limit state “Punch” of concrete design
codes for columns punching through floors, because there is no firm attachment to the
reinforcement steel. A safe approach is to only consider the transfer of forces from the floor
to the pile in the contact areas between the pile and the underwater concrete.
In the case of a steel pile the force from the floor is transferred in the entire contact area,
assuming a good attachment of the pile and the concrete. The slide plane in the concrete is
assumed to be the smallest envelope around the pile.
In the case of prefab concrete piles, one usually assumes that the shear force in the floor is
transferred to the pile along the ribbed sides of the pile.
The limit state involved assumes that the calculation value of the occurring shear force in the
mentioned slide plains equals the maximum acceptable shear force in the underwater
concrete.

48.4 Fracture of the floor


Fracture of the floor is possible when the ultimate bearing capacity of the arch is exceeded.
The determination of the ultimate bearing capacity of the arch created in the underwater
concrete floor is not simple because the geometry of the arch is not set beforehand. To
determine the maximum bearing capacity, one often assumes an arch with a height of 75% of
the floor thickness and with an arc thickness of 10% of the floor thickness. Generally, this
limit state is not normative for the thickness of the underwater concrete.

2-D Arch effect


The shape and height of the pressure arches depend on the type of load and the arch thrust
that occur.
Before we look more closely at the (3-D) arch that is created between the supports, the
general distribution of forces in a two-dimensional arch should be considered first.
The value of the arch thrust can be found by considering the balance of moments in the
centre. The vertical reaction force and the uniform distributed load to the left of the centre
cause an anticlockwise moment equal to 1/8 q l2. The arch thrust in the left support causes a
clockwise moment equal to H f.
The balance of moments leads to ( M  18 ql 2  Hf  0 ):

ql 2 ql 2
H or f  . (48.3)
8f 8H

221
The optimum shape of an arch with a uniform distributed load is derived from the balance of
moments for every point on the arch:
1
H y  q x (l  x) . (48.4)
2
The position of the median of a pure pressure arch, loaded by an equally spread load is:
q x (l  x)
y . (48.5)
2H

Figure ‎48-6. Schematisation of a pressure arch

For a horizontal non-reinforced concrete beam, in which a uniform distributed load leads to a
pressure arch, the height of the arch is determined by the point of application of the
horizontal arch thrust and the centroid of the stress diagram in the middle of the beam. The
shape of the stress diagram depends on the deformations in the cross section of the beam and
on the - diagram of concrete. Figure 48-7 shows a number of possible stress diagrams.

-1.75 ‰ -3.5

-1.75 ‰ -1.75 ‰ -3.5 ‰ -3.5 ‰ -3.5 ‰

Figure ‎48-7. Centroid of the pressure arch in concrete, dependent on the deformation

The position of the line of action of the horizontal arch thrust depends, amongst other factors,
on the way the pressure is applied to the side of the beam. In this case the position of the line

222
of action depends on the angle of rotation of the plane on which the force acts. For an
underwater concrete floor in an excavated building site, this angle of rotation occurs when
the wall and floor deflect after the water has been removed from the site.

3-D Dome effect


The previous section only discussed the arch. In an underwater concrete floor with tension
piles, pressure arches will occur between the piles in different directions. Between these
pressure arches, the load is also transferred by means of arches. Between four piles the arches
will form an inverted dome. Figure 48-8 schematically shows the main load bearing
structure, drawn in thick lines.

Figure ‎48-8. Dome effect between four piles

The horizontal arch thrusts on the dome can be calculated by considering the balance of
moments around a horizontal axis through the centre of the dome.

Figure ‎48-9. Balance of moments around an axis through the centre

The equilibrium leads to:


q l2 l12 q l1 l2 2
H1  and H 2 
8f 8f , (48.6)

Consequently:
H1 l1
 , (48.7)
H 2 l2

223
The vertical reaction force is:
V q l1 l2

4 4 , (48.8)

Example pile joint design:


Starting values:
 thickness of floor 1.50 m
 concrete quality B 22.5
 piles 0.4 m, c-c. 3.05 m (in both directions)
 groundwater pressure underneath floor 100 kN/m2

Calculated value for the load in the limit state of collapse:


q = 1.7(100 – 1.523) = 111.35 kN/m2

(N.B.  = 1.2 for unit weight and 1.5 for water pressure lead to a lower q; see also VB 1974/1984,
article A 401.2.2).

Along the circumference plain of the pile (this is the least favourable shear plain due to the join
between old and new concrete) the shear force is:
Td = 3.052111.35 = 1035.8 kN

1035.8
d   431.6 kN/m2  1
4  0.4  1.5

1 = 0.5fb = 675 kN/m2. (fb is the calculation value of the tensile strength of concrete)

The concrete can therefore transfer the shear force. Note that in the case of underwater concrete this
deviates from VB 1974/1984, in which part D states that non-reinforced concrete cannot take on shear
forces unless a normal compression force is present.
Instead, the shear force calculations given in part E (reinforced concrete) are used.

The bending moment in the underwater concrete is initially set at:

1
10
 q  l 2  103.6 kNm/m1
103.6
b =  276.3 kPa
1
6  1  1.52
276.3 < 0.7fb (= 910 kN/m2) and is therefore acceptable (VB 1974/1984, art. D 503.5).

In reality the moment is smaller because it is a mushroom slab spanning two directions. In the final
check one must also review the moment and normal compression force, caused by the sheet piling
supports on the floor.

In use the pile will be subjected to a tensile force of 3.05 2(100 – 1.523) = 609 kN and must therefore
be sufficiently pre-stressed. The limit tensile force T of a pile equals the sum of the local shaft
resistances.

224
X Global stability & failure

225
49 Failure modes
A structure normally can fail at more than one way. It is of the highest importance to
distinguish all possible failure modes before the design of a structure. If one mode is
forgotten, the final results will be most likely catastrophic. Possible failure modes are:
 Failure of a structural part
 (Insufficient) bearing capacity
 Overturning (of a structure)
 Global translation (horizontal, vertical or diagonal sliding)
 Global rotation (circular sliding)

The failure of a structural part (anchor, strut, wall, anchor plate, tendon, tension pile, etc.) has
been discussed in the previous chapters. Failure of a structural part leads mostly to
overturning of the remaining structure due to ground and water pressures or to heave of the
total structure in case of failing tension piles.

Figure ‎49-1. Loading of tunnel floor after tension pile failure


(Flake tunnel, Zeeland, the Netherlands)

The failure due to insufficient bearing capacity of the ground below the structure can be
checked with the equation of Meyerhof, see Chapter 22.

The overturning of a structure can be checked by controlling the total moment due to all the
forces on the structure. This is especially important for structures like gravity walls and L-
walls.

The global translation can mostly be checked with the Coulomb criterion
 max  c   n  tan   , especially for horizontal translation (gravity walls and L-walls) and
diagonal translation. Another form of global translation is the vertical uplift (heave) of a
building pit including the tension piles and ground in between, due to the groundwater
pressure. This is, in fact, the checking of the clump criterion. This all will be discussed in
Chapter 50.

The global rotation can be checked with a combination of checking the total moment on the
total structure and the Coulomb criterion, which is in fact the method of Fellenius and
Bishop, see Chapter 51.

226
50 Global translation (sliding)
Horizontal slip planes are used in the calculation of the active and passive earth pressure of
retaining walls and anchor plates. In fact, horizontal slip planes are far easier to calculate than
circular slip planes, but of course nature decides which type of failure plane will occur.

Figure ‎50-1. Horizontal sliding of L-wall

Many structures can be objected to horizontal sliding, for example, L-walls, gravity walls, or
even complete dikes. Even dikes can fail along a horizontal slip surface, but only when the
external load (water pressure) is rather high in comparison to the weight of the dike.
Examples of peat dike failures are in Zoetermeer (NL, 1947), Lisse (NL, 1967), Edenderry
(IRL, Jan. 1998) and Wilnis (NL, Aug. 2003)

Figure ‎50-2. Horizontal sliding of a peat dike (Wilnis, the Netherlands, Aug. 2003).

227
Figure ‎50-3. Peat dike failure (Edenderry, Ireland, Jan.1998).

When slopes and dikes are made of heavier materials (clay or sand), horizontal failure due to
an external force will not occur. Here there is a risk of failure due to its own weight: the
steeper the slope and the higher the object, the higher the risk. This type of failure is caused
by a circular slip surface.
An important form of global translation is the vertical uplift (or heave) of a building pit
including the tension piles and ground in between, due to the groundwater pressure.
Therefore, not only should one check whether a single pile will fail due to sliding and
whether it can be pulled “clean” out of the soil, one should also verify whether all piles,
including the soil, can be pulled out as a “clump”, see Figure 50-4.
The easiest way for a quick design check of the clump criterion goes as follows; the weight
of the ground between the tension piles and the concrete floor (and the weight of the piles
and the rest of the structure, which is mostly far less and therefore neglected) should be
enough to withstand the groundwater pressure below the pile tips. So,
p  v
 w hw   sand hsand   concrete hconcrete (50.1)

Mostly the groundwater table is taken as high as possible (ground level) to be certain.

