You are on page 1of 311

M ASTER ’ S T HESIS

M. S C . IN S TRUCTURAL & C IVIL E NGINEERING

Behaviour of Axially Loaded Bucket Foundations in Sand

Written by
Sorin Grecu

Supervised by
Professor Lars Bo Ibsen

8th of June 2018


School of Engineering and Science
Department of Civil Engineering
Thomas Manns Vej 23
9220 Aalborg Ø
www.civil.aau.dk

Document type:
Master’s thesis

Title:
Behaviour of Axially Loaded Bucket
Foundations in Sand

Project period:
September 2017 - June 2018

Author:
Sorin Grecu

Supervisor:
Lars Bo Ibsen

Main report, no. of pages: 96

Appendices, no. of pages: 202

Sorin Grecu Lars Bo Ibsen


Preface

This thesis reflects the research effort related to the final project within the master’s programme
titled M. Sc. in Structural and Civil Engineering at Aalborg University, Aalborg, Denmark. The
project period started in September 2017 and ended in June 2018. The document comprises the
main report and five appendices.

I would like to express my gratitude to Prof. Lars Bo Ibsen, who was my supervisor. His good
advice and guidance throughout the entire project period are kindly appreciated. I also deeply
acknowledge the support from my friends and especially from my family - their love and care made
me into who I am today.

Sorin Grecu

7th of June 2018

iii
Abstract

The offshore wind industry faces new challenges as cost-efficiency is a key issue for its further
development. Steps towards feasible offshore wind energy are taken by building wind farms farther
into the sea. This strategy, however, raises many economic and technical concerns. Wind turbine
foundations is one of them.

Suction bucket technology is an option that carries great potential in terms of reducing the
overall price of offshore wind turbine installation process. Multi-footed jacket structures mounted on
buckets represent a solution for supporting high capacity wind turbines in transitional water depths.
In this set-up the buckets are mainly subjected to axial loading. Knowledge about their performance
is scarce, so readily-available methods to predict their behaviour are yet to be established.

This study is concerned with the formulation of t − z curves for suction buckets installed in
sand using numerical modelling. Finite element analysis software PLAXIS 2D and PYTHON pro-
gramming language are used to create 100 axisymmetric models. Prescribed vertical displacements
in both directions are applied considering drained and partially drained conditions. The simulated
buckets have diameters from 10 to 20 m and an embedment ratio of 1. Cohesionless soil is assumed
as Frederikshavn sand and is defined with the Hardening Soil small strain model. Compaction
states ranging from loose to very dense are taken into account.

The shear stresses in skirt interfaces are plotted against bucket displacements and the resulting
curves are normalized with respect to peak shear stress and peak displacement. The formulation of
these two quantities is based on regression analysis of data related to each set numerical models that
comprises a given combination of loading direction and drainage condition. Based on normalized
curves and studies of fitting parameters, a mathematical model is created for each of the four sets.

Applying the Winkler method of idealizing soil reaction with springs, the t − z formulations are
employed to calculate the total frictional response of suction buckets. The spring properties are
dependent on vertical stress, bucket diameter, skirt length and friction angle, all of which are basic
parameters in geotechnical design.

The verification of mathematical models highlights some of their limitations and concludes on
the better performance of drained formulations. Suggestions for improvements of t − z expressions
and further research on the topic are presented.

v
Contents

Preface iii

Abstract v

1 Introduction 1

1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1
1.2 Problem formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Purpose of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Literature review 5

2.1 Behaviour of suction buckets under axial loading . . . . . . . . . . . . . . . . . . . 5


2.2 Influential factors on tensile capacity . . . . . . . . . . . . . . . . . . . . . . . . . . .7
2.2.1 Bucket geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7
2.2.2 Pull-out rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7
2.2.3 Soil permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Studies involving numerical modelling . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3.1 Model domain type and size . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3.2 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3.3 Material model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3.4 Interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10
2.3.5 Mesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10

3 Numerical modelling 11

3.1 Software . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
.
3.2 Material modelling and properties . . . . . . . . . . . . . . . . . . . . . . . . . . .12
3.2.1 Soil model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .12
3.2.2 Soil properties: Frederikshavn sand . . . . . . . . . . . . . . . . . . . . . .14
3.2.3 Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .20
3.3 Model domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
.
3.4 Mesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .28
3.5 Interfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .32
3.6 Loading and boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . .35

vii
3.7 Calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .37
.

4 Interpretation of results 39

4.1 Model overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .39


4.2 Data processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .40
4.3 Integration of stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .40
4.3.1 Integration along the skirt . . . . . . . . . . . . . . . . . . . . . . . . . . .40
4.3.2 Integration under the lid . . . . . . . . . . . . . . . . . . . . . . . . . . . .43
4.4 Drained models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .44
4.4.1 Tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .44
4.4.2 Compression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .49
4.5 Partially drained models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
.
4.5.1 Tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
.
4.5.2 Compression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .54

5 Normalization 57

5.1 Data trimming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .57


.
5.2 Bucket under axial tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .59
5.2.1 Drained models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .59
5.2.2 Partially drained models . . . . . . . . . . . . . . . . . . . . . . . . . . . .70
5.3 Bucket under axial compression . . . . . . . . . . . . . . . . . . . . . . . . . . . .73
5.3.1 Drained models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .73
5.3.2 Partially drained models . . . . . . . . . . . . . . . . . . . . . . . . . . . .75
5.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .76

6 Formulation of t-z curves 77

6.1 Drained tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .77


.
6.2 Partially drained tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .78
6.3 Drained compression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .82
6.4 Partially drained compression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .83

7 Verification and load-displacement curves 85

7.1 Comparison with numerical results from the current study . . . . . . . . . . . . . . .85
7.2 Verification with other numerical studies . . . . . . . . . . . . . . . . . . . . . . . .89
7.3 Verification with a physical model . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
.

8 Conclusion and future work 93

Bibliography 96
A HSsmall soil model 1

A.1 The original HS model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1


A.2 Input parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
A.3 Small-strain stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

B Aalborg sand no. 1 7

C Guidelines for PYTHON scripting 11

C.1 Remote scripting server . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11


.
C.2 SciTE editor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .12
C.3 Libraries and connection to PLAXIS application . . . . . . . . . . . . . . . . . . . .12
C.4 PLAXIS 2D Input . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .13
C.4.1 Soil menu . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .13
C.4.2 Structures menu . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .14
C.4.3 Mesh menu . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .17
.
C.4.4 Staged construction menu . . . . . . . . . . . . . . . . . . . . . . . . . . .17
.
C.5 PLAXIS 2D Output . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .19
C.6 Tips & suggestions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .23
C.6.1 Errors when setting material properties . . . . . . . . . . . . . . . . . . . .24
C.6.2 Errors when opening PLAXIS Output application . . . . . . . . . . . . . . .25
C.6.3 Finding object names with PLAXIS user interface . . . . . . . . . . . . . .25
C.6.4 Finding object attributes . . . . . . . . . . . . . . . . . . . . . . . . . . . .25
C.6.5 Saving data for every step . . . . . . . . . . . . . . . . . . . . . . . . . . .26
C.6.6 Fast data writing to files . . . . . . . . . . . . . . . . . . . . . . . . . . . .26
C.6.7 Node data filtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .28
C.6.8 Creating folders with PYTHON . . . . . . . . . . . . . . . . . . . . . . . .30

D Statistical analysis, trimmed and normalized t-z curves 31

D.1 Drained models (Tension) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31


.
D.1.1 Peak shear stress formulation . . . . . . . . . . . . . . . . . . . . . . . . . . 31
.
D.1.2 Peak displacement formulation . . . . . . . . . . . . . . . . . . . . . . . . .36
D.1.3 Trimmed and normalized t-z curves . . . . . . . . . . . . . . . . . . . . . .40
D.2 Partially drained models (Tension) . . . . . . . . . . . . . . . . . . . . . . . . . . .65
D.2.1 Peak shear stress formulation . . . . . . . . . . . . . . . . . . . . . . . . . .65
D.2.2 Peak displacement formulation . . . . . . . . . . . . . . . . . . . . . . . . .69
D.2.3 Trimmed and normalized t-z curves . . . . . . . . . . . . . . . . . . . . . .73
D.3 Drained models (Compression) . . . . . . . . . . . . . . . . . . . . . . . . . . . . .98
D.3.1 Trimmed and normalized t-z curves . . . . . . . . . . . . . . . . . . . . . 103 .
D.4 Partially drained models (Compression) . . . . . . . . . . . . . . . . . . . . . . . 128
.
D.4.1 Trimmed and normalized t-z curves . . . . . . . . . . . . . . . . . . . . . 133
.

E Mathematical models 153


E.1 Drained tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
.
E.2 Partially drained tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
. .
E.2.1 Investigation of post-peak behaviour . . . . . . . . . . . . . . . . . . . . . 167
. .
E.2.2 Peak shear stress and peak displacement . . . . . . . . . . . . . . . . . . . 169
.
E.2.3 Normalized curves vs t-z formulation . . . . . . . . . . . . . . . . . . . . 170
.
E.3 Drained compression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
.
E.4 Partially drained compression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
. .
C HAPTER 1

Introduction

1.1 Background

In the recent years offshore wind is gaining attention as a viable source of energy. Compared to
their on-land analogues, offshore wind turbines (OWT) have higher capacity and are capable of
providing a more reliable power supply due to faster and steadier winds in the open sea. With the
purpose of harvesting more energy and increase feasibility, the current tendency is to place the
OWTs gradually farther from the shore. At the same time, prototypes for OWTs with capacities
as high as 10 MW are projected for the near future. Issues spanning over many disciplines arise
accordingly, and among the most prominent ones are foundations.

Given a fast-paced market, cost efficiency of OWTs becomes an essential matter for achieving
or maintaining the competitive edge. Keeping in mind that offshore foundations account for ca.
30 to 50% of all installation expenses of OWTs, the optimization of geotechnical design plays a
vital role in reducing the overall costs. Deeper waters and larger wind turbines lead to challenges
regarding foundation design. Monopiles are the most popular solution at the moment as they are
preferred for shallow waters, but their performance and economic efficiency in deeper waters is a
questionable matter. All factors considered, the industry is searching for feasible alternatives.

Figure 1.1: Jacket structure mounted on suction buckets. Photo: Ørsted, formerly DONG Energy

One such alternative is the suction bucket, which has been widely used in the oil and gas

1
Chapter 1. Introduction

industry, but represents a promising concept for supporting offshore wind turbines. Suction buckets
bring advantages based on their speed of installation, low environmental impact and ease of
decommissioning, among others. The challenges, however, arise from the fact that wind turbines
are very light structures compared to oil/gas platforms. This means that the loading patterns on
the foundation are of a different nature, respectively significant research effort must be invested in
understanding the behaviour of suction buckets given the new circumstances.

Two approaches for using such foundation type were developed: monobucket and jacket
structure supported by three or four buckets (Figure 1.1). This thesis is placed in the context of the
latter.

1.2 Problem formulation

In an offshore environment the structure is subjected to transient loads from wind, water waves
and currents. The loading direction is predominantly horizontal and causes an overturning moment.
The stability of the system is ensured by axial resistance of suction buckets. This mechanism is
represented schematically in Figure 1.2.

Figure 1.2: The destabilizing overturning moment M is counteracted by tensile resistance RT and compress-
ive resistance RC of individual buckets.

2
1.3 Purpose of the thesis

The tensile resistance comes from skin friction along the skirt and from suction below the
lid, while the compressive one is based on bearing capacity of soil and the same skin friction. It
is critically important to be able to predict the bucket displacements accurately for achieving a
safe design. As mentioned before, suction buckets acting as foundations for wind turbines is a
relatively new concept. This means that standardized readily-available formulas and "best practice"
guidelines for geotechnical design do not exist yet. Naturally, a question follows: how should the
axial response of suction buckets subjected to vertical loading be modelled?

The research presented in this thesis aims to provide a partial answer to the question.

1.3 Purpose of the thesis

The main topic of study is friction between soil and structure when the bucket is loaded axially. To
model this phenomenon, a commonly used approach for pile foundations is the Winkler method.
Its principle lies in idealization of the system by representing the soil reaction through independent
springs, see Figure 1.3.

Figure 1.3: The soil continuum is discretized into springs. The sketch exemplifies a situation of pure axial
tension in drained conditions.

The behaviour of springs that model the soil reaction during vertical loading is described
with load transfer curves or t − z curves, where t - shear stress at the structure’s wall (in case of
suction buckets it is the skirt) and z - vertical displacement of the structure. The total frictional
response of soil can be estimated with load-displacement curves that are computed by integrating
the simultaneous effect of all springs.

Mathematical expressions for t − z curves have been widely used for pile design, with formulas
being provided in documents such as API (2011). Since suction buckets are different in that they
are non-slender structures, the applicability of aforementioned expressions is not proven.

The main purpose of the thesis is to provide a mathematical formulation of t − z curves for
suction buckets installed in sand using finite element modelling. The study aims for a formulation
defined in terms of some readily-available basic parameters, so that it would allow for quick
predictions of foundation behaviour.

3
C HAPTER 2

Literature review

This chapter comprises the outcome of the literature review performed as the first step towards
achieving the goals of the thesis. The review of relevant recent work is essential for identifying
and understanding the state-of-the-art. Thus it is ensured that the current research effort is not
redundant and has the potential to provide meaningful results.

2.1 Behaviour of suction buckets under axial loading

The loading patterns on the suction buckets differ significantly depending on the chosen foundation
concept. Thieken et al. (2014) and Barari & Ibsen (2012) explained that single bucket foundations
are mainly affected by horizontal forces and bending moments, while buckets in multi-footed
structures are dominated by vertical forces. In the latter case, the axial loading comes as a
consequence of the "push-pull" mechanism that resists the overturning moment. Considering the
major difference in how these two systems act, Houlsby et al. (2005) stressed its importance for
foundation design. In this literature review, focus is placed upon the vertical response of buckets
subjected to pure axial loads.

Vaitkunaite (2016) recognized that a bucket foundation under compressive axial loading behaves
like a gravity based foundation, provided that a soil plug is formed. Respectively, given the
extensive experience and well-established methodologies in relation to gravity based foundations,
the assessment of the behaviour of suction bucket foundations under such type of loading should
not pose considerable challenges in terms of design.

The main issue arises arises in connection with suction buckets subjected to tension. Achmus
& Gütz (2016) pointed out the scarcity of research and the lack of knowledge on this topic, even
though the tensile loading aspect is crucial for the design of multi-footed foundations. This means
that there is no readily-available fast and accurate method to quantify the tensile response of such a
system.

Vaitkunaite (2016) made an analogy between the suction bucket system and a syringe. When
the bucket is pulled upwards, a gap tends to form between the lid and the soil surface. The pressure
beneath the lid drops and thus triggers a water flow through the soil body from the surrounding
volume towards the gap. This mechanism is represented schematically in Figure 2.1. It is clear that
drainage conditions have a considerable effect on pull-out resistance.

Senders (2009) categorized the drainage conditions for suction buckets in cohesionless soils in

5
Chapter 2. Literature review

Figure 2.1: Water flow during tensile loading.

two scenarios: drained and undrained, see Figure 2.2. In the first one, the drained pull-out capacity
relies on structure’s self-weight and friction between the skirt and soil. At the same time pore water
fills the gap fast enough to prevent any significant development of pressure difference between
the inside and the outside of the bucket. Failure takes place close to the skirt surface. Senders
(2009) termed this behaviour "frictional". It has been observed that the drained tensile capacity is
considerably smaller than the compressive capacity, Vaitkunaite (2016) stating that in dense sands
the tensile capacity is only approximately 1 % of the latter.

Figure 2.2: Drained (left) and undrained (right) behaviour of cohesionless soil during uplift of suction
buckets. [Senders (2009)]
.

On the other hand, the undrained behaviour is characterized by the occurrence of suction
pressure beneath the lid, which prevents the formation of a gap and thus the soil plug inside the
bucket is uplifted together with the structure, as explained in Thieken et al. (2014). Consequently,
the surrounding soil’s shear resistance is employed. This mechanism is denoted as "reverse end
bearing" in Senders (2009).

In reality, however, the foundation response lies in between these two idealized scenarios. Such

6
2.2 Influential factors on tensile capacity

behaviour is termed "partially drained" or "intermediate". In the presence of seepage, the upward
gradient causes stress relief inside the bucket, which reduces the inner skin friction. On the other
side, an opposite effect takes place: the downward flow towards the skirt toe increases the normal
stresses, leading to larger shear stress at the skirt surface. These effects were documented by
Senders (2009) and noticed in their numerical simulations by Thieken et al. (2014) and Shen et al.
(2017).

2.2 Influential factors on tensile capacity

It was stated in Thieken et al. (2014) and Vaitkunaite (2016) that the tensile capacity is a function
of three main factors: bucket geometry, pull-out rate and soil permeability. The effect of each factor
is discussed in the following sections.

2.2.1 Bucket geometry

By conducting a series of numerical simulations on suction buckets in sand under tensile loading,
Thieken et al. (2014) concluded that the tensile capacity for all drainage conditions increases with
the skirt length, see Figure 2.3. A longer skirt means an increased embedded depth and a larger skirt
area, hence more friction to resist the pull in case of drained behaviour. A larger affected soil mass
increases the undrained capacity, while longer seepage paths directly influence the partly drained
behaviour.

Figure 2.3: Influence of skirt length (left) and bucket diameter (right) on tensile capacity depending on the
pull-out rate [Thieken et al. (2014)].

In the same article, it was observed that the increase of bucket diameter has a major effect
primarily during undrained behaviour. Considerable suction force can be obtained even from
a relatively small increment in diameter, since it is estimated as the product between pressure
differential and inner lid area.

2.2.2 Pull-out rate

There is consistent evidence that the tensile capacity of suction buckets is greatly influenced by the
pull-out rate. A high displacement rate generally leads to undrained behaviour, as water does not
have enough time to flow into the space between the lid and soil surface. The exact pull-out rate at

7
Chapter 2. Literature review

which the behaviour becomes undrained depends on the soil and fluid properties, as well as bucket
dimensions, as explained in Sørensen et al. (2016).

The fact that the tensile resistance is generally higher during undrained loading than during
drained loading is observed and described on basis of numerical simulations (e.g. Thieken et
al. (2014), Achmus & Gütz (2016), Shen et al. (2017)) and laboratory tests (e.g. Feld (2001),
Vaitkunaite (2016)). However, the mobilization of stabilizing suction pressure may require large
upward displacements, which would breach the serviceability limit state, therefore this aspect must
be treated with caution in the design, according to Thieken et al. (2014). It was also acknowledged
that the undrained tensile capacity is limited by cavitation pressure or undrained shear resistance of
the soil, whichever is reached first. The cavitation pressure is estimated as the sum of atmospheric
and hydrostatic pressures, therefore the maximum possible suction depends on the water depth and
water density.

2.2.3 Soil permeability

Thieken et al. (2014) observed that the soil permeability does not influence the tensile capacity
during drained and undrained behaviour. The hydraulic conductivity is mostly relevant for partially
drained behaviour. An increase tensile resistance occurs with a decrease of soil permeability, as
shown in Figure 2.4.

Figure 2.4: Influence of soil permeability on tensile capacity depending on the pull-out rate. [Thieken et al.
(2014)]
.

2.3 Studies involving numerical modelling

Since numerical simulations are to be used as one of the primary tools for assessment of the bucket
foundation response, it is essential to learn how these were implemented in previous works. The
following subsections encompass the most important details that were extracted after the review of
relevant numerical studies.

8
2.3 Studies involving numerical modelling

2.3.1 Model domain type and size

When the tensile capacity was investigated using Finite Element Modelling (FEM), Thieken et al.
(2014), Achmus & Gütz (2016), Sørensen et al. (2016) and Shen et al. (2017) took advantage of
the symmetry and thus defined an axisymmetric domain. In this way the amount of elements was
reduced and the model was simplified, and as a result, the computation time was reduced. In order
to avoid the effects that the boundaries may inflict on the results, appropriate domain size was
defined. In the list below, the dimensions are given as relative to the bucket geometry, where D is
the diameter and L is the skirt length.

• Thieken et al. (2014): 3D and 3L


• Achmus & Gütz (2016): 4D and 3L
• Sørensen et al. (2016): 5D and 5L
• Shen et al. (2017): 2.5D and 5L

2.3.2 Boundary conditions

In the studies mentioned in the previous sections, small divergences are observed in terms of how
the boundary conditions were treated.

The displacement boundary conditions were defined in the same way in Thieken et al. (2014)
and Shen et al. (2017). The radial displacements were restricted along the axis of symmetry and
along the outer edge of the domain. The displacements of the nodes at the bottom were restricted in
both horizontal and vertical displacements. Sørensen et al. (2016) did not fix the bottom nodes in
the horizontal direction.

Thieken et al. (2014) and Sørensen et al. (2016) defined the axis of symmetry as an impermeable
boundary, while the initial pore pressures were applied to the rest of the model edges, which enabled
a free pore fluid flow through the respective boundaries. However, Achmus & Gütz (2016) defined
only the top surface of the soil domain as permeable. Restricting the flow through the rest of
the boundaries proved to have negligible effects on the results, as was revealed by a comparative
calculation.

2.3.3 Material model

Sørensen et al. (2016) used the classical Mohr-Coulomb model in order to describe the soil
behaviour. This proved to be enough to accomplish the goals of the comparative study of various
numerical formulations, as the post-yield zone was not of interest.

Østergaard et al. (2015), who developed p − y curves for bucket foundations in sand, followed
in the same manner by Vethanayagam & Ibsen (2017), implemented the Hardening Soil Small
Strain material model (HSsmall), in order to account for the stress-dependent stiffness of the soil
and to include additional stiffness at very small strains. Østergaard et al. (2015) argued that by
using just the Hardening Soil model (HS) there is a risk of overestimating the deformations, thus
the real soil behaviour at small strains might not be represented accurately.

9
Chapter 2. Literature review

Thieken et al. (2014), Achmus & Gütz (2016) and Sørensen et al. (2016) modelled the bucket
structure as a linear elastic body, attributing typical steel properties to it. Østergaard et al. (2015)
and Vethanayagam & Ibsen (2017) introduced an unrealistically high Young’s modulus for steel
and defined the skirt thickness as ca. 10 times larger than a typical value of about 25 mm. This was
done so the bucket behaves like a rigid body, since only the soil response was of interest.

2.3.4 Interface

Wolf et al. (2013) presented a method of formulation of p − y curves for non-slender monopiles in
cohesionless soil based on numerical analysis with the software PLAXIS 3D. The plate elements
used to model the bucket itself were assigned interface elements, so that the soil-structure interaction
could be modelled properly. Subsequently, after extracting the stresses from the interface and soil
elements, a comparison between the two sets of results revealed a slightly larger subgrade reaction
(p) along the interface. This was not unexpected, as the reaction determined from the soil elements
was based on stress points located in vicinity of the skirt, in contrast with the interface elements
which inherently contain stress points that are closest to the structure.

Brinkgreve et al. (2017) recommended to pay special attention to abrupt changes in boundary
conditions and thus, make use of interfaces in such a way as to prevent non-physical representations
of stresses. Respectively, Wolf et al. (2013) included additional interface elements below the pile
toe as an extension of the plate elements.

2.3.5 Mesh

Wolf et al. (2013) emphasized the importance of refining the mesh around the pile, in order to avoid
non-physical stress concentrations and to obtain a larger amount of stress points in the interface,
thus leading to representative p − y curves.

Østergaard et al. (2015) and Vethanayagam & Ibsen (2017) applied the methodology given in
Wolf et al. (2013) to develop and analyze FE models for deriving p − y curves for suction buckets.
It is considered that that similar ideas, but with the introduction of appropriate modifications, can
be applied to formulate t − z curves.

Note on t-z formulation method

The most significant source of inspiration for reaching the goals of this study is Østergaard et al.
(2015). An extensive amount of information from this paper is used in a direct manner in the current
thesis, therefore a comprehensive list is not given here. Instead references are made throughout the
report whenever deemed necessary. At a conceptual level, the current research and the mentioned
paper may be regarded as parallels.

10
C HAPTER 3

Numerical modelling

This chapter presents how finite element analysis is used to achieve the project goals. Numer-
ical simulations imply a series of critical decisions and choices that are thoroughly explained
in the current chapter. A detailed overview of the procedures and model properties is given.

3.1 Software

The choice of FEM software is based on understanding the physical system and on the desired
outcome. It is clear from the literature review that behaviour of suction buckets subjected to
axial loads can be correctly assessed by taking into account deformations and groundwater flow
simultaneously. The geotechnical software PLAXIS 2017 offers the possibility to model such
behaviour through a so-called fully coupled flow-deformation analysis. This calculation type is
deemed suitable for investigating the problem at hand, following the description given in Brinkgreve
et al. (2017).

A circular bucket foundation installed in homogeneous soil and loaded vertically forms a
system that exhibits cylindrical symmetry, as shown in Figure 3.1. This property can be used
advantageously to reduce the computation time at no cost, by modelling only a fraction of the system
while obtaining information about its entirety. The scope of the three-dimensional problem can be
fully captured by only two dimensions with an axisymmetric model. Therefore it is established that
the FE analysis can be carried out with PLAXIS 2D.

In this project, numerical analysis involves the creation of a relatively large amount of models
(in the order of hundreds). For example, studies of convergence or sensitivity of models require
change of certain input parameters and examination of outcomes, which means that the user faces a
time-consuming repetitive task. By recognizing the iterative nature of such processes, it becomes
obvious that automation is a solution that saves time and increases efficiency of the analysis.

Such a solution is offered in recent versions of PLAXIS 2D and PLAXIS 3D via a wrapper
based on PYTHON programming language. The PYTHON wrapper gives the user advantages
in terms of control of PLAXIS software. No interaction (except for visual inspection) with the
PLAXIS user interface is required, as all the commands can be directly defined by scripting. A
comprehensive set of guidelines for PYTHON scripting is given in Appendix C.

11
Chapter 3. Numerical modelling

Figure 3.1: Axisymmetry of the system.

3.2 Material modelling and properties

The soil-structure interaction is a phenomenon that requires modelling with an appropriate degree
of sophistication. The problem of formulating t − z curves becomes especially complex due to large
deformations, excess pore pressures, changes in hydraulic boundary conditions and non-linear soil
behaviour. The literature review sheds light on which model would be the most suitable. Namely,
Østergaard et al. (2015) serves as the main source of inspiration in this matter.

3.2.1 Soil model

The interest is to observe the soil response both before and after failure. The Hardening Soil Small
Strain model (HSsmall) is able to capture accurately the full behaviour range. This model is chosen
because it bears significant advantages in terms of reproducing real soil behaviour:

• Stiffness dependency on stress level is taken into account.


• The yield surface can expand, thus isotropic hardening is included. See the yield surface
plotted in principal stress space in Figure 3.2. Shear and compression hardening are modelled.
• The model incorporates the realistic hyperbolic stress-strain relationship.
• At very small strains, the material keeps almost all of its initial stiffness (Figure 3.3). With
continued straining the stiffness diminishes non-linearly.

A detailed description of the HSsmall model and its associated parameters is given in Ap-
pendix A. The current section further offers only a general overview on the input parameters.

12
MATERIAL MODELS MANUAL

3.2 Material modelling and properties

σ1

THE HARDENING SOIL MODEL WITH SMALL-STRAIN STIFFNESS (HSSMALL)

7 THE HARDENING SOIL MODEL WITH SMALL-STRAIN STIFFNESS (HSSMALL)

The original Hardening Soil model assumes elastic material behaviour during unloading
and reloading. However, the strain range in which soils can be considered truly elastic,
i.e. where they recover from applied straining almost completely, is very small. With
increasing strain amplitude, soil stiffness decays nonlinearly. Plotting soil stiffness
against log(strain) yields characteristic S-shaped stiffness reduction curves. σ3 Figure 7.1
gives an example of such a stiffness reduction curve. It outlines also the characteristic
shear strains that σ2can be measured near geotechnical structures and the applicable strain
ranges of laboratory tests. It turns out that at the minimum strain which can be reliably
Figure 6.10 Representation
Figure 3.2: Yield surfaceof total yield contour
oflaboratory
the hardening soili.e.
modelofinthe Hardening
principal stress Soil[Brinkgreve
space model in etprincipal stress
al. (2017)].
measured in classical tests, triaxial tests and oedometer tests without
space for cohesionless soil
special instrumentation, soil stiffness is often decreased to less than half its initial value.

shear yield locus can expand up to the ultimate Mohr-Coulomb failure surface. The cap
yield surface expands
1 as a function of the pre-consolidation stress pp .
Shear modulus G/G0 [-]

6.6 STATE PARAMETERS


Very IN THE HARDENING SOIL MODEL
small
In addition to the output
standard of
strains stress and strain quantities, the Hardening Soil
Small strains

model provides output (when being used) on state variables such as the hardening
Larger strains
p
parameter γ and the isotropic
0
-6
pre-consolidation
-5 -4 -3
stress -2pp . These
-1
parameters
Shear strain gS [-] can be
visualised by selecting the1e State parameters
1e 1e
option
1e
from1e
the stresses
1e
menu. An overview
of available state parameters is given below:
Figure 3.3: Characteristic stiffness-strain behaviour of soil [Brinkgreve et al. (2017)].
peq : Equivalent isotropic stress [kN/m2 ]
Figure 7.1 Characteristic stiffness-strain s
behaviour of soil with typical strain ranges for laboratory
The HSsmall parameters
tests mentioned
and structures (after Atkinson
eq
e&2 Sallforslist
in the q
following are the ones which are directly specified
(1991))
2
p =
by the user. The others are kept at their default values
2
+ (p ' )
determined by PLAXIS and are not listed.
The soil stiffness that should be used in M the analysis of geotechnical structures is not the
pp one that relates to the strain
: Isotropic range at the endstress
preconsolidation of construction according to Figure 7.1. [kN/m2 ]
ϕ Instead,
0 effective
very friction anglesoil stiffness and its non-linear dependency on strain amplitude
small-strain
OCR
c0 should be: properly
effective Isotropic over-consolidation
taken
cohesion ratio
into account. In addition OCR
to (all = pofp /p
features theeq )
Hardening [-]
Soil
ψ model, the Hardening
dilatancy angle Soil model with small-strain stiffness offers the possibility to do so.
γp re f : Hardening parameter (equivalent mobilised plastic [-]
E50 The reference
Hardeningsecant modulus
Soil model with small-strain stiffness implemented in PLAXIS is based on
rethe
f Hardening shear strain)
Soil model and usesmodulus
almost entirely the same parameters (see Section
Eoed reference tangent (oedometer)
re6.4). In fact, only two additional parameters
f
Eur Eur reference : Current modulus are
stress-dependent
unloading/reloading needed
elastic to describe
Young's the variation of [-]
modulus
stiffness with strain:
m power for stress-level dependency of stiffness
c re•f : Current depth-dependent cohesion [-]
G0 the initial
reference or very
shear small-strain
modulus shear
at very small strains G0
modulus
γ0.7• threshold
the shearshear
strainstrain
level γ0.7 at which the secant shear modulus Gs is reduced to about
70% of G0
6.7 ON THE USE OF THE HARDENING SOIL MODEL IN DYNAMIC CALCULATIONS
13
When using the Hardening Soil model in dynamic calculations, the elastic stiffness
ref
parameter Eur needs to be selected such that the model correctly predicts wave
velocities in the soil. This generally requires an even larger small strain stiffness rather
Chapter 3. Numerical modelling

3.2.2 Soil properties: Frederikshavn sand

Two types of sands are modelled: Frederikshavn sand and Aalborg sand no. 1. In the current
chapter only Frederikshavn sand is discussed, as this soil type is used for t − z curves formulation
for suction buckets installed in marine sand. The properties of Aalborg sand no. 1 are presented in
Appendix B.

