You are on page 1of 90

Short-term deformations in clay under a

formwork during the construction of a bridge


A design study

Alexander Berglin

Master of science thesis 2017


© Alexander Berglin, 2017
Master of Science Thesis
Royal Institute of Technology (KTH)
Department of Civil and Architectural Engineering
Division of Soil- and Rock Mechanics
ISSN 1652-599X 17:05
Abstract
During the casting of a concrete bridge deck, the temporary formwork is causing the
underlying ground to deform if a shallow foundation solution is used. There are often
demands on the maximum deformation of the superstructure when designing the foundation
for the formwork. To keep the deformations within the desired limits, several ground
improvement methods like deep mixing columns or deep foundation methods like piling
can be used. Permanent ground improvement methods are however expensive, and far from
always needed. To reduce the need for unnecessary ground improvements, it is crucial to
calculate the predicted deformations accurately during the design phase.

The purpose of this thesis was to investigate how short-term deformations in clay under a
formwork during bridge construction should be calculated more generally in future
projects.

Three different calculation models have here been used to calculate the ground
deformations caused by the temporary formwork. A simple analytical calculation and two
numerical calculations based on the Mohr Coulomb and Hardening Soil-Small constitutive
models. The three calculation models were chosen based on their complexity. The
analytical calculation model was the most idealised and the Hardening Soil-Small to be the
most complex and most realistic model.

Results show that the numerical calculation model Mohr Coulomb and the analytical
calculation model gives the best results compared to the measured deformation. One of the
most probable reasons for the result is that both of the models require a few input
parameters that can easily be determined by well-known methods, such as triaxial-, routine-
and CRS-tests. The more advanced Hardening soil small model requires many parameters
to fully describe the behaviour of soil. Many of the parameters are hard to determine or
seldom measured. Due to the larger uncertainties in the parameter selection compared with
the other two models, the calculated deformation also contains larger uncertainties.

Key words: Small-Strain stiffness, Plaxis, Short-term deformations, Elasticity modulus,


Correlations

i
ii
Sammanfattning
Vid gjutning av betongbrodäck kommer den underliggande marken att deformeras av den
temporära formställningen, som tar upp lasterna medan betongen härdar. Det finns oftast
krav på hur stora markdeformationerna maximalt får vara. För att hålla deformationerna
inom gränserna kan diverse markförstärkningsmetoder, så som kalkcementpelare eller
pålar, användas. Permanenta markförstärkningar är oftast väldigt dyra och inte alltid
nödvändiga. Ett alternativ till att använda dyra markförstärkningar skulle kunna vara att
beräkna den förutspådda deformationen med stor exakthet i projekteringsstadiet.

Syftet med det här arbetet var att undersöka hur korttidsstätningar i lera vid en
bronybyggnation ska beräknas mer generellt i framtida projekt.

I detta arbete har tre beräkningsmodeller använts för att beräkna markdeformationerna från
den temporära formställningen. En enklare analytisk modell samt två numeriska
beräkningsmodeller som baseras på Mohr Coulomb och Hardening Soil Small teorierna. De
tre beräkningsmodellerna valdes utifrån deras komplexitet. Den analytiska beräkningen
ansågs vara den mest förenklade modellen medan Hardening Soil-Small var den mest
komplexa och realistiska modellen.

Resultatet visar att trots sin enkelhet så ger den numeriska beräkningsmodellen Mohr
Coulomb och den analytiska beräkningen bäst resultat jämfört med de uppmätta
deformationerna. En möjlig anledning till det goda resultatet är att modellerna endast kräver
ett fåtal ingångsparametrar som kan bestämmas med hjälp av välkända fält- och
laboratoriemetoder så som triaxialförsök, rutinlaboratorieförsök och CRS-försök. Den mer
komplexa modellen Hardening Soil Small kräver flera ingångsparametrar för att kunna
modellera jordens beteende. Många av parametrarna är svåra att bestämma då mätdata
oftast saknas. Osäkerheterna i valet av ingångsparametrar för den mer komplexa hardening
soil small modellen är större än de två andra studerade modellerna, vilekt även ger upphov
till större osäkerheter i dem beräknade deformationerna.

Nyckelord: Small-Strain Stiffness, Plaxis, Korttidssättningar, Elasticitetsmodulen,


Korrelationer

iii
iv
Preface
This thesis concludes my studies in Civil Engineering at KTH. The idea for the thesis was
provided by ELU Konsult where the thesis was written. I would like to thank the extremely
skilled and kind people at the geotechnical division at ELU for their help when it was
needed. A special thanks to my two supervisors at ELU Konsult; Sebastian Addensten and
Anders Beijer-Lundberg for your knowledge, guidance and encouragement!

I would also like to thank Martin Holmén at SGI for providing me with data and expertise
regarding triaxial tests and Dr. Johan Spross at KTH for your valuable comments about the
thesis.

Furthermore I would like to thank Professor Stefan Larsson at KTH. Your knowledge and
enthusiasm about geotechnical engineering has inspired me.

Last but not least, thank you to all the other people who have been supporting me during
my time at KTH.

Stockholm, June 2017

Alexander Berglin

v
vi
Table of Contents

1 Introduction........................................................................................... 1
2 Literature study ..................................................................................... 3
2.1 Brief introduction to consolidation theory ....................................................................... 3
2.2 Brief introduction into elastic theory ............................................................................... 3
2.2.1 Background ................................................................................................................. 3
2.2.2 Theory ......................................................................................................................... 3
2.2.3 Formulation ................................................................................................................. 5
2.3 Elasticity modulus............................................................................................................ 7
2.3.1 Internal factors ............................................................................................................. 7
2.3.2 External factors ............................................................................................................ 7
2.4 Small-strain stiffness........................................................................................................ 8
2.4.1 The influence of diagenesis ........................................................................................... 9
2.4.2 The influence of confining stress ................................................................................. 10
2.4.3 The influence of void ratio .......................................................................................... 11
2.5 Measuring small- strain stiffness .................................................................................... 12
2.5.1 In-Situ tests................................................................................................................ 12
2.5.2 Laboratory tests.......................................................................................................... 14
2.6 Ground investigation methods ....................................................................................... 16
2.6.1 Oedometer tests.......................................................................................................... 16
2.6.2 Triaxial tests .............................................................................................................. 16
2.7 Empirical correlations for determining the soil stiffness ............................................... 21
3 Soil modeling ....................................................................................... 24
3.1 Introduction to numerical modelling ............................................................................. 24
3.2 Mohr coulomb (MC) ...................................................................................................... 24
3.3 Hardening Soil (HS) ...................................................................................................... 26
3.4 Hardening Soil Small (HSS) .......................................................................................... 29
4 Case study: Bridge over Ulvsundavägen ............................................. 31
4.1 Introduction .................................................................................................................... 31
4.2 Studied section ............................................................................................................... 33

vii
4.2.1 Ground conditions ...................................................................................................... 34
4.3 Ground deformation measurements ............................................................................... 36
5 Calculation procedure ......................................................................... 37
5.1 Formwork geometry....................................................................................................... 37
5.2 Empirical correlation study ............................................................................................ 38
5.3 Analytical Calculation.................................................................................................... 39
5.4 2D Numerical simulation ............................................................................................... 41
5.4.1 Assumptions .............................................................................................................. 42
5.4.2 Input parameters......................................................................................................... 42
5.5 3D Numerical simulation ............................................................................................... 45
5.6 Parameter sensitivity analysis ........................................................................................ 46
6 Results ................................................................................................. 47
6.1 Empirical correlation study ............................................................................................ 47
6.2 Calculated deformations ................................................................................................ 50
6.2.1 Analytical calculations ................................................................................................ 50
6.2.2 2D Mohr Coulomb ..................................................................................................... 51
6.2.3 2D HSS Model........................................................................................................... 53
6.2.4 Comparison between 2D and 3D numerical calculations ............................................... 55
6.3 Sensitivity analysis of the HSS model ........................................................................... 56
7 Analysis and discussion ....................................................................... 57
7.1 Empirical correlation for the elasticity modulus............................................................ 57
7.2 Plaxis parameter optimisation function ......................................................................... 57
7.3 Calculated deformations ................................................................................................ 57
7.3.1 Analytical vs measured deformations ........................................................................... 57
7.3.2 MC-calculations ......................................................................................................... 58
7.3.3 HSS-Calculations ....................................................................................................... 58
7.3.4 2D vs 3D ................................................................................................................... 59
7.4 Sensitivity analysis......................................................................................................... 59
7.5 Conclusions and recommendations for practical design ................................................ 60
8 Bibliography ........................................................................................ 61
Appendix A Soil Data................................................................................. 64
Appendix B Numerical input parameters .................................................. 66

viii
Appendix C Performed triaxial tests.......................................................... 70

ix
Notations
Abbreviations Explanation

OCR Over-consolidation ratio


CRS Constant rate of strain
NC Normally consolidated
OC Overconsolidated
HS Hardening soil
HSS Hardening soil-small
MC Mohr Coulomb

Roman letters Explanation Unit

Help paramter [kPa]


b Width of the fictive plate [m]
Cohesion, shear strength [Pa]
Corrected shear strength [Pa]
Coefficient of consolidation [m2 /s]
Void ratio [-]
Young’s modulus [Pa]
Initial Young´s modulus in the elastic range [Pa]
Undrained young´s modulus [Pa]
Secant modulus (50% of peak strength) [Pa]
Secant stiffness in drained triaxial tests [Pa]
Plastic modulus [Pa]
Tangent stiffness for oedometer loading [Pa]
Unloading modulus [Pa]
Unloading/reloading stiffness [Pa]
Young´s modulus in the vertical direction [ ]
Young´s modulus in the horizontal direction [ ]
Yield function [-]
Shear stiffness [Pa]
Specific gravity [-]
Initial Shear modulus for small strains [ ]
Shear modulus in the vertical plane [ ]
Shear modulus in the horizontal plane [ ]
Thickness of soil layers [m]
Drainage distance [m]
Density index [-]
Permeability [m/s

x
Coefficient of lateral earth pressure at rest [-]
Bulk Modulus [Pa]
Elastic modulus [Pa]
Plastic modulus [Pa]
Pore-water pressure [Pa]
Deviatoric stress [Pa]
Q Distributed load [kN/m2 ]
Reference stress [ ]
Force [N]
Deviatoric stress [Pa]
Failure ratio [-]
Sensitivity [-]
Time [s]
Time factor [-]
Velocity of P-wave propagations [m/s]
Velocity of S-wave propagations [m/s]
Wave propagation velocity in soil [m/s]
Plastic limit [%]
Liquid limit [%]
Water content [%]

Greek Symbols Explanation Unit

Deformation [m]
Strain [%]
Friction angle [°]
Lamé constant [-]
Shear strain [Pa]
Shear strain at 30% degradation of small-strain stiffness [Pa]
Dilatancy [°]
Major principal stress [Pa]
Minor principal stress [Pa]
ρ Bulk density [kg/m3 ]
Poissons ratio [-]
Uncorrected undrained shear strength [Pa]
Shear stress at failure [Pa]

xi
1 Introduction
During the casting of a concrete bridge deck, the temporary formwork is causing the
underlying ground to deform if a shallow foundation solution is used. The soil deformations
occur within in the first few days, before the load from the bridge deck can be transferred
through the supports of the bridge when the concrete structure cures. There are often demands
on the maximum deformation of the superstructure when designing the foundation for the
formworks. To keep the deformations within the desired limits, several ground improvement
methods like deep mixing columns or deep foundation methods like piling can be used.
Ground improvement is a possible way to strengthen the soil and therefore reducing the
deformations. Ground improvements are however expensive and sometimes superfluous. To
reduce the need for unnecessary ground improvements it is crucial to be able to predict the
ground deformation accurately during the design phase and adjust the height of the formwork
accordingly.

The deformation response of soil is dependent on many different parameters. These include
the elasticity modulus and the small strain-stiffness of the soil. The modulus of elasticity is
hard to decide in geotechnical engineering, due to the highly non-linear behaviour of soil.
Despite that, the elasticity theory has shown that the calculated deformation of a soil
corresponds well to the measured deformations, if the elasticity modulus is chosen carefully.