Figure ‎50-4. Vertical translation; Left: single pile failure; Right: clump criterion

228
51 Global rotation (circular sliding)
The cheapest way to create a difference in ground level (for a building pit or dike) is not any
kind of retaining structure, but a simple slope. The stability of this slope must be analysed.

Figure ‎51-1. Circular slip surface of dike after rain.

Figure ‎51-2. Circular slip surface at dike (Streefkerk, the Netherlands,1985).

For the analysis of the stability of a slope of arbitrary shape and composition various
approximate methods have been developed. Most of these assume a circular slip surface.

229
Figure ‎51-3. Slope failure (Mondercange, Luxembourg, 2014).

51.1 Safety factor


Using a number of simplifying assumptions a value for the safety factor F, the ratio of
strength (Resistance) and load (Sollicitation), is determined:
strength  Ri
F  .
 Si
(51.1)
load
Instead of F, also SF (safety factor) and FS (Factor of safety) are used. The circle giving the
smallest value 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 are 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 will be presented.

Figure ‎51-4. Circular slip surface.

230
Most methods assume that the soil fails along a circular slip surface, see Figure 51-4. 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
 (c   n tan  ). (51.2)
F
The factor F is assumed to be the same for all slices, an assumption that is common to all
methods. A second assumption is that the failure of the soil can be described with Coulomb
(maximum shear strength depending only on the normal force), while for soil the Mohr-
Coulomb criterion (maximum shear strength depending on two principle stresses) is far more
accurate.
The equilibrium equation to be used in conjunction with a circular slip surface is the equation
of equilibrium of moments with respect to the centre of the circle. This equation gives
 bR
  hbR sin    cos  . (51.3)

Here h is the height of a slice, b its width,  the volumetric weight of the soil in the slice, and
R is the 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 (51.2) and (51.3) that the safety factor
is the sum of the strength (Resistance) divided by the sum of the load (Sollicitation):

F
 R  [(c    tan  ) cos  ] .
i n

S   h sin 
(51.4)
i

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

51.2 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 51-5, are then the weight  bh , a normal stress  n and a shear stress  at the bottom
of the slice.

Figure ‎51-5. Fellenius.

231
The normal stress  n 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
 n   h cos2  , (51.5)
and, because  n   n  p ,

 n   h cos2   p. (51.6)
Substitution into (51.4) finally gives

F
{[c  ( h cos   p) tan  ] cos  } .
2

  h sin 
(51.7)

This is the Fellenius formula.


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 fundamental
base is lacking for all slip surface methods for materials with internal friction, as seen before.
Therefore it is unknown what the ratio is between the stability factor of the most
unfavourable circular slip surface and the real safety of the slope. 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 that,
on the one hand, 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 in the centre, for which   0 . At that slice  n   h , and it is assumed
that there is a shear stress ( n  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 confirms, in a number of special cases, certain limiting values. For
instance, for an infinite slope in a dry frictional material without cohesion, one obtains from
equation (51.7), assuming a straight slip surface at a depth d below the slope, and taking
p  c  0,

F
  d cos  tan   tan  .
  d sin  tan 
This is in perfect agreement with the formula for infinite slopes.
In the case of a slope under water, in the absence of groundwater flow, the limiting value is
not immediately recovered. For such problems, the Fellenius formula might be modified by
using the volumetric weight under water, ( s   w )h rather than  h , and using the excess
water pressure with respect to the hydrostatic water pressure for p. This is somewhat
artificial, however, and for this reason the Fellenius method is rarely used.

232
51.3 Bishop
A method that is frequently used in engineering practice is Bishop’s method. In this method
the forces between the slices are not neglected, but it is assumed that the resultant force is
horizontal, see Figure 51-6. By considering the vertical equilibrium of each slice only,
however, the horizontal forces do not enter into the computations.

Figure ‎51-6. Bishop.

The basic equation again is the equation of moment equilibrium, eq. (51.4). Vertical
equilibrium of a slice now requires that
sin  sin 
 h   n    n  p   .
cos  cos 
If in this equation the value of  is written, in agreement with (51.2), as   (c   n tan  ) F ,
the result is
tan  tan  c
 n (1  )   h  p  tan  . (51.8)
F F
Substitution of  n into (51.4) now leads to the final equation for Bishop’s method,

c  ( h  p) tan 
 cos  (1  tan  tan  F)
F .
  h sin 
(51.9)

Due to the stability factor F also appearing in the right hand side, it must be determined
iteratively, by starting from an initial estimate (for instance F  1), and then calculating an
updated value using (51.9). This must be repeated until the value of F no longer changes. In
general, the procedure converges rather fast.
All possible slip surfaces have to be checked, which means that the x-coordinate, the y-
coordinate and the radius of the slip surface should be varied to find the minimum safety
factor.
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.

233
Figure ‎51-7. Finding lowest safety factor (Sofware Deltares).
If   0 the Bishop and Fellenius methods are identical. If   0 Bishop’s method usually
gives somewhat smaller values. As Bishop’s method is more consistent (vertical equilibrium
is satisfied), and it confirms known results for special cases, it is often used in geotechnical
engineering. Various other methods have been developed, but the results often differ only
slightly from those obtained by Bishop’s method. This may explain its popularity.

51.4 Global failure


Every structure must be checked on global failure as well. One can use software programs
based on several straight slip surfaces, forming together more or less a circular slip surface,
see Figure 51-8.

Figure ‎51-8. Global failure of L -wall

A slip circle can also go through the reinforcement of a terre armée wall, see Figure 6-12
from Chapter 6.2. Only the strip parts outside the slip circle give extra stability against global
failure.

234
O

Fstrip

Figure ‎51-9. Extra stability due to reinforced earth.

Just as for the terre armée wall, one also has to be very careful for a wall supported by several
anchors, to add only the extra force of anchors outside the slip plane into the stability
calculation, see Figure 42-20 from Chapter 42.6.

Figure ‎51-10.Anchors inside slip plane of active zone (Belval, Luxembourg, 2013)

235
The following example is taken from a book which made the mistake of only showing and
checking failure modes a) and b), but forgetting to mention failure mode c).

Figure ‎51-11. Anchors inside and outside the slip plane

Another mistake, see Figure 51-12, is taken from the German handbook “Stahlspundwände
(7), Planung und Anwendung, Dokumentation 598, Bild 15). It shows the incorrect use of an
anchor force RA,K, which should have been excluded because it is an internal force, or it
should be compensated by the neglected reaction force of the anchor on the sheet pile wall
with the same size but opposite direction. This mistake is unfortunately found in many books.

Figure ‎51-12. Warning: Incorrect force figure

236
XI Dewatering

237
52 Groundwater flow
52.1 Hydrostatics
The stress distribution in groundwater at rest follows the rules of hydrostatics. In soil
mechanics the fluid in the soil is usually water, and it can often be assumed that the
groundwater is homogeneous, so that the volumetric weight  w is a constant. In that case the
groundwater pressure can written as
p   w z  C, (52.1)
where C is an integration constant. Equation (52.1) means that the fluid pressure is
completely known if the integration constant C can be found. For this it is necessary, and
sufficient, to know the water pressure in a single point. This may be the case if the phreatic
surface has been observed at some location. In that point the water pressure p  0 for a given
value of z.
The location of the phreatic surface in the soil can be determined from the water level in a
ditch or pond, if it known that there is no, or practically no, groundwater flow. In principle,
the phreatic surface could be determined by digging a hole in the ground, and then waiting
until the water has come to rest. It is much more accurate, and easy, to determine the phreatic
surface using an open standpipe, see Figure 52-1.

Figure ‎52-1. Standpipe.

A standpipe is a steel tube, having a diameter of for instance 2.5 cm, with small holes at the
bottom, so the water can rise in the pipe. Such a pipe can easily be installed into the ground,
by pressing or eventually by hammering it into the ground. The diameter of the pipe is large
enough that capillary effects can be disregarded. After some time, during which the water has
to flow from the ground into the pipe, the level of the water in the standpipe indicates the
location of the phreatic surface, for the point of the pipe. As this water level usually is located
below ground surface, it can be observed with the naked eye. The simplest method to
measure the water level in the standpipe is to drop a small iron or copper weight into the
tube, attached to a flexible cord. As soon as the weight touches the water surface, a sound can
be heard, especially by holding an ear close to the end of the pipe.
The measurement can also be made by accurate electronic measuring devices. Electronic pore
pressure meters measure the pressure in a small cell, by a flexible membrane and a strain
gauge, glued onto the membrane. The water presses against the membrane and the strain
gauge measures the small deflection of the membrane. This can be transformed into the value
of the pressure if the device has been calibrated before.