Where laboratory test data is not available, the parameters are estimated with formulas based
on general experience or theories. Following the method used by Østergaard et al. (2015), most of
the properties of Frederikshavn sand are derived from the angle of internal friction ϕ.

As the first step, the relation between ϕ and relative density ID is found in Bolton (1986):

ϕ = ϕcrit + 3◦ IR − 3◦ ID − ∆ϕ1 (3.1)


 
p0
IR = ID Qmin − ln −R (3.2)
1 kPa

where:
ϕcrit critical friction angle (=33◦ )
∆ϕ1 correction for silt content (2◦ for 5-10 % silt)
IR relative dilatancy index
Qmin soil mineralogy and compressibility coefficient (=10 for quartz)
R fitting coefficient (=1)
p0 reference effective mean stress (=100 kPa)

Eq. (3.1) and Eq. (3.2) are combined and the terms are rearranged to solve for ID :

ϕ − ϕcrit + 3R + ∆ϕ1
ID =   (3.3)
p0
3 Qmin − ln 1 kPa −3

Thus by choosing any ϕ, it is possible to estimate ID with Eq. (3.3). The definition of relative
density ID is:

emax − e
ID = (3.4)
emax − emin

With knowledge of maximum void ratio emax and minimum void ratio emin , the in-situ void
ratio e is calculated with Eq. (3.4). Frederikshavn sand has emax = 1.05 and emin = 0.64, according
to Nielsen et al. (2012).

The unit weight of sand depends on void ratio e, specific gravity of solid particles G and
unit weight of water γw . The grain density of Frederikshavn sand and the water density are

14
3.2 Material modelling and properties

assumed ρs = 2.64 g/cm3 and ρw = 1 g/cm3 respectively. Thus the specific gravity G, which is a
dimensionless quantity, is calculated:

ρs
G= = 2.64 (3.5)
ρw

For the sake of consistency with the assumed water density ρw and following the common
geotechnical practice, the unit weight of water is taken as γw = 10 kN/m3 . The saturated and dry
unit weights of sand are estimated as follows:

(G + e)γw
γsat = (3.6)
1+e
Gγw
γdry = (3.7)
1+e

The oedometer modulus Eoed is estimated in accordance with Janbu’s tangent modulus concept,
presented in DNV (1992):

 1−a
σ0
Eoed = mσa (3.8)
σa

where:

m modulus number
σa atmospheric pressure (=100 kPa)
σ0 current stress level [kPa]
a stress exponent

Since sand falls into the category of soils exhibiting elasto-plastic behaviour, the stress exponent
a = 0.5. This reduces Eq. (3.8) to:

p
Eoed = m σ 0 σa (3.9)

In PLAXIS, the Hardening Soil Small Strain model conveniently requires the stiffness moduli
to be introduced in terms of their reference values, therefore the reference stress level σ 0 is defined
as 100 kPa. At this stage, the only missing component is the dimensionless modulus number m,
which should not be confounded with the power for stress-level dependency of stiffness in HSsmall
model, also denoted by m.

DNV (1992) gives typical ranges of m for Norwegian inorganic sands depending on the
compaction state. The ranges are considered valid for Frederikshavn sand. Similarly to Østergaard

15
Chapter 3. Numerical modelling

400
DNV data
Quadratic fit
350

300

250

200

150

100

50
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 3.4: Modulus number vs relative density, acc. to DNV (1992).

et al. (2015), the modulus number m is formulated as a function of ID by fitting the data provided in
DNV (1992) with a second degree polynomial, as shown in Figure 3.4.

The established relation has the following form:

m = 161.42 ID2 + 199.87 ID + 36.88 (3.10)

By inserting Eq. (3.10) into Eq. (3.9), one obtains the oedometer modulus Eoed as a function of
relative density ID for Frederikshavn sand at the reference stress σ 0 = 100 kPa. To denote the link
to this stress, the superscript "ref" is added. The outcome of Eq. (3.11) is used as one of the input
parameters in numerical models.

re f
Eoed = 16142 ID2 + 19987 ID + 3688 [kPa] (3.11)

DNV (1992) proposes a relation between Poisson’s ratio ν and friction angle ϕ:

1 − sin ϕ
ν= (3.12)
2 − sin ϕ

re f re f
In the context of elasticity, the reference secant modulus E50 is linked to ν and Eoed through
Eq. (3.13).

16
3.2 Material modelling and properties

re f 1 − ν − 2ν 2 re f
E50 = Eoed (3.13)
1−ν

It is acknowledged that by using Eq. (3.13) a very rough assumption is taken. Schanz et al.
(1999) points out that the Hardening Soil model, being elasto-plastic, does not imply a fixed relation
between E50 and Eoed , in contrast to elasticity theory. This allows for the two moduli to be inputted
separately.
re f
The reference unloading/reloading modulus Eur is found according to Brinkgreve et al. (2017).
Eq. (3.14) is the default relation used by PLAXIS.

re f re f
Eur = 3E50 (3.14)

In the absence of relevant advanced laboratory data, the power for stress-level dependency of
stiffness is assumed m = 0.5, which is typical for hard soils, as stated by Schanz et al. (1999).

The HSsmall model involves two specific parameters: Gre f


0 and γ0.7 . The reference shear
modulus at very small strains Gre f
0 is estimated with Eq. (3.15), which was formulated by Hardin &
Black (1969) and recommended by Brinkgreve et al. (2017) as a good approximation.

(2.97 − e)2
Gre f
0 = 33 [MPa] f or pre f = 100 [kPa] (3.15)
1+e

The threshold shear strain γ0.7 is calculated as follows:

1  0 
γ0.7 = 2c (1 + cos (2ϕ)) − σ10 (1 + K0 ) sin (2ϕ) (3.16)
9Gre
0
f

where:

c0 effective cohesion
σ10 effective vertical stress
K0 earth pressure coefficient at rest

To avoid complications, such as numerical instabilities, Brinkgreve et al. (2017) suggests to set
at least c0 = 0.2 kPa. Using a considerably higher value may attribute unrealistic tensile strength to
sands. The effective cohesion is set to 0.1 kPa and K0 = 1 − sin ϕ.

The hydraulic properties of Frederikshavn sand are described in Sjelmo (2012). A series of
tests was carried out using the falling head procedure in the soil laboratory at Aalborg University.

17
Chapter 3. Numerical modelling

The results are extracted in terms of permeability K, which is a quantity that depends only on
the soil skeleton and has units of area. On the other hand, the flow properties of soil may be
described by hydraulic conductivity k, which incorporates fluid properties, see Eq. (3.17). The flow
characteristics are defined through hydraulic conductivity in PLAXIS.

γw
k=K (3.17)
µ

where:

γw unit weight of water


µ dynamic viscosity of water

The influence of fluid temperature is introduced by viscosity µ. The average water temperature
used for testing was about 20 ◦ C, which is not characteristic for water at the sea bottom in the
basin surrounding Denmark. A more realistic value is assumed 9 ◦ C. This water temperature is
set in PLAXIS models. The dependency of water viscosity on temperature is estimated with an
empirically derived exponential function as shown by Eq. (3.18).

µ = 2.414 · 10−8 · 10247.8/(T +133.15) (3.18)

In Eq. (3.18) the temperature T is inserted as ◦ C and the calculated viscosity µ has units of
kN·s/m2 . Figure 3.5 shows how viscosity decreases with temperature. The significant drop within
the marked range justifies the need to account for the temperature effect.

10-6
1.8

1.6

1.4

1.2

0.8

0.6

0.4

0.2

0
0 10 20 30 40 50 60 70 80 90 100

Figure 3.5: Dynamic viscosity of water against temperature. The red lines correspond to 9 and 20 ◦ C.

18
3.2 Material modelling and properties

Sjelmo (2012) present a list of permeability values and the corresponding void ratios at which
they were obtained. After converting permeability K to hydraulic conductivity k and adjusting the
results for 9 ◦ C, the data is fitted with a quadratic function, as shown in Figure 3.6.

10-5
12
Test data
11 Quadratic fit

10

4
0.6 0.62 0.64 0.66 0.68 0.7 0.72 0.74 0.76 0.78 0.8

Figure 3.6: Hydraulic conductivity of Frederikshavn sand as function of void ratio, calibrated for 9 ◦ C. Test
data from Sjelmo (2012).

The established relation is presented by Eq. (3.19).

k = 0.3854 · 10−3 e2 − 0.2006 · 10−3 e + 0.304 · 10−4 (3.19)

During soil deformation, the permeability may change as a result of change in void ratio ∆e.
Brinkgreve et al. (2017) states that this phenomenon is modelled with Eq. (3.20) in PLAXIS.

 
k ∆e
log10 = (3.20)
k0 ck

where k0 is the initial hydraulic conductivity and ck is a parameter that controls the amount of
change in permeability. For HSsmall model, ck adopts values in the order of the compression index
Cc . The compression index is estimated with Eq. (3.21):

2.3(1 + ei )pre f
Cc = re f
(3.21)
Eoed

19
Chapter 3. Numerical modelling

where ei is the initial void ratio and the reference vertical stress pre f = 100 kPa.

The soil-structure friction angle or the interface angle δ is approximated roughly as 2/3 of the
friction angle ϕ, based on indicative values provided in DNV (1992). In PLAXIS, δ is expressed
through the interface strength Rinter .

tan δ
Rinter = (3.22)
tan ϕ

The dilatancy angle ψ is estimated according to Bolton (1986):

ψ = ϕ − ϕcrit (3.23)

In total, 5 sets of Frederikshavn sand geotechnical parameters are used to model the soil. Each
set corresponds to a friction angle chosen from a range of values typical for marine sand in loose,
medium and dense state. Table 3.1 gives an overview on all parameters applied in numerical
modelling.

ϕ [◦ ] 30 33 35 38 40
ID [%] 15.2 37.9 53.1 75.8 91.0
e [-] 0.99 0.89 0.83 0.74 0.68
γsat [kN/m3 ] 18.3 18.7 19.0 19.4 19.8
γdry [kN/m3 ] 13.3 13.9 14.4 15.2 15.7
re f
E50 [kPa] 4727 9717 14044 22129 28622
re f
Eoed [kPa] 7091 13589 18849 28133 35251
re f
Eur [kPa] 14182 29150 42131 66387 85867
ν [-] 0.33 0.31 0.30 0.28 0.26
Gre
0
f
[kPa] 65228 75034 82300 94449 103490
γ0.7 10−3 [m/m] 0.2218 0.1973 0.1813 0.1583 0.1438
K0 [-] 0.5 0.46 0.43 0.38 0.36
k −5
10 [m/s] 20.843 15.946 13.053 9.274 7.128
Rinter [-] 0.63 0.62 0.62 0.61 0.60

emin [-] 0.64


emax [-] 1.05
ϕcrit [◦ ] 33
c0 [kPa] 0.1
m [-] 0.5

Table 3.1: Summary of the geotechnical parameters of Frederikshavn sand for soil models in PLAXIS.

3.2.3 Structure

In order to simplify the analysis, the bucket is attributed properties of a rigid body. This means that
no deformations in the structure are allowed to occur. Since PLAXIS does not provide such option

20
3.3 Model domain

by default, a workaround is found in Østergaard et al. (2015).

The structure is modelled as an elastic body, but instead of using the typical Young’s modulus
for steel (~200 GPa), this value is set to be unrealistically high: E = 600 GPa. Moreover, the plate
thickness is defined as d = 0.3 m, which is about 10 times the usual skirt thickness.

The reasoning behind this procedure is that the main interest lies within the soil behaviour,
rather than the structural response. The rigid body assumption eliminates the deformation of the
structure as an influential factor in the mechanism. Several checks on calculated models confirm
that the plate deflections are indeed very close to zero when the parameters specified above are
used. Therefore, the rigid body assumption is enforced successfully.

3.3 Model domain

There are no restrictions concerning the extents of the model domain. The only rule one has to
keep in mind is that it needs to be large enough for the mechanism to fully develop with minimal
interference from boundaries. When the model constraints affect the area of interest to a significant
degree, the calculation outcome may prove misleading.

As a first trial, the size of the model domain relative to the bucket dimensions is adopted from
Østergaard et al. (2015) and Vethanayagam & Ibsen (2017). This model is labeled "DS0" (Domain
Size 0). Figure 3.7 shows the geometry and size of the domain, including the mesh refinement area,
which is discussed in Section 3.4.

Figure 3.7: Domain size relative to the bucket dimensions in Model DS0. D - diameter, L - skirt length, blue
lines - bucket, orange area - mesh refinement area.

The influence of the model edges is assessed through inspection of shear and effective vertical
stress distribution. The investigated edges are the bottom and the far edge with respect to the bucket

21
Chapter 3. Numerical modelling

(referred to as "right edge"). Ideally, the shear stresses will be equal or almost close to zero along
the specified boundaries, while the vertical stresses will keep a constant value at the bottom.

In Figure 3.8 it is observed that the failure mechanism extends to the bottom of the model, so
the shear stresses have non-zero values along a significant portion of the bottom. Figure 3.8 shows
the results when the bucket is pulled upwards, while in Figure 3.9 the bucket is pushed downwards.
In both cases, the mechanism does not have enough space to fully develop.

Figure 3.8: Shear stress distribution in Model DS0. The bucket is loaded in tension.

As mentioned before, it is of interest to investigate the vertical effective stresses as well.


Figure 3.10 and Figure 3.11 reveal that in Model DS0 the vertical stresses are disturbed at the
bottom.

It is clear that Model DS0, that has the dimensions used by Østergaard et al. (2015), is not
suitable for the formulation of t − z curves. The main problem is that the bottom of the model
affects the failure mechanism considerably. At the same time, the right edge is sufficiently far and
does not have a negative impact.

A series of tests is performed on models with increasing length along Y-axis. An overview of
the models and their dimensions relative to the bucket size is given in Table 3.2.

Model ID DS0 DS1 DS2 DS3 DS4 DS5 DS6 DS7 DS8 DS9 DS10
Height 3L 3.5L 4L 4.5L 5L 5.5L 6L 6.5L 7L 7.5L 8L
Width 4D 4D 4D 4D 4D 4D 4D 4D 4D 4D 4D

Table 3.2: Models with varying heights used for the domain size assessment.

22
3.3 Model domain

Figure 3.9: Shear stress distribution in Model DS0. The bucket is loaded in compression.

Figure 3.10: Effective vertical stress distribution in Model DS0. The bucket is loaded in tension.

Figure 3.12 and Figure 3.13 show the shear stress magnitudes along the bottom of each model at
the last calculation step. As expected, the shear stress decreases with increasing height.

23
Chapter 3. Numerical modelling

Figure 3.11: Effective vertical stress distribution in Model DS0. The bucket is loaded in compression.

1.6
Model DS0 : 3L
Model DS1 : 3.5L
1.4 Model DS2 : 4L
Model DS3 : 4.5L
Model DS4 : 5L
1.2 Model DS5 : 5.5L
Model DS6 : 6L
Model DS7 : 6.5L
Shear stress [kPa]

1 Model DS8 : 7L
Model DS9 : 7.5L
Model DS10 : 8L

0.8

0.6

0.4

0.2

0
0 10 20 30 40 50 60
X-coordinate [m]

Figure 3.12: Shear stress along the bottom line. The bucket is loaded in tension.

Considering only the shear stress values at the bottom, it is obvious that the largest model
(DS10) is the best choice. However, using this domain size would slightly increase the quality of
the outcome at the expense of computation time. Therefore the trade-off is not justified. Instead,
Model DS6 is deemed a better option, as it offers an optimal balance between the result quality and
computational demand. Regarding the shear stresses at the bottom, it is accepted that as long as the
values are below 0.2 kPa the edge effects are negligible. Furthermore, the effective vertical stresses

24
3.3 Model domain

4
Model DS0 : 3L
Model DS1 : 3.5L
3.5 Model DS2 : 4L
Model DS3 : 4.5L
Model DS4 : 5L
3 Model DS5 : 5.5L
Model DS6 : 6L
Model DS7 : 6.5L
Shear stress [kPa]

2.5 Model DS8 : 7L


Model DS9 : 7.5L
Model DS10 : 8L

1.5

0.5

0
0 10 20 30 40 50 60
X-coordinate [m]

Figure 3.13: Shear stress along the bottom line. The bucket is loaded in compression.

along the bottom are checked and plotted in Figure 3.14 and Figure 3.15.

1.01

1.005
Normalized effective vertical stress [kPa]

0.995

0.99
Model DS0 : 3L
Model DS1 : 3.5L
0.985 Model DS2 : 4L
Model DS3 : 4.5L
Model DS4 : 5L
0.98 Model DS5 : 5.5L
Model DS6 : 6L
Model DS7 : 6.5L
0.975 Model DS8 : 7L
Model DS9 : 7.5L
Model DS10 : 8L

0.97
0 10 20 30 40 50 60
X-coordinate [m]

Figure 3.14: Effective vertical stress along the bottom line. The bucket is loaded in tension.

As opposed to shear stress, the effective vertical stress is normalized with respect to its highest
value, since the objective is to check how much it changes along the bottom. It is observed that
Model DS6 performs well both in tension and compression, as the vertical stress remains almost
constant.

25
Chapter 3. Numerical modelling

1.01

1.005
Normalized effective vertical stress [kPa]

0.995

0.99
Model DS0 : 3L
Model DS1 : 3.5L
0.985 Model DS2 : 4L
Model DS3 : 4.5L
Model DS4 : 5L
0.98 Model DS5 : 5.5L
Model DS6 : 6L
Model DS7 : 6.5L
0.975 Model DS8 : 7L
Model DS9 : 7.5L
Model DS10 : 8L

0.97
0 10 20 30 40 50 60
X-coordinate [m]

Figure 3.15: Effective vertical stress along the bottom line. The bucket is loaded in compression.

The last check is made in connection with the right edge of the domain. Figure 3.16 and
Figure 3.17 show that shear stress is very close to zero for all models, therefore no further study is
conducted on this issue.

-20

-40
Y-coordinate [m]

-60
Model DS0 : 3L
Model DS1 : 3.5L
Model DS2 : 4L
-80 Model DS3 : 4.5L
Model DS4 : 5L
Model DS5 : 5.5L
Model DS6 : 6L
-100 Model DS7 : 6.5L
Model DS8 : 7L
Model DS9 : 7.5L
Model DS10 : 8L

-120
-0.014 -0.012 -0.01 -0.008 -0.006 -0.004 -0.002 0
Shear stress [kPa]

Figure 3.16: Shear stress along the right edge. The bucket is loaded in tension.

The conclusion of the domain size investigation is that Model DS6, with height = 6L and width
= 4D, is a suitable basis for further numerical simulations. Figure 3.18 shows the shear stress

26
3.3 Model domain

-20

-40
Y-coordinate [m]

-60
Model DS0 : 3L
Model DS1 : 3.5L
Model DS2 : 4L
-80 Model DS3 : 4.5L
Model DS4 : 5L
Model DS5 : 5.5L
Model DS6 : 6L
-100 Model DS7 : 6.5L
Model DS8 : 7L
Model DS9 : 7.5L
Model DS10 : 8L

-120
-0.04 -0.035 -0.03 -0.025 -0.02 -0.015 -0.01 -0.005 0
Shear stress [kPa]

Figure 3.17: Shear stress along the right edge. The bucket is loaded in compression.

distribution at the last calculation step. It is noticeable that the failure mechanism is fully developed.
Figure 3.19 reveals that the vertical stresses towards the domain bottom are undisturbed.

Figure 3.18: Shear stress distribution in Model DS6. The bucket is loaded in tension (left) and in compres-
sion (right).

27
Chapter 3. Numerical modelling

Figure 3.19: Effective vertical stress distribution in Model DS6. The bucket is loaded in tension (left) and in
compression (right).

3.4 Mesh

PLAXIS 2D offers a choice between two soil element types: 6-node and 15-node triangular
elements. Figure 3.20 illustrates the difference between the two.

Figure 3.20: Soil elements available in PLAXIS 2D: (a) 6-node (b) 15-node

The advantage of 15-node elements is that they provide more accuracy due to increased amount
of degrees of freedom and stress points per element. As fourth-order elements, they can distort in
more ways than the second-order 6-node elements. Moreover, with 12 stress points compared to
only 3, they can generate a better representation of stress fields, especially in areas with large stress
gradients. However, this comes at the cost of computation time and power. Having all the required
resources at disposal, 15-node soil elements are chosen for numerical modelling.

28
3.4 Mesh

Figure 3.7 shows a rectangle around the bucket with dimensions 1.5 D × 2 L. This represents
the area of mesh refinement. The target is to obtain an accurate picture of stresses in the region of
interest, where high stress gradients are expected. Refining the mesh globally would not bring any
significant benefits in terms of describing the soil-structure interaction.

The problem is to find the balance between the amount of mesh refinement and accuracy of
results. Therefore a mesh convergence study is conducted. The procedure consists of gradually
reducing the element size (thus increasing their number) and checking the developed shear stress at
a specific point or zone.

In PLAXIS the local mesh refinement is possible by defining a so-called coarseness factor.
This is a quantity that relates the element size within a specific area to the global element size. The
global element size is set to Medium.

In every model involved in the mesh convergence study the bucket is loaded in tension, drained
conditions are assumed and the domain is given appropriate dimensions as established in Section 3.3
. The chosen bucket dimensions are D = 15 m and L = 15 m. The soil parameters set corresponds
to ϕ = 35◦ , see Table 3.1.

The coarseness factor is decreased down to the lowest value allowed by the program (0.03125).
An overview on the models is given in Table 3.3.

Model ID MC1 MC2 MC3 MC4 MC5 MC6 MC7 MC8


CF 1 0.8 0.6 0.4 0.3 0.2 0.18 0.16
Model ID MC9 MC10 MC11 MC12 MC13 MC14 MC15
CF 0.14 0.12 0.1 0.08 0.06 0.04 0.03125

Table 3.3: Models with varying coarseness factors used for the mesh convergence study.

The interface shear stresses averaged around the middle of the bucket skirt are investigated
considering the outer side of the bucket. It is the values at the end of the calculation that are included
in the study, as opposed to peak shear stress, which occurs before the bucket reaches maximum
displacement (provided that the displacement is large enough).

Figure 3.21 shows that convergence is achieved starting with a coarseness factor of 0.08. Further
local refinement does not bring any change to the outcome. Another point of interest is the tip of
the skirt, where high stress concentration is observed. From Figure 3.22 it is seen that shear stress
increases even after a coarseness factor of 0.08 is applied. This goes according to expectations,
since in regions of high stress concentration the value increases asymptotically.

In conclusion, model MC12 with a coarseness factor of 0.08 is deemed acceptable for use in
further FE analysis.

Figure 3.23 and Figure 3.24 display the chosen mesh.

29
Chapter 3. Numerical modelling

10.85

10.8

10.75

10.7

10.65

10.6

10.55
0.12
10.5

10.45
0.1
0.06 0.04 0.03125
10.4
0.08
10.35
0 0.5 1 1.5 2 2.5
105

Figure 3.21: Convergence of shear stress on the outer side of the bucket.

25.8
0.03125
25.7 0.04
0.08
0.06
25.6
0.1

25.5
0.12
25.4

25.3

25.2

25.1
0 0.5 1 1.5 2 2.5
105

Figure 3.22: Convergence of shear stress at the skirt tip.

30
3.4 Mesh

Figure 3.23: Mesh of the entire model. The quality of elements is given on a scale from 0 to 1, where 0
denotes badly shaped elements (skewed) and 1 indicates ideally shaped elements (equilateral
triangles). The area of mesh refinement is in the upper left corner, where a concentration of
elements takes place.

31
Chapter 3. Numerical modelling

Figure 3.24: Close-up of the mesh around the bucket. The same color bar applies as seen in Figure 3.23.

3.5 Interfaces

Figure 3.25: Interfaces around the bucket [Østergaard et al. (2015)].

32
3.5 Interfaces

To model the soil-structure interaction adequately, interfaces are added along contact surfaces
between structure and soil, illustrated by blue lines in Figure 3.25

At locations where changes in geometry of the structure are abrupt (e.g. corner of a footing
or tip of the bucket skirt) numerical instabilities are likely to occur. In order to avoid sudden
nonphysical jumps in stresses, a common solution is to define interfaces around such locations,
as explained by Brinkgreve et al. (2017). As a first trial, the extended interfaces are established
following the example from Østergaard et al. (2015), shown in Figure 3.25.

The results of several test simulations reveal that the introduction of a horizontal interface at
the bottom (magenta line in Figure 3.25) leads to the formation of an unexpected gap when the
bucket is pulled upwards. This interface seems to divide the soil body, thus creating a nonphysical
discontinuity. This phenomenon is presented in Figure 3.26.

Figure 3.26: Deformations at the bottom of the bucket loaded in tension. The thin horizontal "slit" is the
result of the bottom interface. Larger gaps form at the intersection of the two interfaces.

The gap appears even though the interface is made permeable and its strength is not reduced.
As a workaround, it is decided to eliminate the interface at the bottom and only leave the vertical
extended one (red line in Figure 3.25).

A further analysis sheds light onto issues with the length of the extended interface. The shear
stress concentration takes place at the tip of the interface, instead of at the skirt tip. For example, in
Figure 3.27 the offset is 2 m for a bucket diameter of 10 m. The core problem persists: an extended
interface in the soil creates an unrealistic mechanism.

The attempt to completely remove the extended interface leads to calculation failures. This
may be due to the aforementioned numerical instabilities. It is decided to reduce the length of the
extended interface, in order to "pull" the stress concentration point as close as possible to the skirt
tip, thus modelling more realistic behaviour.

33
Chapter 3. Numerical modelling

Figure 3.27: Shear stress distribution at the skirt tip. The bucket is loaded in tension. D = 10 m. Length of
the extended interface = 0.2D. Blue - bucket structure, red - extended interface.

An optimal extended interface length must be found. The criterion is established in the context
of mesh quality. To give a clear insight, two topics are addressed:

• Element shapes. FE results tend to be inaccurate when badly shaped elements are involved,
especially in non-linear analysis, as explained by Cook et al. (2002). For triangular elements,
the perfect shape is an equilateral triangle. One generally tries to avoid extremely skewed
elements, see Figure 3.28.

Figure 3.28: (a) Equilateral triangle (b) Skewed triangle

• Element connections. It appears that no matter how short the extended interface is, PLAXIS
will connect it with an entire edge of an adjacent soil element as pictured in Figure 3.29.
Elongating the interface to accommodate for two soil elements brings back the initial problem
of excessive length.

It is determined that the extended interface length of 0.02 D provides acceptable mesh quality,
while keeping the stress concentration point reasonably close to the skirt tip, see an example in
Figure 3.30. Such a short length allows the elements to have a good shape due to the very small
coarseness factor of 0.08, found in Section 3.4.

34
3.6 Loading and boundary conditions

Figure 3.29: Connection between 15-node soil elements and interface elements. (a) Long interface may
generate well-shaped triangles. (b) Extremely short interface enforces skewed elements.

Figure 3.30: Shear stress distribution at the skirt tip. The bucket is loaded in tension. D = 10 m. Length of
the extended interface = 0.02D. Blue - bucket structure, red - extended interface.

3.6 Loading and boundary conditions

Section 3.2.3 mentions the rigid body condition assumed for the bucket, which is modelled with plate
elements in PLAXIS. Aside from attributing extremely high and unrealistic stiffness to the plate
material, the loading procedure is also carried out in a manner that ensures rigid body behaviour.

Loading is defined as prescribed displacements along the entire structure. All nodes within the
plate elements are thus assigned vertical displacements of equal magnitude. This is sketched in
Figure 3.31. Moreover, in order to avoid bucket rotation the plate nodes are fixed in horizontal
direction. As a result, the response to pure vertical loading can be generated.

Following the literature review, Chapter 2, the boundary conditions are established so as to
respect the condition of axisymmetry.

35
Chapter 3. Numerical modelling

Figure 3.31: Prescribed vertical displacements assigned to every node in the plate elements and the horizontal
restriction imitate a rigid bucket moving purely upwards.

The nodes at the left and right edge of the model are restricted in the horizontal direction, while
the ones at the bottom are fixed in both directions. This is done automatically by PLAXIS and
illustrated in Figure 3.32.

Figure 3.32: Displacement restrictions along model edges.

In terms of hydraulic BCs, the left edge (axis of symmetry) is defined as impermeable. Water is
allowed to seep through the rest of the edges. Hydarulic BCs are shown in Figure 3.33.

36
3.7 Calculation

Figure 3.33: Hydraulic boundary conditions. Red - impermeable (closed boundary), green - fully permeable
(open boundary).

3.7 Calculation

The last step in FE analysis is running the calculation. For this purpose, four consecutive calculation
phases are defined and listed below:

1. Initial phase. The initial soil stresses are generated based on the specified in-situ conditions.
Soil and groundwater flow BCs are activated. The K0 procedure is carried out during this
phase.
2. Installation. The structures and interfaces are activated. The real-life gradual installation of
the bucket is not simulated.
3. Nil-step. The purpose of this phase is to reset the displacements and small strains to zero,
thus reestablishing the equilibrium after the "installation" of the bucket.
4. Loading. The prescribed displacements are activated. The fully coupled flow-deformation
analysis is used. Pore pressures are taken into account. Being a time-dependent calculation
type, it is possible to control the velocity of the structure.

37
C HAPTER 4

Interpretation of results

The outcome of the FE analysis is discussed in this chapter. It begins with a general overview
on the created models, followed by explanations of how data is handled. The interpretation
of results is given with respect to the ability of numerical methods to simulate the realistic
behaviour of suction buckets subjected to monotonic vertical loading.

4.1 Model overview

A total amount of 100 models is generated. This quantity is justified by the purpose of describing
the general behaviour of suction buckets installed in marine sand. The influence of the governing
factors is examined by comparing the outcomes of individual simulations. These factors are listed:

• Bucket geometry. Foundation dimensions have a straightforward effect on performance,


hence the reason to investigate suction buckets with varying sizes. The aspect ratio D/L
(diameter/skirt length) in all models is 1.
• Soil properties. The in-situ characteristics of soil dictate the geotechnical design. Their
properties are rarely uniform even within relatively small volumes. To capture the full range
of expected Frederikshavn sand characteristics, different sets of geotechnical parameters are
applied. Since each set of properties is based on a specific value of the friction angle ϕ, as
seen in Section 3.2.2, this parameter is used as identification for the model that involves the
respective set.
• Loading direction. Suction buckets exhibit different behaviour depending on the direction
of the applied vertical load. Thus, both tensile and compressive loading conditions are
considered.
• Drainage conditions. Two cases are taken into account: drained and undrained. While the
former can be handled by PLAXIS with relative ease, the latter poses significant obstacles in
terms of accurate computing, as discussed in further sections. Moreover, the term "undrained"
is used interchangeably with "partially drained". It appears that the program is not able to
simulate undrained behaviour in all sense of the concept. However, this is not considered a
drawback, since real soil response lies in between the two theoretical states.