Triaxial tests are generally the most suitable and easily available method for investigating the
strength and deformation properties of soil (SGF, 2012). The parameters obtained from the
triaxial tests are used for idealized analytical or more advanced numerical models in order to
calculate the deformation of the soil. However, triaxial tests are far from always performed in
geotechnical projects. An alternative method for the estimation of the elasticity modulus is by
using empirical correlations that are based on parameters that can be obtained from in-situ or
routine laboratory tests.

Recent studies by Benz (2007), Clayton (2011) and Wood (2016) have shown the importance
of small-strain stiffness in soils in the serviceability limit state. At very small strains, the soil
behaves elastically, but with increasing strain the stiffness decays non-linearly. Despite this,
small-strain stiffness is not too common in design for the serviceability limit state in
geotechnical projects.

This thesis is based on a study of the ground deformations during the construction of the
bridge over Ulvsundavägen in Stockholm. Calculations of the predicted ground deformations
were done in the design phase of the bridge in order to determine if piles had to be installed
under the temporary formworks. In-situ tests as well as triaxial and CRS-tests were carried out
to determine the soil profile and soil parameters needed to calculate the deformations
accurately.

1
The purpose of this thesis was to investigate how short-term deformations in clay under a
formwork during bridge construction should be calculated more generally in future projects.
Three different calculation models were chosen, a simple analytical and two numerical based
on the Mohr Coulomb and Hardening soil small constitutive model. The three models were
chosen based on their complexity, where the analytical method was considered to be the most
idealised and the hardening soil-small to be the most complex and realistic model.

2
2 Literature study
This chapter discusses the mechanical response of soils during stress and deformation change,
in order to relate the later chapter on laboratory tests and numerical models to the scientific
literature.

2.1 Brief introduction to consolidation theory


The deformation of soil is a process that involves three stages: The Elastic stage, followed by
primary- and secondary consolidation (Lambe & Whitman, 1979). Elastic deformations occur
instantly when the soil gets exposed to a load. Elastic deformations mainly occur in friction
material such as sand and gravel. Primary consolidation is when the pore water is being
squeezed out from the soil skeleton. Primary consolidation occurs over a longer time span in
cohesive soils. Secondary consolidation is when the soil skeleton gets deformed, a process
occurring over a long period of time (Larsson, 2008).

2.2 Brief introduction into elastic theory


2.2.1 Background
When a body is subjected to changing forces, it will to some extent change its shape or
volume, hence deformations will occur. The body is said to be elastic if the shape goes back
to its original state when the forces are removed. The phenomenon that the deformation
(strain) is related to the force (stress) was formulated by Robert Hooke in the 1676
(Timoshenko, 1983). Today this phenomenon is known as the generalized Hooke´s law
(Davis & Selvadurai, 1996).

Hooke´s law is an example of a constitutive relation (Timoshenko, 1983). A constitutive


relation is an equation that relates the cause and effect. The constitutive equations all involve
at least one parameter which takes different values for different materials.

Hooke´s law works well for isotropic mechanical behaviour, i.e. same properties in all
directions. Soil is very complex and the behaviour normally not considered to be isotropic.
However, to simplify the behaviour of soil, assumptions of soil being isotropic is done in
many geotechnical areas (Davis & Selvadurai, 1996).

2.2.2 Theory
When an elastic bar gets subjected to uniaxial tension stress it elongates in the direction of
the applied stress, resulting in an extensional strain , (Timoshenko, 1940). According to
the generalized hook’s law, is dependent on , Eq. (1).

(1)
When the bar gets elongated in the stress direction it also gives rise to lateral contraction,
causing the bar to become skinnier. The lateral contraction leads to more strains, and

3
in the lateral directions. If the material is assumed to be linear elastic the relationship between
the strains are:

(2)
The results can be generalized by studying a cube subjected to uniform normal stress in all
directions, Figure 2.1. will be dependent on the stresses in all directions and can be
formulated as Eq.(3).

(3)
Equation (3) can be rewritten and by taking and into account the following equations
will be:

[ ( )]

[ ( )]
(4)

[ ( )]

The shear modulus relates the shear stress at any given point in a body to the shear strain
that occurs at that point, Eq. (5)

(5)
is related to the Young´s modulus by the following relationship:

(6)
( )

Figure 2.1 Cube subjected to normal stresses in all directions

4
By combining equations (4) - (6) the generalized hook´s law can be expressed in a matrix
form in a six-dimensional stress-strain vector space, Eq.(7).

(7)

[ ] [ ]

[ ]

There are two more elastic constants, the bulk modulus and the Lamé constant , (Davis &
Selvadurai, 1996). These two constants are related to , and . relates the sum of the
normal stresses to the volumetric strain and can be obtained from Eq.(8).

(8)
( )
can be expressed as Eq.(9)

(9)
( )( )

2.2.3 Formulation
2.2.3.1 Isotropic elasticity
The response of an isotropic material is independent of the orientation (Davis & Selvadurai,
1996). Isotropic materials can be fully described using two of the five elastic constants, and
, and , Eq.(10) and (11).

( )
(10)
( ) (11)

2.2.3.2 Anisotropic elasticity


Anisotropic elasticity is here presented to provide a more realistic description of real soil
behaviour.

The elastic properties in an anisotropic material are dependent on the orientation of the
sample, (Muir Wood & Arroyo, 2004). To fully describe the anisotropic elasticity of a
material a total number of 21 independent parameters are needed (Jamiolkowski et al., 1996).
This can be compared to the isotropic behaviour, where only two parameters are needed.
However, many materials show a more limited version of anisotropy. One example of this is

5
transverse isotropy or cross-anisotropy. The cross-anisotropy have the same elastic parameters
in the horizontal direction, but different parameters in the vertical direction (Piriyakul, 2006).

The cross-anisotropic elasticity can be described by the following matrix, Eq. (12)

(12)

[ ] [ ]
( )
[ ]

Where:

= Poisson´s ratio for horizontal strain due to horizontal strain at right angles
= Poisson´s ratio for vertical strain due to horizontal stress
= Poisson´s ratio for horizontal strain due to vertical stress
= Young´s modulus in the vertical direction
= Young´s modulus in the horizontal direction
= Shear modulus in the vertical plane
= Shear modulus in the horizontal plane

2.2.3.3 Incompressible elasticity


In Soil mechanics, incompressibility is relevant when the response of a fully saturated soil in
undrained conditions needs to be analysed (Atkinson, 2000). Undrained condition is when the
pore fluid cannot move freely inside the soil particles. An incompressible body is
characterized by an infinite bulk modulus, , which implies the following relationships:
, and (Davis & Selvadurai, 1996; Lambe & Whiteman, 2008).

6
2.3 Elasticity modulus
E is a measurement of the soils stiffness and a crucial input when performing calculations to
predict ground deformations. However due to its complexity, E is a difficult parameter to
determine. Both external and internal factors, such as, water content, stress history,
cementation, particle organization along with loading factors all influence the elasticity
modulus (Briaud, 2001).

2.3.1 Internal factors


If the particles in the soil are close to each other, the modulus tends to be higher than if the
particles are more spaced. How close the particles are to one and another can be obtained by
measuring the dry density of the soil. The higher the dry density, the closer the particles are to
each other. How the soil particles are organized is another important factor that affects the
modulus. Depending on the internal structure, two samples with the same dry density may
have different elasticity moduli.

The water content in soil is one of the most important factors affecting the modulus. At low
water content the water binds the soil particles and increases the effective stress between
them, leading to higher moduli. However, if the water content is too low the modulus will be
lower.

The previous loading history of the soil also influences the moduli. If the soil previously has
been exposed to stresses it is called overconsolidated. Overconsolidated soils often have a
higher elasticity modulus than soils that have not been exposed to previous stresses, normally
consolidated soils (Briaud, 2001).

2.3.2 External factors


Stresses that are induced by the loading process of soil can be normal stresses, shear stresses
or a combination of them. At any arbitrary spatial point, there will be a set of three principal
normal stresses in the soil. The mean value of these stresses will influence the modulus of the
soil. The phenomena were the mean stresses influences the modulus is known as the
confinement effect. The higher the confinement is, the higher the modulus of the soil will be,
Figure 2.2a.

Stresses are induced when loading a soil, due to the non-linearity of soils, the secant modulus
will depend on the strain level. The secant modulus will generally decrease as the strain
increases, this due to the shape of the stress-strain curve, Figure 2.2b.

The rate at which the soil is being loaded also affects the modulus. Soil is a viscous material,
which means that the faster the soil is loaded the higher the modulus will become. The
exponent b in Figure 2.2c is dependent on the soil type. In clays the exponent b often varies
between 0.02 for stiff clays to 0.1 for very soft clays (Briaud, 2001).

The number of times the soil is being loaded also influences the modulus. The more loading
cycles the soil experiences, the lower the modulus will become, Figure 2.2d. The exponent c
varies, but a value of -0.1 to -0.3 is commonly used (Briaud, 2001).

7
Figure 2.2 Loading factors affecting the modulus: a) Mean stress level. b) S train level in the soil. c) S train rate. d)
Number of cycles experienced by the soil (Briaud, 2001).

2.4 Small-strain stiffness


Small-strain stiffness refers to how soil behaves at small strains ( ). At small strains
the soil behaviour is considered to be truly elastic (Atkinson, 2000; Benz, 2007). The range of
strain at which the soil behaves truly elastic is dependent on the material composition and the
stress-strain history of the soil (Wood, 2016). The small-strain stiffness and its degradation
with increasing strains is often described with (Seed & Idriss, 1969). The modulus
describes both the drained and undrained conditions of the soil. However, the degradation of
the stiffness with increasing strains can also be described with (Thiers & Seed, 1968).
Another common way to present the stiffness degradation is in terms of , where is
the initial stiffness at small strain, Figure 2.3 (Benz, 2007).

Several studies have been made regarding the stiffness at small strains and its degradation at
increasing strain, e.g. Benz (2007); Burland, (1989); Jardine, (1986). The studies have shown
that the parameter is affected by several factors, Table 2.1.

8
Figure 2.3 S tiffness-strain behaviour of soil (Benz, 2007).

Table 2.1 Factors affecting the stiffness at small strains (Benz, 2007; Burland, 1989)

Parameter Importance to for Cohesive


soils
Strain amplitude Very Important
Confining stress Very Important
Void ratio Very Important
Plasticity Index Very Important
Overconsolidation ratio Very Important
Diagenesis Less Important
Strain History Relatively Unimportant
Strain rate Relatively Unimportant
Effective material strength Less Important
Grain Characteristics (size, shape) Less Important
Degree of saturation Very Important
Dilatancy Relatively Unimportant

2.4.1 The influence of diagenesis


Diagenesis is a process involving seawater, subsurface brines or meteoric water that alters the
sediments up to the point of metamorphism. The diagenesis process alters the interparticle
structure and therefor also alters the stiffness of the soil with time (Benz, 2007).
The diagenesis process that have a large influence are cementation and aging, which
according to (Terzaghi K, 1996;Mitchell & Soga, 2005) are defined as change in various
mechanical properties resulting from a secondary compression under an external load.

9
2.4.2 The influence of confining stress
(Hardin & Richart, 1963) proposed a relationship between and the effective confining
stress :

( ) (13)
Where is a factor accounting for the type of soil.

For cohesive soils the exponent was previously set to 0.5. However, the value is very
sensitive and dependent on the liquid limit and the plasticity index . Figure 2.4 shows a
compilation of the exponent as a function of and for different clays at very small
strains (Benz, 2007).

Figure 2.5 shows how the stiffness decays with decreasing .

Figure 2.4 exponent m as a function of plasticity index and liquid limit (Benz, 2007)

Figure 2.5 Influence of the confining stress on the decay of small -strain stiffness (Benz, 2007)

10
2.4.3 The influence of void ratio
(Hardin & Richart, 1963) proposed another relationship between the propagation velocity
and void ratio for Ottawa sand:

( ) (14)
Where and are material constants.