238
52.2 Groundwater head
The concept of groundwater head can be illustrated by considering a standpipe in the soil, see
Figure 52-2. The water level in the standpipe, measured with respect to a certain horizontal
level where z  0 , is the groundwater head h in the point indicated by the open end of the
standpipe. In the standpipe the water is at rest, and therefore the pressure at the bottom end of
the pipe is p  (h  z ) w .

Figure ‎52-2. Groundwater head.

In civil engineering many problems are concerned with a single fluid, mostly fresh water, and
the volumetric weight can then be considered as constant. In that case it is convenient to
introduce the groundwater head h, defined as
p
h z . (52.2)
w

52.3 Darcy
The hydrostatic distribution of pore pressures is valid when the groundwater is at rest. When
the groundwater is flowing through the soil the pressure distribution will not be hydrostatic,
because then the flow of groundwater through the pore space is accompanied by a friction
force between the flowing fluid and the soil skeleton, and this must be taken into account.
This friction force (per unit volume) depends on  which is is the dynamic viscosity of the
fluid, and on  , which is the permeability of the porous medium and on qx, qy and qz which
are the components of the specific discharge, that is the discharge per unit area. The precise
definition of qx is the discharge (a volume per unit time) Q through a unit area A
perpendicular to the x-direction, qx  Q A . This quantity is expressed in m3 s per m 2 , a
discharge per unit area. In the SI-system of units this reduces to m/s.
Using these relations Darcy’s law can be written as
h
qx   k ,
x
h
q y  k , (52.3)
y
h
qz   k .
z
The quantity k in these equations is the hydraulic conductivity, defined as

239
 w
k . (52.4)

It is sometimes denoted as the coefficient of permeability. The permeability κ then should be
denoted as the intrinsic permeability to avoid confusion.
Darcy’s law can be written in an even simpler form if the direction of flow is known, for
instance if the water is flowing through a narrow tube, filled with soil. The water is then
forced to flow in the direction of the tube. If that directions is the s-direction, the specific
discharge in that direction is, similar to (52.3),
dh Q
q  ki with i and q  . (52.5)
ds A
This is the form of Darcy’s law as it is often used in simple flow problems. The quantity
dh ds is the increase of the groundwater head per unit of length, in the direction of flow.
The quantity dh/ds is called the hydraulic gradient i and represents the inclination of the
phreatic surface. The minus sign expresses that the water flows in the direction of decreasing
head.

52.4 Permeability
Darcy performed tests as shown in Figure 52-3 to verify his law. For this purpose, he
performed tests with various values of h , and indeed found a linear relation between Q and
h . The same test is still used very often to determine the hydraulic conductivity (coefficient
of permeability) k.

Figure ‎52-3. Permeability test.

It consists of a glass tube, filled with soil. The two ends are connected to small reservoirs of
water, the height of which can be adjusted. In these reservoirs a constant water level can be
maintained. Under the influence of a difference in head h between the two reservoirs, water
will flow through the soil. The total discharge Q can be measured by collecting the volume of
water in a certain time interval. If the area of the tube is A, and the length of the soil sample is
L , then Darcy’s law gives

240
Q h
k . (52.6)
A L
For sand normal values of the hydraulic conductivity k range from 106 m s to 103 m s .
For clay the hydraulic conductivity is usually several orders of magnitude smaller, for
instance k  109 m s , or even smaller. This is because the permeability is approximately
proportional to the square of the grain size of the material, and the particles of clay are about
100 or 1000 times smaller than those of sand. An indication of the hydraulic conductivity of
various soils is given in Table 52-1.

Type of soil k ( m s)
gravel 103  101
sand 106  103
silt 108  106
clay 1010 108

Table ‎52-1: Hydraulic conductivity k.

As mentioned before, the permeability also depends upon properties of the fluid. Water will
flow more easily through the soil than thick oil. This is expressed in equation (52.4),
 w
k , (52.7)

where  is the dynamic viscosity of the fluid. The quantity  (the intrinsic permeability)
depends upon the geometry of the grains skeleton only.

52.5 Flow in a vertical plane


In principle, the flow can be determined if the distribution of the pressure or the head is
known. In order to predict or calculate this pressure distribution, Darcy’s law in itself is
insufficient. A second principle is needed, which is provided by the principle of conservation
of mass. In this chapter, only the simplest cases will be considered, assuming isotropic
properties of the soil, and complete saturation with a single homogeneous fluid (fresh water).
It is also assumed that the flow is steady, which means that the flow is independent of time.
Suppose that the flow is restricted to a vertical plane, with a Cartesian coordinate system of
axes x and z. The z-axis is supposed to be in upward vertical direction, or, in other words,
gravity is supposed to act in negative z-direction. The two relevant components of Darcy’s
law now are
h
qx   k ,
x
h (52.8)
qz   k .
z
Conservation of mass now requires that no water can be lost or gained from a small element,
having dimensions dx and dz in the x, z-plane, see Figure 52-4. In the x-direction, water flows
through a vertical area of magnitude dydz, where dy is the thickness of the element
perpendicular to the plane of flow.

241
Figure ‎52-4. Continuity.

The difference between the outflow from the element on the right end side and the inflow
into the element on the left end side is the discharge
q x
dxdydz.
x
In the z-direction water, flows through a horizontal area of magnitude dxdy . The difference
of the outflow through the upper surface and the inflow through the lower surface is
q z
dxdydz.
z
The sum of these two quantities must be zero, and this gives, after division by dxdydz ,
q x q z
  0. (52.9)
x z
The validity of this equation, the continuity equation, requires that the density of the fluid
is constant, so that conservation of mass means conservation of volume. Equation (52.9)
expresses that the situation shown in Figure 52-4, in which both the flow in x-direction
and the flow in z-direction increase in the direction of flow, is impossible. If the flow in
x-direction increases, the element loses water, and this must be balanced by a decrease
of the flow in z-direction.
Substitution of (52.8) into (52.9) leads to the differential equation
2h 2h
  0, (52.10)
x 2 z 2
where it has been assumed that the hydraulic conductivity k is a constant. Equation (52.10) is
often denoted as the Laplace equation. This differential equation governs, together with the
boundary conditions, the flow of groundwater in a plane, if the porous medium is isotropic
and homogeneous, and if the fluid density is constant. It has also been assumed that no water
can be stored. The absence of storage is valid only if the soil does not deform and is
completely saturated.
The mathematical problem is to solve equation (52.10), together with the boundary
conditions. For a thorough discussion of such problems many specialized books are available,
both from a physical point of view (on groundwater flow) and from a mathematical point of
view (on potential theory). Here only some particular solutions will be considered, and an
approximate method using a flow net.

242
53 Flow net
53.1 Potential and stream function
Two-dimensional groundwater flow through a homogeneous soil can often be described
approximately in a relatively simple way by a flow net, that is, a net of potential lines and
stream lines. The principles will be discussed briefly in this chapter.
The groundwater potential, or just simply the potential,  is defined as
  kh, (53.1)
where k is the permeability coefficient (or hydraulic conductivity), and h is the groundwater
head. It is assumed that the hydraulic conductivity k is a constant throughout the field. If this
is not the case the concept of a potential cannot be used. Darcy’s law, see (52.8), can now be
written as

qx   ,
x
 (53.2)
qz   ,
z
or, using vector notation,
q  . (53.3)
In mathematical physics any quantity whose gradient is a vector field (for example forces or
velocities), is often denoted as a potential. For this reason, in groundwater theory  is also
called the potential. In some publications the groundwater head h itself is sometimes called
the potential, but strictly speaking that is not correct, even though the difference is merely the
constant k.
The equations (53.2) indicate that no groundwater flow will flow in a direction in which the
potential  is not changing. This means that in a figure with lines of constant potential (these
are denoted as potential lines) the flow is everywhere perpendicular to these potential lines,
see Figure 53-1.

Figure ‎53-1. Potential lines and stream lines.