Each model involves a unique combination of these factors as input parameters. Therefore,
the index used to identify a model has the structure Diameter [m] - Friction angle [◦ ] - Drainage
condition [d = drained, u = undrained] - Loading direction [t = tension, c = compression]. For

39
Chapter 4. Interpretation of results

example, Model 10-35-d-t implies a bucket diameter D and skirt length L of 10 m, friction angle
ϕ = 35◦ , drained conditions and tensile loading. Table 4.1 displays all the models generated with
PLAXIS and used for the formulation of t − z curves.

ϕ [◦ ]
30 33 35 38 40
D & L [m]
10 10-30-i-j 10-33-i-j 10-35-i-j 10-38-i-j 10-40-i-j
13 13-30-i-j 13-33-i-j 13-35-i-j 13-38-i-j 13-40-i-j
15 15-30-i-j 15-33-i-j 15-35-i-j 15-38-i-j 15-40-i-j
18 18-30-i-j 18-33-i-j 18-35-i-j 18-38-i-j 18-40-i-j
20 20-30-i-j 20-33-i-j 20-35-i-j 20-38-i-j 20-40-i-j

Table 4.1: Model indices. i = "d" or "u" ; j = "t" or "c".

Finally, due to the nature of the analysis to follow in the next chapters, a reference model is
required. The chosen reference model is 15-35-i-j (Dre f = Lre f = 15 m, ϕre f = 35◦ ). The raw t − z
curves for all models may be found in Appendix D.

4.2 Data processing

At the beginning of Chapter 3 arguments are given in favour of the use of PYTHON scripting for
generating models in PLAXIS. The same idea stands behind extraction of FE results: automation
by scripting offers advantages in terms of speed. Appendix C explains how data is extracted and
stored into files with PYTHON scripting.

Data is further imported in MATLAB, where filtering, sorting, calculation and plotting take
place.

4.3 Integration of stresses

The response of the bucket is calculated by integrating stresses in interface elements. Wolf et al.
(2013) in their study prove that stresses in interfaces give a clearer picture of the soil response, in
contrast to stresses extracted from adjacent soil elements. The reason is that stress points in soil
elements are located at a certain distance from the structure, and thus provide smaller values than
stress points in the interfaces. A study to confirm this phenomenon is conducted. However, for the
sake of good order, the outcome is presented only later in Section 4.4.

4.3.1 Integration along the skirt

Inspiration on the stress integration method along the skirt comes from Wolf et al. (2013). Since
the objective of their study was p − y curves and involved 3D modelling of a laterally loaded pile,
necessary adjustments are made in the context of axisymmetry.

Keeping in mind the cylindrical symmetry, it is clear that stresses at a given depth and radial
distance do not vary irrespective of the angle relative to the central axis, see Figure 4.1.

40
4.3 Integration of stresses

Figure 4.1: Any stress σ acting on the bucket at a given depth is constant no matter what the angle θ is with
respect to a reference line (red).

The bucket skirt is split into layers of approximately equal thickness. The layer thickness must
be large enough so that it encompasses enough stress points from the interface. Should the thickness
be too large though, there is a risk that a given layer will mix the general soil response with more
complex mechanisms at the tip or top of the skirt. An example is sketched in Figure 4.2.

Figure 4.2: Examples of skirt discretization into layers. Interrupted line - layer boundary; circles - interface
stress points; pink area - complex mechanism; green area - simple mechanism. (a) The layers are
too thick. Mechanism mixing occurs. (b) The layers are too thin. Not enough stress points are
captured by individual layers. (c) The layer thickness is balanced. Mechanisms can be separated.
Every layer comprises an acceptable number of stress points.

The layer-defining algorithm in MATLAB is designed so that no stress point is shared by two
layers. The average shear stress per layer is then computed.

41
Chapter 4. Interpretation of results

Layer-wise averaging of stresses leads to "smoothening" the results. This is considered to have
a positive impact on the quality of general behaviour assessment. Numerical methods are inherently
discrete and prone to errors, which makes it impossible to draw accurate conclusions by accounting
for nodes or stress points individually.

The layer boundaries in 2D delimit the thickness of a circular strip around the bucket skirt in
3D, as shown in Figure 4.3.

Figure 4.3: Strip around the bucket skirt. Its area multiplied by the averaged stress in the respective layer
gives the force.

The vertical friction force resulting from one distinct layer is estimated with Eq. (4.1).

! !
1 n 1 n
FF,k = πDout lk ∑ τi,out
n i=1
+ πDin lk ∑ τi,in
n i=1
(4.1)

where

FF,k resulting vertical friction force in layer k


lk thickness of layer k
Dout|in outer/inner diameter of the bucket
τi,out|in shear stress from outer/inner i-th interface stress point
n amount of interface stress points within layer k

It is no coincidence that the amount of interface stress points in a layer is the same on outer and
inner side. When interfaces are defined on both sides of plate elements, PLAXIS assigns identical
nodal coordinates to the respective interface elements. The total friction force FF on the skirt is
calculated with Eq. (4.2).

42
4.3 Integration of stresses

FF = FF,1 + FF,2 + FF,3 + · · · + FF,m (4.2)

where m - amount of layers.

4.3.2 Integration under the lid

The contribution of the lid to the bucket’s axial resistance is based on different forces depending on
the direction of the vertical load. When compressive loads are applied, earth pressure under the
lid is considered. On the contrary, during tensile loading, provided that excess pore pressure is
generated, resistance is gained from the difference between ambient pressure and absolute pressure
under the lid.

Stresses from the lid interface are integrated similarly to ones from the skirt interfaces, with
considerations given to the geometric entity that is formed in 3D when a 2D body is rotated around
a central axis. Examine Figure 4.4.

slice in 2D

bucket lid

Figure 4.4: Splitting the lid into slices in 2D space and rotating them around the central axis generates
annuli. One annulus is marked with black.

As discussed in the previous section, the slice thickness must be balanced and capture enough
stress points. Multiplying the average stress in one slice with the area of the affiliated annulus
results in the force per given annulus.

Whether the bucket is subjected to compression or tension, the concept of estimating the
resultant force from FE data is the same, see Eq. (4.3) and Eq. (4.4).

43
Chapter 4. Interpretation of results

!
m n
1 j 
FP = π ∑ ∑ σN,i
n j i=1
R2j − r2j for compression (4.3)
j=1

!
m n
1 j 
FS = π ∑ ∑ ∆ui
n j i=1
R2j − r2j for tension (4.4)
j=1

where

FP total vertical force under the lid due to soil pressure


FS total force under the lid due to pressure differential
σN,i normal stress in the i-th stress point within a slice
∆u difference between hydrostatic pressure above the lid and pore water pressure in the
i-th stress point within a slice
Rj big radius of the j-th annulus
rj small radius of the j-th annulus
nj amount of interface stress points in the j-th slice
m amount of slices

4.4 Drained models

Drained conditions imply that no excess pore pressures are generated. The output inspection starts
with the drained models, as the system behaviour is expected to be more elementary than the one in
undrained models.

The presentation of results and the discussion take place mainly with respect to the reference
model 15-35-d-j (Dre f = Lre f = 15 m, ϕre f = 35◦ , drained, tension/compression).

4.4.1 Tension

For simulating drained tensile loading, the bucket is pulled at a rate of v = 1 mm/s. Since the weight
of the structure is zeroed, resistance comes only from friction between soil and inner/outer skirt
face.

The first assessment of raw results, shown in Figure 4.5 and Figure 4.6, meets the expectations:
shear stress climbs linearly towards a peak, after which it becomes steady. In the figures’ legends,
the parameter d indicates the mid-layer depths in meters.

Some critical observations are commented:

• The response becomes stiffer with increasing depth. This is due to the increase in vertical
overburden stress, which leads to larger confining pressure. With a larger force acting
perpendicularly to the skirt face, more friction is mobilized. This is reflected in higher peak

44
4.4 Drained models

25
d = 0.7
d = 2.3
d = 3.8
d = 5.2
20 d = 6.7
d = 8.3
d = 9.8
d = 11.3
d = 12.8
d = 14.3
15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.5: Raw t − z curves for the reference model (outer side).

25
d = 0.7
d = 2.3
d = 3.8
d = 5.2
20 d = 6.7
d = 8.3
d = 9.8
d = 11.3
d = 12.8
d = 14.3
15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.6: Raw t − z curves for the reference model (inner side).

shear stress at larger depths. Another consequence is the bucket displacements at which
peaks occur. They seem to depend on vertical overburden stress in a non-linear manner.
• On the outer side, the difference between peak and residual shear stress is larger at smal-
ler depths, decreasing gradually as depth increases. Thus another relation to the vertical
overburden pressure is noticed. However this effect does not take place on the inner skirt
side.

45
Chapter 4. Interpretation of results

• The edge effects are more pronounced at the top than at the bottom. The curves corresponding
to d = 0.7 and d = 2.3 exhibit slight deviations from the general trend. These layers would
be ignored in t − z formulation.
• Comparing the two sides, the general behaviour is very similar, with peak shear stress being
reached at smaller displacements on the inner side. As long as excess pore pressures are not
involved, differences between sides arise because soil is confined in the bucket on one side
and "free" on the other.

To give another viewpoint, the t − z curves are plotted in 3D without averaging the shear stresses.
In Figure 4.7 each curve represents one stress point from the interface.

30

25

20

15

10

0.04 15
10
0.02
5
0 0

Figure 4.7: Raw t − z curves for each stress point (outer side). The red dots mark the peaks.

The effect of complex unpredictable mechanisms is evident at the skirt top (depth 0 m to app. 1
m) and bottom (depth app. 14 m to 15 m). Curves in these zones follow a general trend, but they
are considered outliers nonetheless. The phenomenon is more emphasized in Figure 4.8.

At the beginning of Section 4.3 the correctness of extracting the stresses from interfaces rather
than from soil was mentioned, with reference to Wolf et al. (2013). The same study is undertaken
here, and the results are plotted in Figure 4.9. It is seen that stresses from soil elements indeed adopt
lower values than the ones from interface elements. In conclusion, data extraction from interfaces
is the appropriate method to visualize soil-structure interaction.

46
4.4 Drained models

30

25

20

15

10

0.04 15
10
0.02
5
0 0

Figure 4.8: Raw t − z curves for each stress point (inner side). The red dots mark the peaks.

25
Interface
Soil

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.9: t − z curves, interface vs soil (outer side). The legend containing depths is purposely omitted.

The total average shear stress per layer at a given vertical displacement is found simply by

47
Chapter 4. Interpretation of results

summing the outer and inner friction. Figure 4.10 displays the t − z curves for the total response.

50
d = 0.7
d = 2.3
45
d = 3.8
d = 5.2
40 d = 6.7
d = 8.3
d = 9.8
35 d = 11.3
d = 12.8
d = 14.3
30

25

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.10: Raw t − z curves for the reference model.

The superposition of effects from both sides shows a hyperbolic trend just before the peak is
reached. Comparing with Figure 4.5 and Figure 4.6, it is noticed that the peaks are less pronounced.
Moreover, the difference between peak and residual stress observed on the outer side is almost
completely cancelled out.

30

25

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.11: t − z curves for varying ϕ.

48
4.4 Drained models

It is noted that both peak shear stress and the displacement at which it occurs are linked to
friction angle, see Figure 4.11. According to the way Frederikshavn sand properties were established
(Section 3.2.2), friction angle is directly related to relative density. It is apparent that the state of
compaction plays a critical role in the development of shear stress between soil and structure.

4.4.2 Compression

A significant difference between drained tension and drained compression is that during the latter
the axial resistance is a sum of friction force along the skirt and soil pressure under the lid. The raw
t − z curves for the reference model are shown in Figure 4.12 and Figure 4.13.

30
d = 0.7
d = 2.3
d = 3.8
25 d = 5.2
d = 6.7
d = 8.3
d = 9.8
d = 11.3
20 d = 12.8
d = 14.3

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.12: Raw t − z curves for the reference model (outer side).

On the outer side the curves are similar to the tension case in that there is a clear peak. However,
the post-failure shear stress does not become constant, but continues with a climb. The bigger
difference appears on the inner side, where no peak is present. Figure 4.14 illustrates the sum of
both sides. The total response exhibits a hyberbolic trend.

The perpetual growth in shear stress without a clearly defined peak is explained by the increase
in confining pressure on the inner side due to the lid that compresses the soil body. To visualize this
phenomenon, normal stresses on the skirt are plotted in Figure 4.15.

49
Chapter 4. Interpretation of results

35
d = 0.7
d = 2.3
d = 3.8
30 d = 5.2
d = 6.7
d = 8.3
d = 9.8
25 d = 11.3
d = 12.8
d = 14.3

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.13: Raw t − z curves for the reference model (inner side).

70
d = 0.7
d = 2.3
d = 3.8
60 d = 5.2
d = 6.7
d = 8.3
d = 9.8
50 d = 11.3
d = 12.8
d = 14.3

40

30

20

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.14: Raw t − z curves for the reference model.

50
4.5 Partially drained models

load step load step


increase increase

-5

outer side
inner side
-10

-15
80 60 40 20 0 20 40 60 80

Figure 4.15: Normal stresses on inner and outer skirt face. Stress increases with each load step. Numerical
instabilities occur on account of complex mechanisms at top and bottom. The blue line
represents the skirt.

4.5 Partially drained models

The bucket is vertically displaced at the speed of v = 10 mm/s to simulate partially drained loading.
The purpose is to generate excess pore pressures and assess their effect on the axial response of the
system. Similarly to previous section, the interpretation of results revolves around the reference
model 15-35-u-j (Dre f = Lre f = 15 m, ϕre f = 35◦ , partially drained, tension/compression).

4.5.1 Tension

As the bucket is pulled up groundwater flow is triggered, where water from surrounding soil flows
towards the inner side of the lid. The fully coupled flow-deformation analysis in PLAXIS models
this mechanism successfully. See Figure 4.16.

One proof that partially drained behaviour is simulated can be found by looking at the space
between the lid and the soil surface underneath. In a pure undrained case, the soil surface would
follow the lid without a gap forming between them. This is almost achieved, as revealed in
Figure 4.17.

Contrary to drained models, significant differences are noticed between development of shear
stresses inside and outside the skirt, see Figure 4.18 and Figure 4.19.

On the outer side, the t − z curves resemble the ones presented in drained models. The
continuous slight increase in shear stress after failure is explained by the downwards flow affecting
the vertical overburden pressure.

51
Chapter 4. Interpretation of results

Figure 4.16: Velocity field of groundwater flow when the bucket is pulled upwards. The gap at skirt tip
represents deactivated view of soil elements where velocity vectors are too large due to the
singularity point.

0.05
Lid
0.045 Soil surface

0.04

0.035

0.03

0.025

0.02

0.015

0.01

0.005

0
0 2 4 6 8 10 12 14

Figure 4.17: Vertical displacements of the bucket lid and of the soil surface underneath.

52
4.5 Partially drained models

70
d = 0.6
d = 1.7
d = 2.9
60 d = 4.1
d = 5.2
d = 6.4
d = 7.5
50 d = 8.7
d = 9.8
d = 11
d = 12.1
40 d = 13.2
d = 14.4

30

20

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.18: Raw t − z curves for the reference model (outer side).

6
d = 0.6
d = 1.7
d = 2.9
d = 4.1
5
d = 5.2
d = 6.4
d = 7.5
d = 8.7
4 d = 9.8
d = 11
d = 12.1
d = 13.2
d = 14.4
3

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.19: Raw t − z curves for the reference model (inner side).

53
Chapter 4. Interpretation of results

No peaks occur on the inner side, where the upwards flow is considerably stronger. Notice that
the values shear stresses adopt are in the order of 10 times lower than on the opposite side. The
soil displacements relative to skirt displacements are smaller than on the outer side, as soil moves
together with the bucket. Consequently, less friction is mobilized.

The fastest flow happens around the skirt tip, resulting in a dramatic increase of shear stress on
the outer side, while exerting opposite effect on the other face. The results linked to these areas are
discarded for general t − z formulation.

Superimposing the results from two sides, one observes that the total response is biased by the
outer friction, see Figure 4.20.

70
d = 0.6
d = 1.7
d = 2.9
60 d = 4.1
d = 5.2
d = 6.4
d = 7.5
50 d = 8.7
d = 9.8
d = 11
d = 12.1
40 d = 13.2
d = 14.4

30

20

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 4.20: Raw t − z curves for the reference model.

4.5.2 Compression

The partially drained behaviour when the bucket is compressed proves to be similar to the one
in drained compressive case, at least with the applied loading rate of 10 mm/s. Therefore it is
considered redundant to expand on the current topic.

Note on undrained modelling

Generally, PLAXIS experiences difficulties in terms of converging to a solution when extreme


excess pore pressures arise. In this study, such situations were created when a fast loading rate (e.g.
v = 50 mm/s) was applied in combination with a very low permeability (e.g. k = 10−5 m/s). The
"soil body collapse" error is accompanied by severe numerical instability across the entire model,
i.e. complete misrepresentation of a realistic mechanism.

54
4.5 Partially drained models

A workaround was found by gradually altering several model parameters:

• Load step size. Depending on the goal of the analysis, it may be useful to force PLAXIS to
apply the loads in smaller increments in order to observe the development of deformations/-
pore pressures. However, increments that are too small (i.e. too many steps) might cause
instability. In such a case, decreasing the amount of steps could improve the calculation
procedure. This is controlled with the numerical parameter Max load fraction per step. See
Brinkgreve et al. (2017) for instructions on how to use the parameter.
• Extended interface length. When the length of the extended interface is increased, numerical
stability in terms of pore pressure becomes more achievable. Although considerations must
be taken with regards to realistic stress distribution, as explained in Section 3.5.
• Surface water level. Raising the height of the water column increases the hydrostatic pressure,
which seems to have a stabilizing effect. This parameter has the largest positive impact
compared to the following ones.
• Effective cohesion. The soil body gains stability with increasing cohesion. However, this
value must be kept to an appropriate minimum considering cohesionless soils. Cohesion no
larger than 0.4 kPa was used.
• Hydraulic conductivity. Another option is to make the soil more permeable, in order to
reduce the level of excess pore pressures. For the sake of simulating realistic groundwater
flow through a certain soil type, this is the last parameter one should consider changing.

55
C HAPTER 5

Normalization

In the current chapter the normalization of t-z curves is presented. Various approaches are
discussed and their effects are outlined. Since success in terms of normalization varies from
one method to another, choices of best ones are made. The process of selection is discussed
while arguing the final decisions.

5.1 Data trimming

The results of the study must be expressed in dimensionless terms in order to be able to formulate
general t − z curves. Moreover, normalization allows the assessment of differences between models
and the comparison between the results of the current study and other ones.

As discussed in Chapter 4, the t − z curves reflecting the edge mechanisms at the top and tip of
the skirt should be ignored, as they do not follow the general trend, see Figure 5.1.

Figure 5.1: The curves generated from stress points around top and bottom of the skirt are discarded due
to more complex mechanisms (pink areas). The isolation is done by defining appropriate layer
boundaries (interrupted lines). The curves that follow a specific trend are accepted. In this
sketch the number and thickness of layers, and extents and shapes of complex mechanisms are
only indicative of the concept.

57
Chapter 5. Normalization

The following figures give an example of data trimming:

70
d = 0.7
d = 2.3
d = 3.8
60 d = 5.2
d = 6.7
d = 8.3
d = 9.8
50 d = 11.3
d = 12.8
d = 14.3

40

30

20

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 5.2: Non-trimmed t − z curves. The effect of the mechanism at skirt tip is especially strong.

70
d = 2.3
d = 3.8
d = 5.2
60 d = 6.7
d = 8.3
d = 9.8
d = 11.3
50 d = 12.8

40

30

20

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

Figure 5.3: Trimmed t − z curves. The edge effects are eliminated by discarding the curves corresponding
to the the top and bottom layers.

58
5.2 Bucket under axial tension

5.2 Bucket under axial tension

5.2.1 Drained models

The first attempt to normalize the data is carried out in a similar manner to Østergaard et al. (2015),
who used Rankine pressure and bucket diameter for normalization. An example is illustrated in
Figure 5.4.
Model 9, D = 15, L = 7.5, φ = 40 Model 9, D = 15, L = 7.5, φ = 40
11000 3
z = 2.25
z = 2.75
10000 z = 3.25
z = 3.75
z = 4.25 2.5
9000 z = 4.75
z = 5.25
z = 5.75
8000 z = 6.25
2

7000

[−]
p [kN/m]

rankine
6000 1.5

5000 p/p
z = 2.25
1
4000 z = 2.75
z = 3.25
z = 3.75
3000 z = 4.25
0.5 RMSE = 0.086654 z = 4.75
z = 5.25
2000 z = 5.75
z = 6.25
Best fit
1000
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0
y [m] 0 0.01 0.02 0.03 0.04 0.05
y/D [−]

Figure 9: Step 1 - Raw results from the model are plotted in a p − y Figure 10: (right).
Step 3 andDividing
4 - The results
Figurediagram.
5.4: Trimmed p −already
This data have y curves
been (left) and
trimmed, cf. normalized
section results thearelateral
trimmedsoil
so edge effects inby
pressure the
skirt top and bottom are removed. The best fit is shown.
6.3. Rankine pressure and the lateral displacement by bucket diameter provides acceptable results,
as the curves become almost aligned. From Østergaard
3 et al. (2015).
6.2 Normalization of Soil Pressure
Since Rankine pressure can be easily computed, it represents a convenient option to normalize
To eliminate the depth dependency of the p − y curves, 2.5
p − y curves. An analogue formula related to
the soil pressure p is normalized by the Rankine pres-
t − z curves is found in API (2011) and presented by
Eq. p(5.1).
sure R which is a linear function of the depth z. The 2
displacement y is normalized by the bucket diameter D.
The Rankine pressure is calculated as,
[−]
rankine

1.5
p
pR = γ 0 zD(K
0 γ
− Kγa ) (12)
p/p

τ = σvo K tan δ  (5.1)


1 + sin(ϕ) 1 − sin(ϕ)
= γ 0 zD − , 1
1 − sin(ϕ) 1 + sin(ϕ)
Model 1, D = 10, L = 5, φ = 30
γ 0 is theτspecific
wherewhere - unit soil
skinweight,
friction, σvo
z is the0 depth and ϕ isvertical 0.5
- effective overburden pressure at Model a specific depth, K -
4, D = 10, L = 10, φ = 30
Model 7, D = 15, L = 7.5, φ = 30
the internal angle of friction of the soil. By normalizing
coefficient of lateral earth pressure and δ - friction angle between soil and structure. Model 10, D = 15, L = 15, φ = 30
the pressure p by the Rankine pressure that is a linear Model 13, D = 20, L = 10, φ = 30
Model 16, D = 20, L = 20, φ = 30
function of the depth z, the p − y curves for each depth 0
layer The average shear stress per layer isp isnormalized with respect to the peaky/Dshear stress0.04 calculated
0 0.01 0.02 0.03 0.05
will turn into one curve, if the pressure also a [−]

withfunction
linear Eq. (5.1). of z.The results for the reference model are plotted in Figure 5.5. It is evident◦ that this
Figure 11: Best fit for models with ϕ = 30 .
approach is not suitable for the given case, as the curves diverge to a level that does not allow the
6.3 Trimming andof
generalization Fitting of Data behaviour. Unsatisfactory results are present across all models.
the system
equation (13) it is evident that,
It is evident from investigating the data, that the normal-
izationNormalization of displacements
is not suitable is an top p
additional issue. K0 observations
y
→ β1One
+ β3 of+ the → ∞,made (14)in
by the Rankine pressure in the
p for
and the bottom of the bucket. These variations are con- p
Chapter 4 is that the slopes of lines before the peak show a Rdependency on Kγ depth:a D
− Kγ stiffness increases
sidered to be edge effects and are disregarded. After the
with stress level. It is clear that normalizing withmeaning
data has been trimmed, the results are fitted with a func-
respectthat
to any
β1 andconstant
β3 controlwillthenot not yield
maximum an
relative
acceptable
tion of the type,outcome, i.e. the slopes will vary from one
soil curve to
pressure. another.
The fitting The dependency
function consists of on depth
three terms,
must be introduced. the first two enabling the fitting function to fit both the
p y initial and end slope, while the third term involving K0
= β1 tanh β2 (13)
pR D takes into account the soil pressure at rest at y = 0. The
 y K0 59 trimmed data with the fitted function is shown in figure
+ β3 tanh β4 + p ,
D Kγ − Kγa 10. This procedure is done for all 18 models, and the
fitted functions are gathered - one diagram for ϕ = 30◦ ,
where β2 and β4 are shape coefficients of the fitting func- ϕ = 35◦ and ϕ = 40◦ . These results are shown in figures
tion that controls the climb rate of the function in the 11, 12 and 13. From this initial study of the p − y curves
Chapter 5. Normalization

0.9

0.8

0.7

0.6

0.5

d = 2.3
0.4
d = 3.8
d = 5.2
0.3
d = 6.7
d = 8.3
0.2
d = 9.8
d = 11.3
0.1 d = 12.8

0
0 0.5 1 1.5 2 2.5 3 3.5
10-3

Figure 5.5: Normalized t − z curves with respect to peak shear stress τAPI estimated acc. to API (2011) and
bucket diameter D. Neither the peak shear stresses nor the displacements at which they occur
are aligned.

Two new expressions are required: one for the peak shear stress τ p and one for the displacement
at which it takes place z p , further referred to as peak displacement. Ideally, the curves representing
any depth in any model will have their peaks at coordinates (1,1) in the z/z p − τ/τ p plane.

Keeping in mind the practical aspects of the geotechnical design, τ p and z p ought to be expressed
in terms of readily available parameters related to the system in question. These parameters are
chosen: D, L - bucket diameter and skirt length; σvo
0 - effective overburden vertical stress; ϕ - angle

of internal friction. This selection lays ground for a formulation that combines bucket geometry
and basic soil properties.

Formulation of peak shear stress

The dependency of peak shear stress on the chosen variables is sought through statistical analysis
of FE data from all models. Firstly, the chosen parameters are combined to form a single variable
that has the unit of pressure:

σvo
0  
D F/L2
L tan ϕ

The peak shear stress extracted from each curve in each model is plotted against the established
independent variable, as illustrated in Figure 5.6. Fitting the entire data set with one function would
create an inaccurate model on account of large data spread. One observes the linear trends inclined
at different angles depending on ϕ. Therefore each ϕ-set is fitted with a linear function of the form
f (x) = ax, see Figure 5.7.

60
5.2 Bucket under axial tension

55

50

45

40

35

30

25

20

15

10

5
0 50 100 150 200 250

Figure 5.6: Peak shear stress vs combined parameters. Linear trends at varying slopes are noticed. Legend
key: M X −Y = Model with D = X m and ϕ = Y ◦ .

60

50

40

30

20

10

0
0 50 100 150 200 250

Figure 5.7: Fitting each set of ϕ with a linear function. The lines are forced through the center of the
coordinate system. a - slope of the fitting line.

61
Chapter 5. Normalization

There is no constant term in the fitting function, as the intention is to force the lines to pass
through the origin. This is justified by the physics of the system. Examine Figure 5.8.

Figure 5.8: A y-intercept means that for σvo


0 = 0 kPa there is a positive value of τ , i.e. friction is gained in
p
the absence of normal force on the skirt wall. On the other hand, an x-intercept implies that for
values of x between 0 and the point of intersection with the abscissa, the model predicts τ p < 0
kPa. This leads to the idea that the friction force vector changes to the opposite direction, even
though the mechanics of the system remain the same.

Since the slope of the fitting line seems to depend on the friction angle, a relationship between
these two quantities is sought. See Figure 5.9.

0.3

0.28

0.26

0.24

0.22

0.2

0.18
0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9

Figure 5.9: Relationship between the fitting coefficient a and friction angle ϕ is defined by a linear function.

The peak shear stress as a function of combined parameters is expressed as follows:

62
5.2 Bucket under axial tension

!
σvo
0
τ p = 0.331 tan ϕ D (5.2)
L tan ϕ

Considering that D/L = 1, the equation reduces to just τ p = 0.331 σvo


0 . It turns out that a similar

linear relationship as the one provided by API (2011) is obtained. It has a better effect in terms of
normalization since it is applied on the same models it was derived from. See Figure 5.10.

1.2

0.8

0.6
d = 3.7
d = 4.8
d = 5.9
0.4
d=7
d=8
d = 9.1
0.2 d = 10.2
d = 11.3
d = 12.4
0
0 0.5 1 1.5 2 2.5 3 3.5
-3
10

Figure 5.10: Normalized t − z curves using the found linear expression for peak shear stress. The results are
improved, cf. Figure 5.5.

A different fitting function is used with the purpose of achieving a more accurate expression.
The peak shear stresses are fitted with a power function of the type f (x) = axb . Just as before, sets
belonging to each ϕ are treated independently, as displayed in Figure 5.11

The link of coefficients a and b to the friction angle is defined in Figure 5.12.

The plots where the mathematical model is compared with the best fits and with the FE data are
presented in Appendix D.

63
Chapter 5. Normalization

60

50

40

30

20

10

0
0 50 100 150 200 250

Figure 5.11: Power law is used to fit each ϕ-set. a, b - fitting coefficients of the power function.

0.21 1.095

0.2

1.09
0.19

0.18
1.085
0.17

0.16
1.08
0.15

0.14
1.075

0.13

0.12 1.07
0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9

Figure 5.12: Both coefficients are expressed as quadratic functions of friction angle ϕ.

The new equation for peak shear stress takes the following form:
!(−0.55 tan2 ϕ+0.814 tan ϕ+0.793)
 σvo
0
τ p = 0.381 tan2 ϕ − 0.342 tan ϕ + 0.199 D (5.3)
L tan ϕ

Applying Eq. (5.3) results in a significant improvement, as illustrated in Figure 5.13.

64
5.2 Bucket under axial tension

1.2

0.8

0.6
d = 3.7
d = 4.8
d = 5.9
0.4
d=7
d=8
d = 9.1
0.2 d = 10.2
d = 11.3
d = 12.4
0
0 0.5 1 1.5 2 2.5 3 3.5
10-3

Figure 5.13: Normalized t − z curves using a power function for peak shear stress.

Since the peaks are not aligned around τ/τ p = 1 and the trend related to depth is still present, it
is understood that systematic error persists in the mathematical model. Therefore, it is decided to
improve the model by incorporating the dependency on bucket diameter.

Data from each model is fitted with a power function. Due to the high concentration of data in
one single figure, the plots are separated per friction angle and shown in Appendix D.