Based on equation (14) (Hardin & Richart, 1963) derived a formula for how is dependent
on the :

( )
(15)
Equation (15) has proven to correspond reasonably well for clays with low surface activity.
For higher surface activity the coefficient 2.97 is replaced by a higher one (Benz, 2007).

Other relationships between and that are frequently used is, (Benz, 2007; Burland,
1989):

(16)
Where for sand and clay and in the range of for various clays.

Figure 2.6 shows how the stiffness decays with the soil relative density for tests performed at
a confining stress of 80 kPa. The soil relative density is expressed by the density index:

Figure 2.6 Influence of soil relative density on the decay on small-strain stiffness (Benz, 2007)

11
2.5 Measuring small-strain stiffness
2.5.1 In-Situ tests
In-situ tests are not measuring the small-strain stiffness directly but rather the elastic soil
mechanical propagation properties, e.g. (Mayne et al., 1999). Instead other parameters are
measured and related to the stiffness by mathematical relationships (Jardine et al., 1986). One
of the most common, indirect, methods of measuring the small-strain stiffness in the elastic
domain is by using wave propagation velocities.

The wave propagation velocity is dependent on the stiffness and density of the material that
the waves are propagating through. The higher the stiffness of a material is, the higher the
velocity will become (Kramer, 1996).

√ (17)

√ (18)

Where and are the wave propagation velocities for Primary (compression) and
Secondary (shear) waves, Figure 2.7.

( ) (19)

( )( ) (20)

By combining equation (17),(18),(19) and (20) the following relationship can be derived:

(21)

( )
(22)
( )

12
Figure 2.7 Primary and secondary waves (Kramer, 1996)

2.5.1.1 Cross hole seismic


To perform a cross hole seismic survey at least two vertically drilled bore holes are needed. In
one of the boreholes an energy source is lowered to the desired depth. In the other holes
receivers are placed at the same depth, Figure 2.8. The distance between the energy source
and the receivers must be known exactly, which often demands inclinometer reading in each
hole. By knowing the distance and by measuring the time it takes for the signal to reach the
receive it is possible to calculate the wave propagation velocities. By using the measured
velocity, can be obtained by Eq.(21) (Kramer, 1996).

The cross hole seismic survey is the most expensive testing method for in-situ small strain
stiffness, however it is also the most reliable (Benz, 2007).

Figure 2.8 Cross hole seismic survey (Clayton, 2011)

13
2.5.1.2 Continuous surface wave
A vibrator sends out single-frequency sinusoidal force at the ground surface. The waves travel
through the ground and are measured by geophones at a certain distance from the vibrator,
Figure 2.9. By using different frequencies, a profile of phase velocity against wavelength is
obtained. From the phase velocity – wavelength profile it is possible to calculate a stiffness-
depth profile (Clayton, 2011).

2.5.2 Laboratory tests


Several laboratory testing methods have been developed to determine the static and dynamic
small-strain stiffness. Laboratory testing tends to give a lower small strain-stiffness value than
field tests (Jardine et al., 1986). The two main reasons for this are explained by sample
disturbances and errors related to interpretation, such as assumptions or idealisations. For the
laboratory results to be closer to the in-situ values, great care must be taken to both of the
possible error sources (Wood, 2016).

2.5.2.1 Bender elements


The bender element method was first introduced at the end of the 1970s (Viggiani &
Atkinson, 1995). The bender element method has become more popular over the past decade,
mostly due to its perceived simplicity. The bender element consists of two thin piezo-ceramic
plates that are bonded together, with the soil sample in-between them, Figure 2.10.

By applying a voltage at one of the plates, it will contract or extend, generating seismic waves
in the soil. The element on the opposite side will register the incoming waves. By knowing
the distance between the elements and the time obtained for a wave to travel through the soil,
between the plates, the velocity of the waves can be calculated (Clayton, 2011).With the
velocities known, can be obtained by using Eq.(21).

Figure 2.9 Continuous surface wave survey (Clayton, 2011)

14
Figure 2.10 S ketch of a bender element (Clayton, 2011)

2.5.2.2 Resonant column testing


The resonant column testing has been used for more than 40 years. It is a method for
determining and at very small strains. The method also estimates the rate of stiffness
degradation with increasing strain (Clayton, 2011). The method works by vibrating the soil
specimen with a certain frequency. The frequency is increased until the first-mode of
vibration for the specimen is reached. At this frequency, the resonance frequency and
amplitude is measured. By knowing the geometry of the specimen, the measured data is used
to calculate the wave propagation velocity. The velocity is then used to calculate , Eq.(21).

Figure 2.11 shows a schematic drawing of the resonant column testing apparatus.

Figure 2.11 S chematic drawing of a resonant column apparatus (Clayton, 2011)

15
2.6 Ground investigation methods
Ground investigations are crucial to obtain knowledge about the soil and groundwater
conditions but also to determine the soils properties, (Lambe & Whitman, 1979) Ground
investigations can be divided into two sections, in-situ tests or laboratory tests.

The advantage with laboratory testing is that there is a high degree of control over the
conditions compared to the field tests (Wood, 2016). The disadvantage with laboratory tests is
that the soil sample is being taken out of its original conditions, which can lead to different
results. Great care must be taken when soil samples are being extracted from the field.
Experience has shown that many soil types, including clay, are very sensitive to sampling
disturbance. If the sampling disturbance is significant it may lead to the laboratory test being
almost worthless (Davis & Selvadurai, 1996).

In this section, the focus will be on the laboratory test method triaxial testing.

2.6.1 Oedometer tests


The oedometer test is a common method of determining the soils deformation properties
(Larsson, 2008). The oedometer test can be performed in two variants, incremental load steps
or constant rate of strain (CRS). In Sweden, the CRS test is the more common of the two
methods. The CRS method determines several parameters such as, the pre-consolidation
pressure , coefficient of consolidation , permeability and the modulus .

2.6.2 Triaxial tests


The triaxial test is one of the most common and versatile performed geotechnical laboratory
tests for determining the shear strength and stiffness of soil (SGF, 2012). In the triaxial test a
soil specimen gets loaded in its axial and horizontal direction (Lambe & Whitman, 1979). The
deformations and strains due to the loading are measured and evaluated to determine the
parameters of the soil. The primary parameters that are determined from the test are: Angle of
shearing resistance ϕ΄, cohesion c and the undrained shear strength , depending if the
shearing is carried out in a drained in an undrained way. Other parameters such as the
compression index , and can also be determined (GDS, 2013). The advantage of the
triaxial test compared to other methods is the ability to simulate the original stresses and pore
water pressures in the soil (SGF, 2012).

Figure 2.12 shows an illustration of a triaxial cell.

16
Figure 2.12 Illustration of a triaxial cell (GDS , 2013)

2.6.2.1 Standard triaxial tests


A triaxial test usually consists of four stages: specimen and system preparation, saturation,
consolidation and shearing (Jardine, Symes, & Burland, 1984). In the first stage the soil
sample taken from the field is prepared and put into the triaxial cell. It is important to keep the
disturbance of the sample to a minimum during the preparation since sample disturbances can
affect the results. The purpose of the saturation stage is that all voids within the test sample
are filled with water. Before the specimen gets sheared, the specimen is brought up to the
desired effective stress. In the fourth and last stage of the test, the specimen is sheared by
applying an axial strain at a constant rate. The rate at which the specimen is being sheared is
dependent on which triaxial test that is conducted.

Figure 2.13 S tress state during triaxial compression (GDS , 2013)

17
There are three types of standard triaxial tests that can be performed (Lambe & Whitman,
1979):

 Unconsolidated Undrained test (UU)


 Consolidated Undrained test (CU)
 Consolidated Drained test (CD)

The Unconsolidated undrained test is the fastest and simplest test procedure used for
evaluating the short-term soil stability. During loading of the specimen, the total stresses are
measured which allows the undrained shear strength to be determined.

The consolidated drained test describes the long-term response of the soil and is used for
determining the angle of shearing resistance and the cohesion. It is a more time-consuming
process for testing cohesive soils than the unconsolidated undrained test. The reason for this is
that the rate of strain must be slower to allow for small pore water pressure changes.

The consolidated undrained test is the most common triaxial test (Larsson, 2008). It
determines the same parameters as the consolidated drained test, but at a shorter time. During
the consolidated drained test the change in the excess pore pressure can be measured within
the specimen as shearing takes place, leading to the ability of using a higher rate of strain
(GDS, 2013).

The standard triaxial tests can be performed as active or passive tests. During active triaxial
tests the specimen is being loaded by a higher axial than horizontal load. During passive tests,
the horizontal load applied is higher than the axial. The purpose of the active and passive tests
is to simulate real stress behaviours that would occur in the field.

Figure 2.13 illustrates the stress state during triaxial compression.

2.6.2.2 Triaxial test presentation


Performed triaxial tests are often presented graphically through several plots. The most
common ones are:

 Stress against axial strain ( )


 Change in pore pressure against axial strain ( )
 Stress path with effective stress ( )

Figure 2.14 illustrates results from an active undrained triaxial test performed on clay.

18
Figure 2.14 Presentation of an active undrained triaxial test: a) b) c) (S GF, 2012)

2.6.2.3 Evaluation of the elasticity modulus in triaxial tests


The elasticity modulus of a material can be determined by the slope of a line in the stress-
strain curve, Figure 2.15. However, the elasticity modulus will vary depending on how the
line is defined. The line can be drawn as a tangent or as a secant to the strain-stiffness curve.
The line can be defined in the beginning of the stress-strain curve, when the strains are small
or at the end of the curve, when the stresses have decreased. There is no rule of thumb for
deciding where this line should be placed in order to get a realistic value of the elasticity
modulus (Briaud, 2001).

The elasticity modulus can be decided through several equations, following the chapter 2.2.2:

[ ( )] ( ( ))

[ ( )] ( ( ))
(23)
[ ( )] ( ( ))

(24)

(25)
( )

19
(26)

Depending on where the line is drawn, several elasticity moduli can be determined (Schanz,
Vermeer, & Bonnier, 1999). , Unloading modulus , secant modulus at 50% of the shear
strength reloading modulus and cyclic modulus are examples of moduli that can be
determined, Figure 2.16.

Figure 2.15 Determination of the elasticity modulus (Briaud, 2001)

Figure 2.16 Definition of different slopes used to evaluate the corresponding elasticity modulus (Briaud, 2001)

20
2.7 Empirical correlations for determining the soil stiffness
In the absence of laboratory tests for determining the elasticity modulus empirical correlations
can be used (Duncan & Bursey, 2013). The correlations are derived from parameters that can
be obtained from in-situ, CRS or routine laboratory tests, , or . The results from the
correlations are often more useful and effective than direct measurements of the elasticity
modulus (Duncan & Bursey, 2013).

The correlation seen in Eq. (28) below is dependent on the plasticity index, , Eq. (27). In
Sweden, the plastic limit is not determined in routine laboratory tests, however Table 2.2
shows correlations between the and (IEG, 2011)

(27)

Table 2.2 Correlation between and for clay

Explanation Liquid limit [%] Plasticity index


Low plasticity clay 15-30
Medium plasticity clay 30-50 10-25
High plasticity clay 50-80 25-50
Very high plasticity clay

(28)

(29)

Eq. (29) is based on an assumption often used in practical design (Trafikverket, n.d.)

Where is calculated according to (Trafikverket, 2014)

( ) (30)

The small-strain stiffness is measured, as mentioned earlier, through dynamical tests


performed in a laboratory or in the field, (Atkinson, 2000; Burland, 1989). However, the
small-strain stiffness is not too often used in routine design and therefore far from always
measured. In Sweden, it is common to relate to the undrained shear strength through
several empirical correlations. Table 2.3 below is originally found in the thesis by Wood,
(2016).