243
The flow can also be described in terms of a stream function. This can best be introduced by
noting that the flow must always satisfy the equation of continuity, see (52.9), i.e.
qx qz
  0. (53.4)
x z
This means that a function  must exist such that

qx   ,
z
 (53.5)
qz   .
x
By the definition of the components of the specific discharge in this way, as being derived
from this function  , the stream function, the continuity equation (53.4) is automatically
satisfied, as can be verified by substitution of eqs. (53.5) into (53.4).
It follows from (53.5) that the flow is precisely in x-direction, if the value of  is constant in
x-direction. This can be checked by noting that the condition qz  0 can only be satisfied if
 x  0 . Similarly, the flow is in z-direction only if  is constant in z-direction, because
it follows that qx  0 if  x  0 . This suggests that in general the stream function  is
constant in the direction of flow. Along the stream lines in Figure 53-1, the value of  is
constant. Formally this property can be proved on the basis of the total differential
 
d  dx  dz  qz dx  qx dz. (53.6)
x z
This will be zero if z x  qz qx , and that means that the direction in which   0 is
given by qz qx , which is precisely the direction of flow. It can be concluded that in a mesh
of potential lines and stream lines, the value of  is constant along the stream lines.
If the x-direction coincides with the direction of flow, the value of qz is 0. It then follows
from (53.2) and (53.5) that in this case  is constant in z-direction, and  is constant in x-
direction. Furthermore, in that case one may write, approximately
 
 . (53.7)
x z
It now follows that if the intervals  and  are chosen to be equal, then x  z , i.e. the
potential line and the stream line locally form a small square. This is a general property of the
system of potential lines and streamlines (the flow net): potential lines and stream lines form
a system of ”squares”.
The physical meaning of  can be derived immediately from its definition, see equation
10.2. If the difference in head between two potential lines, along a stream line, is h , then
  k h . The physical meaning of  can best be understood by considering a point in
which the flow is in x-direction only. In such a point q  qx   z , or   qz . In
general, one may write
  qn, (53.8)
where n denotes the direction perpendicular to the flow direction, with the relative orientation
of n and s being the same as for z and x. If the thickness of the plane of flow is denoted by B,
the area of the cross section between two stream lines is nB . It now follows that
Qn   B. (53.9)

244
The quantity Qn appears to be equal to the discharge being transported between two stream
lines. It will appear that this will enable to determine the total discharge through a system.

53.2 Flow under a structure


As an example the flow under a structure will be considered, see Figure 53-2. In this case, a
sluice has been constructed into the soil. It is assumed that the water level on the left side of
the sluice is a distance H higher than the water on the right side. At a certain depth, the
permeable soil rests on an impermeable layer. To restrict the flow under the sluice, a sheet
pile wall has been installed on the upstream side of the sluice bottom. It reduces the risk of
erosion of soil particles below the sluice. This effect is called piping and can dangerously
undermine the structure. By placing the sheet pile at the upstream side, also the pore
pressures below the sluice are reduced, in order to maximise the friction capacity of the
sluice.

Figure ‎53-2. Flow net.

The flow net for a case like this can be determined iteratively. The best procedure is by
sketching a small number of stream lines, say 2 or 3, following an imaginary water particle
from the upstream boundary to the downstream boundary. These stream lines must follow the
direction of the constraining boundaries at the top and the bottom of the flow field. The
knowledge that the stream lines must be perpendicular everywhere to the potential lines, can
be used by drawing the stream lines perpendicular to the horizontal potential lines to the left
and to the right of the sluice. After sketching a tentative set of stream lines, the potential lines
can be sketched, taking care that they are perpendicular to the stream lines. In this stage, one
should try to take the distance between the potential lines as equal to the distance between the
stream lines. In the first trial this will not be successful, at least not everywhere, which means
that the original set of stream lines must be modified. This then must be done, perhaps using
a new sheet of transparent paper superimposed onto the first sketch. The stream lines can
then be taken in a way that a better approximation of squares is obtained.
The entire process must be repeated a few times, until finally a satisfactory system of squares
is obtained, see Figure 53-2. Near the corners in the boundaries some special ”squares” may
be obtained, sometimes having 5 sides. This must be accepted, because the boundary imposes
the bend in the boundary. In the case of Figure 53-2 at the right end of the net, one half of a
square is left. It turns out that there are 12.5 intervals between potential lines, which means
that the interval between two potential lines is
kH
  . (53.10)
12.5
As the flow net consists of squares it follows that    , so that
kH
  . (53.11)
12.5

245
As there appear to be 4 stream bands, the total discharge, according to (53.9), is now
4
Q kHB  0.32kHB, (53.12)
12.5
in which B is the width perpendicular to the plane of the figure. The value of the discharge Q
must be independent of the number of stream lines that has been chosen, of course. This is
indeed the case, as can be verified by repeating the process with 4 interior stream lines rather
than 3. It will then be found that the number of potential intervals will be larger, about in the
ratio 5 to 4. The ratio of the number of squares in the direction of flow to the number of
squares in the direction perpendicular to the flow remains (approximately) constant.
From the completed flow net, the groundwater head in every point of the field can be
determined. For instance, it can be observed that between the point at the extreme left below
the bottom of the sluice and the exit point at the right, about 6 squares can be counted (5
squares and 2 halves). This means that the groundwater head in that point is
6
h H  0.48H , (53.13)
12.5
if the head is measured with respect to the water level on the right side.
The pore water pressure can be derived if the head is known, as well as the elevation, because
h  z  p  w . The evaluation of the water pressure may be of importance for the structural
engineer designing the concrete floor, and for the geotechnical engineer who wishes to know
the effective stresses, so that the deformations of the soil can be calculated.
In the case illustrated in Figure 53-2, it can be observed that at the right hand exit, next to the
structure, in the last (half) square h   H (2 12.5) and z  0.3d , if d is the depth of the
structure into the ground. Then, approximately, iz  h z  0.133H d . The particles at the
soil surface are also acted upon by gravity, which leads to a volume force of magnitude
( s   w ) , negative because it is acting in downward direction. It seems tempting to
conclude that there is no danger of erosion of the soil particles if the upward force is smaller
than the downward force. This would mean, assuming that  s  w  2 , so that
icr   ( s   w )  w  1 , that the critical value of H d would be about 7.5. Only if the value
of H d would be larger than 7.5, erosion of the soil would occur, with the possible loss of
stability of the floor foundation at the right hand side.
In reality the danger may be much greater. If the soil is not completely homogeneous, the
gradient h z at the downstream exit may be much larger than the value calculated here.
This will be the case if the soil at the downstream side is less permeable than the average. In
that case a pressure may build up below the impermeable layer, and the situation may be
much more dangerous. On the basis of continuity one might say, very roughly, that the local
gradient will vary inversely proportional to the value of the hydraulic conductivity, because
k1i1  k2i2 . This means that locally, the gradient may be much larger than the average value
that will be calculated on the basis of a homogeneous average value of the permeability.
Locally soil may be eroded, which will then attract more water, and this may lead to further
erosion. The phenomenon is called piping, because a pipe may be formed, just below the
structure. Piping is especially dangerous if a structure is built directly on the soil surface. If
the structure of Figure 53-2 were built on the soil surface, and not into it, the velocities at the
downstream side would be even larger (the squares would be very small), with a greater risk
of piping.
Prescribing a safe value for the gradient is not so simple. For that reason, large safety factors
are often used. In the case of vertical outflow, as in Figure 53-2, a safety factor 2, or even

246
larger, is recommended. In cases with horizontal outflow the safety factor must be taken
much larger, because in that case there is no gravity to oppose erosion. In many cases piping
has been observed, even though the maximum gradient was only about 0.1, assuming
homogeneous conditions. Technical solutions are reasonably simple, although they may be
costly. A possible solution is that on the upstream side, or near the upstream side, the
resistance to flow is enlarged, for instance by putting a blanket of clay on top of the soil, or
into it. Another class of solutions is to apply a drainage at the downstream side, for instance
by the installation of a gravel pack near the expected outflow boundary. In the case of
Figure 53-2, a perfect solution would be to make the sheet pile wall longer, so that it reaches
into the impermeable layer. A large dam built upon a permeable soil should be protected by
an impermeable core or sheet pile wall, and a drain at the downstream side. The large costs of
these measures are easily justified when compared to the cost of losing the dam.

Figure ‎53-3. Flow net under construction with computer program.

247
54 Flow towards well
For the theoretical analysis of groundwater flow several computational methods are available,
both analytical and numerical. Studying groundwater flow is of great importance for soil
mechanics problems, because the influence of the groundwater on the behaviour of a soil
structure is very large. Many dramatic accidents have been caused by higher pore water
pressures than expected. For this reason, the study of groundwater flow requires special
attention, much more than given in the few chapters of this book. In this chapter one more
example will be presented: the flow caused by wells. Direct applications include the drainage
of a building pit, or the production of drinking water by a system of wells.
Groundwater flow can be divided into three types: groundwater in an unconfined aquifer,
groundwater in a confined aquifer and groundwater in a semi-confined aquifer. In an
unconfined aquifer, the flow profile equals the piezometric height. The groundwater level is
therefore free. In a confined aquifer the flow profile is determined by an impermeable layer.
A badly sealed layer that leaks is called a semi-confined aquifer.

54.1 Confined aquifer


The solutions to be given here apply to a homogeneous sand layer, confined between two
impermeable clay layers, see Figure 54-1. This is denoted as a confined aquifer, assuming
that the pressure in the groundwater is sufficiently large to ensure complete saturation in the
sand layer.

Figure ‎54-1. Single well in aquifer.