The fitting function has the same form f (x) = axb , but the coefficients a and b are defined
as functions of two variables: bucket diameter D and friction angle ϕ. Their found values are
presented in Table 5.1.

ϕ
30◦ 33◦ 35◦ 38◦ 40◦ 30◦ 33◦ 35◦ 38◦ 40◦
D
10 m 0.113 0.115 0.124 0.14 0.15 1.12 1.14 1.141 1.143 1.141
13 m 0.106 0.11 0.12 0.136 0.146 1.127 1.141 1.14 1.141 1.14
15 m 0.101 0.107 0.116 0.134 0.144 1.131 1.143 1.143 1.139 1.139
18 m 0.1 0.106 0.116 0.133 0.143 1.129 1.14 1.139 1.137 1.136
20 m 0.096 0.105 0.116 0.133 0.142 1.134 1.14 1.136 1.133 1.133

Table 5.1: Values of a-coefficients (left section) and b-coefficients (right section).

The values above are used for surface fitting, see Figure 5.14 and Figure 5.15.

65
Chapter 5. Normalization

NRMSE = 3.85 %
0.16

0.14

0.12

0.1

0.08
0.9
0.8 1.4
0.7 1.2
1
0.6
0.8
0.5 0.6

Figure 5.14: Fitting the a-coefficients with a quadric surface. Dre f = 15 m.

NRMSE = 12.24 %
1.145

1.14

1.135

1.13

1.125

1.12
0.9
0.8 1.4
0.7 1.2
1
0.6
0.8
0.5 0.6

Figure 5.15: Fitting the b-coefficients with a quadric surface. Dre f = 15 m.

66
5.2 Bucket under axial tension

The dependencies of both a and b on ϕ and D are modelled with bivariate quadratic functions of
the form f (x, y) = p1 + p2 x + p3 y + p4 xy + p5 y2 . Since there is stronger non-linearity with respect
to ϕ, only its corresponding term is raised to the second degree. The coefficients are calculated
according to Eq. (5.4) and Eq. (5.5).
   
D D
a = α1 + α2 + α3 tan ϕ + α4 tan ϕ + α5 tan2 ϕ (5.4)
Dre f Dre f
   
D D
b = β1 + β2 + β3 tan ϕ + β4 tan ϕ + β5 tan2 ϕ (5.5)
Dre f Dre f

where:

α1 α2 α3 α4 α5
0.183 -0.045 -0.2616 0.0432 0.2721
β1 β2 β3 β4 β5
0.7989 0.0733 0.8381 -0.1079 -0.4947

By including the expressions for coefficients a and b, the peak shear stress can be estimated as
follows:
!{β }{P}T
σ 0
τ p = {α}{P}T D vo (5.6)
L tan ϕ

where:

{α} {α1 α2 α3 α4 α5 }
{β } {β1 β2 β3 β4 β5 }
{P} {1 D/Dre f tan ϕ (D/Dre f ) tan ϕ tan2 ϕ}

The formulation acc. to Eq. (5.6) yields acceptable normalization of peak shear stresses, see
Figure 5.16. The normalized curves for all models can be found in Appendix D.

The next step is seeking an expression for the peak displacements z p , in order to normalize the
data on x-axis.

67
Chapter 5. Normalization

1.2

0.8

0.6

d = 4.8
d = 5.9
0.4
d=7
d=8
d = 9.1
0.2 d = 10.2
d = 11.3
d = 12.4
0
0 0.5 1 1.5 2 2.5 3 3.5
10-3

Figure 5.16: Normalized t − z curves using a power function in which the coefficients depend on bucket
diameter and friction angle.

Formulation of peak displacement

Similarly to peak shear stresses, the link between peak displacements and an independent variable
composed of some basic parameters is investigated. The unit of this variable must be length:

σvo
0
D [L]
σre f

where σre f = 100 kPa.

The FE results from all models are plotted in Figure 5.17. One observes the curving trend,
which leads to the conclusion that fitting with a linear function would create an inaccurate model.
Moreover, treating the data as individual sets of friction angles would not generate acceptable
results either, because the influence of bucket diameter is significant.

Thus data from each model is fitted separately with a power function f (x) = cxd and a study of
its coefficients is carried out in the same manner as the one for peak shear stress. Their values and
surface plots are presented in Appendix D.

The coefficients c and d are calculated with Eq. (5.7) and Eq. (5.8) respectively. Contrary to
peak shear stress formulation, the exponent d is fitted with a first-degree polynomial. Since the use
of higher order polynomial functions does not minimize the regression residuals (they become even
higher!), the simpler relation is preferred.

68
5.2 Bucket under axial tension

0.05

0.045

0.04
M 10-30
0.035 M 10-33
M 10-35
M 10-38
M 10-40
0.03 M 13-30
M 13-33
M 13-35
0.025 M 13-38
M 13-40
M 15-30
M 15-33
0.02 M 15-35
M 15-38
M 15-40

0.015 M 18-30
M 18-33
M 18-35
M 18-38
0.01 M 18-40
M 20-30
M 20-33
M 20-35
0.005 M 20-38
M 20-40

0
0 5 10 15 20 25 30 35

Figure 5.17: Peak displacement vs combined parameters. The relations appear to adopt a hyperbolic trend.
There is less linearity and stronger dependency on bucket diameter in comparison to peak shear
stress, cf. Figure 5.6. Legend key: M X −Y = Model with D = X m and ϕ = Y ◦ .

   
D D
c = γ1 + γ2 + γ3 tan ϕ + γ4 tan ϕ + γ5 tan2 ϕ (5.7)
Dre f Dre f
 
D
d = δ1 + δ2 + δ3 tan ϕ (5.8)
Dre f

where:

γ1 γ2 γ3 γ4 γ5
0.0833 0.016 -0.2243 -0.0136 0.1512
δ1 δ2 δ3
0.7019 -0.0067 -0.5357

The equation for peak displacements is established:


 {δ }{Pδ }T
T σvo
0
z p = {γ}{Pγ } D (5.9)
σre f

where:

69
Chapter 5. Normalization

{γ} {γ1 γ2 γ3 γ4 γ5 }
{δ } {δ1 δ2 δ3 }
{Pγ } {1 D/Dre f tan ϕ (D/Dre f ) tan ϕ tan2 ϕ}
{Pδ } {1 D/Dre f tan ϕ}

Satisfactory normalization of data on x-axis is achieved by applying Eq. (5.9), see Figure 5.18.

1.2

0.8

0.6

d = 4.8
d = 5.9
0.4
d=7
d=8
d = 9.1
0.2 d = 10.2
d = 11.3
d = 12.4
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

Figure 5.18: Normalized t − z curves using formulations for τ p and z p in terms of basic parameters. All the
peaks are concentrated around (1,1).

Comparison plots between FE data, best fits and mathematical model, together with normalized
curves for all models can be found in Appendix D.

5.2.2 Partially drained models

As seen Section 4.5.1, where the results related to a bucket is subjected to tension in partially
drained conditions are discussed, the post-peak curves to not become constant, but rather continue
with a slight increase. This means maximum friction is achieved after the "peak" occurs. For the
sake of consistency of terminology, the "peak" is redefined as the point of largest curvature.

Peak displacement

In Figure 5.19 it is noticeable that models with ϕ = 30◦ deviate from the other ones. Moreover,
they contain outliers that introduce strong bias in the mathematical model. The respective ϕ-set is
discarded from the analysis.

70
5.2 Bucket under axial tension

0.05

0.045

0.04
M 10-30
0.035 M 10-33
M 10-35
M 10-38
M 10-40
0.03 M 13-30
M 13-33
M 13-35
0.025 M 13-38
M 13-40
M 15-30
M 15-33
0.02 M 15-35
M 15-38
M 15-40

0.015 M 18-30
M 18-33
M 18-35
M 18-38
0.01 M 18-40
M 20-30
M 20-33
M 20-35
0.005 M 20-38
M 20-40

0
0 5 10 15 20 25 30 35

Figure 5.19: Peak displacement vs combined parameters. Dependency on friction angle as well as bucket
diameter is observed.

Each FE model is fitted individually with a power function f (x) = cxd and the analysis of c-
and d-coefficients is performed similarly to previous section. See Appendix D for the details of the
analysis in question.

One major difference from peak displacement formulation in drained conditions is that the
exponent d is taken as the average of all models. The outcome of the regression analysis is reflected
in Eq. (5.10) and Eq. (5.11).

   2  
D D D tan ϕ
c = 0.0048 − 0.0037 + 0.0099 tan ϕ + 0.0096 − 0.0204
Dre f Dre f Dre f
(5.10)

d = 0.5913 (5.11)

NOTE! It was not possible to observe clear trends of peak displacement based on the first
batch of FE models. The issue was the size of the load step PLAXIS applied in the calculation.
Within the range of bucket displacements where the peaks occurred, the intervals between
data points were too large and thus did not represent the development around the peaks clearly
enough. Consequently, the peak displacements were of equal magnitude for 3-4 neighbouring
layers. As a solution, all FE models were run again while forcing PLAXIS to apply more
calculation steps by lowering the "Max load fraction per step" parameter from the default of
0.5 to a range between 0.05 and 0.02.

71
Chapter 5. Normalization

Peak shear stress

The same combination of basic parameters as in Section 5.2.1 is used. The peaks shear stresses are
plotted against the independent variable in Figure 5.20.

40

35

30

25

20

15

10

0
20 40 60 80 100 120 140 160 180 200 220

Figure 5.20: Peak shear stress vs combined parameters. Dependency on friction angle is present, while the
bucket diameter does not have significant influence. The value represented by the outlying red
circle is discarded from the regression analysis.

An expression for peak stress f (x) = axb is sought in an identical manner to drained models.
The analysis of fitting coefficients may be found in Appendix D. The coefficients a and b are
established using the same function type as seen in Eq. (5.4) and Eq. (5.5). Quantities α and β are
found:

α1 α2 α3 α4 α5
-0.4675 -0.0753 1.3639 0.0806 -0.8737
β1 β2 β3 β4 β5
2.9391 0.2263 -4.854 -0.2933 3.3032

Using the formulation for peak shear stress and peak displacement, an example of normalized
curves is given in Figure 5.21.

It is apparent that in context of partial drainage, the shear stress development after the peak
undergoes significant changes if compared to the drained case. The stress decrease/increase and the
curving trends are effects of excess pore pressure arising during undrained loading. With the taken
approach, the post-peak part cannot be normalized so that the curves align, but this behaviour is
accounted for in t − z curves formulation, as described in Chapter 6.

72
5.3 Bucket under axial compression

1.2

0.8

0.6

d = 3.1
d = 4.4
0.4
d = 5.7
d = 6.9
d = 8.2
0.2 d = 9.4
d = 10.6
d = 11.9
0
0 1 2 3 4 5 6

Figure 5.21: Normalized t − z curves using the established formulations for τ p and z p . The pre-peak curves
are fully aligned, while after the peak the behaviour varies depending on depth.

The trimmed and normalized t − z curves for all models are located in Appendix D.

5.3 Bucket under axial compression

5.3.1 Drained models

The development of friction on the skirt during compression is different than during tension, as
discussed in Chapter 4. The main reason is the lid action that gradually increases the normal force
on the skirt wall. In Figure 5.22 it is observed that no peaks occur, even though the bucket is
displaced 0.3 m downwards. The notion of "peak shear stress" is still used to mark the point of
maximum curvature.

NOTE! The first series of simulations involved bucket displacements until 0.05 m. Since the
peaks were not captured, another batch of models was created and the displacements were
set to 0.3 m. This proved to only elongate the curves, without reaching any point where the
shear stresses would start to decline. Displacing the bucket even further resulted in calculation
failures, so it was decided to continue with the existing set of models. It is acknowledged
that by obtaining the peaks, more information can be gathered about the system and a better
mathematical model can be created. In the current state, only the initial response is described.

73
Chapter 5. Normalization

120
d = 4.2
d = 5.2
d = 6.1
100 d=7
d = 7.9
d = 8.9
d = 9.8
d = 10.8
80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3

Figure 5.22: Example of trimmed t − z curves for a model in drained compression. The maximum curvature
takes place within a few load steps, afterwards the stresses increase linearly.

The stresses at maximum curvature are plotted against the combined basic parameters to
discover any trends. See Figure 5.23.

60

50

40

30

20

10

0
20 40 60 80 100 120 140 160 180 200 220

Figure 5.23: The data spread is considerably larger than in FE models subjected to tension, cf. Figure 5.6
and Figure 5.20. There is strong dependency on diameter and friction angle at the same time.
The data sets cannot be grouped per ϕ.

It is clear that a power law has to be applied while fitting each FE model, just as shown in

74
5.3 Bucket under axial compression

previous sections. Appendix D gives details on the analysis, including information on coefficients
 0 b
a and b in τ p = a σvo / DL tan ϕ .

An example of normalized curves is seen in Figure 5.24. A formulation for "peak" displacements
is not required, as it is sufficient to use the bucket diameter.

4.5

3.5

2.5

2
d = 4.2
d = 5.2
1.5
d = 6.1
d=7
1 d = 7.9
d = 8.9
0.5 d = 9.8
d = 10.8
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

Figure 5.24: Normalized t − z curves for drained compression. The stresses are normalized with respect to
the stress value at maximum curvature, while displacements with bucket diameter D. After a
certain point, the curves begin to diverge as their slopes vary.

5.3.2 Partially drained models

The applied normalization procedure consists of the same steps as taken for drained compression
models. See Appendix D for details. Generally, the amount of discarded layers in this batch of
models is very high. On the basis of large complex mechanisms triggered by the groundwater
flow, about 60 % of data from each model are not used in the analysis. Moreover, the FE models
with ϕ = 30◦ are ignored, since they do not follow the trend set by the rest of the models. Their
inclusion leads to a strong bias that decreases the descriptive abilities of the mathematical model.

Figure 5.25 illustrates an example of normalized curves. It is evident that the curves are aligned
only around the initial part and then continue with a hyperbolic trend, the amplitude of which
depends on the vertical stress. Similarly to the case of undrained tension, excess pore water pressure
introduces non-linearity in the shear stress development along the skirt walls. This phenomenon is
treated further in Chapter 6.

75
Chapter 5. Normalization

1.6

1.4

1.2

0.8
d = 3.4
d = 4.1
0.6
d = 4.9
d = 5.6
d = 6.4
0.4
d = 7.1
d = 7.9
0.2
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

Figure 5.25: Normalized t − z curves for partially drained compression. The stresses are normalized with
respect to the stress value at maximum curvature, while displacements with bucket diameter D.
Upon reaching τ/τ p = 1, the curves start arching with varying radii.

5.4 Summary

The normalization of shear stresses in every model requires a peak shear stress formulation based
on statistical analysis. Additionally, expressions for peak displacements are found for models
involving tension. The normalization of displacements with bucket diameter is acceptable for
compression models.

Power functions are used in all models to fit FE data. Studies of fitting coefficients are carried
out, so that peak shear stresses and peak displacements can be expressed in terms of basic parameters
D, L, σvo0 and ϕ.

Best results are achieved in drained tension models, as the system behaviour is relatively
elementary. Respectively, the mathematical formulation of t − z curves is the simplest when
compared to other cases, see Chapter 6. For partially drained tension, the post-peak shear stresses
depend on depth, therefore a more complex mathematical model is required to account for this
behaviour.

Compression models are similar in that the t − z curves have a hyperbolic trend. The normaliz-
ation of their FE results is somewhat less reliable, because the peaks are not captured given the
prescribed bucket displacements, and thus the "peak shear stress" is redefined as the "stress at
maximum curvature". The location of largest curvature is unstable due to the distance between
discrete data points. This hinders the establishment of clear trends, hence the reduced reliability of
the relations found with regression analysis.

76
C HAPTER 6

Formulation of t-z curves

Following the normalization of FE data, the formulation of t − z curves is undertaken in this


chapter. The establishment of mathematical models that predict the skin friction along the
skirt wall is carried out in different manners depending on drainage conditions and loading
direction. The concept remains the same for all models: the behaviour of each t − z spring is
dictated by a combination of vertical stress, bucket geometry and friction angle.

6.1 Drained tension

The general concept of how shear stresses develop with vertical displacements during drained
tension is illustrated in Figure 6.1.

Figure 6.1: The shear stress increases linearly until the peak stress is reached. Then it becomes constant.

The expressions for peak shear stress τ p and peak displacement z p derived in Section 5.2.1 are
sufficient to establish the general mathematical model, which consists of two parts:

77
Chapter 6. Formulation of t-z curves

(
z z
τ zp if zp <1
= (6.1)
τp 1 if z
zp >1

The model implies direct proportionality between stresses and displacements until the peak
is reached, afterwards stresses remain unchanged irrespective of any given displacements. The
equations for τ p and z p , together with comparisons between the normalized curves and the ones
predicted by the mathematical model, are presented in Appendix E. An example is given in
Figure 6.2.

1.2

0.8

0.6

d = 4.8, Normalized Data


d = 5.9, Normalized Data
0.4 d = 7, Normalized Data
d = 8, Normalized Data
d = 9.1, Normalized Data
d = 10.2, Normalized Data
0.2 d = 11.3, Normalized Data
d = 12.4, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

Figure 6.2: Comparison plot between the normalized t − z curves and the response predicted with the
mathematical model. Overall there is good correlation, although the model can be improved to
take into account the slightly non-linear behaviour before the peak is almost reached and the
difference between peak and residual stress given a small vertical overburden stress.

6.2 Partially drained tension

In Chapter 5 it was observed that the normalized post-peak t − z curves in undrained models do not
fully align as in drained models. Their non-linear behaviour is characterized by convex curves, see
Figure 6.3.

In order to account for this phenomenon, the general concept of t − z curves for drained tension
is modified as shown in Figure 6.4.

78
6.2 Partially drained tension

1.4

1.2

0.8

0.6
d = 3.1
d = 4.4
d = 5.7
0.4
d = 6.9
d = 8.2
d = 9.4
0.2
d = 10.6
d = 11.9
0
0 1 2 3 4 5 6 7

Figure 6.3: Example of normalized t − z curves for partially drained tension. The shape of the curves after
the peak resemble a quadratic function. The arc radii seem to depend on vertical stress.

Figure 6.4: Modified t − z concept to incorporate non-linear stress-dependent behaviour after reaching the
peak. The pre-peak part remains unchanged.

The first attempt at modelling the post-peak part involves fitting each curve in every model with
a polynomial function of the form f (x) = p2 x2 + p1 x + p0 . Then individual studies of p2 , p1 and
p0 are conducted to express these parameters as functions of σvo 0 . As an example, Figure 6.5 shows

the distribution of p2 .

79
Chapter 6. Formulation of t-z curves

0.04

0.03

0.02

0.01

-0.01

-0.02 p2 values
linear fit
-0.03
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

Figure 6.5: Values of p2 coefficients when the post-peak curves are fitted with f (x) = p2 x2 + p1 x + p0 .
There is hardly any clear trend when data from all models is aggregated. In this case a linear fit
is the only option.

Since coefficients p1 and p0 do not exhibit any better trends and are fitted by linear regression,
the outcoming mathematical model can mostly predict the average post-peak behaviour, see
Figure 6.6.

1.5

d = 2.7, Normalized Data


d = 3.8, Normalized Data
d = 4.9, Normalized Data
d = 5.9, Normalized Data
d = 7, Normalized Data
d = 8.1, Normalized Data
d = 9.2, Normalized Data
0.5 d = 10.3, Normalized Data
d = 2.7, Mathematical Model
d = 3.8, Mathematical Model
d = 4.9, Mathematical Model
d = 5.9, Mathematical Model
d = 7, Mathematical Model
d = 8.1, Mathematical Model
d = 9.2, Mathematical Model
d = 10.3, Mathematical Model

0
0 1 2 3 4 5 6 7 8

Figure 6.6: The shear stresses until just after the peak are represented accurately, however the deviation of
curves is not modelled appropriately.

80
6.2 Partially drained tension

An issue with the current model is that given larger displacement, the peak shear stress is
overestimated due to the growth rate of the quadratic polynomial. To solve the problem, another
fitting function is chosen: f (x) = p2 /x2 + p1 x + p0 . At the same time its coefficients are defined in
terms of σvo0 , D and ϕ. The details of this study are given in Appendix E.

The mathematical model for partially drained tension is defined as follows:


z z

 if <1
τ  zp zp
z
= 1 if zp =1 (6.2)
τp     
p / z 2 + p z + p

if z
>1
2 zp 1 zp 0 zp

The new fitting function combined with the study of p-coefficients yields improved results, see
Figure 6.7.

1.4

1.2

0.8 d = 4.7, Normalized Data


d = 5.4, Normalized Data
d = 6.1, Normalized Data
d = 6.8, Normalized Data
0.6 d = 7.6, Normalized Data
d = 8.3, Normalized Data
d = 9, Normalized Data
d = 9.8, Normalized Data
d = 4.7, Mathematical Model
0.4 d = 5.4, Mathematical Model
d = 6.1, Mathematical Model
d = 6.8, Mathematical Model
d = 7.6, Mathematical Model
0.2 d = 8.3, Mathematical Model
d = 9, Mathematical Model
d = 9.8, Mathematical Model

0
0 1 2 3 4 5 6

Figure 6.7: The new mathematical model takes into account the development of post-peak shear stresses
depending on vertical stress, bucket diameter and friction angle.

In the example given above there is good agreement between normalized FE data and mathem-
atical model, however a drawback of the given t − z formulation is evident by comparing other set
of data, such as presented in Figure 6.8.

Since the expressions for τ p and z p are based on results from a statistical analysis and already
contain certain amount of error, the mathematical model further adds to it. This makes predictions
for some combinations of parameters erroneous.

At this point one has to choose between using a more elementary model such as the one
described at the beginning of the section, or improve the current model by at least smoothing
the peaks. Alternatively a different fitting function may be applied. Indeed, using an exponential
function of the type f (x) = a exp (bx)+c exp (dx) proves to be more suitable for fitting the post-peak

81
Chapter 6. Formulation of t-z curves

1.8

1.6

1.4

1.2

d = 3.6, Normalized Data


1 d = 4.1, Normalized Data
d = 4.6, Normalized Data
d = 5.3, Normalized Data
0.8 d = 5.9, Normalized Data
d = 6.4, Normalized Data
d = 6.9, Normalized Data
0.6 d = 7.5, Normalized Data
d = 3.6, Mathematical Model
d = 4.1, Mathematical Model
0.4 d = 4.6, Mathematical Model
d = 5.3, Mathematical Model
d = 5.9, Mathematical Model
d = 6.4, Mathematical Model
0.2 d = 6.9, Mathematical Model
d = 7.5, Mathematical Model

0
0 1 2 3 4 5 6 7

Figure 6.8: The sharp peaks at z/z p = 1 mean that the peak shear stress is overestimated at that point. The
effect is stronger at smaller depths. Higher friction is predicted overall given this combination
of parameters.

parts of the normalized t − z curves in the given range. However, upon extrapolating by increasing
the displacements, the shear stresses increase exponentially, thereby adding an unrealistic amount
of stress between soil and structure.

6.3 Drained compression

When the bucket is loaded vertically in compression, no sudden stress peaks appear. This behaviour
is described in previous chapters and is exemplified in Figure 5.25. In the context of formulating
a mathematical model, this means that the curves need not be split into distinct parts, as in the
case of tensile loading. Furthermore, since the ultimate shear stress is not reached with the current
FE models (see the note in Section 5.3.1), the t − z curves can be formulated with monotonically
increasing functions.

Østergaard et al. (2015) fitted the normalized p − y curves with a function of the type f (x) =
A1 tanh (B1 x) + A2 tanh (B2 x) +C. This expression is applied in the current study, but with omission
of the free term C, since zero mobilized friction is modelled at zero vertical displacement. The
mathematical model adopts the general form:
τ  z  z
= A1 tanh B1 + A2 tanh B2 (6.3)
τp D D

REMINDER! For compression models, the "peak" shear stress τ p is in fact the stress at
maximum curvature.

82
6.4 Partially drained compression

Using Eq. (6.3) to fit the data proves to be a suitable approach, as the curves are followed
closely, see Figure 6.9.

6
Normalized Data
Best fit
5

0
0 0.005 0.01 0.015

Figure 6.9: Normalized t − z curves using a tanh-function. There is good agreement between the observed
values and the best fit, especially around the initial part. The deviation of tail values depending
on the stress level is not taken into account, thus the mean behaviour is modelled.

A study of coefficients with the purpose of examining their dependency on D and ϕ reveals
that they have high variance and only a mean trend can be described. It turns out this affects the
mathematical model to a degree that it cannot describe the system behaviour. In extreme cases
it predicts negative shear stresses or values in the order of 5 times higher than the observed ones.
Therefore it is decided to use another fitting function:
τ  z B
=A (6.4)
τp D

The power function does not capture the shear stress development as accurately as the tanh-
function, however, based on its simplicity and lower coefficient variance the mathematical model is
stable and can predict the average behaviour, as shown in Figure 6.10.

Details of the coefficient investigation and comparison plots between the mathematical model
and normalized data for all models can be found in Appendix E.

6.4 Partially drained compression

The similarities between drained and partially drained compression are outlined in the previous
chapters. Considering that normalization is performed in the same way, it follows naturally that
establishing the mathematical model comprises identical steps. Respectively, the fitting function
shown by Eq. (6.3) is used as first trial.

83
Chapter 6. Formulation of t-z curves

6
Normalized Data
Power fit
5 Mathematical Model: t-z formulation

0
0 0.005 0.01 0.015

Figure 6.10: The mathematical model follows the best fit closely. The model can be improved by including
the slopes of individual curves as functions of vertical stress.

The outcome of fitting the data with the tanh-function is similar: an inaccurate mathematical
model resulting from high variance of A1 , A2 , B1 and B2 coefficients. An example of how the
mathematical model performs is given in Figure 6.11.

3.5

2.5

1.5

Normalized Data
0.5 tanh - fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015 0.02 0.025

Figure 6.11: The mathematical model overestimates the shear stresses when the tanh-function is used for
fitting, even though the fit itself agrees with FE data at an acceptable level.

It is decided to recur to fitting with the power function, see Eq. (6.4), thus enabling the model
to represent roughly the mean behaviour. Appendix E presents information related to the study of
coefficients and mathematical model for partially drained compression.

84
C HAPTER 7

Verification and load-displacement curves

The mathematical formulations of t − z curves are used for modelling the total frictional
response of bucket foundations. The performance of mathematical models is observed by
comparing the obtained force-displacement curves with FE data. Verification is carried out
based on results from physical modelling.

7.1 Comparison with numerical results from the current study

104
-4
Mathematical model
FE results D = 20 m
-3.5

-3

-2.5 D = 18 m

-2

-1.5
D = 15 m

-1
D = 13 m

-0.5
D = 10 m

0
0 -0.005 -0.01 -0.015 -0.02 -0.025 -0.03 -0.035 -0.04 -0.045 -0.05

Figure 7.1: The ultimate friction force is predicted with high accuracy by mathematical models. There are
certain discrepancies before peak friction is reached.

The friction force is computed directly from FE models by aggregating the effect of every
interface stress point along the skirt at every load step. In this process no stress points are
disregarded, contrary to the approach taken when deriving expressions for t − z curves. At the

85
Chapter 7. Verification and load-displacement curves

same time, the load due to friction is also estimated with the established mathematical models. The
soil-structure interaction is modelled by discretizing the skirt into springs, the properties of which
depend on vertical stress, bucket diameter/skirt length and friction angle. Firstly, comparison is
made for drained tension case, see Figure 7.1.

Confidence is gained in relation to drained tension formulation as it agrees closely with the FE
results. The mathematical models predict lower stiffness because they imply an idealized pre-peak
state where there is direct proportionality between shear stresses and displacements. On the other
hand, the situation is different for undrained tension, as shown in Figure 7.2.

104
-3
D = 10, Mat. mod.
D = 10, FE
D = 13, Mat. mod.
-2.5 D = 13, FE
D = 15, Mat. mod.
D = 15, FE
D = 18, Mat. mod.
-2 D = 18, FE
D = 20, Mat. mod.
D = 20, FE

-1.5

-1

-0.5

0
0 -0.005 -0.01 -0.015 -0.02 -0.025 -0.03 -0.035 -0.04 -0.045 -0.05

Figure 7.2: The disagreement between mathematical models and FE results is evident. The predicted friction
force magnitudes are about 50% lower and their continuous growth is barely taken into account.
The only common trait is the displacement at which the initial slopes start curving.

The inability of the undrained formulation to capture the frictional response accurately may
be explained by two factors. First of all, it is the non-physical stress discontinuities that the
mathematical model creates around peak displacement given a certain combination of D and ϕ,
particularly for small diameters and large friction angles.

The other reason may be found in the amount of discarded data due to edge effects. On average,
only ca. 45% of data from interface points is used to obtain normalized curves that form the basis
for the mathematical model. It is apparent that assigning characteristic behaviour from the middle
of the skirt to its top and bottom regions is a flawed approach. Figure 7.3 illustrates the same
comparison plot, but the FE friction load is calculated from the skirt region used for mathematical
formulation.

Comparison plots for drained and undrained compression are presented in Figure 7.4 and
Figure 7.5 respectively. Agreement with FE data is achieved in loose terms. While for drained

86
7.1 Comparison with numerical results from the current study

-16000
D = 10, Mat. mod.
D = 10, FE
-14000 D = 13, Mat. mod.
D = 13, FE
D = 15, Mat. mod.
-12000 D = 15, FE
D = 18, Mat. mod.
D = 18, FE
-10000 D = 20, Mat. mod.
D = 20, FE

-8000

-6000

-4000

-2000

0
0 -0.005 -0.01 -0.015 -0.02 -0.025 -0.03 -0.035 -0.04 -0.045 -0.05

Figure 7.3: Considering only the stress points around the middle of the skirt, the FE friction load is closer to
the one predicted by the mathematical models, cf. Figure 7.2.

models the use of a power law seems to be suitable, in the undrained case it severely misrepresents
the initial slopes. This cannot be acceptable, since the initial stiffness is critically important for
the foundation design. It is clear that the t − z formulation must be improved by either redefining
the fitting function type or by splitting the behaviour into pre-peak and post-peak parts in a similar
manner as for the tension cases.

The load-displacement curves are plotted together in Figure 7.6. It is emphasized that only
friction is considered. The initial stiffness during undrained loading is smaller compared to the
drained one. With the current t − z formulation this phenomenon is represented correctly only for
tensile loading.

During undrained tension, less friction is mobilized on the inner side of the skirt, where the
upward groundwater flow and the movement of the soil plug together with the bucket cause vertical
stress relief. This is not the case for drained tension, where inner and outer friction contribute to the
total load in almost equal measures.

When the bucket is compressed, the continuous stress increase on the inner side due to lid
action gradually enables more friction between soil and structure. This effect is stronger for the
drained case.

Since only the formulation for drained tension is deemed successful in agreeing with FE data,
it is further compared with other studies.