In the same thesis by Wood, (2016), the correlations were compared with field measurements
of for two different geological deposits at Gothenburg and Uppsala, site 1 and site 10

21
respectively. The spheres and rings in Figure 2.17 and Figure 2.18 represent the degree of
sample disturbance in the laboratory test and the quality of the field test respectively. Large
spheres indicates large sample disturbances and large rings represents very good quality of the
field test (Wood, 2016).

Figure 2.17 and Figure 2.18 illustrates the accuracy of the correlations given in Table 2.3 for
determining .
Table 2.3 Empirical correlations for the small-strain shear modulus (Wood, 2016).

Source Soils Correlation


Andréasson (1979) High plasticity post glacial soft
(uncorrected shear vane) clays (Gothenburg Area)

Stokoe (1980) (average Su


from CAUC, CAUE & DSS) Clays ( )

Larsson & Mulabdic (1991) High- low plasticity soft clays


(corrected shear vane and at (Western and central Sweden ( )
some sites SuDSS) and Norway)
Larsson & Mulabdic (1991) Low plastic and varved or
(corrected shear vane and at otherwise inhomogeneous soils ( )
some sites SuDSS) & organic clays
Bråten et al. (1991) Medium and low plasticity soft
(SuDSS) clays (Norway)
( )

Long et al. (2013) Medium plasticity firm clays


(Su from CAUC tests) (Ireland)

22
Figure 2.17 Comparison between the empirical correlation for and field measurements from site 1, Gothenburg.
(Wood, 2016).

Figure 2.18 Comparison between the empirical correlation for and field measurements from site 10, Uppsala
(Wood, 2016).

23
3 Soil modeling

3.1 Introduction to numerical modelling


During the past 40 years, the fast development of computers and numerical modelling
software has made it possible for sophisticated geotechnical problems to be analysed by most
engineering practices. The numerical software offers a variety of constitutive models that the
engineers can choose from. The constitutive models are often based on different theories and
require different input parameters (Brinkgreve et al., 2015).

Plaxis is an advanced numerical simulation program that was created during the 1970’s at the
University of Delft, in the Netherlands (Brinkgreve et al., 2015). It was originally used for
elastic-plastic calculations for axi-symmetrical problems based on higher-order elements.
Today Plaxis offers a wide range of advanced soil models and simulations in both 2D and 3D
(Brinkgreve et al., 2015).

In this thesis, the constitutive models Mohr Coulomb and Hardening Soil Small have been
studied in both 2D and 3D models of the case study presented later. The Mohr-Coulomb
model is generally used for drained analysis in geotechnical engineering, and the hardening
soil model is similar to other types of small strain model used in numerical analysis, e.g. in
Jardine et al., (1986)

3.2 Mohr coulomb (MC)


The Mohr Coulomb model is one of the most generally used plasticity constitutive models,
(Lambe & Whitman, 1979). It assumes the soil to be elastic perfectly-plastic, Figure 3.1, and
normally non-associated plastic strain is assumed (Vermeer & de Borst, 1984). The Mohr
Coulomb model is a straight-forward constitutive model with few parameters method that can
be used as a first analysis of the problem. The model only includes a small number of features
that the soil behaviour shows in reality. The model does not take stress, stress-path nor strain
dependency of stiffness or anisotropic stiffness into account (Brinkgreve et al., 2015).

The first part of the Mohr Coulomb constitutive model, the linear elastic part, is based on
Hooke’s law of isotropic stiffness, Eq.(7). The second part, the perfectly-plastic part, is based
on the Mohr Coulomb failure criterion. The Mohr Coulomb Failure Criterion can be
expressed as Eq.(31) and illustrated in Figure 3.2.

(31)

Where is the shear stress and is the normal effective stress on the failure plane.

24
Figure 3.1 Illustration of the elastic perfectly-plastic model (Brinkgreve et al., 2015)

Figure 3.2 Mohr Coulomb failure criterion for undrained case (Brinkgreve et al., 2015)

Since the plastic state is an important part of the Mohr Coulomb, it is important to know when
plastic yield occurs. For plastic yield to occur, development of irreversible strains needs to
take place. To evaluate if irreversible strains have taken place a yield function is introduced
as a function of the stress and strain. Eq. (32) shows the yield function as formulated in
terms of principal stresses (Brinkgreve et al., 2015).

The Mohr Coulomb yield criterion is represented by a hexagonal cone in the principal stress
space, Figure 3.3.

( ) ( ) ( ) (32)

The Mohr Coulomb model requires five input parameters that can be obtained from basic soil
testing, Table 3.1.

25
Figure 3.3 Mohr Coulomb yield surface in the principal stress space (c=0) (Brinkgreve et al., 2015)

Table 3.1 Basic Input parameters for the Mohr Coulomb soil model

Parameter Description
Internal friction angle
Cohesion
Dilatancy
Young’s modulus
Poisson’s ratio

3.3 Hardening Soil (HS)


The hardening soil model is more advanced than the Mohr Coulomb model for simulating the
behaviour of soil, (Schanz et al., 1999), and is primarily based on the double hardening
model presented by Vermeer in 1978, (Vermeer, 1978). The total strains are calculated using
a stress-dependent stiffness. The stress-dependent stiffness is different for both loading and
unloading or reloading (Surarak et al., 2012).

In the HS model, the stress-strain relationship due to primary loading is assumed to be a


hyperbolic curve, Figure 3.4. The hyperbolic function for the undrained triaxial test was stated
by (Kondner, 1963) and can be formulated as:

(33)

Where is the deviatoric stress and is the axial strain.


is the ultimate deviatoric stress at failure and derived from the Mohr Coulomb failure
criterion involving the strength parameters and and is defined as:

( ) (34)

26
Figure 3.4 Hyperbolic S tress-strain relationship in primary loading for a standard drained triaxial test (Brinkgreve et
al., 2015)

And is defined as:

(35)

Where is the failure ratio, which in PLAXIS has the standard value of 0.9.

in Figure 3.4, is the initial stiffness, and is due to the highly non-linearity of soil the is
used instead of for calculations. is the confining stress dependent stiffness modulus, at
50 % of the shear strength, for primary loading and given by Eq.(36).

( ) (36)

Where is the reference stiffness modulus corresponding to the reference stress, which in
PLAXIS is set to as a default value.

Equation (36) shows that the actual stiffness for the HS model depends on the minor principle
stress , which is the effective confining pressure in a triaxial test (Brinkgreve et al.,
2015). The amount of stress dependency is given by the power-law coefficient (Surarak,
2010). For clay soils the value of is in the range of 0.7-1.0 (Benz, 2007).

The stress dependent stiffness modulus for unloading and reloading stress paths can be
calculated by Eq.(37).

( ) (37)

is the reference modulus for unloading and reloading corresponding to the reference
pressure . In PLAXIS is set equal to due to practical
reasons.

27
In the HS model the shear yielding function is defined as:

̅ (38)

Where ̅ is given by Eq. (39) and is given by Eq. (40).

( )
̅ { } (39)
( )

(40)

By looking at Eq. (38)- (40) it can be seen that the parameters, , obtained
from triaxial test controls the shear hardening yield surface.

Another important stiffness parameter to control the magnitude of the plastic strains is the
reference oedometer modulus . Similar to the Unloading modulus and secant
modulus , the oedometer modulus can be calculated by Eq. (41).

( ) (41)

The HS model needs the following input parameters, Table 3.2.


Table 3.2 Input parameters for the Hardening soil model

Parameter Description
Internal friction angle
Cohesion
Dilatancy
Secant stiffness in standard drained triaxial tests
Unloading/reloading stiffness
Tangent stiffness for primary oedometer loading
Exponent for stress-level dependency
Poisson’s ratio for unloading-reloading
Reference stress for stiffness
value for normal consolidation
Failure ratio

28
3.4 Hardening Soil Small (HSS)
The Hardening Soil Small Model is an addition to the Hardening Soil model, with the
capacity of modelling the soil behaviour at very small strains, and is similar to other types of
soil models presented in, e.g. Jardine et al, (1986). The HSS model uses the same input
parameters as the HS model, Table 3.2, except for two additional parameters (Benz, 2007):

 which is the initial or very small-strain shear modulus


 which is the shear strain level at which the secant shear modulus is reduced to
about 70% of

is, similarly to , dependent on the confining pressure, Eq. (42)


(Brinkgreve et al., 2015).

( ) (42)

The parameter defines the level of shear strain where has been reduced to 70% of its
initial value, Figure 3.5. One of the most used models in soil dynamics is the Hardin-Drnevich
relationship, the hyperbolic law for larger strains, Eq. (43).

(43)
| |

Where is the threshold shear strain and is quantified as:

(44)

Where is the shear stress at failure.

Eq.(43) relates to large strain behaviour of soil and has been modified in a study by
(Santos & Gomes Correia, 2001) in order to fit small strain behaviour of soil, Eq.(45)

(45)
| |
Where is a factor set to 0.385.

can be approximated with Eq.(46).

[ ( ( )) ( ) ( )
(46)

Where is negative.

29
Figure 3.5 Illustration of the parameter (Brinkgreve et al., 2015)

30
4 Case study: Bridge over
Ulvsundavägen
4.1 Introduction
The project studied in this thesis is a new construction of an 85 meters long and 16 meters
wide bridge for public transport, cyclist and pedestrians located between Bromma airport and
Ulvsundavägen in Stockholm, Sweden, Figure 4.1.

During the casting of the bridge decks temporary formworks were used as supports, Figure
5.1. The formwork transfers the entire load to the ground before the bridge deck has hardened
and the loads can be taken up by the supports of the bridge. During this time, the temporary
formwork is causing the underlying soil to deform.

The whole bridge rests on ten supports, creating nine spans in between them. A study was
carried out during the design phase of the bridge. The purpose of the study was to investigate
if the loads from the formworks could be transferred directly to the ground or if piling was
necessary to keep the deformations within desired limits. The purpose was also to calculate
the deformations caused by the temporary formworks, so that the height of the formworks
could be adjusted accordingly.

Several ground investigations have been made in the area. The investigations consisted of
both field test and laboratory tests. The purpose of the investigations was to determine the
ground conditions and soil parameters in order to calculate the predicted ground deformations
under the temporary formwork.

Piston sampling, Soil-rock probing, weight sounding and groundwater measurements were
carried out during the site investigation. Clay samples that were extracted from the field by
the piston sampler were sent to a laboratory for routine, triaxial and CRS-tests to determine
the soil properties.

The results from field and laboratory test for a studied section will be presented in Appendix
A and used for analytical and numerical calculations in chapter 5.
Material parameters that could not be determined by field or laboratory tests were obtained
from TK Geo 13 (Trafikverket, 2014).

31
Figure 4.1 S atellite picture of the location of the bridge. The highlighted area shows the studied section in this thesis.
Picture taken from (www.eniro.se 2014-04-20)

32
4.2 Studied section
The ground deformations in one of the nine spans were chosen to be investigated in this
thesis, namely the deformations between supports 6-7.

The span between supports 6-7 was chosen due to two main reasons:

 The ground deformations were measured prior to and after the casting of the concrete
bridge deck at two locations close to support seven.
 Triaxial- as well as CRS-tests have been performed on clay samples taken from
borehole 15E03, Figure 4.2.

Figure 4.2 S tudied span and location of borehole 15E03

33
4.2.1 Ground conditions
The soil profile between supports 6-7 was determined through weight soundings, soil-rock
probing, CRS-, triaxial- and routine laboratory tests. Table 4.1 show the results from the
routine laboratory and CRS-tests. Figure 4.3 shows the result from a sounding in borehole
15E03.

The ground water level was measured at two locations, support 4 and support 10, Table 4.2.
Based on the measurements the ground water level was assumed to be 0.9 meters below the
ground surface at point 15E03.

The soil between supports 6-7 initially consisted of an approximately 1 m thick layer of fill
material, mostly silt, sand and gravel. Underneath the fill layer there is a 0.6 m thick layer of
stiffer clay. The stiffer clay is resting on softer clay with the thickness of approximately 10
meters. The softer clay rests on an up to two-meter-thick layer of moraine resting on solid bed
rock. During the construction of the bridge the initial 1-2 meters were excavated and replaced
by crushed rock for increased bearing capacity. In the studied section an assumption was
made that the upper two meters were excavated and replaced by crushed rock.