In this case the groundwater flows in a horizontal plane. In this plane the cartesian coordinate
axes are denoted as x and y. The groundwater flow is described by Darcy’s law in the
horizontal plane,
h
qx   k ,
x
h (54.1)
q y  k ,
z
and the continuity equation for an element in the horizontal plane,
qx q y
  0. (54.2)
x y
It now follows, if it is assumed that the hydraulic conductivity k is constant, that the partial
differential equation governing the flow is

248
2h 2h
  0. (54.3)
x 2 y 2
This is again Laplace’s equation, but this time in a horizontal plane.
The problem to be considered concerns the flow in a circular region, having a radius R, to a
well in the centre of the circle. This is an important basic problem of groundwater mechanics.
The boundary conditions are that at the outer boundary (for r  R ), the groundwater head is
fixed: h  h0 , and that at the inner boundary, the centre of the circle, a discharge Q0 is being
extracted from the soil. This is called a well or a source.

Figure ‎54-2. Hydraulic head of a confined aquifer in a circular island

It is postulated that the solution of the hydraulic head due to the source is
Q0 r
h0  h   ln   , (54.4)
2 kH  R 
where Q0 is the discharge of the well, k the hydraulic conductivity of the soil, H the thickness
of the layer, h0 the value of the given head at the outer boundary ( r  R ), and r is a polar
coordinate,
r x2  y 2 . (54.5)
That the expression (54.4) indeed satisfies the differential equation (54.3) can be verified by
substitution of this solution into the differential equation. The solution also satisfies the
boundary condition at the outer boundary, as for r  R the value of the logarithm is 0 (
ln(1)  0 ). The boundary condition at the inner boundary can be verified by first
differentiating the solution (54.4) with respect to r. This gives
dh Q0
 . (54.6)
dr 2 kHr
This means that the specific discharge in r-direction is, using Darcy’s law,
dh Q
qr  k  0 . (54.7)
dr 2 Hr
The total amount of water flowing through a cylinder of radius r and height H is obtained by
multiplication of the specific discharge qr by the area 2 rH of such a cylinder,
dh
Q  2 rHqr  2 kHr  Q0 (54.8)
dr

249
This quantity appears to be constant, independent of r, which is in agreement with the
continuity principle. It appears that through every cylinder, whatever the radius, an amount of
water –Q0 is flowing in the positive r-direction. That means that an amount of water +Q0 is
flowing towards the centre of the circle. That is precisely the required boundary condition,
and it can be conclude that the solution satisfies all conditions, and therefore must be correct.
The flow rate very close to the centre is especially large, because at this point the discharge
Q0 must flow through a very small surface area. At the outer boundary the surface area is
very large, so there the flow rate will be very small, and therefore the gradient will also be
small. This makes it plausible that the precise form of the outer boundary is not so important.
The solution (54.4) can also be used, at least as a first approximation, for a well in a region
that is not precisely circular, for instance a square. Such a square can then be approximated
by a circle, taking care that the total circumference is equal to the perimeter of the square.
It may be noted that everywhere in the aquifer r  R . This means the logarithm in eq. (54.4)
is negative, and therefore h  h0 , as could be expected. This is confirmed by the pumping of
the groundwater head, which will indeed be lowered.

In the same way, a solution can be found for a sink, which is the opposite of a source.

Figure ‎54-3. Hydraulic head for a sink in a confined aquifer

The solution of the hydraulic head due to the sink is the opposite of a source:
Q0 r
h0  h   ln   . (54.9)
2 kH  R 

54.2 Unconfined aquifer


The example of Figure shows a confined aquifer. In case there is no clay layer on top, but
only sand, the height for groundwater transport is not the same everywhere as the hydraulic
head (h < H). This is called an unconfined aquifer.

250
Figure ‎54-4. Hydraulic head of an unconfined aquifer in a circular island

One is expected to be familiar with the terms "unconfined aquifer" (free groundwater with a
free groundwater level) and "confined aquifer" (groundwater trapped between impermeable
ground layers).

First it should be determined what phreatic surface will be established in the circular island,
with a drainage well in the middle, shown in Figure 54-4. The subsoil below the island and
the surrounding free water is impermeable, contrary to the soil that constitutes the island.
Before the pump is turned on, the groundwater level is equal to the water level around the
island.
The groundwater level drops gradually until a stationary situation is achieved. Darcy’s law
applies to the discharge of the pump at the filter:
qk i
Q
in which: q = specific discharge at filter: q  [m/s]
F
Q = discharge [m3/s]
F = area of filter [m2]
k = coefficient of permeability [m/s]
dh
i = slope = [-]
dr

dh Q dr
For the island, this means: Q  k 2  h r or: 2 h dh 
dr k r
Q
where the solution is: h2 = ln r + C
k
2 Q
If r=R then: h0 = ln R + C
k
If we subtract the equations from one another, the result is:
Q0 r
h02  h2   ln . (54.10)
k R
Or with a rainfall per area of N:

h02  h2  
2k
 R  r    k ln R .
N 2 2 Q0 r
(54.11)

251
54.3 Semi-confined aquifer
Usually the sealing layer is not entirely impermeable and the water is in a semi-confined
aquifer. Figure 54-5 shows a common profile: two very permeable layers separated by an
almost impermeable layer of clay with thickness d. We assume that any withdrawal of water
from the bottom layer does not cause a reduction of the groundwater level in the top layer
(for instance, this is realised by polder ditches in the top layer). This means that the bottom
layer is fed by the top layer through the layer of clay.

Figure ‎54-5. Dewatering of semi-confined aquifer in a circular island

According to Darcy this supply equals


H -h
L = k
d
per unit of area, in which k' is the vertical coefficient of permeability of the clay layer. In this
situation the following relation can be derived:
Q
h0  h   K0  r   (54.12)
2 kD
in which:
K0  r    a modified Bessel-function of the first type, order zero.
(see following table, or see Abramowitz & Stegun: Handbook of
mathematical functions). (For r /  : K0  r /    ln 1.123  r /   )
  the leakage factor = k DC [-]
d
C  the resistance of the clay layer = [s] or [d]
k

r/ 0.1 0.5 1.0 1.5 2.0 2.5 3.0 4.0


K0(r/) 2.43 0.92 0.42 0.21 0.11 0.062 0.035 0.011

Table ‎54-1 Values for the Bessel function

252
54.4 Superposition
It is important to note that the differential equation of Laplace (54.3) is linear, which means
that solutions can be added. This is the superposition principle. Using this principle solutions
can be obtained for a system of many wells, for instance, a drainage system. All wells should
be operating near the centre of a large area, the outer boundary of which is schematised to a
circle of radius R. For a system of n wells the solution is
n Qj rj
h0  h   ln( ). (54.13)
j 1 2 kH R
Here Qj is the discharge of well j, and rj is the distance to that well. The influence of all wells
has simply been added to obtain the solution. The discharge Qj may be positive if the well
extracts water, or negative, for a recharging well.
At the outer boundary of the system all the values rj are approximately equal to R, the radius
of the area, provided that the wells are all located in the vicinity of the centre of that area.
Then all logarithms are 0, and the solution satisfies the condition that h  h0 at the outer
boundary, at least approximately.

In Figure 54-6, the potential lines and the stream lines have been drawn for the case of a
system of a single well and a single recharge well in an infinite field, assuming that the
discharges of the well and the recharge well are equal. In mathematical physics, these
singularities are often denoted as a sink and a source.

Figure ‎54-6. Sink and source.

If the discharge of the sink and the source are of the same magnitude, but opposite (so a sink
and a source), the hydraulic head will be constant on the vertical in the middle. This means
that such a situation is just like a river in the middle. Alternatively, one can calculate the
effect of a source at a distance, a, from a river, by assuming instead of the river a sink at a
distance, a, at the other side of the river. Due to the same distance from the vertical line to
both wells there will be equilibrium: the effect down on the hydraulic head caused by the sink
will erase the effect up on the hydraulic head caused by the source.

253
Figure ‎54-7. Well near a river: position of the original well (source) and image well (sink).
(Picture: Barends & Uffink)

Figure ‎54-8. Two wells (2 sources).

254
Figure ‎54-9. Well near impermeable wall: position of the original well (source) and image
well (source). (Picture: Barends & Uffink)
In the same way, one can calculate the effect of a well (source) near an impermeable wall, for
example a sheet pile wall. One simply assumes, instead of the wall, another well (source); an
image well, at the other side of the wall in the calculation. Due to the symmetry, water will
not flow to the right source nor to the left source, but will go along the wall (as long as both
wells are of the same magnitude).

In addition, cases that are more complex can be calculated, by using more than one image
well, for example a bend of a sheet pile wall, or a bend in a river (see Figure 54-10).