87
Chapter 7. Verification and load-displacement curves

104
12
D = 10, Mat. mod.
D = 10, FE
D = 13, Mat. mod.
10 D = 13, FE
D = 15, Mat. mod.
D = 15, FE
D = 18, Mat. mod.
8 D = 18, FE
D = 20, Mat. mod.
D = 20, FE

0
0 0.05 0.1 0.15 0.2 0.25 0.3

Figure 7.4: Overall friction is slightly overpredicted with the current t − z formulation. The initial slopes in
D = 18m and D = 20m are not captured as accurately as in other FE models. The mean trend is
nonetheless represented correctly.

104
4.5
D = 10, Mat. mod.
D = 10, FE
4
D = 13, Mat. mod.
D = 13, FE
3.5 D = 15, Mat. mod.
D = 15, FE
D = 18, Mat. mod.
3 D = 18, FE
D = 20, Mat. mod.
D = 20, FE
2.5

1.5

0.5

0
0 0.05 0.1 0.15 0.2 0.25 0.3

Figure 7.5: The power function is not suitable for representation of undrained compressive behaviour, as the
"best fit" cannot account for the initial and tail values at the same time.

88
7.2 Verification with other numerical studies

104
5
Drained compression - Mat. mod.
Undrained compression - Mat. mod.
4 Drained tension - Mat. mod.
Undrained tension - Mat. mod.
Drained compression - FE
Undrained compression - FE
3 Drained tension - FE
Undrained tension - FE

-1

-2
-0.1 -0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.08 0.1

Figure 7.6: Load due to friction vs vertical displacement for all 4 cases. The drained conditions are better
represented by mathematical models than the undrained ones.

7.2 Verification with other numerical studies

The t − z formulation for drained tension is compared with the numerical results obtained by
Thieken et al. (2014). The input parameters in the mathematical model are selected in accordance
with the ones used for numerical simulations described in the cited article. See Figure 7.7. Very
close agreement is evident in terms of initial stiffness as well as ultimate friction.

Another comparison is made with Shen et al. (2017), as shown in Figure 7.8. The mathematical
model does not predict the frictional response adequately in this case. A simulation with PLAXIS
2D is performed with the purpose of determining whether the reason is a shortcoming of the t − z
formulation or differences in terms of how numerical simulations are approached. It turns out that
PLAXIS 2D can generate comparable results, therefore the issue lies in the mathematical model.

A limitation of the current t − z expression is revealed. Soil-structure interaction is not mod-


elled accurately for D/L 6= 1. In Chapter 5 the peak shear stress τ p is defined as a function of
σvo
0 /(D/L) tan ϕ. If D/L = 2 as in the case of Shen et al. (2017), the value of the variable is halved,

thus the predicted peak shear stress is reduced considerably. The effect becomes more significant
with increasing ϕ and σvo0 . The mathematical model is overly sensitive to D/L ratio and without a

further parametric study its performance is bound by D/L = 1.

89
Chapter 7. Verification and load-displacement curves

-5500

-5000

-4500

-4000

-3500

-3000
Thieken et al. (2014)
t-z formulation
-2500

-2000

-1500

-1000

-500

0
0 -0.02 -0.04 -0.06 -0.08 -0.1 -0.12 -0.14 -0.16 -0.18 -0.2

Figure 7.7: There is an almost ideal match between the two data sets. In the compared numerical study,
D and L were assumed as 10 m and a very dense sand was modelled. These quantities fall
within the range used for the definition of the mathematical model, which may explain the good
correlation.

-900

-800

-700

-600

-500 Shen et al. (2017)


t-z formulation
-400 PLAXIS 2D

-300

-200

-100

0
0 -0.005 -0.01 -0.015 -0.02 -0.025 -0.03 -0.035 -0.04 -0.045 -0.05

Figure 7.8: No agreement is found between the predicted response and numerical results from Shen et al.
(2017). The simulation with PLAXIS 2D shows similar initial stiffness, while the post-peak
offset may be explained by the use of a different soil model and input parameters.

90
7.3 Verification with a physical model

7.3 Verification with a physical model

A series of tests on a steel bucket with D = 1 m and L = 0.5 m subjected to drained monotonic
tensile loading was carried out at Aalborg University by Vaitkunaite et al. (2016). The physical
model comprised a circular steel box with 1.5 m in height and 2.5 m in diameter. The bucket
was installed in Aalborg sand no.1. Additional confining pressure was obtained by increasing the
vertical overburden pressure. Tests were performed with an applied pressure of 20, 40 and 70 kPa.
Figure 7.9 displays the results of the tests compared to the established mathematical model.

-100

-90

-80

-70
Test 0 kPa
-60 Test 20 kPa
Test 40 kPa
-50 Test 70 kPa
t-z formulation 0 kPa
-40 t-z formulation 20 kPa
t-z formulation 40 kPa
-30 t-z formulation 70 kPa

-20

-10

0 -0.01 -0.02 -0.03 -0.04 -0.05 -0.06 -0.07

Figure 7.9: The mathematical model does not match the test results. The difference is by a factor of app. 4.
However the model seems to be able to predict the peak displacements.

One of the reasons for the mismatch could be the D/L ratio mentioned in the previous section.
Besides this, the dimensions of the physical model lie beyond the range of the ones considered in the
numerical study. In the course of statistical analysis on parameters, the choices of fitting functions
were based on their descriptive abilities within the given range (high R-square and low error). In
many cases such functions were second degree polynomials. This means that extrapolation is
questionable on account of possible overfitting of data. Comparison with larger physical models
would shed more light on this issue.

91
C HAPTER 8

Conclusion and future work

The frictional response of suction buckets subjected to pure vertical loading was investigated
by means of finite element analysis with PLAXIS 2D. The numerical simulation procedures
were enhanced by scripting in PYTHON programming language. The assumed soil type was
Frederikshavn sand. Its geotechnical parameters were estimated from empirical relations and
laboratory test data. Hardening Soil with small strain stiffness was applied to model realistic soil
behaviour, especially at the initial loading stages.

A total amount of 100 numerical models were created, with bucket diameters ranging from 10
to 20 meters and sand compaction state from loose to dense. Tensile and compressive loading types
were assessed in drained and partially drained conditions.

Formulations of peak shear stress at soil-structure contact regions and vertical bucket displace-
ment at which it occurs were based on regression analysis. Statistical studies of observed stress
and displacement values, as well as the fitting function parameters, resulted in expressions of the
investigated quantities in terms of chosen basic parameters: bucket diameter, skirt length, vertical
overburden pressure and angle of internal friction. The obtained expressions were used to normalize
the t − z curves.

In accordance with normalized data trends, mathematical models for each combination of
loading type and drainage condition were established. Force-displacement curves were created by
discretizing the bucket skirt into springs and aggregating their individual contributions per vertical
displacement increment. The properties of each spring vary with given combinations of bucket
geometry, soil properties, depth below mud line, loading and drainage conditions.

Verification with numerical models revealed that drained formulations generally agree with FE
results. With the approaches adopted in the current study, the mathematical models proved to be
limited in their ability to account for higher non-linearity induced by excess pore water pressure
during undrained loading. Furthermore, using parameter values outside the studied range could
lead to considerable errors, as understood from a comparison with a physical model.

Generally the outcome of the thesis represents the starting step towards potentially more detailed
and generalized formulations of t − z curves for suction buckets. Readily available and accurate
expressions would bring advantages in terms of geotechnical design, and as a consequence, increase
the cost efficiency of offshore wind power.

93
Chapter 8. Conclusion and future work

Future work

A list of possible ways to improve/expand the current research is given:

• The influence of other soil parameters could be assessed and implemented in the t − z
formulation. However the number of additional factors should to be kept to a reasonable
minimum, since each extra piece of information adds layers of complexity that might increase
the error and undermine the practicality of a simple formulation.
• Numerical simulations of a wider range of bucket dimensions, including various diamet-
er/skirt length ratios, would give more insight and provide a larger amount of data points for
statistical analysis. Thus the versatility of mathematical models would be increased.
• Similar studies could be performed with respect to other soil types, such as clay or silt.
• The calibration of fitting parameters with laboratory or field tests would add to the reliability
of the t − z formulation. This is especially important since some essential parameters such as
interface friction angle were defined based on general experience.
• The total axial response of bucket foundations cannot be described solely by friction along the
skirt faces, except during drained tension. In the context of the Winkler method, additional
springs would be required to model the lid and skirt toe contributions to the overall stiffness
and bearing capacity of the system.

94
Bibliography

Achmus, M. & Gütz, P. (2016). Numerical modeling of the behavior of bucket foundations in
sand under cyclic tensile loading. In Insights and Innovations in Structural Engineering,
Mechanics and Computation (pp. 2085–2091). Taylor & Francis Group, London.
Andersen, A. T., Madsen, E. B. & Schaarup-Jensen, A. L. (1998). Eastern Scheldt Sand, Baskarp
sand no. 15: Part 1 (Data Report No. 9701). Aalborg: Geotechnical Engineering Group.
API. (2011). Geotechnical and Foundation Design Considerations: ANSI/API Recommended
Practice 2GEO. American Petroleum Institute.
Barari, A. & Ibsen, L. B. (2012). Undrained response of bucket foundations to moment loading.
Applied Ocean Research, (36), 12–21.
Bolton, M. D. (1986). The strength and dilatancy of sands. Géotechnique, (36), 65–78.
Brinkgreve, R. B. J., Kumarswamy, S. & Swolfs, W. M. (2017). Plaxis 2D 2017 Manual. PLAXIS
bv. Delft, Netherlands.
Cook, R. D., Malkus, D. S., Plesha, M. E. & Witt, R. J. (2002). Concepts and Applications of Finite
Element Analysis (4th ed.). John Wiley & Sons, Inc.
DNV. (1992). Foundations. Classification Notes no. 30.4. Oslo, Norway: Det Norske Veritas.
Feld, T. (2001). Suction buckets: A new innovative concept, applied to offshore wind turbines (PhD
Thesis, Aalborg University, Aalborg, Denmark).
Hardin, B. O. & Black, W. L. (1969). Closure to vibration modulus of normally consolidated clays.
In Journal of the Soil Mechanics and Foundations Division (pp. 1531–1537). ASCE 95:6.
Hedegaard, J. & Borup, M. (1993). Klassifikationsforsøg med Baskarp sand no. 15. Department of
Civil Engineering, Aalborg University. Aalborg, Denmark.
Houlsby, G. T., Ibsen, L. B. & Byrne, B. W. (2005). Suction caissons for wind turbines. In Frontiers
in Offshore Geotechnics I: Proceedings of the First International Symposium in Frontiers
in Offshore Geotechnics (ISFOG), Perth, Australia (pp. 75–93). Taylor & Francis Group,
London.
Ibsen, L. B., Hanson, M., Hjort, T. & Thaarup, M. (2009). MC-parameter calibration of Baskarp
sand no. 15 (DCE Technical Reports No. 62). Department of Civil Engineering, Aalborg
University. Aalborg, Denmark.
Nielsen, S. D., Shajarati, A., Sørensen, K. W. & Ibsen, L. B. (2012). Behaviour of dense Fre-
derikshavn sand during cyclic loading (DCE Technical Memorandum No. 15). Department
of Civil Engineering, Aalborg University. Aalborg, Denmark.
Østergaard, M. U., Knudsen, B. S. & Ibsen, L. B. (2015). P-y curves for bucket foundations in sand
using finite element modeling. In Frontiers in Offshore Geotechnics III: Proceedings of the
Third International Symposium in Frontiers in Offshore Geotechnics (ISFOG), Oslo, Norway
(pp. 343–348). Taylor & Francis Group, London.
Ottosen, N. & Ristinmaa, M. (2005). The mechanics of constitutive modeling. Elsevier Science.

95
BIBLIOGRAPHY

Schanz, T., Vermeer, P. A. & Bonnier, P. G. (1999). The hardening soil model: Formulation and
verification. Beyond 2000 in Computational Geotechnics - 10 years of PLAXIS, Balkema,
Rotterdam, 281–290.
Senders, M. (2009). Suction caissons in sand as tripod foundations for offshore wind turbines (PhD
Thesis, The University of Western Australia, Perth, Australia).
Shen, K., Zhang, Y., Klinkvort, R. T., Sturm, H., Jostad, H. P., Sivasithamparam, N. & Guo, Z.
(2017). Numerical simulation of suction bucket under vertical tension loading. Zhejiang
University. Hangzhou, China.
Sjelmo, A. (2012). Soil-structure interaction in cohesionless soils due to monotonic loading (MSc
Thesis, Aalborg University, Aalborg, Denmark).
Sørensen, E. S., Clausen, J. & Damkilde, L. (2016). Comparison of numerical formulations for the
modeling of tensile loaded suction buckets. Computers and Geotechnics, (83), 198–208.
Thieken, K., Achmus, M. & Schröder, C. (2014). On the behavior of suction buckets in sand under
tensile loads. Computers and Geotechnics, (60), 88–100.
Vaitkunaite, E. (2016). Physical modelling of bucket foundations subjected to axial loading (PhD
Thesis, Aalborg University, Aalborg, Denmark).
Vaitkunaite, E., Ibsen, L. B. & Nielsen, B. N. (2016). Testing of axially loaded bucket foundations
with applied overburden pressure (DCE Technical Reports No. 209). Department of Civil
Engineering, Aalborg University. Aalborg, Denmark.
Vaitkune, E., Nielsen, B. N. & Ibsen, L. B. (2016). Bucket foundation response under various
displacement rates. International Journal of Offshore and Polar Engineering, 26(2), 116–
124.
Vethanayagam, V. & Ibsen, L. B. (2017). Determination of p-y curves for bucket foundations in silt
and sand using finite element modelling (DCE Technical Memorandum No. 64). Department
of Civil Engineering, Aalborg University. Aalborg, Denmark.
Wolf, T. K., Rasmussen, K. L., Hansen, M., Roesen, H. R. & Ibsen, L. B. (2013). Assessment of
p-y curves from numerical methods for a non-slender monopile in cohesionless soil (DCE
Technical Memorandum No. 24). Department of Civil Engineering, Aalborg University.
Aalborg, Denmark.

96
A PPENDIX A

HSsmall soil model

The Hardening Soil model with small-strain stiffness (HSsmall) is based on the original Hardening
Soil model (HS) developed by Schanz et al. (1999). This chapter gives an overview of the original
model and outlines the additional terms of the HSsmall model. The information given here is
largely founded on author’s interpretation of the explanations offered in Brinkgreve et al. (2017).

A.1 The original HS model

As the name implies, the HS model takes into account soil hardening during plastic loading. A
hardening plasticity model captures real soil behaviour more accurately than an elastic perfectly-
plastic model, e.g. Mohr-Coulomb. This is due to the fact that in reality, the relation between
stresses and strains exhibits a hyperbolic trend and the yield stress can increase. See Figure A.1.

Figure A.1: (a) Elastic-perfectly plastic behaviour. (b) Hardening plasticity.

While the Mohr-Coulomb model assumes a fixed yield surface, the HS model allows isotropic
hardening. In the principal stress space, isotropic hardening is reflected by the change in size of the
yield surface, while its position and shape remain unaffected [Ottosen & Ristinmaa (2005)]. See
the yield surface of the HS model in Figure A.2.

Two types of hardening are accounted for:

• Shear hardening happens due to plastic shearing during deviatoric loading. A non-associated
flow rule is adopted. This process continues until the ultimate shear strength according to the
Mohr-Coulomb model is reached.

1
MATERIAL MODELS MANUAL

Chapter A. HSsmall soil model

σ1

σ3
σ2
Figure 6.10
FigureRepresentation
A.2: Yield surfaceofof total yield contour
the hardening of inthe
soil model Hardening
principal Soil [Brinkgreve
stress space model in etprincipal stress
al. (2017)].
space for cohesionless soil

• Compression hardening takes place during isotropic compression and is related to plastic
shear yield locus can expand up to the ultimate Mohr-Coulomb failure surface. The cap
volumetric strain. The cap (cf. Figure A.2) is used to model compression hardening. Since the
yield surfacecapexpands as a function of the pre-consolidation stress pp . strain development
yield surface is also used as plastic potential, the flow rule for plastic
is associated.

6.6 STATE PARAMETERS IN THE HARDENING SOIL MODEL

InA.2 Inputtoparameters
addition the output of standard stress and strain quantities, the Hardening Soil
model provides output (when being used) on state variables such as the hardening
The stiffness
parameter γ p andparameters required
the isotropic for HS model are discussed
pre-consolidation stress ppin. detail,
Theseasparameters
they outline some
can core
be
characteristics of the HS model concept.
visualised by selecting the State parameters option from the stresses menu. An overview
of available state parameters is given below:
re f
eq
• Secant stiffness in standard drained triaxial test, E50 . This modulus represents the secant
2
p stiffness at 50 % of the maximum deviatoric stress at a reference cell pressure pre f ,[kN/m
: Equivalent isotropic stress which is]
s depends on the current stress level, the HS model
by default 100 kPa. As the soil stiffness
offers a considerable advantageeq
e2 it possible
q
by making to account for this phenomenon. The
p =
actual stiffness, E50 , depends on the minor + (p') 2
2 principal stress, σ3 . However, the user does not
0
M
have to specify this value, since it is automatically calculated and kept track of by Eq. (A.1).
pp : Isotropic preconsolidation stress [kN/m2 ]
p eq
OCR : Isotropic c cos ϕ − σ30 sin ϕ m ratio (OCR = p /p )
 over-consolidation  [-]
re f
E50 = E50 (A.1)
γp : Hardening ϕ + pre f sin ϕ(equivalent mobilised plastic
c cosparameter [-]
shear strain)
where c is the effective cohesion, ϕ is the effective angle of internal friction and m is the
Eur : Current
power for the amountstress-dependent
of stress dependency.elastic
re f
E50 is Young's modulus
linked to plastic [-]defined
shearing and is
in Figure A.3
c : Current depth-dependent cohesion re f
[-]
• Tangent stiffness for primary oedometer loading, Eoed . This parameter is the modulus
obtained from an oedometer test, thus it is associated with compressibility of soils.The
6.7 ON THE USE OF THE HARDENING SOIL MODEL IN DYNAMIC CALCULATIONS
2
When using the Hardening Soil model in dynamic calculations, the elastic stiffness
ref
parameter Eur needs to be selected such that the model correctly predicts wave
velocities in the soil. This generally requires an even larger small strain stiffness rather
A.2 Input parameters

reference modulus is linked to the reference vertical stress pre f . The actual value, Eoed is
determined by Eq. (A.2).

 m
σ30
c cos ϕ − sin ϕ
re f 
 K0nc 

Eoed = Eoed  c cos ϕ + pre f sin ϕ  (A.2)

where K0nc is the initial coefficient of lateral earth pressure for normal consolidation. The
re f
definition of Eoed is given in Figure A.4
re f
• Unloading/reloading stiffness, Eur . The unloading and reloading is elastic in the HS model
re f
and Eur is the reference Young’s modulus. The stress-level dependent stiffness, Eur , is found
in a similar way as E50 by Eq. (A.3).

 m
re f c cos ϕ − σ30 sin ϕ
Eur = Eur (A.3)
c cos ϕ + pre f sin ϕ

re f
See Figure A.3 for the graphical representation of Eur .

re f re f
Figure A.3: Definition of E50 and Eur for drained triaxial test results at the reference confining pressure
pre f . q = deviatoric stress and ε1 = axial strain. Reproduced from Brinkgreve et al. (2017).

Other input parameters are not discussed here, but they are mentioned as follows:

ψ dilatancy angle
σt tensile strength
νur Poisson’s ratio for unloading/reloading
Rf failure ratio
cinc increase of cohesion with depth

3
Chapter A. HSsmall soil model

THE HARDENING SOIL MODEL WITH SMALL-STRAIN STIFFNESS (HSSMALL)

7 THE HARDENING SOIL MODEL WITH SMALL-STRAIN STIFFNESS (HSSMALL)

The
Figure original
A.4: Hardening
Definition re f Soil model assumes elastic material behaviour during unloading
of Eoed in oedometer test results. σ1 = vertical stress. Reproduced from Brinkgreve
and reloading. However,
et al. (2017). the strain range in which soils can be considered truly elastic,
i.e. where they recover from applied straining almost completely, is very small. With
increasing strain amplitude, soil stiffness decays nonlinearly. Plotting soil stiffness
A.3 against log(strain)
Small-strain yields characteristic S-shaped stiffness reduction curves. Figure 7.1
stiffness
gives an example of such a stiffness reduction curve. It outlines also the characteristic
shear strains that can be measured near geotechnical structures and the applicable strain
Withranges
increasing strain amplitude,
of laboratory the soil
tests. It turns outstiffness decays
that at the non-linearly,
minimum so the
strain which cansoil
be is truly elastic
reliably
onlymeasured
for very small
in classical laboratory tests, i.e. triaxial tests and oedometer tests without stiffness.
strains. Therein lies the motivation for HS model with small-strain
Figure A.5 shows
special the non-linear
instrumentation, soil decay of the
stiffness sheardecreased
is often modulus with increasing
to less than halfshear strains.
its initial value.

1
Shear modulus G/G0 [-]

Very
small
strains Small strains

Larger strains
0 Shear strain gS [-]
-6 -5 -4 -3 -2 -1
1e 1e 1e 1e 1e 1e

Figure A.5: Characteristic stiffness-strain behaviour of soil [Brinkgreve et al. (2017)].

The HSsmall model is able to simulate such behaviour by addition of two parameters on top of
Figure 7.1 Characteristic stiffness-strain behaviour of soil with typical strain ranges for laboratory
the ones presented in the previous section. These parameters are:
tests and structures (after Atkinson & Sallfors (1991))

re f geotechnical structures is not the


•The soil modulus
Shear stiffness that should
at very smallbestrains
used in(εthe
< analysis
10−6 ), Gof
0 . Similarly to E50 and Eur , the stress-
one thatdependency
level relates to the strain
of G 0 is range
modeledat the
by end
Eq. of construction
(A.4) according to Figure 7.1.
Instead, very small-strain soil stiffness and its non-linear dependency on strain amplitude
should be properly taken into account. In addition to all features of the Hardening Soil
 Soil model0 with small-strain
model, the Hardening  stiffness offers the possibility to do so.
c cos ϕ − σ sin ϕ m
G = Gre f
0 0
3
(A.4)
The Hardening Soil c cos ϕwith
model re f sin ϕ
+ psmall-strain stiffness implemented in PLAXIS is based on
the Hardening Soil model and uses almost entirely the same parameters (see Section
6.4). In fact, only two additional parameters are needed to describe the variation of
4
stiffness with strain:
• the initial or very small-strain shear modulus G0
• the shear strain level γ0.7 at which the secant shear modulus Gs is reduced to about
70% of G
A.3 Small-strain stiffness

The reference value is estimated according to Hardin & Black (1969) by Eq. (A.5).

(2.97 − e)2
Gre f
0 = 33 [MPa] f or pre f = 100 [kPa] (A.5)
1+e

where e is the void ratio.


• Threshold shear strain, γ0.7 . This is the value of the shear strain at which the reference secant
shear modulus, Gre f re f
s , is decayed to 72.2 % of G0 . The threshold shear strain is related to
the Mohr-Coulomb failure parameters through Eq. (A.6).

1  0 
γ0.7 = 2c (1 + cos (2ϕ)) − σ10 (1 + K0 ) sin (2ϕ) (A.6)
9Gre
0
f

5
A PPENDIX B

Aalborg sand no. 1

Ibsen et al. (2009) states the unit weight of Aalborg sand no. 1 as γ ≈ 20 kN/m3 .

Hedegaard & Borup (1993) established the minimum and maximum void ratio of Aalborg sand
no.1: emin = 0.549 and emax = 0.858. Considering the definition of relative density ID [Eq. (3.4)],
these void ratio limits are used to estimate the actual void ratio at any compaction state, provided
that knowledge of relative density ID exists. In Vaitkunaite et al. (2016) all tests were performed on
sand with ID = 81 % and in Vaitkune et al. (2016) the values were in the range from 79 % to 90 %.

The Mohr-Coulomb strength parameters were derived from drained triaxial tests and calibrated
by Ibsen et al. (2009). The dependency of the friction angle ϕ on relative density ID and confining
pressure σ3 is emphasized in the mentioned study. In this project, focus is placed upon the triaxial
test series that involved sand samples with ID ≈ 80 %.

By fitting the Curved Coulomb Criterion to test data, Ibsen et al. (2009) suggest Eq. (B.1) to
estimate the friction angle.

 
δqf
−1  δ σ3
ϕ= sin δq
 (B.1)
2 + δ σ3f

where q f is the deviatoric stress at failure. The dependency on ID was introduced through
the asymptotic friction angle ϕa with Eq. (B.2). In the context of Curved Coulomb Criterion, the
friction angle ϕ approaches ϕa as the confining pressure σ3 tends towards infinity.

ϕa = 0.091 ID + 30.6 (B.2)

where ID must be expressed in %. Using Eq. (B.1) the triaxial data is plotted and fitted as shown
in Figure B.1.

Figure B.1 illustrates a power function of the form f (x) = axb + c, where the c-term is replaced
by ϕa . In this way, one obtains Eq. (B.3) that is able to estimate the friction angle ϕ based on ID
and σ3 .

7
Chapter B. Aalborg sand no. 1

58
Triaxial data
56
Power fit
54 Asymptote

52

50

48

46

44

42

40

38

36

0 100 200 300 400 500 600 700 800 900

Figure B.1: Friction angle as function of confining pressure for ID = 80 %. Triaxial data extracted from
Ibsen et al. (2009).

ϕ = 34.88 σ3−0.4606 + 0.091 ID + 30.6 (B.3)

Ibsen et al. (2009) calibrated other parameters using triaxial data: dilatancy angle ψ [Eq. (B.4)]
re f
and secant modulus E50 for a reference confining pressure of 100 kPa [Eq. (B.5)].

ψ = 0.195 ID + 14.9 σ3−0.0976 − 9.95 (B.4)


re f
E50 = 0.6322 ID2.507 + 10920 [kPa] (B.5)

where ID is inserted in % and σ3 in kPa. Eq. (B.5) is especially useful as it directly provides
one of the stiffness moduli required for the HSsmall model. Ibsen et al. (2009) determined the
power for stress-level dependency of stiffness as m = 0.58.

The Poisson’s ratio ν of Aalborg sand no.1 was found by Andersen et al. (1998) to be ν = 0.25.

The flow properties are taken from Sjelmo (2012), see Figure B.2. In contrast to Frederikshavn
sand [Section 3.2.2], data for Aalborg sand no. 1 is extracted directly in terms of hydraulic
conductivity k. As the intention is to simulate numerically an experimental procedure done in
laboratory conditions, no adjustment for water temperature effect is required. Water had T ≈ 20 ◦ C
in permeability tests performed by Sjelmo (2012) and in all bucket loading tests.

The empirical relation between hydraulic conductivity and void ratio is determined and dis-
played by Eq. (B.6).

8
10-4
1.8
Sjelmo (2012)
Quadratic fit
1.6

1.4

1.2

0.8

0.6

0.4
0.5 0.55 0.6 0.65 0.7 0.75 0.8

Figure B.2: Hydraulic conductivity of Aalborg sand no.1 as function of void ratio.

k = 0.8822 · 10−3 e2 − 0.7360 · 10−3 e + 0.1963 · 10−3 (B.6)

All parameters of Aalborg sand no. 1 that are not mentioned in this section are estimated in the
same manner as for Frederikshavn sand in Section 3.2.2.

9
A PPENDIX C

Guidelines for PYTHON scripting

All the components needed for PYTHON scripting are automatically installed with the PLAXIS
software. The user may write the script with a text editor of choice, but the one provided by
default is SciTE. In the current appendix no attempt is made at explaining the underlying details
related strictly to the discipline of computer science and programming, as it would be outside the
project scope. Thus, the reader is reminded that the following sections of this appendix present the
developed PYTHON scripting methods and how they were applied to PLAXIS 2D.

In order to be able to access the scripting features, a list of prerequisites must be fulfilled, as
defined in Brinkgreve et al. (2017).

C.1 Remote scripting server

Before running a script, it is always necessary to activate the remote scripting server in PLAXIS
Input or PLAXIS Output (or both, depending on the task). The process of activation is the same for
both applications, so PLAXIS Input is used as an example.

• Start the PLAXIS Input application.


• Start a new project.
• Go to the Expert menu and select Configure remote scripting server. A window as shown
Figure C.1 in will open.
• Find a port that is available. The default is given as 10000.
• Depending on the PLAXIS version, the user may be requested to define a password for
security reasons.
• Start the server and check that all the tick marks on the left side are active, then close the
window.

The same procedure must be carried out in PLAXIS Output program. If PLAXIS Input and
PLAXIS Output are running at the same time, they cannot use the same port, therefore it needs to
be ensured that the ports are different.

11
Chapter C. Guidelines for PYTHON scripting

Figure C.1: Configure remote scripting server window.

C.2 SciTE editor

If the server is active, no further interaction with PLAXIS user interface is required, since all the
procedures can be defined programmatically in SciTE editor. As previously mentioned, the editor
comes directly with the installation of PLAXIS software.

When the editor is open, a file must be created and saved with the extension ".py". Furthermore,
the programming language must be defined as "Python" under the menu Language. Assuming that
the requirements above are met, the scripts can be run by pressing the F5 key.

C.3 Libraries and connection to PLAXIS application

The first step in using PYTHON wrapper for PLAXIS is importing the scripting libraries and
establishing the connection with the application. In order to do this, the following boilerplate code
must be inserted at the beginning of the script. The reader is invited to search for the definition of a
"boilerplate code".

# Define the path to the scripting libraries


plaxis_path = r'C:\Program Files\Plaxis\PLAXIS 2D\python\Lib\site-packages'
import imp
found_module = imp.find_module('plxscripting', [plaxis_path])
plxscripting = imp.load_module('plxscripting', *found_module)
from plxscripting.easy import *

# Port info
localhostport = 10000
localhostport_output = 10001

# Connect to PLAXIS application


s_i, g_i = new_server('localhost', localhostport, password = 'your password')
s_o, g_o = new_server('localhost', localhostport_output, password = 'your password')

12
C.4 PLAXIS 2D Input

The path to the scripting libraries might vary from one computer to another, therefore it
should be changed accordingly if it is different from the one presented above through the variable
plaxis_path. The correct path is the one which leads to the directory that contains the folder
named "plxscripting". It is important to remember the absence of trailing backslashes when defining
file paths.

In this example it is assumed that 10000 and 10001 are the port numbers for PLAXIS Input and
PLAXIS Output, respectively. These numbers must be the same as the ones defined earlier in the
Configure remote scripting server window. The same applies to the password.

The variables s_i and s_o are bound to objects that represent the PLAXIS application and they
allow the user to control project files, e.g. open projects. On the other hand, g_i and g_o relate to
the model. These are the variables that must be used in order to manipulate the current model.

C.4 PLAXIS 2D Input

This section deals with the part of the script where the model is defined and the calculation is
performed with PLAXIS Input. The full list of commands and objects that were designed for easy
scripting can be accessed from the PLAXIS application in the menu Help » Command reference.

The following code opens a new project and defines its properties. The properties are set by
using name-value pair arguments. If a certain property is not explicitly addressed in the script, the
default value is assumed in the project.