Figure 4.4 shows the interpreted soil profile that have been used for the analytical and
numerical calculations in chapter 5.

The depths in Figure 4.4 are based on the distance from the ground surface +0.0

Table 4.1 Results from the routine laboratory and CRS -tests taken from borehole 15E03

Depth k [m/s]
[ ] [ [ [ [ [ [ [ M´
[m]
1.5 1.67 69 67 17 19 15.56 54 531 88 12.4 3.7E-10
2.5 1.58 80 63 22 12 10.11 40 266 55 12.2 5.9E-10
3.5 1.54 93 72 26 12 9.52 44 234 57 13.2 6.9E-10
5.5 1.64 65 61 10 12 10.25

Table 4.2 Conducted ground water measurements in the area

Measuring Measuring period Ground water level Depth below ground surface (m)
point
09R187GV 26/7-2009 – 30/1- +2.9 - +1.9 1.8-2.8
(+4.7) 2015
13W018G 8/28-2013 – 30/1- +3.3 - +2.9 0.7-1.1
+(4.0) 2015

34
Figure 4.3 Result from sounding in borehole 15E03

Figure 4.4 Interpreted original soil profile(left) and soil profile after replacing the upper two meters during
construction(right)

35
4.3 Ground deformation measurements
Ground deformation measurements were conducted at a total of 12 points along the bridge.
The measurements were done directly and 21 days after the casting of the concrete bridge
deck.

Figure 4.5 shows where deformations have been measured between support 6-7.

The measured deformations for the studied section are presented in Table 4.2 and are used as
a reference to the calculated deformations in chapter 5. The accuracy of the measured
deformations is around .

Figure 4.5 Conducted ground deformation measurements between support 6-7

Table 4.3 Measured ground deformations between supports 6-7

Point Load group Directly after 21 days after Total


casting [mm] casting [mm] deformation
[mm]
11 1 4 9 13 ( )
12 2 5 9 14 ( )

36
5 Calculation procedure
5.1 Formwork geometry
The formwork supporting the bridge deck during construction consists of 153 steel rods
placed on double rectangular wooden plates of sizes 0.1 m*0.4*0.4 m, Figure 5.1. The
spacing between the steel rods and the rod forces can be seen in Figure 5.2.

Figure 5.1 Temporary formwork for casting the bridge deck

1 2

Figure 5.2 Geometry and load distribution in the steel rods

37
5.2 Empirical correlation study
Empirical correlations are very important in geotechnical engineering. One of the most
important correlation used in the design phase is the correlation for the elasticity modulus.
Many correlations exist for determining the elasticity modulus, as shown in section 2.7. The
correlations were compared to the initial elasticity modulus and the secant modulus at 50%
of the peak strength ,evaluated from triaxal test from several sites around Sweden, Figure
5.3, to study the accuracy of the correlations.

Figure 5.3 Approximate location of the performed triaxial and laboratory tests used in the correlation study.

38
5.3 Analytical Calculation
A simplified analytical calculation was made for the span between supports 6-7 as a reference
to the measured deformations and the calculated deformations using more advanced
numerical soil models. Deformations due to consolidation in the clay have been assumed to be
occurring during the first seven days. The assumption is based on that the concrete bridge
deck would have cured enough for the loads to be transferred to the ground through the
permanent supports rather than through the temporary formwork.

The analytical calculation was based on Eq. (47)-(49).

∫ (47)

∫ ( ) (48)

∫ ( ) (49)

Where:

= vertical stress increase caused by the form work.


was calculated using the 2-1 method, which assumed the loads to decrease with increased
depth, due to load spread, Eq. (50). The load spread was assumed only to be occurring in the
fill material.

( ) (50)

Where:
b = Width of a fictive plate, z =Depth below the ground surface, Q = Distributed load from
the form work and h= Thickness of the layer

The distributed loads were calculated by summing the forces in the steel rods and dividing
them by an area, which is illustrated by the two larger blue rectangles 1 and 2 in Figure 5.2.

For simplicity an average distributed load of 29 kN/m2 was used for both areas.

39
Equations (47)-(49) calculates the deformations after an infinite period of time. To be able to
calculate the deformations after seven days, as assumed previously, a time factor was
implemented, Eq. (51).

(51)

Where:
= Time factor connected to the degree of consolidation for the clay
= Coefficient of consolidation (evaluated from the performed CRS-test, Appendix A -
Figure 1)
k = Permeability
= Drainage distance in the clay
= Time in second
= Unit weight of water

The modulus used in Eq. (47)-(49) has for simplicity been set to , as previously seen in
Eq.(29).

The deformation in the clay after seven days were calculated as ∑


Where: √

The clay was divided into a single layer and the deformation properties obtained from CRS-
test at 3.5 m depth was used for the calculation of the deformation, Table 5.1.

Table 5.2 shows the evaluated stress state in the soil.

Table 5.1 Results from the CRS -test performed at 3.5 m depth

Depth
[ ] [ [ [ [ [ [ M´ k [m/s] cv [m2 /s]
[m]
3.5 1.54 93 72 9.52 44 234 57 13.2 6.9E-10

Table 5.2 Evaluated stress state in the soil

Depth
[ [kPa] [kPa] [kPa] [kPa]
[m]
0 0 0 0 39 40
0.9 18 0 18 24.6 42.6
2.5 50.9 16 34.9 21.4 56.3
3.5 66.3 26 40.3 21.4 61.7
The deformations occurring in the fill material were calculated using theory of elasticity, Eq.
(52).

(52)

40
5.4 2D Numerical simulation
The 2D numerical calculations have been performed using two constitutive models, the
simpler Mohr Coulomb- and the more refined Hardening soil small model both with the
Undrained-A condition in Plaxis.

Figure 5.4 shows the geometry used in the 2D numerical simulations.

Figure 5.4 2D geometry used in the MC and HS S simulations

41
5.4.1 Assumptions
The loads had to be simplified due to the 2D geometry. The point loads from each steel rod
were transformed into line loads by dividing the forces with the distances between the rods.

The two load cases that were used in calculations, will be referred to as load group 1 and load
group 2, see the orange rectangular in Figure 5.2.

Where:

In the numerical simulations, the settlements have been assumed to be occurring during the
first seven days, see section 5.3.

5.4.2 Input parameters


The input parameters for the MC and HSS-model can be chosen from empirical correlations,
field data or customized to fit laboratory test results by the help of the Plaxis “SoilTest”
program. The SoilTest program is based on a single point algorithm and allows the user to
simulate soil lab tests quickly. The input parameters can be manipulated manually in the
SoilTest program until the simulated test corresponds well to the actual performed test.
However, Plaxis also has a built-in function, “parameter optimisation”, that allows the user to
choose which and how much the parameters should get manipulated for the simulated test to
match the performed laboratory test.

Two ways of obtaining the input parameters for the Mohr Coulomb calculations have been
investigated:

 Strictly empirical correlations


 Parameters obtained from triaxial test combined with correlations for the other
parameters

Three ways of obtaining the input parameters for the Hardening soil small calculations have
been investigated:

 Using the Plaxis Parameter Optimisation function - Allowing Plaxis to automatically


match the fictive stress-strain curve obtained from Plaxis “SoilTest” program to the
real curve obtained from the triaxial test.
 Strict empirical correlations – All input parameters are based on empirical
correlations.
 Customized input – Parameters obtained from the triaxial test were mixed with
parameters obtained from the Plaxis parameter optimisation function and from
empirical correlations.

42
The Plaxis parameter optimisation requires a laboratory test data to be used. The strict
empirical correlation method can be used when simpler ground investigation methods
have been performed, e.g. routine laboratory tests.

5.4.2.1 Stiffness moduli


The elasticity moduli used in the Mohr Coulomb model are the same as previously stated in
section 5.3.1.

The Hardening soil small model requires three different elasticity moduli as input parameters,
.

was assumed to have no effect on the actual deformation since no unloading or reloading
occurred. Therefor the value of was set to a standard value of . The value for
varies depending on the method that was used to obtain the input parameters.
was set to for the customized input and automatically determined by Plaxis for the
parameter optimisation case. For the strict empirical method was set to

. The value for was automatically determined in Plaxis for the parameter optimisation
and the customized input methods. For the strict empirical correlation method, the value for
was set to .

5.4.2.2
and are two important parameters that can be determined by correlations or by
performing triaxial test. Equations (53)-(55) were used for determining empirically
(Trafikverket, n.d.). Eq.(56) was used for evaluating from the performed triaxial test. Due
to the OCR of 1.37 the correlations for both normally consolidated and overconsolidated clay
was studied.

(53)

( ) (54)

(55)

(56)

43
5.4.2.3
and are two important parameters in the HSS model. As mentioned earlier, is not
frequently measured in Sweden.

In this thesis, has been determined by the parameter optimisation function in Plaxis and by
the correlation ( ) .

Wood (2016) showed that the correlation is very conservative compared to the measured field
values, Figure 2.17 and Figure 2.18. To make the correlation less conservative and closer to a
realistic value a factor 3 was used with the correlation for , Eq.(57).

[( ) ]
(57)

The factor 3 was determined based on the results from field measurements of at a site in
Uppsala, Sweden, in the thesis by (Wood, 2016).

The correlation for is dependent on the plasticity index. For the determination of the
plasticity index was set to 30 %, based on the average liquid limit obtained from routine
laboratory tests and the correlation, Table 2.2, resulting in a value of .

The parameter was determined in three different ways:

 Automatically by the Plaxis optimisation function


 Approximated by Eq.(58) for the customized parameter method
 Assumed value for the strict empirical method.

[ ( ( )) ( ) ( ) (58)

44
5.5 3D Numerical simulation
The 3D numerical calculations have been performed using the same two constitutive models
as for the 2D case. The calculations done in the 3D are based on the same parameters as in the
2D case and will therefore not be discussed here. Just like the 2D model, the 3D model takes
consolidation settlements into account. The settlements have been assumed to be occurring
during the seven first days, see section 5.3.

The permeability of the clay, obtained from the CRS-tests, was assumed to be the same in all
directions. The assumption that the permeability is the same in all direction is probably not
correct, but there are no measurements in the horizontal direction. The geometry that was
used in the 3D numerical simulation was based on the soil profiles stated in section 4.2.1.
Figure 5.5 shows the initial geometry and loads used for the 3D simulation.

Figure 5.5 Initial geometry for the 3D numerical simulation

45
5.6 Parameter sensitivity analysis
Due to the complexity of the Hardening Soil Small constitutive model and due to the
uncertainties of determining some of the input parameters, e.g. a simplified
sensitivity analysis was performed. The purpose of the sensitivity analysis was to investigate
how large effect each of the input parameters would have on the calculated deformations. The
result of the simplified sensitivity analysis would give an indication of which of the
parameters that needed to be determined with greater caution.

The customized parameters were used to calculate a deformation which constituted as a


reference deformation in the sensitivity analysis. Most of the input parameters were varied by
increasing or decreasing the reference value by 5,10 and 15% and thereafter calculating a new
deformation. However, some parameters could not be varied as much due to regulations in
Plaxis.

Table 5.3 shows which parameters that were studied.


Table 5.3 Parameter variation

Parameter

46
6 Results
6.1 Empirical correlation study
The results from the comparison show that the correlations for determining the elasticity
modulus show large variations. The standard correlation used for high plasticity clays in
Sweden, gives a very low elasticity modulus compared to the evaluated, for
the area where the bridge was built, as seen in Figure 6.1-6.3. The standard correlation
is also underestimating the elasticity modulus from the Norrköping test.
However, the correlation seems to be reasonable for the Uppsala test, Figure 6.6.

In general, the less commonly used correlation in Sweden seems to give the
best result compared to the triaxial test, Figure 6.1-6.3.

Triaxial tests have only been performed at one depth for the Stockholm, Ulvsunda area. The
moduli are considered constant at the increasing depth, as indicated by the black lines.