Figure ‎54-10. Well near a river bend: position of the original well (source) and 3 image
wells (2 sinks and 1 source). (Picture: Barends & Uffink)

255
Notes:
 One has to distinguish well between an unconfined and a confined aquifer.
 In all places on the construction site, the water level must be about 0.50 metres below ground level, to avoid
problems with soggy ground, freezing and the accumulation of rainwater.
 Vacuum pumping is possible for water level reductions of up to 6 metres. To achieve a greater reduction one
must switch to underwater pumps (pressure instead of suction).
 The capacities of vacuum pumps vary from 1 to 2 m 3/hour.
 The capacities of underwater pumps vary from 5 to 250 m3/hour.
 The formulas have all been derived for so-called perfect wells, i.e. their filters run along the full height of
the layer from which water is extracted. This is not always the case. In an unconfined aquifer the following
applies: If the bottom of the well does not coincide with an impermeable layer, the formula for unconfined
aquifer is still used, as it has been shown that the water beneath the bottom of the well hardly participates in
the flow (i.e. assume that the impermeable layer does not extend beyond the bottom of the well). If the
impermeable layer is very deep, Sichardt advises us to increase the calculated discharge by 20%.
 The formulas have been derived for a circular island, not an everyday situation. In reality the excavated site
that needs to be dewatered is usually on land. R is understood to be the distance from the well to a point
where the groundwater level reduction approaches zero. For small dewatering systems, R is sometimes
approximated with Sichardt’s empirical formula:
R = 3000 (H - h) k (Units are metres and seconds).
 For large scale dewatering, one uses experience or pump tests. The influence of a “wrong” estimate of R is
limited by the use of the (natural) logarithm of r/R.
 The formulas are derived for one well in the centre. In practice, a number of wells will be placed around the
building site. In this case, the principle of superposition applies, i.e. the water level reduction in any one
point is the sum of the reductions caused by each individual well. In an unconfined aquifer this is not the
case, as this requires the superposition of the differences of the squares of H and h.
 The formulas are derived for a stationary situation. Other formulas apply for the preceding non-stationary
phases. However, dewatering systems are generally designed for the stationary situation, though it is
important to start pumping well in time because it takes some time to reach the stationary situation (and the
desired situation). Furthermore, the pump system needs to have sufficient overcapacity to deal with
precipitation (a greater reduction of the groundwater level than technically necessary can create a buffer for
extreme rainfall) and the possible failure of one of the pumps.
 Certain parameters, such as the coefficient of permeability k, can be found by carrying out laboratory tests
on undisturbed samples. However, they are usually not representative for the extensive area in which the
groundwater flow occurs. It is therefore desirable, if not indispensable, to carry out a pump test; a
dewatering well surrounded by a number of measuring wells. The pump test can determine parameters such
as the coefficient of transmissibility T or kD (usually [m2/day]) and the leakage factor  ([m]).
 As a rule, the building sites are close to waterways (rivers, canals), which are to take on the hydraulic
engineering work some time later. This waterway can act as a “supplying” border. At this border, the
piezometric height is not affected by the dewatering. Mathematically, this can be approximated by
introducing a fictitious “mirror well”: drawing a line from the dewatering well perpendicular to the
supplying border and, on the opposite side of the border, projecting an injection well, which injects as much
water as the first well discharges, on this line. By making the distances to the border the same, the
piezometric height remains the same along the border, provided it is straight (the reduction of the water level
caused by the drainage well is compensated by the injection of water in the other well). With both of these
wells, one can calculate the water level reduction at any point on the side of the drainage well. For example,
the following applies to a given point in a confined aquifer on the line between the well and the fictitious
injection well (given that the distance from the drainage well –and hence also from the injection well- to the
border is l ):
Q  r 2l -r
H - h=-  ln - ln 
2 k D  R R 
 Note that there is not normally a perfect supplying border. In wide, deep rivers (e.g. the Nieuwe Maas at
Pernis, measuring approximately 600 m by 13 m) the influence of the dewatering system was still detectable
on the other side. A layer of silt on the riverbed creates a large entry resistance, which means the river does
not qualify as a perfect supplying border: no reduction of the piezometric height at the border as a result of
dewatering. In that case the border should really be moved, from the waterway to further inland, on the
fictitious injection side.
 If the river has just been dredged, the layer of silt will have been removed and the river can act as a perfect
supplying border.
 The following section gives a number of global calculation examples for the preliminary design of a
dewatering system.

256
54.5 Examples

First calculation example


Consider an excavated building site which measures 50100 m2 and which is located in a 25
m thick layer of very permeable soil found, in turn, on an impermeable layer. The
groundwater table is 1 metre below ground level (H = 24 m).

Figure ‎54-11. Situation first mathematical example

At the site we want to reduce the groundwater level by 8 metres (h = 16 m), so to 0.5 m
below the bottom of the site. For the soil: k = 510-4 m/s applies.
For comparison, the following table gives an overview of order of magnitude of the
permeability k, it is certainly is not absolute:

Soil type k [m/s]


gravel 10-2
coarse sand 10-3
moderately sand 10 to 10-4
-3

fine sand 10-4 to 10-5


clay 10-9 to 10-11

Table ‎54-2 Permeability

We schematise the rectangular site (diagonal = 112 m) as a circle with a radius of 56 m.


Replacement the drainage wells can be imagined, which are placed along the sides of the
rectangular site, by one single large well in the middle of the circle (strictly speaking this is
only justifiable if the real wells are divided evenly around the circumference of a circle).

We estimate R to be 1000 m and then find:


Q 56
242  162   ln
 5 10 4
1000
This results in: Q = 0.174 m3/s or 628 m3/h. This means 16 underwater pumps, each capable
of 40 m3/hour. These drainage wells are projected around the site (they will be installed from
ground level outside the excavated site) and are to be spread evenly round the circumference.
Subsequently one has to check whether the groundwater level is at least 0.50 m below the
bottom of the excavation everywhere on the site.

If that is not the case one has to look into a different distribution of the pumps (or possibly
different pump capacities).
This will also be necessary if the groundwater table has been reduced too much. Some spare
capacity is welcome in case of the failure of one of the pumps, etc.

257
What influence does the choice of radius have?
If we use Sichardt’s empirical formula, which is only used for small drainages and is
certainly not valid here, we find a radius R = 537 m, from which we derive a discharge Q
=800 m3/h (27.4% more than with a radius R = 1000 m). For a radius R = 2000 m, the
discharge becomes Q = 506 m3/h (19.4% less). In other words, the discharge is relatively
insensitive to the size of the radius, which was nearly halved and doubled in the above.

Second mathematical example


Assume the composition of the soil is different. The top 13 metres are impermeable, whilst
the soil below remains unchanged and the piezometric height is - 1.00 m.

Figure ‎54-12. Situation second mathematical example

The underside of this impermeable layer is subject to an upward water pressure of 120
kN/m2. After the excavation, a layer of only 4.5 m remains above the -13 m mark. For a
volumetric weight is 16 kN/m3 this leaves a downward pressure of 72 kN/m2, too little to find
an equilibrium, which will lead to the bottom bursting open. If we want to include a safety
factor of 1.2, the upward pressure may not exceed 72/1.2 = 60 kN/m2. In other words: the
piezometric height must be reduced by 6 m. The thickness of the water-containing layer is D
= 25 - 13 =12 m. As this concerns a prefect confined aquifer, we use the following formula:
Q 56
6 ln
2  5 10 12 1000
4

This gives: Q = 0.078 m3/s or 283 m3/h.


If 8 underwater pumps of each 40 m3/hour are projected around the site, again, one must
check whether the piezometric height is reduced by at least 6 metres everywhere below the
excavated site.

What changes if there is a “supplying border” 300 metres from the centre of the site (this
would have to be a deep river as it would have to make direct contact with the layer carrying
water, which ends at -13 m)?

The following approximation:


Q  56 2  300  56 
6  ln  ln 
2  5 104 12  1000 1000 
results in: Q = 0.099 m3/s or 358 m3/h.
If one does not add at least 2 underwater pumps, there is a risk that the site bottom will burst
open. (1 underwater pump is not enough, because there it leaves insufficient spare capacity
for precipitation, etc.).

258
55 Dewatering
55.1 Open dewatering and wellpoint drainage

Theoretically two methods of dewatering are possible:


 open dewatering, where drainage ditches along the foot of a slope carry the water to wells;
from there the water is removed from the building site using dirty water pumps.
 wellpoint drainage, for which the following systems can be distinguished:
 vacuum drainage with closed wells, where flexible pipes connect the riser of every
well to a header. The suction pressure in the header is created using a centrifugal
or plunger pumps.
 vacuum drainage with open wells. The system is the same as the vacuum drainage
mentioned above, though here each well has a draw pipe suspended in it. The
advantage over the system mentioned above is that the water level in the wells can
be reduced maximally, provided the draw pipes are long enough.
 a pumping system with underwater pumps, in which a pump is placed in every
well.

Figure ‎55-1. Types of wellpoint drainage

Dewatering in the open is only possible for small water level reductions in impermeable soil
types (at most 4 to 5 metres in clay).

As the suction pressure can never be less than atmospheric pressure ( 10 m column of
water), the water level reduction that can be achieved by means of vacuum drainage is rarely
more than 5 to 6 metres. In deep excavations a vacuum drainage therefore must be carried out
in stages. The method is applicable a range of soil types, from silty to fine sand sills. The
yield per well is approximately 1.5 to 2 m3/hour.