# Start a new project


s_i.new()

# Set model and elements properties


g_i.setproperties("ModelType", "Axisymmetry",
"ElementType", "15-Noded")

# General format
g_i.setproperties(Name1, Value1, Name2, Value2, ... NameN, ValueN)

The boundaries of the model are set independently with the following command:

g_i.SoilContour.initializerectangular(Xmin, Ymin, Xmax, Ymax)

C.4.1 Soil menu

The snippet of the code below creates a soil material set and specifies some of its properties. Unlike
model properties, some soil characteristics (e.g. stiffness, unit weight) do not have default values
and thus must be defined by the user, otherwise PLAXIS gives a warning and the execution of the
script stops.

13
Chapter C. Guidelines for PYTHON scripting

sand = g_i.soilmat() # Create a soil material set

sand.setproperties("MaterialName", "Aalborg Sand no.1",


"Colour", 964844,
"SoilModel", 4,
"DrainageType", "Undrained (A)",
"gammaUnsat", 18,
"gammaSat", 20,
"DilatancyCutOff", True,
"einit", 0.99,
"emin", 0.64,
"emax", 1.05,
"E50ref", 4920,
"EoedRef", 7390,
"EurRef", 14770,
"powerm", 0.5,
"cref", 0.1,
"phi", 35,
"psi", 0,
"gamma07", 0.22,
"G0ref", 60000,
"K0nc", 0.5,
"InterfaceStrength", "Rigid")

Material properties are set in the same manner (name-value pairs) as model properties. From
the technical point of view, it does not make a difference whether the arguments are written in
one line or in a column, as seen above. However, the latter makes it easier to navigate and apply
changes, should such issue arise. Moreover, it is good practice to assign objects to meaningfully
named variables when possible. In the example above, sand is the entity that represents the soil
material in question. The advantage, if not necessity, of such a method becomes apparent in projects
that involve multiple soil types.

The next step is to define the boreholes and the soil strata:

borehole = g_i.borehole(X) # X = X-coordinate of the borehole


borehole.setproperties("Head", Y) # Y = Y-coordinate of the water head

g_i.soillayer(Y) # Define a layer


# Y = Y-coordinate of the lower boundary of the layer

g_i.Soils[0].Material = sand # Assign the soil material to the layer

C.4.2 Structures menu

The structural entities, boundary conditions, loads/prescribed displacements and soil polygons are
created under the Structures menu. The starting point is to select the actual menu and create the
lines which define the geometry of the structure, the interfaces and the distribution of prescribed
displacements.

14
C.4 PLAXIS 2D Input

g_i.gotostructures() # Move to STRUCTURES tab

# Create lines
g_i.line(X1, Y1, X2, Y2) # Line_1: lid of the bucket
g_i.line(X1, Y1, X2, Y2) # Line_2: skirt of the bucket
g_i.line(X1, Y1, X2, Y2) # Line_3: extended interface
# X1, Y1 = coordinates of the first point
# X2, Y2 = coordinates of the second point

# Assign lines to variables


lid_line = g_i.Line_1
skirt_line = g_i.Line_2
ext_int_line = g_i.Line_3

In this example three lines are created and PLAXIS names the line objects according to the
order in which they are defined. These objects are further assigned to variables for easy handling.
It is not a necessity, but it grants advantages in terms of scripting when these objects are called later
in the script.

Since the function line() is only a command that instructs the PLAXIS application to draw
a line, it cannot be assigned to a variable right away; i.e. variable1 = g_i.line() would
lead to PYTHON raising an error when variable1 is eventually called. On the other hand,
writing variable1 = g_i.Line_1 links the variable to an existing line object, therefore making
it possible to manipulate this object through variable1.

In the following code, plate and interface objects are assigned to lines. The techniques of using
variables and the alternative method of using direct object names are presented.

# USING VARIABLES

# Create plates
g_i.plate(lid_line) # Plate_1: assign plate to lid line
g_i.plate(skirt_line) # Plate_2: assign plate to skirt line

# Create interfaces
g_i.neginterface(lid_line) # Negative interface under the lid

g_i.neginterface(skirt_line) # Negative skirt interface


g_i.posinterface(skirt_line) # Positive skirt interface

g_i.neginterface(ext_int_line) # Negative extended interface


g_i.posinterface(ext_int_line) # Positive extended interface

# USING OBJECT NAMES AS SEEN IN PLAXIS APPLICATION

# Create plates
g_i.plate(g_i.Line_1) # Plate_1: assign plate to lid line
g_i.plate(g_i.Line_2) # Plate_2: assign plate to skirt line

# Create interfaces
g_i.neginterface(g_i.Line_1) # Negative interface under the lid

g_i.neginterface(g_i.Line_2) # Negative skirt interface

15
Chapter C. Guidelines for PYTHON scripting

g_i.posinterface(g_i.Line_2) # Positive skirt interface

g_i.neginterface(g_i.Line_3) # Negative extended interface


g_i.posinterface(g_i.Line_3) # Positive extended interface

The use of variables is better in that their name carries a meaning relevant to the user, thus
eliminating the need to check the PLAXIS user interface and search for the name of the object itself.
However, all the guidelines that follow are given with reference to the object names in PLAXIS
application, for the purpose of learning. See Section C.6.3 for instructions of how to find the names
that PLAXIS gives to the generated objects.

The last step in defining the plate structure is attributing material properties. This is done in the
same way as for the soil material.

steel = g_i.platemat() # Create a plate material set

steel.setproperties("MaterialName", "Steel",
"Colour", 16711680,
"MaterialNumber", 0,
"Elasticity", 0,
"IsIsotropic", True,
"IsEndBearing", False,
"EA", 180000000,
"EI", 1350000,
"nu", 0.3)

# Assign material properties to plates


g_i.Plate_1.Material = steel
g_i.Plate_2.Material = steel

Prescribed displacements are applied to line objects as follows:

# Assign prescribed displacements to lines


g_i.linedispl(g_i.Line_1)
g_i.linedispl(g_i.Line_2)

# Set properties of line displacements


g_i.LineDisplacement_1.setproperties("Displacement_x", "Fixed",
"uy_start", 0.3)

g_i.LineDisplacement_2.setproperties("Displacement_x", "Fixed",
"uy_start", 0.3)

Groundwater flow boundary conditions may be defined in the same manner as the line objects.

# Create groundwater flow BCs

g_i.gwfbc(Xmin, Ymax, Xmin, Ymin) # Left edge of the domain (GWFlowBC_1)

g_i.gwfbc(Xmin, Ymin, Xmax, Ymin) # Bottom of the domain

g_i.gwfbc(Xmax, Ymin, Xmax, Ymax) # Right edge of the domain

16
C.4 PLAXIS 2D Input

g_i.gwfbc(Xmax, Ymax, Xmin, Ymax) # Top of the domain

g_i.GWFlowBC_1.Behaviour = "Closed" # Make the gwfbc on the left edge impermeable

The function gwfbc() takes the same arguments as line objects: coordinates of the first and of
the second point. In this example, the boundaries of the project domain are used as the arguments.
Default properties of groundwater flow BCs are assumed as long as they are not changed, as seen
in the example above, where the properties of only one boundary are manipulated.

If the model requires local mesh refinement, a soil polygon that encompasses the refinement
area must be drawn. PLAXIS then resets the material of the polygon to "none". In many situations
the material of the polygon is the same as that of the surrounding soil and it must be specified.

g_i.rectangle(X1, Y1, X2, Y2) # Draw a soil polygon

# NOTE! The soil within the polygon is named Soil_2

g_i.Soil_2.Material = sand # Assign soil material to the soil within the polygon

C.4.3 Mesh menu

The following part of the script is responsible for generating the mesh. Both the global mesh
coarseness and the local mesh coarseness factor can be adjusted. It is important to notice that the
coarseness factor is not applied to the soil object, but rather to the polygon itself. If the global
coarseness is not specified, the default value of "medium" is assumed.

g_i.gotomesh() # Move to MESH tab

# Define the local coarseness factor for Polygon_1


g_i.BoreholePolygon_1_Polygon_1_1.CoarsenessFactor = 0.3

g_i.mesh() # Generate the mesh

g_i.viewmesh() # Open a view of the generated mesh

The command viewmesh() is optional and relevant only if visual inspection is desired. The
execution of the command results in an error if PLAXIS Output is closed or if its local server is not
activated by the user prior to running the script.

C.4.4 Staged construction menu

The initial stage is present by default and thus does not need to be added by the user. However, if
necessary, its properties can be changed as seen below.

17
Chapter C. Guidelines for PYTHON scripting

g_i.gotostages() # Move to STAGED CONSTRUCTION tab

# Initial phase
g_i.InitialPhase.Deform.IgnoreSuction = False # Ignore suction: OFF

The process of defining a phase follows a specific pattern:

• Add a phase. At the same time the previous phase must be stated.
• Make the phase current.
• Set the properties of the phase.

This pattern can be observed in the following code, where three phases are created.

# Phase_1 (Installation)
g_i.phase(g_i.InitialPhase) # Add Phase_1
g_i.setcurrentphase(g_i.Phase_1) # Make Phase_1 current

g_i.Phase_1.Identification = "Installation" # Name the phase

g_i.Plates.activate(g_i.Phase_1) # Activate plates


g_i.Interfaces.activate(g_i.Phase_1) # Activate interfaces
g_i.GroundwaterFlowBCs.activate(g_i.Phase_1) # Activate groundwater flow BCs

# Don't reduce the strength of the extended interfaces


g_i.set(g_i.NegativeInterface_3_1.ApplyStrengthReduction, (g_i.Phase_1), False)

g_i.set(g_i.PositiveInterface_2_1.ApplyStrengthReduction, (g_i.Phase_1), False)

# Make the extended interfaces fully permeable


g_i.set(g_i.NegativeInterface_3_1.ActiveInFlow, (g_i.Phase_1), False)

g_i.set(g_i.PositiveInterface_2_1.ActiveInFlow, (g_i.Phase_1), False)

# Phase_2 (Nil-step)
g_i.phase(g_i.Phase_1) # Add Phase_2
g_i.setcurrentphase(g_i.Phase_2) # Make Phase_2 current

g_i.Phase_2.Identification = 'Nil-step' # Name the phase

# Reset displacements to zero and reset small strains


g_i.Phase_2.Deform.ResetDisplacementsToZero = True

# Phase_3 (Loading)
g_i.phase(g_i.Phase_2) # Add Phase_3
g_i.setcurrentphase(g_i.Phase_3) # Make Phase_3 current

g_i.Phase_3.Identification = "Loading" # Name the phase

# Define the calculation type


g_i.Phase_3.DeformCalcType = "Fully coupled flow-deformation"

# Set the time interval for fully coupled f-d analysis


g_i.Phase_3.TimeInterval = 2

18
C.5 PLAXIS 2D Output

# Activate line displacements


g_i.LineDisplacements.activate(g_i.Phase_3)

The addition of a new phase is carried out by the function phase(), the argument of which
is the previous phase. In order to change the properties of objects during a certain phase, the
set(object.attribute, (phase), value) command is used. The properties of the phase
itself (e.g. ID, calculation type) can be accessed directly through the dot symbol. When the phases
are set, the model is ready for calculation. This is done simply with one command:

g_i.calculate()

In practical terms, it is beneficial to save the project after the calculation is finished.

# Save the project file


save_path = r'your path\project_name'
g_i.save(save_path)

When writing strings that represent file paths it is critical to remember the "r" flag preceding
the string and the absence of the trailing backslash. In the case above, project_name is the name
of the saved project and it is not required to add the PLAXIS-specific extension ".p2dx".

At this point the project is ready for data extraction, which can be carried out with relative ease
by scripting.

C.5 PLAXIS 2D Output

Generally, visualizing results as contour lines, arrows or shadings is done directly in PLAXIS
Output. These techniques are useful when the goal is to investigate phenomena such as stress
distribution, flow net, failure mechanism etc., because they offer a qualitative description of the
behaviour of the system. In these terms scripting does not offer any considerable advantages.

It is often the case that data needs to be plotted as curves in order to analyze certain behavioural
aspects of the model. This requires precise quantities to be extracted and plotted. It is possible to
do it with Curves manager in PLAXIS Output, but the process becomes notably time-consuming
when working with large amount of various physical quantities related to different nodes and stress
points. Moreover, studies that involve comparison of models can be carried out only by extracting
data manually from each of them and by exporting it to files. Cumbersome manual data extraction
can be avoided with the PYTHON wrapper.

The process of data extraction is divided into two steps:

1. Requesting data from PLAXIS Output and storing it into PYTHON objects, such as list.
2. Writing data to a file.

19
Chapter C. Guidelines for PYTHON scripting

The options of how to carry out the first step are rather limited, as opposed to the amount of
choices the user has at the second step. The manner in which data is going to be handled dictates
the format and the structure of the file to which data is written. For example, the user might choose
to use Microsoft Excel for data processing, in which case ".xlsx" files have to be created.

The degree of manual effort the user has to apply depends on the automation level of the
PYTHON script and on the choice of software for data processing. This section describes only one
of the many possible ways to handle data. The primary purpose of the method is to reduce the need
of manual interaction to the minimum.

Keeping in mind that data processing is undertaken by using MATLAB as the program of
choice, information from PLAXIS Output is stored in ".txt" files. All aspects of the method lead
towards "MATLAB-friendly" data handling.

Assuming all the requirements for scripting are met and the calculation is complete, the first
step is opening the project with the command s_o.open().

open_path = r'your path\project_name'


s_o.open(open_path)

General information about the project can be accessed through object attributes. These attributes
contain scalar quantities. In the example below, the amount of nodes and elements is extracted.

NodeNo = g_o.GeneralInfo.NodeCount
ElemNo = g_o.GeneralInfo.SoilElementCount

The variables NodeNo and ElemNo are scalars that represent the node number and the element
number, respectively. At this point the first step in data extraction is complete, since the two
PYTHON objects that hold information about the project are defined. Next, the information is
written to a text file:

# Create the text file


file_info = open(r'your path\general_info.txt', 'w')

# Write the information to the file


file_info.write('NodeNo: ' + str(NodeNo) + '\r\n' + 'ElemNo: ' + str(ElemNo))

# Close the file


file_info.close()

The function open() takes two arguments. The first one is the path, which ends with the name
of the file that is created. Unlike saving or opening PLAXIS projects, the file extension must be
specified. The second argument is the mode, which tells the system how to open the file. The mode
"w" stands for "write" and it instructs the system to create the file if it does not exist and open it in
writing mode.

The actual writing is carried out by using write(), the argument of which is data in string
format. Since NodeNo and ElemNo contain integer or floating-point format, they must be converted

20
C.5 PLAXIS 2D Output

to strings with the str() function. The string '\r\n' tells the system to create a new row in the
text file and move to the start of the row. It is good practice to close the file when data insertion is
finished. The generated file can be viewed in Figure C.2.

Figure C.2: Text file containing the number of nodes and elements in the project.

It may be of interest to extract information about each step of the calculation. The following
code shows how to obtain stepwise data, namely the step identification and time stamp.

stepID = [] # Create empty list for step identification


stepTime = [] # Create empty list for time stamps

# Loop over steps in Phase 3


for step in g_o.Phase_3.Steps:
stepID.append(step.ID) # Add ID for step N
stepTime.append(step.Reached.Time.value) # Add time stamp for step N

Contrary to the example about general information where scalar quantities are obtained, stepwise
data is retrieved into a list object. The first two lines of code preallocate empty lists to which data
will be added (or appended). The for-loop instructs the system to go over each step in the third
phase and append data to lists. As a result, both lists contain as many elements as there are steps.
Within the loop, the entity step is an incrementing index variable that increases by a value of 1
with each iteration, thus keeping track of the iterative procedure.

Writing data from lists to a file also takes place in a different manner:

# Create the text file


file_step_time = open(r'your path\step_time.txt', 'w')

# Add a header as the first row in the file


file_step_time.write('StepID TimeStamp[s]' + '\r\n')

# Insert data from lists into the file


# Loop over each element in the list
for step, time in zip(stepID, stepTime):
file_step_time.write(str(step) + ' ' + str(time) + '\r\n')

21
Chapter C. Guidelines for PYTHON scripting

# Close the file


file_step_time.close()

Looping over two lists simultaneously requires two index variables: one for each list. Addition-
ally, the list elements must be aggregated with the zip() function. The order in which variables
are declared matters, as the i-th index variable relates to the i-th list. In the example above, step
is linked to the stepID list and time to stepTime respectively. Care must be taken with lists of
unequal sizes, because the iterative procedure stops when the last element in the shortest list is
reached.

At each iteration a new row in the file is added: step ID, space delimiter, time stamp and
relocation to the beginning of a new row. The output file looks like the one shown in Figure C.3

Figure C.3: Text file containing the step IDs and time stamps.

Data related to nodes and stress points can be arranged in the same way. Ultimately, it is up
to the user to decide how information is structured in a file. In the following example, a similar
principle is applied to data from soil nodes:

# Loop over steps in Phase 3


for step in g_o.Phase_3.Steps:
# Nodal coordinates on X-axis
Xcoor = g_o.getresults(step, g_o.ResultTypes.Soil.X, 'node')
# Nodal coordinates on Y-axis
Ycoor = g_o.getresults(step, g_o.ResultTypes.Soil.Y, 'node')
# Displacements in X-direction
Ux = g_o.getresults(step, g_o.ResultTypes.Soil.Ux, 'node')
# Displacements in Y-direction
Uy = g_o.getresults(step, g_o.ResultTypes.Soil.Uy, 'node')
# Total displacements
Utot = g_o.getresults(step, g_o.ResultTypes.Soil.Utot, 'node')

# Create a .txt file


file_node = open(r'your path' + str(step.Name) + '.txt', 'w')

# Add a header as the first row in the file


file_node.write('X-coor[m] Y-coor[m] Ux[m] Uy[m] Utot[m]' + '\r\n')

22
C.6 Tips & suggestions

# Insert data from lists into the file


# Loop over each element in the list
for X, Y, Ux, Uy, Utot in zip(Xcoor, Ycoor, Ux, Uy, Utot):
file_node.write(str(X)+' '+str(Y)+' '+str(Ux)+' '+str(Uy)+' '+str(Utot)+'\r\n')

# Close the newly created file after each iteration


file_node.close()

The purpose of the code above is to create a new text file with each iteration, i.e. one file per
step. This means each file must have a unique name, otherwise the same file is overwritten after
every iteration. To uniquely identify each file, the step ID is included in its name. Data is structured
like a two-dimensional array, where rows stand for nodes and columns represent various physical
quantities.

There is no need to create empty lists before data extraction. The function getresults()
returns a list of floating-point data type. What makes it possible to store data in a two-dimensional
array fashion is that PLAXIS does not change the order of list elements. In other words, the i-th
element in any list, regardless of the requested physical quantity, is related to the i-th node or stress
point. This rule holds true among steps as well. This means that if files were superimposed, a
three-dimensional array would be obtained. For clarification reasons, an analogy is made with a
book where a page represents a step and where every page contains an equally sized matrix (see
Figure C.4).

Figure C.4: Schematic representation of how data is arranged in text files. Q = any physical quantity.

In order to extract data from stress points, the third argument of the getresults() function
must be changed from 'node' to 'stresspoint'. Similarly, to obtain information about other
entities, such as plates or interfaces, changes to the second argument of the same function must be
applied, e.g. from g_o.ResultTypes.Soil.X to g_o.ResultTypes.Plate.X.

The method of writing data to files described above is generally slow and the time consumption
increases with the amount of information. In order to optimize the process, a faster method is
applied. The procedure is presented in Section C.6.6.

C.6 Tips & suggestions

This section describes errors encountered while working with PYTHON wrapper for PLAXIS,
including the developed solutions. Another focus is placed on recommendations aimed at improving

23
Chapter C. Guidelines for PYTHON scripting

scripting procedures and suggestions for handling the PLAXIS software.

C.6.1 Errors when setting material properties

Section C.4.2 shows how to set the properties of a plate material set. A problem might appear when
setting the axial stiffness (EA) or the bending stiffness (EI). Since neither of the quantities can be
zero, PLAXIS does not allow the creation of a plate material set unless both values are non-zero
numbers. The error window is shown in Figure C.5.

Figure C.5: Error window that pops up when the axial stiffness is given a zero value.

The issue arises when these properties are defined by scripting with the PYTHON wrapper.
Regardless of what value is inserted as EA or EI, the error occurs and thus the execution of the
script stops. Similar issues have been encountered with soil material sets, where several properties
(e.g. minimum and maximum void ratio, or activation of dilatancy cut-off) remain unchanged
despite the commands given in the script. The reason for the error is unknown and most probably it
lies within a bug in the PLAXIS application.

The suggested manner in which this issue can be dealt with is a workaround, rather than a
solution to the problem.

A new project is created manually with the PLAXIS user interface. The material sets for soil
and plates and their properties are created in the same way. No other changes shall be applied to
the model. The project is saved, so that it can be used as a template. Having a template project with
predefined material sets means that it can be opened every time by scripting and the model may be
built on top of it, instead of starting a new project. The final model should be saved as a project
with a different name, so that the template remains intact. The command for opening a project is:

s_i.open(r'your path\template_project.p2dx')

It is important to notice that the file extension ".p2dx" must be specified in order to run the
command successfully.

24
C.6 Tips & suggestions

C.6.2 Errors when opening PLAXIS Output application

When an iterative procedure of generating models is running, it is often the case that PLAXIS
Output suddenly crashes after several iterations. This makes the PYTHON script stop the execution.
This issue is solved partially by closing the project after data extraction. PLAXIS Output may
still crash, but only after a considerable amount of models have been run. Closing the project is
performed at the end of each iteration by applying the following command:

s_o.close()

There is no need to use any arguments in this function, because it refers automatically to the
currently open project.

C.6.3 Finding object names with PLAXIS user interface

As discussed in Section C.4.2, it is possible to refer to objects by using their names as given by the
PLAXIS application. In order to find these names in the PLAXIS user interfaces, it is enough to
click on the object and use the Selection explorer. An example is shown in Figure C.6.

Figure C.6: The Selection explorer tab.

C.6.4 Finding object attributes

Knowledge of object attributes is necessary to manipulate object properties. The most straight-
forward way to find the syntax for a specific attribute is using the Command reference given in

25
Chapter C. Guidelines for PYTHON scripting

PLAXIS. However, it appears that the documentation is not complete, as some attributes are not
listed. As an alternative to documentation, the attribute list can be found with the dir() function
in PYTHON. An example of how to find the attributes for a plate object is given below:

print(dir(g_i.Plate_1))

The function print() displays the list of valid attributes found by dir(). This method is of
the "trial and error" kind, especially if the situation involves navigation through more than one
attribute. For example, if one wants to find how to access the attribute related to axial stiffness of a
plate, the following procedure might be carried out:

print(dir(g_i.Plate_1)) # List of attributes for Plate_1

# Identify the suitable attribute. In this case it is "Material"

print(dir(g_i.Plate_1.Material)) # List of attributes for Plate_1.Material

# In the second list, one of the attributes is "EA"


# With the obtained knowledge, the axial stiffness can be accessed

g_i.Plate_1.Material.EA = 1300000

C.6.5 Saving data for every step

By default, PLAXIS saves information about the last calculation step at the end of each phase.
The program stores stepwise data only for points that are selected prior to the calculation with the
feature Select points for curves. The upper limit of selected points is 10 nodes and 10 stress points.
This amount might prove to be insufficient if the scope of the analysis is broader.

The parameter that controls how many steps are saved is called Max number of steps stored.
The default value of 1 can be changed as follows:

# Maximum number of steps stored during Phase 3


g_i.Phase_3.MaxStepsStored = 10000

Inserting a very large number ensures that data will be saved for all steps and for all nodes
and stress points. At the same time, the limitation concerning the number of selected points is
overcome. As a consequence, the project file becomes considerably larger: in the order of hundreds
of megabytes. This also means the amount of data to be handled may increase to an extent that
makes the process of writing to text files very slow. The following section proposes a fast method
to deal with large data files.

C.6.6 Fast data writing to files

The standard method of writing data to files is presented in Section C.5. However, using JSON
(JavaScript Object Notation) format proves to be approximately four times faster. JSON data is
written in key:value pairs (e.g. "city":"Aalborg").

26
C.6 Tips & suggestions

The procedure consists of the following steps:

1. Request data from PLAXIS Output. For i number of requested quantities, there will be an i
number of lists.
2. Assemble the lists into a "matrix" structure. In pythonic terms the "matrix" is essentially a
list of lists.
3. Put the "matrix" into a dictionary object. Dictionaries are sets of key-value pairs, so at this
step data takes a format suitable for applying JSON encoding.
4. Write JSON-formatted files.

These steps are represented in scripting terms below. The same example as in Section C.5 for
soil nodes is given.

# Import the scripting libraries


# These lines can be placed at the beginning of the script
import json
import numpy as np

# STEP 1 (NOTE! This step is the same as in previous method)


# Loop over steps in Phase 3
for step in g_o.Phase_3.Steps:
# Nodal coordinates on X-axis
Xcoor = g_o.getresults(step, g_o.ResultTypes.Soil.X, 'node')
# Nodal coordinates on Y-axis
Ycoor = g_o.getresults(step, g_o.ResultTypes.Soil.Y, 'node')
# Displacements in X-direction
Ux = g_o.getresults(step, g_o.ResultTypes.Soil.Ux, 'node')
# Displacements in Y-direction
Uy = g_o.getresults(step, g_o.ResultTypes.Soil.Uy, 'node')
# Total displacements
Utot = g_o.getresults(step, g_o.ResultTypes.Soil.Utot, 'node')

# STEP 2
# Create a matrix that contains the extracted data from the current step
mtrx = np.append(([Xcoor[:], Ycoor[:], Ux[:], Uy[:]]), [Utot[:]], axis=0).tolist()

# STEP 3
# Put the matrix into a dictionary
dataDict = {'matrix':mtrx}

# STEP 4
# Insert the dictionary into a file using JSON encoding
with open (r'your path' + str(step.Name) + '.txt', 'w') as json_file:
json_file.write(json.dumps(dataDict, json_file, separators=(',',':')))

# NOTE! No need to "manually" close the file

The output text files are almost incomprehensible to the naked eye. Therefore further processing
with other software is required, e.g. MATLAB. There are built-in functions in MATLAB that decode
JSON-encoded text files and convert it into array structures. The description of how JSON-files are
imported and handled with MATLAB is outside the scope of this chapter.

27
Chapter C. Guidelines for PYTHON scripting

C.6.7 Node data filtering

Referring back to Figure C.2, the number of soil nodes and elements is 10931 and 1287, respectively.
However, when requesting soil node data from PLAXIS and writing it to text files according to the
structure described in Figure C.4, one will observe that the amount of rows (nodes) is larger than
10931. Precisely, the number of rows can be calculated as:

no. of rows = no. of elements × no. of nodes per element

Therefore, if 15-noded elements are used, then the amount of rows will be 19305. There
are indeed 10931 unique soil nodes, but the discrepancy appears due to sharing of nodes among
elements. The conclusion is that PLAXIS returns information about nodes in an element-wise
manner. To explain this phenomenon, Figure C.7 shows an example of how data is given for plate
elements, since they share nodes as well.

Figure C.7: A plate consisting of 2 elements. The node marked with red is a shared node. Information
related to the node is given as two identical rows in the text file.

It is clear that the data files contain identical rows that refer to the same node. Thus, the
redundant rows must be filtered out as part of data processing.

The situation becomes more complicated in case of interfaces on both sides of a structural
object. Since interfaces do not have real thickness, PLAXIS attributes the same coordinates to a
node that belongs to a negative interface and to a positive interface (see Figure C.8).

Regarding shared nodes, the result is that four rows contain the same coordinates, but describe
two different nodes: one shared node from the negative interface and one from the positive interface.
Thus it becomes obvious that filtering interface node data should not be based on coordinates alone,
as one of the nodes might be disregarded completely.

28
C.6 Tips & suggestions

Figure C.8: A plate with 2 interfaces. Each object consists of 2 elements. Nodes highlighted with orange
and purple have the same coordinates.

A general logical algorithm for distinguishing between shared nodes in the negative and positive
interfaces has not been established. This means that a unique filtering procedure must be found
for each project individually. The following list is a step-by-step method that was applied in the
current project, but which could prove to be useful in others as well.

• Open PLAXIS Output.


• Open a project.
• Expand the Interfaces branch in the model explorer on the left hand side.
• Expand each interface object down to the interface elements. Interface elements are numbered
with integer values: "Interface n". One interface element contains 5 nodes if 15-noded soil
elements are used.
• Toggle the view of an interface object or of an interface element, and compare with the model
view in order to match the object with its label.

29
Chapter C. Guidelines for PYTHON scripting

• In the file to which data was written, every 5 consecutive rows (nodes) represent one interface
element. The order of elements in the file corresponds to the integer value n, that appears in
the element label. See Figure C.9.

Figure C.9: The interface element labels in the rectangle appear as in the PLAXIS Output model explorer.
Node data in the text file is structured as seen in the schematic table on the right.

This procedure enables the user to make a connection between the interface elements and
the order in which node data appears in the text file. The obtained knowledge can be applied for
building a filtering and sorting algorithm which is able to process the data from any model, as long
as the layout remains unchanged, i.e. dimensions may vary, but not the amount of interface objects.

All the issues described above do not occur with stress points, since they are uniquely defined
and not shared among elements.

C.6.8 Creating folders with PYTHON

If the script is designed to extract information from multiple models with a for-loop, it is advant-
ageous to store results for each model in a separate folder. As opposed to creating folders manually
before running the script, the folders can be created programmatically with PYTHON by using the
following code:

# Import the scripting library


import os

# Create a folder
os.mkdir('folder name specific for each model')

A folder is created in the directory where the current PYTHON file is located, therefore there is
no need to specify the path or to add the "r" flag at the beginning of the string.

30
A PPENDIX D

Statistical analysis, trimmed and


normalized t-z curves

D.1 Drained models (Tension)

D.1.1 Peak shear stress formulation

60

50

40

30

20

10

0
0 50 100 150 200 250

Figure D.1: Comparison between the linear fit and mathematical model.

31
Chapter D. Statistical analysis, trimmed and normalized t-z curves

60

50

40

30

20

10

0
0 50 100 150 200 250

Figure D.2: FE data and the mathematical model established by linear fitting.

60

50

40

30

20

10

0
0 50 100 150 200 250

Figure D.3: Comparison between the power fit and mathematical model.

32
D.1 Drained models (Tension)

60

50

40

30

20

10

0
0 50 100 150 200 250

Figure D.4: FE data and the mathematical model established by power fitting.

30

25

20

15

10

0
0 20 40 60 80 100 120

Figure D.5: FE data, power fit and mathematical model for each FE model.

33
Chapter D. Statistical analysis, trimmed and normalized t-z curves

40

35

30

25

20

15

10

0
0 20 40 60 80 100 120 140 160

Figure D.6: FE data, power fit and mathematical model for each FE model.