Stockholm, Ulvsunda 15E01


Elasticity modulus [kPa]

0 2000 4000 6000 8000 10000 12000


1

1.5 250*cu

2 Correlation 30% IP
Depth below ground surface [m]

2.5 Correlation 35% IP

3 Correlation 40% IP

3.5 Correlation 45% IP

4 Correlation 50% IP
E50 EI
4.5 EI

5 E50

5.5

Figure 6.1 Comparison between correlations and triaxial results from borehole 15E01 taken from the construction
site.

47
Stockholm, Ulvsunda 15E09
Elasticity modulus [kPa]
0 5000 10000 15000
2

2.5 250*cu
Depth below ground surface [m]

Correlation 30% IP
3 Correlation 35% IP
Correlation 40% IP
3.5
Correlation 45% IP
E50 EI
4 Correlation 50% IP
EI
4.5 E50

5.5

Figure 6.2 Comparison between correlations and triaxial results from borehole 15E09 taken from the construction
site.

Stockholm, Ulvsunda 15E03


Elasticity Modulus [kPa]
0 5000 10000 15000
1 250*cu

Correlation 30% IP
1.5
Correlation 35% IP
Depth below ground surface [m]

2 Correlation 40% IP

Correlation 45% IP
2.5
Correlation 50% IP
E50 EI
3 EI

E50
3.5

Figure 6.3 Comparison between correlations and triaxial results from borehole 15E03 taken from the construction
site.

48
Norrköping, Location 1
Elasticity Modulus [kPa]
0 5000 10000 15000 20000 25000
4

6 250*cu
Depth below ground surface [m]

Correlation 30% IP
8 Correlation 35% IP
Correlation 40% IP
10 Correlation 45% IP
Correlation 50% IP
12 EI
E50
14

16

Figure 6.4 Comparison between correlations and triaxial results from the Norrköping Area

Norrköping, Location 2
Elasticity modulus [kPa]
0 5000 10000 15000 20000 25000
4

6
Depth below the ground surface [m]

250*cu
Correlation 30% IP
8 Correlation 35% IP
Correlation 40% IP
10 Correlation 45% IP
Correlation 50% IP
12
EI
E50
14

16

Figure 6.5 Comparison between correlations and triaxial results from the Norrköping Area

49
Uppsala
Elasticity Modulus [kPa]
0 5000 10000 15000 20000 25000
0

250*cu
10
Depth below ground surface [m]

Correlation 30% IP
20
Correlation 35% IP

30 Correlation 40% IP

Corelation 45% IP
40
EI
50
E50

60

Figure 6.6 Comparison between correlations and triaxial results from the Uppsala Area

6.2 Calculated deformations


The results from the calculations have been normalized, i.e. . Where are the
calculated deformations and are the measured deformations. It is important to have in
mind that the measured deformations are very small (13-14 mm). Due to the small
deformations, even a small change of will lead to deformation changes of .
The change represents the resolution of the measurements.

6.2.1 Analytical calculations


The analytical calculations give good result compared to the measured deformation. The
difference between them is approximately 5-6 mm, which is considered to be very good due
to the simplicity of the analytical calculation.

Table 6.1 Results from the analytical calculation

Point Total Total Degree of Deformation after Normalized


deformation in deformation consolidation seven days result
fill layer in clay
11 1.6 mm 690 mm 2.56 % 19.23 mm 147 %
12 1.6 mm 690 mm 2.56 % 19.23 mm 137 %

50
6.2.2 2D Mohr Coulomb
The Mohr Coulomb calculations in 2D show that the results are strongly dependent on the
choice of soil modulus and the magnitude of the loads. For load group 1 the deformation
ranges between 140-500% where the initial elasticity modulus gives the most accurate results,
Figure 6.7. The correlation that is commonly used in Sweden for high plasticity clays,
overestimates the deformations by a factor of approximately 5.

As the load decreases for load group 2, the correlations show better accuracy. For load group
2 the deformation ranges between 109-340%, Figure 6.8. The standard correlation
overestimates the deformation by a factor of approximately 3.4 for load group 2.

Figure 6.9 and Figure 6.10 shows the deformations and strain levels in the soil using 2D MC
with the initial elasticity modulus obtained from triaxial test.

600%
2D MC Normalized deformations. Load group 1
486%
Normalized Deformations [%]

500% 408%
374%
400% 338%
303%
267%
300%
174%
200% 143%
100%
100%
0%
1

Measured deformation EI E50


Correlation 30% IP Correlation 35% IP Correlation 40% IP
Correlation 45% IP Correlation 50% IP Correlation 250*Cu

Figure 6.7 Influence of the elasticity modulus for load group 1

2D MC Normalized deformations. Load group 2


400% 335%
Normalized deformations [%]

350% 287%
300% 264%
250% 213% 218%
189%
200% 133%
150% 100% 109%
100%
50%
0%
1

Measured deformation EI E50


Correlation 30% IP Correlation 35% IP Correlation 40% IP
Correlation 45% IP Correlation 50% IP Correlation 250*Cu

Figure 6.8 Influence of the elasticity modulus for load group 2

51
Figure 6.9 Calculated deformations using 2D MC model, Load group 1, with the initial elasticity modulus obtained
from triaxial test

Figure 6.10 S train level in the soil profile for load group 1 using 2D MC model with initial elasticity modulus obtained
from triaxial test

52
6.2.3 2D HSS Model
The results from the 2D HSS simulation show that the deformation varies depending on the
method of determining the parameters, Figure 6.11 and Figure 6.12 The most accurate result
is obtained when the Plaxis Parameter optimisation was used for determining the input
parameters based on matching the fictive to the real triaxial test curve. The 2D HSS model
overestimates the deformations for all cases except for the Plaxis parameter optimisation in
load group 2. Figure 6.13 and Figure 6.14 shows the deformations and strain levels in the soil
using 2D HSS model with the customized parameter input method.

2D HSS Model. Load group 1


350%
Normalized deformations [%]

293%
300%

250% 220%
200% 174%
143%
150% 125%
100%
100%
50%
0%
1

Measured deformation 2D MC (EI)


2D HSS Customized Parameters 2D HSS Plaxis Parameter Optimisation
2D HSS Correlation NC-Clay 2D HSS Correlation OC-Clay

Figure 6.11 2D HS S Results for load group 1

2D HSS Results. Load group 2


250%
212%
Normalized deformations [%]

200%
154% 157%
150%
100% 109%
98%
100%

50%

0%
1

Measured deformations 2D MC (EI)


2D HSS Customized Parameters 2D HSS Plaxis Parameter Optimisation
2D HSS Correlation NC-Clay 2D HSS Correlation OC-CLay

Figure 6.12 2D HS S Results for load group 2

53
Figure 6.13 Calculated deformations for load group 1 using 2D HS S model with the customized parameters method

Figure 6.14 S train level in the soil profile for load group 1 using 2D HS S model with the customized parameter
method

54
6.2.4 Comparison between 2D and 3D numerical calculations
Both the 2D and 3D numerical calculations overestimate the deformations in most of the
simulations. However, the results from the 2D and 3D numerical calculations are relatively
similar to each other, with the 3D giving slightly lower results for most of the simulations for
both load group 1 and 2, Figure 6.15and Figure 6.16 respectively.

Comparison between 2D and 3D Calculations


350%
293% 293%
Normalized deformations [%]

300%
250% 220%
195%
200% 174% 160%
143% 141%
150% 125% 124%
100%
100%
50%
0%
1

Measured deformations 2D MC
3D MC 2D HSS Customized Parameters
3D HSS Customized Parameters 2D HSS Plaxis Parameter Optimisation
3D HSS Plaxis Parameter Optimisation 2D HSS Correlation NC Clay
3D HSS Correlation NC Clay 2D HSS Correlation OC Clay
3D HSS OC Clay

Figure 6.15 2D vs 3D S imulations for load group 1

2D vs 3D Load group 2
250%
212% 210%
Normalized Deformations [%]

200%
154% 157% 157%
150% 122% 130%
100% 109% 101%
98%
100%

50%

0%
1

Measured deformation 2D MC 3D MC
2D HSS Customized parameters 3D HSS Customized parameters 2D HSS Parameter optimisation
3D HSS Parameter Optimisation 2D HSS Correlation NC-Clay 3D HSS Correlation NC-Clay
2D HSS Correlation OC-Clay 3D HSS Correlation OC-Clay

Figure 6.16 2D vs 3D S imulations for load group 2

55
6.3 Sensitivity analysis of the HSS model
The sensitivity analysis for the 2D HSS model shows that many of the studied parameters had
little effect on the calculated deformation. The stiffness moduli and the shear strength show
little influence on the deformations. A 15% decrease of the modulus gave rise to an
approximate deformation increase of 2%. The shear strength of the clay showed an even less
influence, the 15% change only influenced the deformations by approximately 0.5%.

The parameters showing the largest influence on the deformations were the lateral earth
pressure coefficient , the power law coefficient and the initial small strain stiffness
By decreasing by 15%, the deformations increased by approximately 26%. The parameter
also had a large influence on the deformations, by decreasing the reference value by 15%,
the deformations decreased by approximately 12%.

SEN SI TI VI TY A N A LYSIS R ESU LTS

130.00%

125.00%

120.00%

NORMALIZED DEFORMATIONS [%]


115.00%

110.00%

105.00%

100.00%

95.00%

90.00%

85.00%

80.00%
-20.00% -15.00% -10.00% -5.00% 0.00% 5.00% 10.00% 15.00% 20.00%
PARAMETER VARIATION [%]
E50ref Eoed m K0nc c´ref Psi γ0,7 G0 K0

Figure 6.17 Results from the sensitivity analysis

56
7 Analysis and discussion
7.1 Empirical correlation for the elasticity modulus
The comparison between performed triaxial tests and existing correlations give varying results
depending on the geographical location. Another problem with determining the stiffness
based on the correlations is that they do not take the strain level in the soil into account. The
stiffness, as mentioned earlier, is dependent on the strain level in the soil. The smaller the
strains, the higher the stiffness. To obtain realistic values of the elasticity modulus, triaxial
tests needs to be performed with minimal disturbances.

7.2 Plaxis parameter optimisation function


The Plaxis parameter optimisation function shows good similarities with the actual performed
triaxial test, see Appendix B - Figure 1 and Appendix B - Figure 2. However, by slightly
changing the parameter range in the optimisation function a totally different result can be
obtained. This means that great care must be taken when using the parameter optimisation
function. The Plaxis parameter optimisation function uses the Plaxis program SoilTest, which
simulates real life soil tests. The SoilTest program is based on a single point algorithm. Soil
behaviour is very complex and it is highly unlikely that a simple mathematical algorithm can
reproduce the actual behaviour of soil.

7.3 Calculated deformations


In geotechnical engineering many simplifications are often required due to the complexity of
soil. Soil is often simplified by assuming homogeneity and linear elasticity, while in reality
soil is not a homogenous material and also far from linear elastic. Due to the required
simplifications, a natural deviation between measured deformation and calculated is common.

Another important factor that needs to be discussed when it comes to deformations in clay is
time. Deformations increase over time, due to consolidation. In the three calculation models
the deformations were assumed to be occurring during the first seven day, before the concrete
had hardened. The concrete may very well harden after three days, which would result in
slightly lower deformations. The uncertainties in the time it takes for the concrete to harden
also gives rise to uncertainties in the calculated deformation.

7.3.1 Analytical vs measured deformations


The analytical calculation, which was the simplest of the three studied models, was based on
well-known formulas and easily determined parameters obtained from the commonly used
CRS-test. Despite its simplicity the analytical calculation produces a really good result
compared to the measured deformation, with a difference of approximately 5 mm.

57
7.3.2 MC-calculations
The Mohr Coulomb constitutive model is an idealization of the soil response and does not
take the stress nor the strain dependency into account. Due to the simplifications, the method
requires fewer input parameters. Most of the parameters that affect the deformations, such as
the modulus of elasticity and the shear strength of the soil can easily be measured in the field
or in a laboratory.