Underwater pumps are suspended in the groundwater. Thus the limit of the suction head is
not relevant (mainly compression instead of suction) and large groundwater level reductions
can be realised. This makes this type of pump attractive for the construction of hydraulic
engineering works. The installation is more reliable than a vacuum drainage system, in which
air leaks can lead to problems (air suction in one well can stop the water yield of all other
wells connected to the same pump). One must ensure that the pump is suspended at sufficient
depth to prevent it from falling dry. Underwater pumps are used in soil types ranging from

259
fine sand to coarse gravel. The capacities of the pumps vary between 5 and 250 m3/hour (in
hydraulic engineering pumps with capacities between 40 and 120 m3/hour are usually
employed).
There are two ways to dewater deep building pits:
1. with stacked well points (vacuum pumps) or
2. with a deep well, which is a pump deep in the well which does not suck up, but
pushes up the groundwater.

Figure ‎55-2. Building pit with stacked well points (left) or a deep well (right)

For a shallow open building pit, also a simple dewatering system with slopes at the bottom
and a simple pump pit in the lowest corner can also be enough.

Figure ‎55-3. Building pit with ring drainage and single pump pit

260
55.2 Dewatering problems

A pump system (e.g. consisting of wells with underwater pumps) can start to function less
efficiently or even fail entirely due to a number of reasons. Causes could include:
 electricity failure (power station, transformer station, cables etc.)
 mechanical or electrical defects of the pump(s)
 breakage, tearing and leakage of the pipes
 blocked pipes (for instance due to iron oxide deposits) and/or filters encasings and
blocked perforations in the casings.
 air in the drainage pipes or suction of air via the pumps. The air can accumulate in high
points of a non-level discharge pipe, increasing the resistance.

Reduced efficiency, or even failure, can have the following consequences:


 in an unconfined aquifer:
the excavation can be partly or entirely flooded, which will lead to a work delay and
possibly to damaged equipment, certainly with regards to electrical or electronic
items.
 in a confined aquifer:
as for an unconfined aquifer, but also bursting open of the bottom of the excavation,
which can cause disturbance of the foundation soil of the structure under construction.

Depending on the extent of the possible damage, measures must be taken such as placing
emergency generators, which partially or fully take over the power supply in case of power
failure, and double cables on the building site. The examples above are based on the
assumption that the building site uses power from the national grid. However, if the building
site is in an isolated location, it can be worthwhile to use generators for the daily power
supply. In this case one could also consider backup generators.
Depending on the extent of the possible damage and on the duration of the drainage, the
following matters will also be necessary:
 regular inspection.
 preventive maintenance.
 monitoring operations.
 emergency procedures for calamities etc.

The last two points particularly depend greatly on the situation. In small projects stagnation
of the drainage need not be very grave (though disruption of the foundation base is also
serious in these projects). However, in large projects, it can be beneficial to register certain
parameters continuously, and to have alarms sound and trigger certain procedures
automatically when certain limits are exceeded. Limits that could be exceeded are the
groundwater level measured in a well (upper limit), and the minimum drainage discharge
(lower limit). The alarm can be sounded in a permanently staffed area, in somebody’s home,
close to the site or with an occupant who is on guard duty outside normal working hours or
by phone. Procedures that can be initiated automatically include starting emergency
generators if the grid fails.

55.3 Hindrance for the surroundings

Often an excavated building site (with slopes and a dewatering system) presents the lowest
building costs, even if one includes the possible damage to the surroundings. Disadvantages
can be the dewatering and the large use of space (the slopes require a lot more space than is

261
necessary for the completed structure, which can cause problems in built-up areas and if the
new structure is to be built next to the an existing structure).
Very often dewatering causes no damage at all or only limited damage.

In agricultural areas, dewatering dries out the soil, thereby ruining crops. Nature reserves and
recreational areas can also suffer badly from a lowered groundwater level. Sometimes the
surface layer will “retain” sufficient amounts of precipitation to counter the effect of the
lowered groundwater level. Furthermore, precipitation wells for agriculture and horticulture
can dry up.

Another important point is the possible settlement as a result of the temporary lowering of the
groundwater level. This causes an increase of the effective ground pressure. In soil types that
are susceptible to this, it will lead to compression of the soil and settlement of the ground
level as well as to an increase of negative shaft resistance on piles. If houses or buildings
settle evenly across the whole of their foundation, this does not have to result in damages.
It can be worse if settlements are uneven for one building. The breaking of unfounded sewer
pipes and cables at the point where they connect to a house or building on a pile foundation is
not uncommon. Wooden piles can start to rot above groundwater level.

Even settlements of ground level in agricultural areas are not a problem as long as the water
level in the drainage ditches can be adjusted accordingly. In old agricultural areas with very
different water levels in the ditches, this is not always possible and a smaller difference
between ground level and groundwater level can cause a permanent reduction of crop
productivity (if this difference used to be too great, productivity can be increased). Uneven
settlements can be a nuisance for sowing and harvesting. Drains can get the wrong angle.

In water extraction areas dewatering is usually forbidden, because, as a rule, drainage water
is pumped out to open water and is therefore lost. Return pumping can offer a solution: the
water pumped out of the site is returned to the ground through injection wells, at some
distance from the building site.
Sometimes the water acquired by means of drainage is of bad quality and may not be
discharged into open water, e.g. to prevent it from becoming brackish (protection of
agriculture, horticulture, livestock industry).
Groundwater flow (however slow) can lead to unwanted diffusion of harmful substances
from toxic waste dumps.
Good observation (groundwater levels, height measurements, descriptions of buildings and
crops) before, during and after dewatering is necessary to identify the damages. Photos can
be used to describe houses and buildings.

262
XII Safe design

263
56 Limit states and design rules
56.1 Limit states

The previous chapters only discussed the plane loads on and strengths of a civil engineering
work and its parts. This is not yet sufficient to determine the dimensions of structural
elements of a civil engineering work. One also needs structural design rules. Nearly all
design rules are derived from failure modes and describe a certain limit state.
The terms ‘failure mode’ and ‘limit state’ possibly need to be elaborated.
A failure mechanism is a description of the way in which a structure is no longer able to fulfil
its function. Not being able to fulfil a function can relate to a temporary or permanent
situation. If the construction collapses permanent failure is involved.
The state just before failure of the structure is called a limit state and the function that
describes this situation is a so-called reliability function.

Two types of limit states are distinguished according to the nature of the failure, namely:
 ultimate limit states (Type 1)
 serviceability limit states (Type 2)

In a situation of the ultimate limit state (Type 1) there is failure of the construction.
In a situation of the serviceability limit state (Type 2) there is no failure of the construction,
but the displacements lead to unacceptable loss of serviceability, damage or high
maintenance costs.
The ultimate limit state (Type 1) can be split into:
Type 1A : Failure of the construction due to a failure mechanism.
Type 1B : Displacement higher than Type 2, leading to insufficient safety

Examples of Type 1A:


 failure of pile foundation by insufficient bearing capacity of the soil
 failure of shallow foundation by insufficient bearing capacity of the soil
 landslide of slope
 failure of sheet pile wall with soil collapsing

Examples of Type 1B:


 failure of foundation beam due to large displacement of pile foundation
 failure of floor due to large displacements of shallow foundation
 failure of adjacent building due to large displacements of sheet pile wall

264
56.2 Design rules

In the design rules there is a distinction between settlement (W), rotation (), relative rotation
() and tilt ( ), see the following figure.

Figure ‎56-1. Definitions of differential settlement factors.

Every country has its own specific geotechnical design rules. For European countries, these
are listed in the national annex of Eurocode 7 “ Geotechnical Design”. In several countries
the rules are more or less like in the table below.

Design rules: Limit State


Type 1B Type 2
Settlement W - 0.15 m
Rotation  - 1:300
Relative rotation  1:100 1:300

Table ‎56-1. Design rules foundations

For the deformations of shallow foundations see chapters 14 to 18.


For the deformations of piles, see chapter 27.7.

265
57 Material and load factors
57.1 Design theory

During the design process one must take both the ultimate limit state and the serviceability
limit state into account. In this case, the ultimate limit state refers to the stability of the
structure and the subsoil whilst the serviceability limit state creates serviceability
requirements for the geometry of the design.

The general form of a reliability function is:


Z  RS (57.1)
in which: Z = reliability function
R = the resistance to failure (Résistance), or the strength
S = the load (Sollicitation)

The terms load and strength are to be interpreted broadly. For example, in the above case of
the serviceability limit state of the breakwater, the strength is defined as the maximum
allowed wave height in the harbour and the load is the occurring wave height in the harbour
basin, which is influenced by the geometry of the structure. If Z<0 the construction will fail
according to the given mode.