45

40

35

30

25

20

15

10

0
0 20 40 60 80 100 120 140 160 180

Figure D.7: FE data, power fit and mathematical model for each FE model.

34
D.1 Drained models (Tension)

50

45

40

35

30

25

20

15

10

0
0 50 100 150 200 250

Figure D.8: FE data, power fit and mathematical model for each FE model.

60

50

40

30

20

10

0
0 50 100 150 200 250

Figure D.9: FE data, power fit and mathematical model for each FE model.

35
Chapter D. Statistical analysis, trimmed and normalized t-z curves

D.1.2 Peak displacement formulation

ϕ
30◦ 33◦ 35◦ 38◦ 40◦ 30◦ 33◦ 35◦ 38◦ 40◦
D
10 m 0.0095 0.0062 0.0051 0.0041 0.0038 0.4058 0.3409 0.3160 0.2901 0.2592
13 m 0.0106 0.0075 0.006 0.0056 0.005 0.4228 0.3268 0.3244 0.2563 0.2428
15 m 0.0124 0.0081 0.0077 0.0061 0.0056 0.3824 0.3394 0.2764 0.2696 0.257
18 m 0.0132 0.0086 0.0082 0.0072 0.007 0.407 0.3627 0.3073 0.2760 0.2407
20 m 0.0176 0.0085 0.0085 0.0082 0.0074 0.3223 0.3963 0.3240 0.2657 0.2637

Table D.1: Values of c-coefficients (left section) and d-coefficients (right section) used for regression analysis
in the context of formulating the peak displacement.

NRMSE = 7.07 %
0.02

0.015

0.01

0.005

0
0.9
0.8 1.4
0.7 1.2
1
0.6
0.8
0.5 0.6

Figure D.10: Fitting c-coefficients with a quadric surface.

36
D.1 Drained models (Tension)

NRMSE = 13.31 %
0.45

0.4

0.35

0.3

0.25

0.2
0.9
0.8 1.4
0.7 1.2
1
0.6
0.8
0.5 0.6

Figure D.11: Fitting d-coefficients with a plane.

0.025

0.02

0.015

0.01

0.005

0
0 1 2 3 4 5 6 7 8 9

Figure D.12: FE data, power fit and mathematical model for each FE model.

37
Chapter D. Statistical analysis, trimmed and normalized t-z curves

0.035

0.03

0.025

0.02

0.015

0.01

0.005

0
0 2 4 6 8 10 12 14

Figure D.13: FE data, power fit and mathematical model for each FE model.

0.04

0.035

0.03

0.025

0.02

0.015

0.01

0.005

0
0 2 4 6 8 10 12 14 16 18 20

Figure D.14: FE data, power fit and mathematical model for each FE model.

38
D.1 Drained models (Tension)

0.05

0.045

0.04

0.035

0.03

0.025

0.02

0.015

0.01

0.005

0
0 5 10 15 20 25 30

Figure D.15: FE data, power fit and mathematical model for each FE model.

0.06

0.05

0.04

0.03

0.02

0.01

0
0 5 10 15 20 25 30 35

Figure D.16: FE data, power fit and mathematical model for each FE model.

39
Chapter D. Statistical analysis, trimmed and normalized t-z curves

D.1.3 Trimmed and normalized t-z curves

25

20

15

10

d = 3.2
d = 3.9
d = 4.6
5 d = 5.3
d = 6.1
d = 6.8
d = 7.5
d = 8.2

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 3.2
d = 3.9
0.4
d = 4.6
d = 5.3
d = 6.1
0.2 d = 6.8
d = 7.5
d = 8.2
0
0 0.5 1 1.5 2 2.5 3 3.5 4

40
D.1 Drained models (Tension)

25

20

15

10

d = 3.2
d = 3.9
d = 4.6
5 d = 5.3
d = 6.1
d = 6.8
d = 7.5
d = 8.2

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 3.2
d = 3.9
0.4
d = 4.6
d = 5.3
d = 6.1
0.2 d = 6.8
d = 7.5
d = 8.2
0
0 1 2 3 4 5 6

41
Chapter D. Statistical analysis, trimmed and normalized t-z curves

30

25

20

15

10
d = 3.2
d = 3.9
d = 4.6
d = 5.3
5 d = 6.1
d = 6.8
d = 7.5
d = 8.2

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 3.2
d = 3.9
0.4
d = 4.6
d = 5.3
d = 6.1
0.2 d = 6.8
d = 7.5
d = 8.2
0
0 1 2 3 4 5 6 7 8

42
D.1 Drained models (Tension)

30

25

20

15

10
d = 3.2
d = 3.9
d = 4.6
d = 5.3
5 d = 6.1
d = 6.8
d = 7.5
d = 8.2

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 3.2
d = 3.9
0.4
d = 4.6
d = 5.3
d = 6.1
0.2 d = 6.8
d = 7.5
d = 8.2
0
0 1 2 3 4 5 6 7 8 9 10

43
Chapter D. Statistical analysis, trimmed and normalized t-z curves

30

25

20

15

10
d = 3.2
d = 3.9
d = 4.6
d = 5.3
5 d = 6.1
d = 6.8
d = 7.5
d = 8.2

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 3.2
d = 3.9
0.4
d = 4.6
d = 5.3
d = 6.1
0.2 d = 6.8
d = 7.5
d = 8.2
0
0 1 2 3 4 5 6 7 8 9

44
D.1 Drained models (Tension)

35

30

25

20

15

d = 4.2
10 d = 5.1
d = 6.1
d=7
d = 7.9
5 d = 8.8
d = 9.7
d = 10.7

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.2
d = 5.1
0.4
d = 6.1
d=7
d = 7.9
0.2 d = 8.8
d = 9.7
d = 10.7
0
0 0.5 1 1.5 2 2.5 3

45
Chapter D. Statistical analysis, trimmed and normalized t-z curves

35

30

25

20

15

d = 4.2
10 d = 5.1
d = 6.1
d=7
d = 7.9
5 d = 8.8
d = 9.7
d = 10.7

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.2
d = 5.1
0.4
d = 6.1
d=7
d = 7.9
0.2 d = 8.8
d = 9.7
d = 10.7
0
0 0.5 1 1.5 2 2.5 3 3.5 4

46
D.1 Drained models (Tension)

35

30

25

20

15

d = 4.2
10 d = 5.1
d = 6.1
d=7
d = 7.9
5 d = 8.8
d = 9.7
d = 10.7

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.2
d = 5.1
0.4
d = 6.1
d=7
d = 7.9
0.2 d = 8.8
d = 9.7
d = 10.7
0
0 1 2 3 4 5 6

47
Chapter D. Statistical analysis, trimmed and normalized t-z curves

35

30

25

20

15

d = 4.2
10 d = 5.1
d = 6.1
d=7
d = 7.9
5 d = 8.8
d = 9.7
d = 10.7

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.2
d = 5.1
0.4
d = 6.1
d=7
d = 7.9
0.2 d = 8.8
d = 9.7
d = 10.7
0
0 1 2 3 4 5 6 7

48
D.1 Drained models (Tension)

40

35

30

25

20

15
d = 4.2
d = 5.1
10 d = 6.1
d=7
d = 7.9
5 d = 8.8
d = 9.7
d = 10.7

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.2
d = 5.1
0.4
d = 6.1
d=7
d = 7.9
0.2 d = 8.8
d = 9.7
d = 10.7
0
0 1 2 3 4 5 6 7

49
Chapter D. Statistical analysis, trimmed and normalized t-z curves

40

35

30

25

20

15
d = 4.8
d = 5.9
10 d=7
d=8
d = 9.1
5 d = 10.2
d = 11.3
d = 12.4

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.8
d = 5.9
0.4
d=7
d=8
d = 9.1
0.2 d = 10.2
d = 11.3
d = 12.4
0
0 0.5 1 1.5 2 2.5

50
D.1 Drained models (Tension)

40

35

30

25

20

15
d = 4.8
d = 5.9
10 d=7
d=8
d = 9.1
5 d = 10.2
d = 11.3
d = 12.4

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.8
d = 5.9
0.4
d=7
d=8
d = 9.1
0.2 d = 10.2
d = 11.3
d = 12.4
0
0 0.5 1 1.5 2 2.5 3 3.5

51
Chapter D. Statistical analysis, trimmed and normalized t-z curves

40

35

30

25

20

15
d = 4.8
d = 5.9
10 d=7
d=8
d = 9.1
5 d = 10.2
d = 11.3
d = 12.4

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.8
d = 5.9
0.4
d=7
d=8
d = 9.1
0.2 d = 10.2
d = 11.3
d = 12.4
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

52
D.1 Drained models (Tension)

45

40

35

30

25

20

15
d = 4.8
d = 5.9
d=7
10
d=8
d = 9.1
d = 10.2
5 d = 11.3
d = 12.4

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.8
d = 5.9
0.4
d=7
d=8
d = 9.1
0.2 d = 10.2
d = 11.3
d = 12.4
0
0 1 2 3 4 5 6

53
Chapter D. Statistical analysis, trimmed and normalized t-z curves

45

40

35

30

25

20

15
d = 4.8
d = 5.9
d=7
10
d=8
d = 9.1
d = 10.2
5 d = 11.3
d = 12.4

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.8
d = 5.9
0.4
d=7
d=8
d = 9.1
0.2 d = 10.2
d = 11.3
d = 12.4
0
0 1 2 3 4 5 6

54
D.1 Drained models (Tension)

45

40

35

30

25

20

15
d = 5.8
d=7
d = 8.4
10
d = 9.7
d = 10.9
d = 12.2
5 d = 13.5
d = 14.8

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 5.8
d=7
0.4
d = 8.4
d = 9.7
d = 10.9
0.2 d = 12.2
d = 13.5
d = 14.8
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

55
Chapter D. Statistical analysis, trimmed and normalized t-z curves

45

40

35

30

25

20

15
d = 5.8
d=7
d = 8.4
10
d = 9.7
d = 10.9
d = 12.2
5 d = 13.5
d = 14.8

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 5.8
d=7
0.4
d = 8.4
d = 9.7
d = 10.9
0.2 d = 12.2
d = 13.5
d = 14.8
0
0 0.5 1 1.5 2 2.5

56
D.1 Drained models (Tension)

50

45

40

35

30

25

20

15 d = 5.8
d=7
d = 8.4
10 d = 9.7
d = 10.9
d = 12.2
5 d = 13.5
d = 14.8

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 5.8
d=7
0.4
d = 8.4
d = 9.7
d = 10.9
0.2 d = 12.2
d = 13.5
d = 14.8
0
0 0.5 1 1.5 2 2.5 3 3.5

57
Chapter D. Statistical analysis, trimmed and normalized t-z curves

50

45

40

35

30

25

20

15 d = 5.8
d=7
d = 8.4
10 d = 9.7
d = 10.9
d = 12.2
5 d = 13.5
d = 14.8

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 5.8
d=7
0.4
d = 8.4
d = 9.7
d = 10.9
0.2 d = 12.2
d = 13.5
d = 14.8
0
0 0.5 1 1.5 2 2.5 3 3.5 4

58
D.1 Drained models (Tension)

50

45

40

35

30

25

20

15 d = 5.8
d=7
d = 8.4
10 d = 9.7
d = 10.9
d = 12.2
5 d = 13.5
d = 14.8

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 5.8
d=7
0.4
d = 8.4
d = 9.7
d = 10.9
0.2 d = 12.2
d = 13.5
d = 14.8
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

59
Chapter D. Statistical analysis, trimmed and normalized t-z curves

50

45

40

35

30

25

20

15 d = 6.5
d = 7.9
d = 9.3
10 d = 10.7
d = 12.1
d = 13.6
5 d = 15
d = 16.4

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 6.5
d = 7.9
0.4
d = 9.3
d = 10.7
d = 12.1
0.2 d = 13.6
d = 15
d = 16.4
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4

60
D.1 Drained models (Tension)

50

45

40

35

30

25

20

15 d = 6.5
d = 7.9
d = 9.3
10 d = 10.7
d = 12.1
d = 13.6
5 d = 15
d = 16.4

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 6.5
d = 7.9
0.4
d = 9.3
d = 10.7
d = 12.1
0.2 d = 13.6
d = 15
d = 16.4
0
0 0.5 1 1.5 2 2.5

61
Chapter D. Statistical analysis, trimmed and normalized t-z curves

60

50

40

30

20
d = 6.5
d = 7.9
d = 9.3
d = 10.7
10 d = 12.1
d = 13.6
d = 15
d = 16.4

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 6.5
d = 7.9
0.4
d = 9.3
d = 10.7
d = 12.1
0.2 d = 13.6
d = 15
d = 16.4
0
0 0.5 1 1.5 2 2.5 3

62
D.1 Drained models (Tension)

60

50

40

30

20
d = 6.5
d = 7.9
d = 9.3
d = 10.7
10 d = 12.1
d = 13.6
d = 15
d = 16.4

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 6.5
d = 7.9
0.4
d = 9.3
d = 10.7
d = 12.1
0.2 d = 13.6
d = 15
d = 16.4
0
0 0.5 1 1.5 2 2.5 3 3.5

63
Chapter D. Statistical analysis, trimmed and normalized t-z curves

60

50

40

30

20
d = 6.5
d = 7.9
d = 9.3
d = 10.7
10 d = 12.1
d = 13.6
d = 15
d = 16.4

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 6.5
d = 7.9
0.4
d = 9.3
d = 10.7
d = 12.1
0.2 d = 13.6
d = 15
d = 16.4
0
0 0.5 1 1.5 2 2.5 3 3.5 4

64
D.2 Partially drained models (Tension)

D.2 Partially drained models (Tension)

D.2.1 Peak shear stress formulation

ϕ
33◦ 35◦ 38◦ 40◦ 33◦ 35◦ 38◦ 40◦
D
10 m 0.0464 0.0947 0.0729 0.0886 1.2397 1.0972 1.1972 1.1662
13 m 0.0493 0.0629 0.0848 0.0791 1.2123 1.1816 1.1505 1.1801
15 m 0.0474 0.0623 0.0680 0.0836 1.2143 1.1761 1.1903 1.1658
18 m 0.0410 0.0566 0.0655 0.0848 1.2401 1.1933 1.1917 1.1532
20 m 0.0446 0.0638 0.0765 0.0818 1.2173 1.1633 1.1582 1.1588

Table D.2: Values of a-coefficients (left section) and b-coefficients (right section) used for regression analysis
in the context of formulating the peak shear stress.

NRMSE = 16.66 %
0.1

0.09

0.08

0.07

0.06

0.05

0.04
0.9
1.4
0.8
1.2
0.7 1
0.8
0.6 0.6

Figure D.17: Fitting a-coefficients with a quadric surface.

65
Chapter D. Statistical analysis, trimmed and normalized t-z curves

NRMSE = 19.86 %

1.25

1.2

1.15

1.1

1.05
0.9
1.4
0.8
1.2
0.7 1
0.8
0.6 0.6

Figure D.18: Fitting b-coefficients with a quadric surface.

18

16

14

12

10

0
0 20 40 60 80 100 120

Figure D.19: FE data, power fit and mathematical model for each FE model.

66
D.2 Partially drained models (Tension)

25

20

15

10

0
0 20 40 60 80 100 120 140

Figure D.20: FE data, power fit and mathematical model for each FE model.

30

25

20

15

10

0
0 20 40 60 80 100 120 140 160

Figure D.21: FE data, power fit and mathematical model for each FE model.

67
Chapter D. Statistical analysis, trimmed and normalized t-z curves

35

30

25

20

15

10

0
0 20 40 60 80 100 120 140 160 180 200

Figure D.22: FE data, power fit and mathematical model for each FE model.

40

35

30

25

20

15

10

0
0 50 100 150 200 250

Figure D.23: FE data, power fit and mathematical model for each FE model.

68
D.2 Partially drained models (Tension)

D.2.2 Peak displacement formulation

ϕ
33◦ 35◦ 38◦ 40◦ 33◦ 35◦ 38◦ 40◦
D
10 m 3.99 5.96 4.06 3.71 602.79 534.15 569.69 537.1 ×10−3
13 m 4.15 3.16 2.5 2.08 588.75 597.93 577.49 633.81 ×10−3
15 m 4.16 3.4 2.32 1.88 593.2 566.34 604.46 630.25 ×10−3
18 m 4.87 4.39 2.86 2.55 637.38 575.83 615.24 592.02 ×10−3
20 m 5.5 5.2 2.92 2.72 586.66 515.13 610.28 573.9 ×10−3

Table D.3: Values of c-coefficients (left section) and d-coefficients (right section) used for regression analysis
in the context of formulating the peak displacement.

NRMSE = 9.21 %
10-3
6

1
0.9
1.4
0.8
1.2
0.7 1
0.8
0.6 0.6

Figure D.24: Fitting c-coefficients with a quadric surface.

69
Chapter D. Statistical analysis, trimmed and normalized t-z curves

NRMSE = 26.48 %

0.65

0.6

0.55

0.5
0.9
1.4
0.8
1.2
0.7 1
0.8
0.6 0.6

Figure D.25: Fitting d-coefficients with a plane. NOTE! The average value of d is used in the formulation
for undrained models in tension.

0.018

0.016

0.014

0.012

0.01

0.008

0.006

0.004

0.002

0
0 1 2 3 4 5 6 7 8

Figure D.26: FE data, power fit and mathematical model for each FE model.

70
D.2 Partially drained models (Tension)

0.02

0.018

0.016

0.014

0.012

0.01

0.008

0.006

0.004

0.002

0
0 2 4 6 8 10 12 14

Figure D.27: FE data, power fit and mathematical model for each FE model.

0.025

0.02

0.015

0.01

0.005

0
0 2 4 6 8 10 12 14 16 18

Figure D.28: FE data, power fit and mathematical model for each FE model.

71
Chapter D. Statistical analysis, trimmed and normalized t-z curves

0.035

0.03

0.025

0.02

0.015

0.01

0.005

0
0 5 10 15 20 25

Figure D.29: FE data, power fit and mathematical model for each FE model.

0.04

0.035

0.03

0.025

0.02

0.015

0.01

0.005

0
0 5 10 15 20 25 30 35

Figure D.30: FE data, power fit and mathematical model for each FE model.

72
D.2 Partially drained models (Tension)

D.2.3 Trimmed and normalized t-z curves

25
d = 2.1
d = 2.9
d = 3.8
d = 4.6
20 d = 5.5
d = 6.3
d = 7.1
d = 7.9

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.8

1.6

1.4

1.2

0.8 d = 2.1
d = 2.9
0.6 d = 3.8
d = 4.6
0.4 d = 5.5
d = 6.3
0.2 d = 7.1
d = 7.9
0
0 1 2 3 4 5 6 7 8

73
Chapter D. Statistical analysis, trimmed and normalized t-z curves

25
d = 2.1
d = 2.9
d = 3.8
d = 4.6
20 d = 5.5
d = 6.3
d = 7.1
d = 7.9

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.8

1.6

1.4

1.2

0.8
d = 2.1
d = 2.9
0.6
d = 3.8
d = 4.6
0.4 d = 5.5
d = 6.3
0.2 d = 7.1
d = 7.9
0
0 1 2 3 4 5 6 7 8 9

74
D.2 Partially drained models (Tension)

25
d=2
d = 2.9
d = 3.8
d = 4.5
20 d = 5.4
d = 6.3
d=7
d = 7.9

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.5

d=2
d = 2.9
0.5
d = 3.8
d = 4.5
d = 5.4
d = 6.3
d=7
d = 7.9
0
0 1 2 3 4 5 6 7 8 9 10

75
Chapter D. Statistical analysis, trimmed and normalized t-z curves

25
d=2
d = 2.9
d = 3.8
d = 4.5
20 d = 5.4
d = 6.3
d=7
d = 7.9

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.6

1.4

1.2

0.8

d=2
0.6 d = 2.9
d = 3.8
0.4 d = 4.5
d = 5.4
d = 6.3
0.2 d=7
d = 7.9
0
0 2 4 6 8 10 12

76
D.2 Partially drained models (Tension)

25
d=2
d = 2.9
d = 3.8
d = 4.5
20 d = 5.4
d = 6.3
d=7
d = 7.9

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.6

1.4

1.2

0.8

d=2
0.6 d = 2.9
d = 3.8
0.4 d = 4.5
d = 5.4
d = 6.3
0.2 d=7
d = 7.9
0
0 5 10 15

77
Chapter D. Statistical analysis, trimmed and normalized t-z curves

25
d = 2.7
d = 3.7
d = 4.9
d = 5.9
20 d=7
d = 8.1
d = 9.2
d = 10.2

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.5

d = 2.7
d = 3.7
0.5
d = 4.9
d = 5.9
d=7
d = 8.1
d = 9.2
d = 10.2
0
0 1 2 3 4 5 6

78
D.2 Partially drained models (Tension)

25
d = 2.7
d = 3.8
d = 4.9
d = 5.9
20 d=7
d = 8.1
d = 9.2
d = 10.3

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.4

1.2

0.8

0.6
d = 2.7
d = 3.8
d = 4.9
0.4
d = 5.9
d=7
d = 8.1
0.2
d = 9.2
d = 10.3
0
0 1 2 3 4 5 6

79
Chapter D. Statistical analysis, trimmed and normalized t-z curves

30
d = 2.7
d = 3.8
d = 4.9
25 d = 5.9
d=7
d = 8.1
d = 9.2
d = 10.3
20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.4

1.2

0.8

0.6
d = 2.7
d = 3.8
d = 4.9
0.4
d = 5.9
d=7
d = 8.1
0.2
d = 9.2
d = 10.3
0
0 1 2 3 4 5 6 7

80
D.2 Partially drained models (Tension)

30
d = 2.7
d = 3.8
d = 4.9
25 d = 5.9
d=7
d = 8.1
d = 9.2
d = 10.3
20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.4

1.2

0.8

0.6
d = 2.7
d = 3.8
d = 4.9
0.4
d = 5.9
d=7
d = 8.1
0.2
d = 9.2
d = 10.3
0
0 1 2 3 4 5 6 7 8 9

81
Chapter D. Statistical analysis, trimmed and normalized t-z curves

30
d = 2.7
d = 3.7
d = 4.9
25 d = 5.9
d=7
d = 8.1
d = 9.2
d = 10.2
20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.4

1.2

0.8

d = 2.7
d = 3.7
0.6
d = 4.9
d = 5.9
d=7
0.4 d = 8.1
d = 9.2
d = 10.2
0.2
0 2 4 6 8 10 12

82
D.2 Partially drained models (Tension)

25
d=3
d = 4.3
d = 5.6
d = 6.8
20 d = 8.1
d = 9.4
d = 10.5
d = 11.8

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.4

1.2

0.8

0.6
d=3
d = 4.3
d = 5.6
0.4
d = 6.8
d = 8.1
d = 9.4
0.2
d = 10.5
d = 11.8
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

83
Chapter D. Statistical analysis, trimmed and normalized t-z curves

30
d = 3.1
d = 4.4
d = 5.7
25 d = 6.9
d = 8.2
d = 9.4
d = 10.6
d = 11.9
20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.4

1.2

0.8

0.6
d = 3.1
d = 4.4
d = 5.7
0.4
d = 6.9
d = 8.2
d = 9.4
0.2
d = 10.6
d = 11.9
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

84
D.2 Partially drained models (Tension)

30
d = 3.1
d = 4.4
d = 5.7
25 d = 6.9
d = 8.2
d = 9.4
d = 10.6
d = 11.9
20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 3.1
d = 4.4
0.4
d = 5.7
d = 6.9
d = 8.2
0.2 d = 9.4
d = 10.6
d = 11.9
0
0 1 2 3 4 5 6

85
Chapter D. Statistical analysis, trimmed and normalized t-z curves

30
d = 3.1
d = 4.4
d = 5.7
25 d = 6.9
d = 8.2
d = 9.4
d = 10.6
d = 11.9
20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.4

1.2

0.8

0.6
d = 3.1
d = 4.4
d = 5.7
0.4
d = 6.9
d = 8.2
d = 9.4
0.2
d = 10.6
d = 11.9
0
0 1 2 3 4 5 6 7

86
D.2 Partially drained models (Tension)

35
d = 3.1
d = 4.4
d = 5.7
30 d = 6.9
d = 8.2
d = 9.4
d = 10.6
25 d = 11.9

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.4

1.2

0.8

0.6
d = 3.1
d = 4.4
d = 5.7
0.4
d = 6.9
d = 8.2
d = 9.4
0.2
d = 10.6
d = 11.9
0
0 1 2 3 4 5 6 7 8 9

87
Chapter D. Statistical analysis, trimmed and normalized t-z curves

25
d = 3.7
d = 5.2
d = 6.8
d = 8.2
20 d = 9.7
d = 11.3
d = 12.7
d = 14.2

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.4

1.2

0.8

0.6
d = 3.7
d = 5.2
d = 6.8
0.4
d = 8.2
d = 9.7
d = 11.3
0.2
d = 12.7
d = 14.2
0
0 0.5 1 1.5 2 2.5 3 3.5

88
D.2 Partially drained models (Tension)

30
d = 3.7
d = 5.2
d = 6.8
25 d = 8.2
d = 9.7
d = 11.3
d = 12.7
d = 14.2
20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 3.7
d = 5.2
0.4
d = 6.8
d = 8.2
d = 9.7
0.2 d = 11.3
d = 12.7
d = 14.2
0
0 0.5 1 1.5 2 2.5 3 3.5 4

89
Chapter D. Statistical analysis, trimmed and normalized t-z curves

35
d = 3.7
d = 5.2
d = 6.8
30 d = 8.2
d = 9.7
d = 11.3
d = 12.7
25 d = 14.2

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 3.7
d = 5.2
0.4
d = 6.8
d = 8.2
d = 9.7
0.2 d = 11.3
d = 12.7
d = 14.2
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

90
D.2 Partially drained models (Tension)

35
d = 3.7
d = 5.2
d = 6.8
30 d = 8.2
d = 9.7
d = 11.3
d = 12.7
25 d = 14.2

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.1

0.9

0.8

0.7

0.6

0.5 d = 3.7
d = 5.2
0.4 d = 6.8
d = 8.2
0.3 d = 9.7
d = 11.3
0.2 d = 12.7
d = 14.2
0.1
0 1 2 3 4 5 6

91
Chapter D. Statistical analysis, trimmed and normalized t-z curves

35
d = 3.7
d = 5.2
d = 6.8
30 d = 8.2
d = 9.7
d = 11.3
d = 12.7
25 d = 14.2

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 3.7
d = 5.2
0.4
d = 6.8
d = 8.2
d = 9.7
0.2 d = 11.3
d = 12.7
d = 14.2
0
0 1 2 3 4 5 6 7

92
D.2 Partially drained models (Tension)

25
d = 4.2
d = 5.8
d = 7.6
d = 9.2
20 d = 10.7
d = 12.5
d = 14.2
d = 15.7

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.2
d = 5.8
0.4
d = 7.6
d = 9.2
d = 10.7
0.2 d = 12.5
d = 14.2
d = 15.7
0
0 0.5 1 1.5 2 2.5 3

93
Chapter D. Statistical analysis, trimmed and normalized t-z curves

35
d = 4.2
d = 5.8
d = 7.6
30 d = 9.2
d = 10.7
d = 12.5
d = 14.2
25 d = 15.7

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.2
d = 5.8
0.4
d = 7.6
d = 9.2
d = 10.7
0.2 d = 12.5
d = 14.2
d = 15.7
0
0 0.5 1 1.5 2 2.5 3 3.5

94
D.2 Partially drained models (Tension)

35
d = 4.2
d = 5.8
d = 7.6
30 d = 9.2
d = 10.7
d = 12.5
d = 14.2
25 d = 15.7

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.2

0.8

0.6

d = 4.2
d = 5.8
0.4
d = 7.6
d = 9.2
d = 10.7
0.2 d = 12.5
d = 14.2
d = 15.7
0
0 0.5 1 1.5 2 2.5 3 3.5 4

95
Chapter D. Statistical analysis, trimmed and normalized t-z curves

40
d = 4.2
d = 5.8
35 d = 7.6
d = 9.2
d = 10.7
d = 12.5
30 d = 14.2
d = 15.7

25

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.1

0.9

0.8

0.7

0.6

0.5 d = 4.2
d = 5.8
0.4 d = 7.6
d = 9.2
0.3 d = 10.7
d = 12.5
0.2 d = 14.2
d = 15.7
0.1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

96
D.2 Partially drained models (Tension)

40
d = 4.2
d = 5.8
35 d = 7.6
d = 9.2
d = 10.7
d = 12.5
30 d = 14.2
d = 15.7

25

20

15

10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05

1.1

0.9

0.8

0.7

0.6

0.5 d = 4.2
d = 5.8
0.4 d = 7.6
d = 9.2
0.3 d = 10.7
d = 12.5
0.2 d = 14.2
d = 15.7
0.1
0 1 2 3 4 5 6

97
Chapter D. Statistical analysis, trimmed and normalized t-z curves

D.3 Drained models (Compression)

ϕ
30◦ 33◦ 35◦ 38◦ 40◦ 30◦ 33◦ 35◦ 38◦ 40◦
D
10 m 0.6780 0.8532 0.9526 1.8339 1.6053 0.7426 0.8046 0.8235 0.7400 0.7612
13 m 0.4294 0.8601 1.0788 1.3660 1.4406 0.6325 0.7320 0.7409 0.7351 0.7590
15 m 0.3695 0.6285 1.0016 1.2432 1.4663 0.6355 0.6427 0.7205 0.7180 0.7196
18 m 0.1984 0.5047 1.0195 1.2034 1.2328 0.6411 0.6512 0.6322 0.7002 0.7281
20 m 0.1789 0.2810 0.8958 1.1867 1.2172 0.6409 0.6651 0.6382 0.6377 0.7232

Table D.4: Values of a-coefficients (left section) and b-coefficients (right section) used for regression analysis
in the context of formulating the shear stress at maximum curvature

NRMSE = 8.2 %

1.5

0.5

0
0.9
0.8 1.4
0.7 1.2
1
0.6
0.8
0.5 0.6

Figure D.31: Fitting a-coefficients with a quadric surface.

98
D.3 Drained models (Compression)

NRMSE = 18 %
0.85

0.8

0.75

0.7

0.65

0.6

0.55
0.9
0.8 1.4
0.7 1.2
1
0.6
0.8
0.5 0.6

Figure D.32: Fitting b-coefficients with a quadric surface.

The expressions for fitting coefficients are established as:

   
D D
a = α1 + α2 + α3 tan ϕ + α4 tan ϕ + α5 tan2 ϕ (D.1)
Dre f Dre f
   
D D
b = β1 + β2 + β3 tan ϕ + β4 tan ϕ + β5 tan2 ϕ (D.2)
Dre f Dre f

where:

α1 α2 α3 α4 α5
-4.6423 -0.8192 13.7692 0.2501 -6.8894
β1 β2 β3 β4 β5
0.5868 -0.3899 0.8862 0.3148 -0.672

99
Chapter D. Statistical analysis, trimmed and normalized t-z curves

60

50

40

30

20

10

0
0 20 40 60 80 100 120

Figure D.33: FE data, power fit and mathematical model for each FE model.