The results from the MC calculation show that the calculated deformations vary between
109 % to 500 % depending on the load group and choice of correlation for the elasticity
modulus. The best result is obtained when the elasticity modulus is evaluated from a triaxial
test and the worst result is when using the approximate correlation . The result
from the MC-model is heavily dependent on the choice of elasticity modulus. By evaluating
the triaxial test and choosing an elasticity modulus based on the predicted strain level in the
soil, the results match the measured deformations well.

7.3.3 HSS-Calculations
The more advanced HSS-model takes the stress and strain dependency of the soil into
account. Due to its complexity, the model requires more input parameters to fully capture the
true behaviour of soil.

The calculated deformations vary depending on how the input parameters were determined.
The best result was obtained when the Plaxis parameter optimisation function was used to
match a fictive triaxial test to the actual performed test. By using optimised parameters, the
results were almost identical to the measured deformations. The optimisation function is
however believed to be wrongly determining some parameters, such as the stiffness
degradation parameter . The parameter obtained from the optimisation function is
approximately a factor 100 times higher than the value obtained from the correlation. The
larger value means that the soil is a lot stiffer even at higher strains, which in turn gives rise to
smaller deformations, which is observed in the results.

The customized parameter method which was a mix between correlations, Plaxis parameter
optimisation and data based from triaxial test overestimated the deformations with
approximately 150-180 % compared to the measured deformations.

The HSS calculations based on strict empirics show varying results depending on if the clay is
considered to be normally consolidated or overconsolidated. The OCR influences the lateral
earth pressure coefficient , which has a large effect on the calculated deformations, Figure
6.17. Clay is considered to be normally consolidated if the and overconsolidated if
the . With an OCR of 1.37 the clay can be considered to be slightly
overconsolidated. By assuming slightly overconsolidated clay, the results from the strict
empirical correlation overestimates the deformations by a factor 2 for load group 1,Figure
6.11 and approximately a factor of 1.5 for load group 2, Figure 6.12.

Overall the three different methods of determining the input parameters studied in this thesis
have uncertainties. To decrease the uncertainties and thereby increasing the accuracy of the

58
calculated ground deformations more advanced ground investigation methods, such as cross
hole seismic tests or resonant column tests, needs to be conducted to asses the small-strain
stiffness properties.

7.3.4 2D vs 3D
The results obtained from the 2D and 3D numerical simulations are very similar. The small
difference that exists may come from different mesh sizes and the required load simplification
in the 2D geometry. The 2D numerical simulations use very a fine mesh while the 3D
simulations are using a coarse mesh. The main reason for the mesh difference is the
computation time, where both converged during the model testing. A 2D simulation with
very fine mesh took about 10-15 min for the simple geometry meanwhile the 3D simulation
with coarse mesh took approximately 6-7 hours. The 2D geometry in Plaxis is also faster to
set up than the 3D geometry.

7.4 Sensitivity analysis


Many parameters in the HSS model are intercorrelated, for example the stiffness degradation
parameter is affected by the small strain stiffness. By varying only one of the parameters
while keeping the other parameter constant the sensitivity analysis fails to show the true
importance of some parameters. Therefor it is important to keep in mind that the results from
the simplified sensitivity only gives an indication of the importance of some parameters.

The simplified sensitivity analysis shows that many of the parameters studied had a negligible
effect on the deformation. The elasticity modulus that is important in the MC model has
almost no effect on the deformations in the HSS model according to the analysis. The main
reason for this is believed to be that the elasticity modulus is correlated to the level of strains.
At lower strain levels, the elasticity modulus increases, therefor the starting input stiffness has
less of an affect compared to the MC model.

The three parameters that had the largest effect on the deformations were .

The power law coefficient , had a great impact on the deformations, however, determining
the coefficient is not easy. (Surarak et al., 2012) mentions that m should be set to 0.9-1.0 for
clay soils while (Benz, 2007) recommends a value between 0.7-1.0. The coefficient is also
dependent on the liquid limit and the plasticity index on the clay (Benz, 2007).

The lateral earth pressure coefficient also had a large influence on the deformations. A
small decrease of 5% gave rise to an 8% deformation increase. Therefor it is important to
determine accurately.

One of the more difficult and yet most important parameters to determine is the small-strain
stiffness . The parameter is dependent on a lot of variables and seldom measured for
practical design cases. Due to its dependency of many variables the span for the small-strain
stiffness is large. The value of in this thesis is assumed to be in the range of 15 000-50 000
kPa. The chosen value of used as a reference value in the sensitivity
analysis lies in the mid of the large span. Based on the obtained deformations the true value of

59
the small-strain stiffness may however be in the range of 40 000 – 45 000 kPa from back
analysis of the deformations.

7.5 Conclusions and recommendations for practical design


Soil is a complex material and calculating the predicted deformations is tricky due to the
many simplifications that need to be done during the process. There are many models for
calculating the predicted ground deformations when the soil is subjected to external loads.
The different models often require different parameters obtained from different types of field
or laboratory tests.

The result from this thesis show that despite its simplicity the analytical calculation based on
parameters obtained from CRS-test manages to predict the ground deformation with great
accuracy.

The numerical Mohr Coulomb constitutive model managed to predict the ground deformation
with greater accuracy than the analytical calculation, however the MC model is heavily
dependent on having a realistic value on and E. In order to determine the elasticity
modulus realistically, triaxial test must be performed and evaluated correctly. Triaxial tests
are however seldom performed in smaller projects due to the cost.

The Hardening Soil Small constitutive model, which simulates the soil behavior more
accurately than the simpler MC model, failed to give more accurate deformations than the MC
model. The main reason is believed to be the difficulty of determining its parameters.

In order to use the HSS-model with greater accuracy many of the parameters, such as and
needs to be measured by dynamic tests in a laboratory or out in the field.

The difference between 2D and 3D numerical calculations showed minor differences. The
difference that occurred may have been caused by different mesh qualities in Plaxis.
For geometries similar to those in the thesis, the 2D model is recommended due to its
simplicity and faster computation time. The 3D model can be used for complex problems
where the simplifications into 2D may have a significant effect on the result. However,
modeling with 3D often requires other types of simplifications.

For future practical design it is recommended to skip the most advanced model, hardening soil
small, due to the lack of quality input data. This thesis has shown that the analytical
calculation model works very well for calculating the predicted short-term ground
deformations. The method is fast and can be done by parameters obtained from routine- and
CRS-tests, which often are performed even in smaller projects, which makes it very important
in future practical design.

It is also recommended to perform some kind of numerical simulation as a complement to the


analytical calculation. The numerical simulation must however be based on quality input data
obtained from e.g. correctly performed and evaluated triaxial tests.

60
8 Bibliography
Atkinson, J. H. (2000). Non-linear soil stiffness in routine design. Géotechnique, 50(5), 487–
508. https://doi.org/10.1680/geot.2000.50.5.487

Benz, T. (2007). Small-Strain Stiffness of Soils and its Numerical Consequences. University of
Stuttgart.

Briaud, J.-L. (2001). Introduction to Soil Moduli. Geotechnical News, (June), 1–8.
https://doi.org/10.1017/CBO9781107415324.004

Brinkgreve, R. B. J., Engin, E., & Swolfs, W. M. (2015). Material Models Manual. Plaxis
2015, 202.

Burland, J. B. (1989). "Small is beautiful’ - the stiffness of soils at small strains. Canadian
Geotechnical Journal, 26(4), 499–516. Retrieved from
http://www.scopus.com/inward/record.url?eid=2-s2.0-
0024920802&partnerID=40&md5=13f119eae36027c6f9c81e3137f9f837

Clayton, C. R. I. R. I. (2011). Stiffness at small strain: research and practice. Géotechnique,


61(1), 5–37. https://doi.org/10.1680/geot.2011.61.1.5

Davis, R. O., & Selvadurai, A. P. S. (1996). Elasticity and Geomechanics. Retrieved from
https://books.google.com/books?id=4Z11rZaUn1UC&pgis=1
Duncan, J. M., & Bursey, A. (2013). Soil Modulus Correlations. Foundation Engineering in
the Face of Uncertainty, 321–336. https://doi.org/10.1061/9780784412763.026
GDS. (2013). What is triaxial testing? Part One: Introduction to Triaxial Testing, (Cd), 1–4.
Retrieved from http://www.gdsinstruments.com/__assets__/pagepdf/000037/part 1 of
3_.pdf

Hardin, B. O., & Richart, F. E. (1963). Elastic wave velocities in granular soils. Journal of
Soil Mechanics & Foundations Div, 89(SM1), 33–65.

IEG. (2011). Implementeringskommision för europastandarder inom geoteknik .

Jamiolkowski, M., O’Neill, D. a., Bellotti, R., & Presti, D. C. F. Lo. (1996). Anisotropy of
small strain stiffness in Ticino sand. Géotechnique, 46(1), 115–131.
https://doi.org/10.1680/geot.1996.46.1.115
Jardine, R. J., Potts, D. M., Fourie, a. B., & Burland, J. B. (1986). Studies of the influence of
non-linear stress–strain characteristics in soil–structure interaction. Géotechnique, 36(3),
377–396. https://doi.org/10.1680/geot.1986.36.3.377

Jardine, R. J., Symes, M. J., & Burland, J. B. (1984). Measurement of soil stiffness in the
triaxial apparatus. Geotechnique, 34(3), 323–340.
https://doi.org/10.1680/geot.1985.35.3.378
Kondner, R. L. (1963). Hyperbolic stress-strain response: cohesive soils. Journal of the Soil
Mechanics and Foundations Division, 89(1), 115–144. https://doi.org/10.1016/0022-
4898(64)90153-3

61
Kramer, S. L. (1996). Geotechnical Earthquake Engineering. Prentice-Hall, Inc., 6, 653.
https://doi.org/10.1007/ 978-3-540-35783-4

Lambe, T. W., & Whitman, R. V. (1979). Soil Mechanics, SI Version, 553.


https://doi.org/10.1017/CBO9781107415324.004

Larsson, R. (2008). Jords egenskaper, 62.


Mayne, P. W., Schneider, J. a, & Martin, G. K. (1999). Small-and large-strain soil properties
from seismic flat dilatometer tests. Pre-Failure Deformation Characteristics of
Geomaterials, Balkema, M. Jamiolkowski; R. Lancelotta: LoPresti (Eds).

Mitchell, J. K., & Soga, K. (2005). Funamentals of Soil Behaviour. Fundamentals of Soil
Behavior, 143–171.

Muir Wood, D., & Arroyo, M. (2004). Discussion: On the applicability of cross-anisotropic
elasticity to granular materials at very small strains. Géotechnique, 54(1), 75–76.
https://doi.org/10.1680/geot.2004.54.1.75

Piriyakul, K. (2006). Anisotropic stress-strain behaviour of Belgian Boom clay in the small
strain region. PhD Thesis, 183.

Santos, J. A., & Gomes Correia, A. (2001). Reference threshold shear strain of soil. Its
application to obtain an unique strain-dependent shear modulus curve for soil.
Proceedings of the 15th International Conference on Soil Mechanics and Geotechnical
Engineering, 1(1993), 267–270. Retrieved from
http://www.cabdirect.org/abstracts/20023124705.html

Schanz, T., Vermeer, a, & Bonnier, P. (1999). The hardening soil model: formulation and
verification. Beyond 2000 Comput. Geotech. 10 Years PLAXIS Int. Proc. Int. Symp.
beyond 2000 Comput. Geotech. Amsterdam Netherlands 1820 March 1999, 281.
Retrieved from
http://books.google.com/books?hl=en&lr=&id=ylNlhdPdB6cC&oi=fnd&
amp;pg=PA281&dq=The+hardening+soil+model+:+Formulation+and+verification
&ots=niCoPaWQxj&sig=1CXTV74dETqppmKFz0zWM-aoArI

Seed, H. B., & Idriss, I. M. (1969). Influence of soil conditions on ground motions during
earthquakes. American Society of Civil Engineers, Journal of the Soil Mechanics and
Foundations Division, 95(SM1), 99–137.