In practice, to guarantee a certain safety margin, the strength is often tested against the loads,
according to:
Rrep
Rd  Sd    S Srep (57.2)
R
in which: Rrep = representative value for the strength
Srep = representative value for the load
R = partial safety factor for the strength (material factor) = m
S = partial safety factor for the load (load factor) = g, q. R
Rd = calculation value of the strength
Sd = calculation value of the load

Figure ‎57-1. Calculation values for load and strength.

When determining the dimensions of the design, a value for the strength will be selected that
has a calculation value greater than the calculation value of the load.

266
The idea is that, by assuming 95% of the upper limit of the load and by multiplying this with
a load factor, a calculation value is acquired with a small probability of exceedance. If the
95% lower limit of the strength divided by a material factor (giving the calculation value of
the strength) exceeds the calculation value of the load, the area overlapped by both curves is
extremely small (see Figure 57-1). This overlapping area is a measure of the failure
probability, but it is not the actual mathematical probability of failure!

Load combinations (the Turkstra rule)


Generally, more than one load acts on a structure. In such cases one must make a
combination of the loads. If the loads are time dependent, it is too unfavourable to add up the
representative values of all loads and to multiply by the same partial safety coefficient. After
all, the maximum values of the loads do not necessarily all act on the structure at the same
time.
Some norms overcome this by using the Turkstra rule for the variable loads. According to
Turkstra, one load is considered dominant in every combination of loads.

Five load combinations can be distinguished:

Ultimate limit state:


1. fundamental combinations:  f;g Grep   f;qQextr ;rep    f;qQi;mom;rep
i 2

2. fundamental combinations:  f;g Grep


3. exceptional combinations:  f;g Grep   f;a Fa;rep    f;qQi;mom;rep
i 1
Serviceability limit states:
4. incidental combinations: Grep  Qextr ;rep   Qi ;mom;rep
i2

5. instantaneous combinations: Grep    f;gQi;mom;rep


i 1
in which: Grep = the representative value of the permanent load
Qextr;rep = the representative value of the extreme variable load of the
first order
Qi;mom;rep= the representative value of the instantaneous variable load of the
ith order
Fa;rep = the representative value of the exceptional load
f;g = load factor for the permanent loads
f;q = load factor for the variable loads
f;a = load factor for the exceptional loads

For geotechnical structures limit state 1 is subdivided into:


1a. collapse of the ground
1b. the occurrence of deformations of the ground of such magnitude that the safety
requirements are no longer met

This subdivision is really senseless for the load factors for soil, because in the table these are
the same in both limit states (1a and 1b) with one exception (namely of anchor piles with a
test load).

267
For determining the extreme and instantaneous representative value of a loads, one should
know that usually the extreme values relate to a 50-year reference period. If the considered
period is less than the reference period of the loads, the loads can be reduced by a factor:
 1    t 
 t  1   ln   (57.3)
 9   t50 a 
in which: t = considered period in which the load can occur
t50a = 50 years
 = ratio between the representative value of the instantaneous
and extreme loads in the 50 year reference period.

57.2 Load factors

Every country has its own load factors. For European countries, these are listed in the
national annex of Eurocode 7 “ Geotechnical Design”. In several countries these factors are
more or less like in the tables below.

safety load f;g  


class combinations f;q f;a
normal favourable
(not
favourable)
fundamental
combinations
1 1 1.2 0.9 1.2 -
2 1 1.2 0.9 1.3 -
3 1 1.2 0.9 1.5 -
1-2-3 2 1.35 0.9 - -
exceptional
combinations
1-2-3 1.0 1.0 1.0 1.0

Table ‎57-1. Load factors for the ultimate limit states

safety load f;g  


class combinations f;q f;a
normal favourable
(not
favourable)
incidental
combinations
1-2-3 4 1.0 1.0 1.0 -
instantaneous
combinations
1-2-3 5 1.0 1.0 1.0 -

Table ‎57-2. Load factors for the serviceability limit states

268
57.3 Material factors

Material (property) m
Concrete (compression stress) 1.2
Concrete (tensile stress) 1.4
Reinforcement steel 1.15
Pre-stressed steel 1.1
Construction steel 1

Table ‎57-3. Material factors for concrete and steel

limit states
ultimate service-
ability
type of geotechnical partial material factor 1A 1B 1A 1B
structure (m) favourable 1) unfavourable 2)
all geotechnical m;g Volumetric weight of 1.1 1.1 1 1 1
structures soil
Foundations of buildings:
on piles m;b1 without investigations 3) 1.4 1.4 1 1 1
(compression load) m;b2 with test load 1.25 1.25 1 1 1
m;b3 for test loaded piles and
anchors 1.15 1.15 1 1 1
m;b4 from soundings 1.25 1.25 1 1 1

shallow m; tangent of the angle of
internal friction 1.15 1.15 1 1 1

m;c1 cohesion (bearing
capacity of foundations) 1.6 1.6 1 1 1

m;fundr undrained shear strength
1.35 1.35 1 1 1

other structures: m;b1 without investigations 1.4 1.4 1 1 1
piles/anchor (tension, m;b2 with test load 1.4 1.1 1 1 1
horizontal load)
m;b3 for test loaded piles and
anchors 1.25 1.25 1 1 1
m;b4 from soundings 1.4 1.4 1 1 1
m; tangent of the angle of
internal friction 1.2 1.2 1 1 1
m;c2 cohesion (ground
pressures, equilibrium 1.5 1.5 1 1 1
slopes)
m;fundr undrained shear strength
1.5 1.5 1 1 1
deformations:
m;Cc, m;Ca, m;Csw 1 1 0.8 0.8 1
m;Cp, m;Cs 1.3 1.3 1 1 1
m;E 1.3 1.3 1 1 1
1) "Favourable": when an increase of the value of the parameter concerned leads to a more favourable result.
2) "Unfavourable": when an increase of the value of the parameter concerned leads to a less favourable result
3) For this hammer diagrams need to have been used recorded during the driving of piles, otherwise m;b1 = 1.8 is valid for
1A and 1B (favourable):

Table ‎57-4. Material factors for soil properties

269
Examples:
 Uneven settlements: Check rotation ( = h / l = 1 / 300) in usage phase, with m,Cp/s =
1.0 in the settlement calculations.
 The bursting open of the bottom of an excavated building site (ultimate limit state) with
m,g = 1.1 for the weight of the soil.

In Eurocode 7 “Geotechnical Design” the safety factor is not written as F or SF anymore, but
as:


Rrep
S
(57.4)
rep

If an overall safety factor is found of about   min  1, 25 , the structure is usually accepted.
While in this way the partial load and material factors cannot easily be implemented, another
ratio is chosen nowadays, which is the degree of exploitation:

 S d

S
rep  S
R R
(57.5)
d rep /R

This means:
 S  R min
  (57.6)
 
Therefore,   1.0 or 100% degree of exploitation means the structure is just at the allowable
limits of the design codes and   1.1 doesn’t necessarily mean failure, but is not allowed.

57.4 Inadequate standards

Some norms assume a reference period of 50 years and an acceptable probability of failure of
10-4 per year. The demands regarding the reference period and probability of failure of large
hydraulic engineering projects can deviate from these norms. This means that the partial
safety factors (load and material factors) given above cannot be applied. The Eurocode does
give a correction factor for a reference period shorter than 50 years, but it does not give any
calculation rules regarding longer reference periods and deviating acceptable probabilities of
failure.
The Eurocode does not give any characteristic values or load factors for typical hydraulic
engineering loads, such as waves and current. The reason for this is obvious, as the statistics
of waves and currents differ from one location to another and therefore cannot be generalised
in a norm.

For probabilities of failure that deviate from the norm or for the occurrence of loads that are
not described in the norm, one has to resort to probabilistic calculation techniques to
determine the design values of the load and strength. For this, one is referred to books on
‘Probabilistic design’.

270
58 Control sensors
Probably the most important part of a safe design is the control of the loads, stresses, strains
and deformations at crucial parts, during and also after the construction, of the geotechnical
structure. This does not give extra safety by increasing the strength or reducing the load, but
it warns the contractor, hopefully in time, that the design is according to the reality, or not.
If not, one might still have time to respond to this. Only in case a complete “if-then-else”
design has been worked out in advance, for all possible unwanted measurement values, one
can say that the project is designed according to the observational method.

There is a wide range of geotechnical sensors available, for example:


 Pressure cells
 Standpipe piezometers (pore water pressure)
 Pressure transducer piezometers (also pore water pressure)
 Inclinometers (change of angle in depth)
 Tilt meters (change of angle locally)
 Anchor (and strut) load cells (both hydraulic and electric)
 Pendulum with a vertical cable (horizontal displacements in dams and towers)
 Vibrating wire strain gauges (strain and stress measurements of structures)
 Bore hole multiple rod extension meters (vertical displacements).

In Figure 58-1 on the next page, all these type of control sensors are shown in the same order
of the list above.

271
Figure ‎58-1. Different types of control sensors.

See previous page for explanations.


All pictures are from SISGEO, Italy.

272

Вам также может понравиться