SGF. (2012). Triaxialförsök -.


Surarak, C. (2010). Geotechnical Aspects of the Bangkok MRT Blue Line Project,
(September), 400.

Surarak, C., Likitlersuang, S., Wanatowski, D., Balasubramaniam, A., Oh, E., & Guan, H.
(2012). Stiffness and strength parameters for hardening soil model of soft and stiff
Bangkok clays. Soils and Foundations, 52(4), 682–697.
https://doi.org/10.1016/j.sandf.2012.07.009

Terzaghi K, P. R. and M. G. (1996). Soil Mechanics in Engineering Practice.


Thiers, G. R., & Seed, H. B. (1968). Strength and Stress-Strain Characteristics of Clays
Subjected to Seismic Loading Conditions. ASTM Special Technical Publication 450,

62
Vibration Effects of Earthquakes on Soils and Foundations: A Symposium, 3–56.

Timoshenko, S. (1940). Srength of Materials.

Timoshenko, S. S. (1983). History of Strength of Materials: With a Brief Account of the


History of Theory of Elasticity and Theory of Structures. New York: Dover, 452.
Retrieved from
http://books.google.co.in/books/about/History_of_Strength_of_Materials.html?id=tkScQ
myhsb8C&pgis=1
Trafikverket. (n.d.). TR Geo 13 TDOK 20130668 v2, 0–103. https://doi.org/ISBN: 978-91-
7467-114-8
Trafikverket. (2014). Trafikverkets tekniska krav för geokonstruktioner TK Geo 13, 103.
https://doi.org/ISBN: 978-91-7467-114-8

Vermeer, P. a. (1978). A double hardening model for sand. Géotechnique, 28(4), 413–433.
https://doi.org/10.1680/geot.1978.28.4.413

Vermeer, P. A., & de Borst, R. (1984). Non-Associated Plasticity for Soils, Concrete and
Rock. Heron, 29(3). https://doi.org/10.1007/978-94-017-2653-5_10

Viggiani, G., & Atkinson, J. H. (1995). Stiffness of fine-grained soil at very small strains.
Géotechnique, 45(2), 249–265. https://doi.org/10.1680/geot.1995.45.2.249

Wood, T. (2016). On the Small Strain Stiffness of Some Scandinavian Soft Clays and Impact
on Deep Excavations.

63
Appendix A Soil Data
Appendix A presents data obtained from in-situ and laboratory tests performed between
supports 6-7.
Appendix A - Table 1 Results from the routine and CRS laboratory tests from borehole 15E03

Depth k
[ ] [ [ [ [ [ [ [ M´
[m] [m/s]
1.5 1.67 69 67 17 19 15.56 54 531 88 12.4 3.7E-
10
2.5 1.58 80 63 22 12 10.11 40 266 55 12.2 5.9E-
10
3.5 1.54 93 72 26 12 9.52 44 234 57 13.2 6.9E-
10
5.5 1.64 65 61 10 12 10.25

Appendix A - Table 2 Consolidation stage from triaxial test performed on samples from Borehole 15E03

Stage [kPa] [kPa]


1 6 6
2 34 21
3 25 19.5

Appendix A - Table 3 S hearing from triaxial test performed on samples from Borehole 15E03

Back pressure 300 kPa


Dry density 1.59 g/cm3
Liquid limit before test 74 %
Liquid limit after test 68 %
Consolidation strain 2.70 %
Consolidation strain 3.52 %
Shear strength 11.6 kPa
Axial strain failure 1.40 %
Axial failure stress 34 kPa
Radial failure stress 11 kPa

64
Appendix A - Figure 1 Consolidation coefficient cv evaluated from the CRS -test performed at 3.5 m depth

65
Appendix B Numerical input parameters
Appendix B presents the input parameters used for the numerical calculation models as well
as the result from matching the fictive triaxial test to the performed test to obtain input
parameters for the numerical calculations.

Appendix B - Table 1 Input parameters for the numerical Mohr Coulomb model

Varies kN/m2
0.33 -
1 kPa
30 °
0 °
0.78 -
m/day

Appendix B - Table 2 The different elasticity moduli that were studied in the numerical MC model

Method of obtaining the stiffness Elasticity modulus [kPa]


– Triaxial test 10470
– Triaxial test 8149

5053

4330

3790

3369

3032

2526

66
Appendix B - Table 3 Input parameters for the Fill material

20 kN/m3
23 kN/m3
50000 kN/m2
0.2265 -
0.1 kPa
45 °
15 °
0.2929 -

Appendix B - Table 4 Input parameters for the HS S calculations

Parameter Customized Parameter Correlations Correlations Units


parameters optimisation NC-Clay OC-Clay
10470 10230 5053 5053 kN/m2
5282 5427 4043 4043 kN/m2
31410 29780 31410 31410 kN/m2
0.9016 0.9057 1 1 -
0.2 0.2 0.2 0.2 -
0.5130 0.5062 0.5 0.5 -
0.78 0.78 0.632 0.751 -
100 100 100 100 kN/m2
28300 24320 28300 28300 kN/m2
0.04270 -
1 1.476 1 1 kN/m2
30 30 30 30 °
-3.829 -2.930 0 0 °
0.9 0.9 0.9 0.9 -
16.10 16.10 16.10 16.10 kN/m3
16.10 16.10 16.10 16.10 kN/m3
2.020 2.020 2.020 2.020 -

Where .

Parameters not mentioned above were set to standard values according to Plaxis.

67
Stress-Strain diagram
14

12

Triaxial test
10
Shear stress [kPa]

8 Plaxis parameter
optimisation

0
0 2 4 6 8 10 12 14
Axial strain [%]

Appendix B - Figure 1 Results from the Plaxis parameter optimisation compared with triaxial test from borehole
15E03

Stress-Path diagram
14

12

Triaxial test
10

8
τ [kPa]

6 Plaxis paramter
optimisation
4

0
0 5 10 15 20 25 30
s´ [kPa]

Appendix B - Figure 2 Results from the Plaxis parameter optimisation compared with triaxial test from borehole
15E03

68
Appendix B - Table 5 Parameter range for the automatically determined parameters in the customized method

Parameter Min Value Optimal value Max Value Units


- 10470 - kN/m2
4000 5282 10000 kN/m2
- 31410 - kN/m2
0.7 0.9016 1.0
- 0.2 -
0.5 0.513 0.62
- 100 - kN/m2
- 28300 - kN/m2
- - -
- 1 - kN/m2
- 30 -
-7 -3.829 -2 °
- 0.9 °
- 16.10 - kN/m3
- 16.10 - kN/m3

Appendix B - Table 6 Plaxis parameter optimisation parameter range

Parameter Min Value Optimal value Max Value Units


8000 10230 10470 kN/m2
4000 5427 10000 kN/m2
- 29780 - kN/m2
0.8 0.9057 1.0
- 0.2 -
0.5 0.5062 0.62
- 100 - kN/m2
10000 24320 30000 kN/m2
0.01 -

1 1.476 2 kN/m2
- 30 -
-7 -2,930 -3 °
- 0.9 °
- 16.10 - kN/m3
- 16.10 - kN/m3

69
Appendix C Performed triaxial tests
Appendix C presents data from different triaxial-, routine- and CRS-tests are here presented
for clay samples taken from Stockholm, Norrköping and Uppsala, Figure 5.3.

Stockholm, Ulvsunda. Borehole 15E01


16

14

12
Shear stress[kPa]

10

0
0 2 4 6 8 10 12 14

Axial strain [%]

Appendix C - Figure 1 Results from the triaxial test from borehole 15E01, Ulvsunda, S tockholm

14
Stockholm, Ulvsunda. Borehole 15E03
12

10
Shear stress[kPa]

0
0 2 4 6 8 10 12 14
Axial strain [%]

Appendix C - Figure 2 Results from the triaxial test from borehole 15E03, Ulvsunda, S tockholm

70
Stockholm, Ulvsunda. Borehole 15E09
14

12
Shear stress [kPa]

10

0
0 2 4 6 8 10 12 14
Axial Strain [kPa]

Appendix C - Figure 3 Results from the triaxial test from borehole 15E09, Ulvsunda, S tockholm

Norrköping, Loaction 1. Depth 5m


40

35

30
Shear Stress [kPa]

25

20

15

10

0
0 2 4 6 8 10 12 14 16 18 20
Axial Strain [%]

Appendix C - Figure 4 Results from triaxial test from Norrköping, Location 1 Clay. 5m Depth

71
Norrköping, Location 1. Depth 8m

40

35

30
Shear stress [kPa]

25

20

15

10

0
0 2 4 6 8 10 12 14 16 18 20
Axial Strain [%]

Appendix C - Figure 5 Results from triaxial test from Norrköping, Location 1 Clay. 8m Depth

Norrköping, Location 1. Depth 13 m


45

40

35
Shear stress [kPa]

30

25

20

15

10

0
0 2 4 6 8 10 12 14 16 18 20
Axial Strain [%]

Appendix C - Figure 6 Results from triaxial test from Norrköping, Location 1. 13m Depth

72
Appendix C - Table 1 Results from routine and CRS -test for Norrköping, Location 1 Clay

Depth
ρ [t/m3] [ [ [ [ [
[m]
4 1.61 72.2 82 15 40 29.9 -
5 1.56 76.4 83 14 31 23.1 97
6 1.59 83.9 93 14 33 23.3 98
8 1.57 78.3 82 18 32 23.9 110
9 1.56 81.9 86 18 33 24.2 128
13 1.55 80.4 84 15 28 20.7 116
14 1.56 78.9 83 16 29 21.5 136

Norrköping, Location 2. Depth 5m


25

20
Shear stress [kPa]

15

10

0
0 2 4 6 8 10 12 14 16 18 20
Axial Strain [%]

Appendix C - Figure 7 Results from triaxial test from Norrköping, Location 2. 5m Depth

73
Norrköping, Location 2. Depth 8m
30

25

20
Shear stress [kPa]

15

10

0
0 2 4 6 8 10 12 14 16 18 20
Axial Strain [%]

Appendix C - Figure 8 Results from triaxial test on Norrköping, Location 2. 8m Depth

Norrköping, Location 2. Depth 13m


30

25
Shear stress [kPa]

20

15

10

0
0 2 4 6 8 10 12 14 16 18 20
Axial Strain [%]

Appendix C - Figure 9 Results from triaxial test on Norrköping, Location 2. 13m Depth

74
Appendix C - Table 2 Results from routine and CRS -test for Norrköping, Location 1 Clay

Depth
ρ [t/m3] [ [ [ [ [
[m]
4 1.48 92 89 15 19 13.7 -
5 1.57 86.9 84 15 18 13.3 62
6 1.53 83.9 77 18 20 15.4 64
8 1.52 78.3 93 14 27 19.1 74
9 1.51 81.9 94 15 20 14.1 68
13 1.61 80.4 87 17 19 13.8 88
14 1.55 78.9 86 11 19 13.9 93

Uppsala. Depth 3m
25

20
Shear Stress [kPa]

15

10

0
0 2 4 6 8 10 12 14 16 18 20
Axial Strain [%]

Appendix C - Figure 10 Results from triaxial test from Uppsala 3m Depth

75
Appendix C- Table 3 Results from routine and CRS -test for Uppsala

Depth ρ (t/m3) [ [ τ [kPa] cu [kPa] [


[m]
3 1,6 64 70 12 26 20,88046 62
5 1,54 80 93 8,6 25 17,66784 67
9 1,53 79 93 9,3 37 26,1484 90
12 1,51 80 94 7,9 46 32,35274 103
14 1,56 75 88 7,8 44 31,87838 109
22 1,64 61 71 7,1 41 32,71737 144
28 1,66 66 74 7,7 31 24,28109 160
35 1,81 46 53 6,7 32 29,12639 185
42 1,83 48 56 7,1 40 35,51699 248
55 1,89 38 45 6,2 46 45,06849 294

76

You might also like