60

50

40

30

20

10

0
0 20 40 60 80 100 120 140

Figure D.34: FE data, power fit and mathematical model for each FE model.

100
D.3 Drained models (Compression)

50

45

40

35

30

25

20

15

10

0
0 20 40 60 80 100 120 140 160

Figure D.35: FE data, power fit and mathematical model for each FE model.

50

45

40

35

30

25

20

15

10

0
0 20 40 60 80 100 120 140 160 180 200

Figure D.36: FE data, power fit and mathematical model for each FE model.

101
Chapter D. Statistical analysis, trimmed and normalized t-z curves

60

50

40

30

20

10

0
0 50 100 150 200 250

Figure D.37: FE data, power fit and mathematical model for each FE model.

102
D.3 Drained models (Compression)

D.3.1 Trimmed and normalized t-z curves

60
d = 2.8
d = 3.4
d = 4.1
50 d = 4.7
d = 5.3
d=6
d = 6.6
d = 7.2
40

30

20

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2
d = 2.8
d = 3.4
1.5
d = 4.1
d = 4.7
1 d = 5.3
d=6
0.5 d = 6.6
d = 7.2
0
0 0.005 0.01 0.015 0.02 0.025 0.03

103
Chapter D. Statistical analysis, trimmed and normalized t-z curves

90
d = 2.8
d = 3.4
80 d = 4.1
d = 4.7
d = 5.3
70 d=6
d = 6.6
d = 7.2
60

50

40

30

20

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3

3.5

2.5

d = 2.8
1.5 d = 3.4
d = 4.1
1 d = 4.7
d = 5.3
d=6
0.5 d = 6.6
d = 7.2
0
0 0.005 0.01 0.015 0.02 0.025 0.03

104
D.3 Drained models (Compression)

120
d = 2.8
d = 3.4
d = 4.1
100 d = 4.7
d = 5.3
d=6
d = 6.6
d = 7.2
80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3

3.5

2.5

d = 2.8
1.5 d = 3.4
d = 4.1
1 d = 4.7
d = 5.3
d=6
0.5 d = 6.6
d = 7.2
0
0 0.005 0.01 0.015 0.02 0.025 0.03

105
Chapter D. Statistical analysis, trimmed and normalized t-z curves

150
d = 2.8
d = 3.5
d = 4.1
d = 4.6
d = 5.3
d=6
d = 6.6
d = 7.2
100

50

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2 d = 2.8
d = 3.5
1.5 d = 4.1
d = 4.6
1 d = 5.3
d=6
0.5 d = 6.6
d = 7.2
0
0 0.005 0.01 0.015 0.02 0.025 0.03

106
D.3 Drained models (Compression)

140
d = 2.8
d = 3.5
d = 4.1
120 d = 4.7
d = 5.3
d = 5.9
d = 6.6
100 d = 7.2

80

60

40

20

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2

4.5

3.5

2.5

2
d = 2.8
d = 3.5
1.5
d = 4.1
d = 4.7
1 d = 5.3
d = 5.9
0.5 d = 6.6
d = 7.2
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

107
Chapter D. Statistical analysis, trimmed and normalized t-z curves

70
d = 3.7
d = 4.5
d = 5.3
60 d = 6.1
d = 6.9
d = 7.7
d = 8.6
50 d = 9.3

40

30

20

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3

d = 3.7
d = 4.5
2
d = 5.3
d = 6.1
d = 6.9
1 d = 7.7
d = 8.6
d = 9.3
0
0 0.005 0.01 0.015 0.02 0.025

108
D.3 Drained models (Compression)

90
d = 3.7
d = 4.5
80 d = 5.3
d = 6.1
d = 6.9
70 d = 7.7
d = 8.6
d = 9.3
60

50

40

30

20

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2
d = 3.7
d = 4.5
1.5
d = 5.3
d = 6.1
1 d = 6.9
d = 7.7
0.5 d = 8.6
d = 9.3
0
0 0.005 0.01 0.015 0.02 0.025

109
Chapter D. Statistical analysis, trimmed and normalized t-z curves

120
d = 3.7
d = 4.5
d = 5.3
100 d = 6.1
d = 6.9
d = 7.7
d = 8.6
d = 9.3
80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2
d = 3.7
d = 4.5
1.5
d = 5.3
d = 6.1
1 d = 6.9
d = 7.7
0.5 d = 8.6
d = 9.3
0
0 0.005 0.01 0.015 0.02 0.025

110
D.3 Drained models (Compression)

150
d = 3.7
d = 4.5
d = 5.3
d = 6.1
d = 6.9
d = 7.7
d = 8.6
d = 9.3
100

50

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2
d = 3.7
d = 4.5
1.5
d = 5.3
d = 6.1
1 d = 6.9
d = 7.7
0.5 d = 8.6
d = 9.3
0
0 0.005 0.01 0.015 0.02 0.025

111
Chapter D. Statistical analysis, trimmed and normalized t-z curves

180
d = 3.7
d = 4.5
160 d = 5.3
d = 6.1
d = 6.9
140 d = 7.7
d = 8.6
d = 9.3
120

100

80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2 d = 3.7
d = 4.5
1.5 d = 5.3
d = 6.1
1 d = 6.9
d = 7.7
0.5 d = 8.6
d = 9.3
0
0 0.005 0.01 0.015 0.02 0.025

112
D.3 Drained models (Compression)

70
d = 4.2
d = 5.2
d = 6.1
60 d=7
d = 7.9
d = 8.9
d = 9.8
50 d = 10.8

40

30

20

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3

d = 4.2
3 d = 5.2
d = 6.1
2 d=7
d = 7.9
d = 8.9
1 d = 9.8
d = 10.8
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

113
Chapter D. Statistical analysis, trimmed and normalized t-z curves

100
d = 4.2
d = 5.2
90
d = 6.1
d=7
80 d = 7.9
d = 8.9
d = 9.8
70 d = 10.8

60

50

40

30

20

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2 d = 4.2
d = 5.2
1.5 d = 6.1
d=7
1 d = 7.9
d = 8.9
0.5 d = 9.8
d = 10.8
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

114
D.3 Drained models (Compression)

120
d = 4.2
d = 5.2
d = 6.1
100 d=7
d = 7.9
d = 8.9
d = 9.8
d = 10.8
80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2
d = 4.2
d = 5.2
1.5
d = 6.1
d=7
1 d = 7.9
d = 8.9
0.5 d = 9.8
d = 10.8
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

115
Chapter D. Statistical analysis, trimmed and normalized t-z curves

150
d = 4.2
d = 5.2
d = 6.1
d=7
d = 7.9
d = 8.9
d = 9.8
d = 10.8
100

50

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2 d = 4.2
d = 5.2
1.5 d = 6.1
d=7
1 d = 7.9
d = 8.9
0.5 d = 9.8
d = 10.8
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

116
D.3 Drained models (Compression)

180
d = 4.2
d = 5.2
160 d = 6.1
d=7
d = 7.9
140 d = 8.9
d = 9.8
d = 10.8
120

100

80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2 d = 4.2
d = 5.2
1.5 d = 6.1
d=7
1 d = 7.9
d = 8.9
0.5 d = 9.8
d = 10.8
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

117
Chapter D. Statistical analysis, trimmed and normalized t-z curves

80
d = 5.1
d = 6.2
70 d = 7.3
d = 8.5
d = 9.6
d = 10.7
60 d = 11.8
d = 12.9

50

40

30

20

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3

14

12

10

6
d = 5.1
d = 6.2
d = 7.3
4
d = 8.5
d = 9.6
d = 10.7
2
d = 11.8
d = 12.9
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

118
D.3 Drained models (Compression)

100
d = 5.1
d = 6.2
90
d = 7.3
d = 8.5
80 d = 9.6
d = 10.7
d = 11.8
70 d = 12.9

60

50

40

30

20

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3

d = 5.1
d = 6.2
2
d = 7.3
d = 8.5
d = 9.6
1 d = 10.7
d = 11.8
d = 12.9
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

119
Chapter D. Statistical analysis, trimmed and normalized t-z curves

120
d = 5.1
d = 6.2
d = 7.3
100 d = 8.5
d = 9.6
d = 10.7
d = 11.8
d = 12.9
80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3

d = 5.1
d = 6.2
2
d = 7.3
d = 8.5
d = 9.6
1 d = 10.7
d = 11.8
d = 12.9
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

120
D.3 Drained models (Compression)

150
d = 5.1
d = 6.2
d = 7.3
d = 8.5
d = 9.6
d = 10.7
d = 11.8
d = 12.9
100

50

0
0 0.05 0.1 0.15 0.2 0.25 0.3

4.5

3.5

2.5

2 d = 5.1
d = 6.2
1.5 d = 7.3
d = 8.5
1 d = 9.6
d = 10.7
0.5 d = 11.8
d = 12.9
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

121
Chapter D. Statistical analysis, trimmed and normalized t-z curves

180
d = 5.1
d = 6.2
160 d = 7.3
d = 8.5
d = 9.6
140 d = 10.7
d = 11.8
d = 12.9
120

100

80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3

d = 5.1
d = 6.2
2
d = 7.3
d = 8.5
d = 9.6
1 d = 10.7
d = 11.8
d = 12.9
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

122
D.3 Drained models (Compression)

80
d = 5.6
d = 6.9
70 d = 8.1
d = 9.4
d = 10.6
d = 11.9
60 d = 13.1
d = 14.4

50

40

30

20

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3

30

25

20

15

d = 5.6
d = 6.9
10
d = 8.1
d = 9.4
d = 10.6
5 d = 11.9
d = 13.1
d = 14.4
0
0 0.005 0.01 0.015

123
Chapter D. Statistical analysis, trimmed and normalized t-z curves

100
d = 5.6
d = 6.9
90
d = 8.1
d = 9.4
80 d = 10.6
d = 11.9
d = 13.1
70 d = 14.4

60

50

40

30

20

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3

d = 5.6
3 d = 6.9
d = 8.1
2 d = 9.4
d = 10.6
d = 11.9
1 d = 13.1
d = 14.4
0
0 0.005 0.01 0.015

124
D.3 Drained models (Compression)

120
d = 5.6
d = 6.9
d = 8.1
100 d = 9.4
d = 10.6
d = 11.9
d = 13.1
d = 14.4
80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3

d = 5.6
d = 6.9
2
d = 8.1
d = 9.4
d = 10.6
1 d = 11.9
d = 13.1
d = 14.4
0
0 0.005 0.01 0.015

125
Chapter D. Statistical analysis, trimmed and normalized t-z curves

150
d = 5.6
d = 6.9
d = 8.1
d = 9.4
d = 10.6
d = 11.9
d = 13.1
d = 14.4
100

50

0
0 0.05 0.1 0.15 0.2 0.25 0.3

d = 5.6
d = 6.9
2
d = 8.1
d = 9.4
d = 10.6
1 d = 11.9
d = 13.1
d = 14.4
0
0 0.005 0.01 0.015

126
D.3 Drained models (Compression)

180
d = 5.6
d = 6.9
160 d = 8.1
d = 9.4
d = 10.6
140 d = 11.9
d = 13.1
d = 14.4
120

100

80

60

40

20

0
0 0.05 0.1 0.15 0.2 0.25 0.3

d = 5.6
d = 6.9
2
d = 8.1
d = 9.4
d = 10.6
1 d = 11.9
d = 13.1
d = 14.4
0
0 0.005 0.01 0.015

127
Chapter D. Statistical analysis, trimmed and normalized t-z curves

D.4 Partially drained models (Compression)

30

25

20

15

10

5
20 40 60 80 100 120 140 160

Figure D.38: Shear stress at maximum curvature for partially drained models in compression. Data set for
each model is fitted with a power function f (x) = axb .

ϕ
33◦ 35◦ 38◦ 40◦ 33◦ 35◦ 38◦ 40◦
D
10 m 0.3514 0.3794 0.5493 0.7328 0.7995 0.8073 0.7748 0.7373
13 m 0.3656 0.4511 0.4625 0.5309 0.8070 0.7920 0.8241 0.8150
15 m 0.3178 0.3664 0.4160 0.5203 0.8324 0.8339 0.8444 0.8252
18 m 0.3331 0.3718 0.4304 0.5039 0.8319 0.8377 0.8455 0.8369
20 m 0.4412 0.3578 0.4225 0.4793 0.7595 0.8352 0.8378 0.8457

Table D.5: Values of a-coefficients (left section) and b-coefficients (right section) used for regression analysis
in the context of formulating the stress at maximum curvature.

128
D.4 Partially drained models (Compression)

NRMSE = 10.39 %
0.8

0.7

0.6

0.5

0.4

0.3
0.9
1.4
0.8
1.2
0.7 1
0.8
0.6 0.6

Figure D.39: Fitting a-coefficients with a quadric surface.

NRMSE = 20.13 %
0.9

0.85

0.8

0.75

0.7
0.9
1.4
0.8
1.2
0.7 1
0.8
0.6 0.6

Figure D.40: Fitting b-coefficients with a quadric surface.

129
Chapter D. Statistical analysis, trimmed and normalized t-z curves

The expressions for fitting coefficients are established as:

   
D D
a = α1 + α2 + α3 tan ϕ + α4 tan ϕ + α5 tan2 ϕ (D.3)
Dre f Dre f
   
D D
b = β1 + β2 + β3 tan ϕ + β4 tan ϕ + β5 tan2 ϕ (D.4)
Dre f Dre f

where:

α1 α2 α3 α4 α5
0.808 1.3935 -3.7744 -2.0392 4.5917
β1 β2 β3 β4 β5
0.2842 -0.5567 2.091 0.8392 -1.9543

16

14

12

10

0
0 10 20 30 40 50 60 70

Figure D.41: FE data, power fit and mathematical model for each FE model.

130
D.4 Partially drained models (Compression)

20

18

16

14

12

10

0
0 10 20 30 40 50 60 70 80 90 100

Figure D.42: FE data, power fit and mathematical model for each FE model.

25

20

15

10

0
0 20 40 60 80 100 120

Figure D.43: FE data, power fit and mathematical model for each FE model.

131
Chapter D. Statistical analysis, trimmed and normalized t-z curves

30

25

20

15

10

0
0 20 40 60 80 100 120 140

Figure D.44: FE data, power fit and mathematical model for each FE model.

30

25

20

15

10

0
0 50 100 150

Figure D.45: FE data, power fit and mathematical model for each FE model.

132
D.4 Partially drained models (Compression)

D.4.1 Trimmed and normalized t-z curves

12

10

d = 2.3
4
d = 2.7
d = 3.3
d = 3.7
2 d = 4.2
d = 4.7
d = 5.3
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.4

1.2

0.8

d = 2.3
0.6
d = 2.7
d = 3.3
d = 3.7
0.4 d = 4.2
d = 4.7
d = 5.3
0.2
0 0.005 0.01 0.015 0.02 0.025 0.03

133
Chapter D. Statistical analysis, trimmed and normalized t-z curves

14

12

10

6
d = 2.3
d = 2.8
4
d = 3.3
d = 3.8
d = 4.3
2
d = 4.7
d = 5.3
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.4

1.2

0.8

d = 2.3
0.6
d = 2.8
d = 3.3
d = 3.8
0.4 d = 4.3
d = 4.7
d = 5.3
0.2
0 0.005 0.01 0.015 0.02 0.025 0.03

134
D.4 Partially drained models (Compression)

16

14

12

10

6 d = 2.3
d = 2.8
4 d = 3.3
d = 3.8
d = 4.3
2 d = 4.7
d = 5.3
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.8

1.6

1.4

1.2

0.8 d = 2.3
d = 2.8
0.6 d = 3.3
d = 3.8
d = 4.3
0.4 d = 4.7
d = 5.3
0.2
0 0.005 0.01 0.015 0.02 0.025 0.03

135
Chapter D. Statistical analysis, trimmed and normalized t-z curves

25

20

15

10
d = 2.2
d = 2.7
d = 3.3
5 d = 3.8
d = 4.2
d = 4.7
d = 5.2
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35

1.8

1.6

1.4

1.2

1 d = 2.2
d = 2.7
0.8 d = 3.3
d = 3.8
d = 4.2
0.6 d = 4.7
d = 5.2
0.4
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035

136
D.4 Partially drained models (Compression)

15

10

d=3
5
d = 3.6
d = 4.3
d = 4.9
d = 5.6
d = 6.2
d = 6.8
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.4

1.2

0.8

d=3
0.6
d = 3.6
d = 4.3
d = 4.9
0.4 d = 5.6
d = 6.2
d = 6.8
0.2
0 0.005 0.01 0.015 0.02 0.025

137
Chapter D. Statistical analysis, trimmed and normalized t-z curves

18

16

14

12

10

d=3
6
d = 3.6
d = 4.3
4 d = 4.9
d = 5.6
2 d = 6.2
d = 6.8
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.6

1.4

1.2

0.8
d=3
d = 3.6
0.6
d = 4.3
d = 4.9
d = 5.6
0.4
d = 6.2
d = 6.8
0.2
0 0.005 0.01 0.015 0.02 0.025

138
D.4 Partially drained models (Compression)

25

20

15

10
d=3
d = 3.6
d = 4.3
5 d = 4.9
d = 5.6
d = 6.2
d = 6.8
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.8

1.6

1.4

1.2

0.8 d=3
d = 3.6
0.6 d = 4.3
d = 4.9
d = 5.6
0.4 d = 6.2
d = 6.8
0.2
0 0.005 0.01 0.015 0.02 0.025

139
Chapter D. Statistical analysis, trimmed and normalized t-z curves

30

25

20

15

d=3
10
d = 3.6
d = 4.3
d = 4.9
5 d = 5.6
d = 6.2
d = 6.8
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.8

1.6

1.4

1.2

1 d=3
d = 3.6
0.8 d = 4.3
d = 4.9
d = 5.6
0.6 d = 6.2
d = 6.8
0.4
0 0.005 0.01 0.015 0.02 0.025

140
D.4 Partially drained models (Compression)

18

16

14

12

10

d = 3.4
6
d = 4.1
d = 4.9
4 d = 5.6
d = 6.4
2 d = 7.1
d = 7.9
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.4

1.2

0.8

d = 3.4
0.6
d = 4.1
d = 4.9
d = 5.6
0.4 d = 6.4
d = 7.1
d = 7.9
0.2
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

141
Chapter D. Statistical analysis, trimmed and normalized t-z curves

25

20

15

10
d = 3.4
d = 4.1
d = 4.9
5 d = 5.6
d = 6.4
d = 7.1
d = 7.9
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.6

1.4

1.2

0.8
d = 3.4
d = 4.1
0.6
d = 4.9
d = 5.6
d = 6.4
0.4
d = 7.1
d = 7.9
0.2
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

142
D.4 Partially drained models (Compression)

30

25

20

15

d = 3.4
10
d = 4.1
d = 4.9
d = 5.6
5 d = 6.4
d = 7.1
d = 7.9
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.8

1.6

1.4

1.2

0.8 d = 3.4
d = 4.1
0.6 d = 4.9
d = 5.6
d = 6.4
0.4 d = 7.1
d = 7.9
0.2
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

143
Chapter D. Statistical analysis, trimmed and normalized t-z curves

35

30

25

20

15
d = 3.4
d = 4.1
10
d = 4.9
d = 5.6
d = 6.4
5
d = 7.1
d = 7.9
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.8

1.6

1.4

1.2

1 d = 3.4
d = 4.1
0.8 d = 4.9
d = 5.6
d = 6.4
0.6 d = 7.1
d = 7.9
0.4
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

144
D.4 Partially drained models (Compression)

25

20

15

10
d=4
d = 4.9
d = 5.9
5 d = 6.8
d = 7.6
d = 8.5
d = 9.4
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.4

1.2

0.8

d=4
0.6
d = 4.9
d = 5.9
d = 6.8
0.4 d = 7.6
d = 8.5
d = 9.4
0.2
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

145
Chapter D. Statistical analysis, trimmed and normalized t-z curves

25

20

15

10
d=4
d = 4.9
d = 5.9
5 d = 6.8
d = 7.6
d = 8.5
d = 9.4
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.6

1.4

1.2

0.8
d=4
d = 4.9
0.6
d = 5.9
d = 6.8
d = 7.6
0.4
d = 8.5
d = 9.4
0.2
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

146
D.4 Partially drained models (Compression)

30

25

20

15

d = 4.1
10
d = 4.9
d = 5.8
d = 6.7
5 d = 7.6
d = 8.5
d = 9.4
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.8

1.6

1.4

1.2

0.8 d = 4.1
d = 4.9
0.6 d = 5.8
d = 6.7
d = 7.6
0.4 d = 8.5
d = 9.4
0.2
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

147
Chapter D. Statistical analysis, trimmed and normalized t-z curves

35

30

25

20

15
d=4
d = 4.9
10
d = 5.8
d = 6.8
d = 7.7
5
d = 8.5
d = 9.4
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.8

1.6

1.4

1.2

1
d=4
d = 4.9
0.8
d = 5.8
d = 6.8
d = 7.7
0.6
d = 8.5
d = 9.4
0.4
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

148
D.4 Partially drained models (Compression)

25

20

15

10
d = 4.4
d = 5.5
d = 6.6
5 d = 7.5
d = 8.5
d = 9.5
d = 10.5
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.4

1.2

0.8

d = 4.4
0.6
d = 5.5
d = 6.6
d = 7.5
0.4 d = 8.5
d = 9.5
d = 10.5
0.2
0 0.005 0.01 0.015

149
Chapter D. Statistical analysis, trimmed and normalized t-z curves

30

25

20

15

d = 4.4
10
d = 5.5
d = 6.6
d = 7.5
5 d = 8.5
d = 9.5
d = 10.5
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.6

1.4

1.2

0.8
d = 4.4
d = 5.5
0.6
d = 6.6
d = 7.5
d = 8.5
0.4
d = 9.5
d = 10.5
0.2
0 0.005 0.01 0.015

150
D.4 Partially drained models (Compression)

35

30

25

20

15
d = 4.4
d = 5.5
10
d = 6.6
d = 7.5
d = 8.5
5
d = 9.5
d = 10.5
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.8

1.6

1.4

1.2

0.8 d = 4.4
d = 5.5
0.6 d = 6.6
d = 7.5
d = 8.5
0.4 d = 9.5
d = 10.5
0.2
0 0.005 0.01 0.015

151
Chapter D. Statistical analysis, trimmed and normalized t-z curves

40

35

30

25

20

15 d = 4.5
d = 5.5
10 d = 6.6
d = 7.5
d = 8.4
5 d = 9.4
d = 10.5
0
0 0.05 0.1 0.15 0.2 0.25 0.3

1.8

1.6

1.4

1.2

1
d = 4.5
d = 5.5
0.8
d = 6.6
d = 7.5
d = 8.4
0.6
d = 9.4
d = 10.5
0.4
0 0.005 0.01 0.015

152
A PPENDIX E

Mathematical models

E.1 Drained tension

Peak shear stress


!{β }{P}T
σvo
0
τ p = {α}{P}T D
L tan ϕ

where:
{α} {α1 α2 α3 α4 α5 }
{β } {β1 β2 β3 β4 β5 }
{P} {1 D/Dre f tan ϕ (D/Dre f ) tan ϕ tan2 ϕ}
Dre f = 15 m

α1 α2 α3 α4 α5
0.183 -0.045 -0.2616 0.0432 0.2721
β1 β2 β3 β4 β5
0.7989 0.0733 0.8381 -0.1079 -0.4947

Peak displacement
 0 {δ }{Pδ }T
T σvo
z p = {γ}{Pγ } D
σre f
where:
{γ} {γ1 γ2 γ3 γ4 γ5 }
{δ } {δ1 δ2 δ3 }
{Pγ } {1 D/Dre f tan ϕ (D/Dre f ) tan ϕ tan2 ϕ}
{Pδ } {1 D/Dre f tan ϕ}
σre f = 100 kPa

γ1 γ2 γ3 γ4 γ5
0.0833 0.016 -0.2243 -0.0136 0.1512
δ1 δ2 δ3
0.7019 -0.0067 -0.5357

153
Chapter E. Mathematical models

1.2

0.8

0.6

d = 3.2, Normalized Data


d = 3.9, Normalized Data
0.4 d = 4.6, Normalized Data
d = 5.3, Normalized Data
d = 6.1, Normalized Data
d = 6.8, Normalized Data
0.2 d = 7.5, Normalized Data
d = 8.2, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3 3.5 4

1.2

0.8

0.6

d = 3.2, Normalized Data


d = 3.9, Normalized Data
0.4 d = 4.6, Normalized Data
d = 5.3, Normalized Data
d = 6.1, Normalized Data
d = 6.8, Normalized Data
0.2 d = 7.5, Normalized Data
d = 8.2, Normalized Data
Mathematical model

0
0 1 2 3 4 5 6

154
E.1 Drained tension

1.2

0.8

0.6

d = 3.2, Normalized Data


d = 3.9, Normalized Data
0.4 d = 4.6, Normalized Data
d = 5.3, Normalized Data
d = 6.1, Normalized Data
d = 6.8, Normalized Data
0.2 d = 7.5, Normalized Data
d = 8.2, Normalized Data
Mathematical model

0
0 1 2 3 4 5 6 7 8

1.2

0.8

0.6

d = 3.2, Normalized Data


d = 3.9, Normalized Data
0.4 d = 4.6, Normalized Data
d = 5.3, Normalized Data
d = 6.1, Normalized Data
d = 6.8, Normalized Data
0.2 d = 7.5, Normalized Data
d = 8.2, Normalized Data
Mathematical model

0
0 1 2 3 4 5 6 7 8 9 10

155
Chapter E. Mathematical models

1.2

0.8

0.6

d = 3.2, Normalized Data


d = 3.9, Normalized Data
0.4 d = 4.6, Normalized Data
d = 5.3, Normalized Data
d = 6.1, Normalized Data
d = 6.8, Normalized Data
0.2 d = 7.5, Normalized Data
d = 8.2, Normalized Data
Mathematical model

0
0 1 2 3 4 5 6 7 8 9

1.2

0.8

0.6

d = 4.2, Normalized Data


d = 5.1, Normalized Data
0.4 d = 6.1, Normalized Data
d = 7, Normalized Data
d = 7.9, Normalized Data
d = 8.8, Normalized Data
0.2 d = 9.7, Normalized Data
d = 10.7, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3

156
E.1 Drained tension

1.2

0.8

0.6

d = 4.2, Normalized Data


d = 5.1, Normalized Data
0.4 d = 6.1, Normalized Data
d = 7, Normalized Data
d = 7.9, Normalized Data
d = 8.8, Normalized Data
0.2 d = 9.7, Normalized Data
d = 10.7, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3 3.5 4

1.2

0.8

0.6

d = 4.2, Normalized Data


d = 5.1, Normalized Data
0.4 d = 6.1, Normalized Data
d = 7, Normalized Data
d = 7.9, Normalized Data
d = 8.8, Normalized Data
0.2 d = 9.7, Normalized Data
d = 10.7, Normalized Data
Mathematical model

0
0 1 2 3 4 5 6

157
Chapter E. Mathematical models

1.2

0.8

0.6

d = 4.2, Normalized Data


d = 5.1, Normalized Data
0.4 d = 6.1, Normalized Data
d = 7, Normalized Data
d = 7.9, Normalized Data
d = 8.8, Normalized Data
0.2 d = 9.7, Normalized Data
d = 10.7, Normalized Data
Mathematical model

0
0 1 2 3 4 5 6 7

1.2

0.8

0.6

d = 4.2, Normalized Data


d = 5.1, Normalized Data
0.4 d = 6.1, Normalized Data
d = 7, Normalized Data
d = 7.9, Normalized Data
d = 8.8, Normalized Data
0.2 d = 9.7, Normalized Data
d = 10.7, Normalized Data
Mathematical model

0
0 1 2 3 4 5 6 7

158
E.1 Drained tension

1.2

0.8

0.6

d = 4.8, Normalized Data


d = 5.9, Normalized Data
0.4 d = 7, Normalized Data
d = 8, Normalized Data
d = 9.1, Normalized Data
d = 10.2, Normalized Data
0.2 d = 11.3, Normalized Data
d = 12.4, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5

1.2

0.8

0.6

d = 4.8, Normalized Data


d = 5.9, Normalized Data
0.4 d = 7, Normalized Data
d = 8, Normalized Data
d = 9.1, Normalized Data
d = 10.2, Normalized Data
0.2 d = 11.3, Normalized Data
d = 12.4, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3 3.5

159
Chapter E. Mathematical models

1.2

0.8

0.6

d = 4.8, Normalized Data


d = 5.9, Normalized Data
0.4 d = 7, Normalized Data
d = 8, Normalized Data
d = 9.1, Normalized Data
d = 10.2, Normalized Data
0.2 d = 11.3, Normalized Data
d = 12.4, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

1.2

0.8

0.6

d = 4.8, Normalized Data


d = 5.9, Normalized Data
0.4 d = 7, Normalized Data
d = 8, Normalized Data
d = 9.1, Normalized Data
d = 10.2, Normalized Data
0.2 d = 11.3, Normalized Data
d = 12.4, Normalized Data
Mathematical model

0
0 1 2 3 4 5 6

160
E.1 Drained tension

1.2

0.8

0.6

d = 4.8, Normalized Data


d = 5.9, Normalized Data
0.4 d = 7, Normalized Data
d = 8, Normalized Data
d = 9.1, Normalized Data
d = 10.2, Normalized Data
0.2 d = 11.3, Normalized Data
d = 12.4, Normalized Data
Mathematical model

0
0 1 2 3 4 5 6

1.2

0.8

0.6

d = 5.8, Normalized Data


d = 7, Normalized Data
0.4 d = 8.4, Normalized Data
d = 9.7, Normalized Data
d = 10.9, Normalized Data
d = 12.2, Normalized Data
0.2 d = 13.5, Normalized Data
d = 14.8, Normalized Data
Mathematical model

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

161
Chapter E. Mathematical models

1.2

0.8

0.6

d = 5.8, Normalized Data


d = 7, Normalized Data
0.4 d = 8.4, Normalized Data
d = 9.7, Normalized Data
d = 10.9, Normalized Data
d = 12.2, Normalized Data
0.2 d = 13.5, Normalized Data
d = 14.8, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5

1.2

0.8

0.6

d = 5.8, Normalized Data


d = 7, Normalized Data
0.4 d = 8.4, Normalized Data
d = 9.7, Normalized Data
d = 10.9, Normalized Data
d = 12.2, Normalized Data
0.2 d = 13.5, Normalized Data
d = 14.8, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3 3.5

162
E.1 Drained tension

1.2

0.8

0.6

d = 5.8, Normalized Data


d = 7, Normalized Data
0.4 d = 8.4, Normalized Data
d = 9.7, Normalized Data
d = 10.9, Normalized Data
d = 12.2, Normalized Data
0.2 d = 13.5, Normalized Data
d = 14.8, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

1.2

0.8

0.6

d = 5.8, Normalized Data


d = 7, Normalized Data
0.4 d = 8.4, Normalized Data
d = 9.7, Normalized Data
d = 10.9, Normalized Data
d = 12.2, Normalized Data
0.2 d = 13.5, Normalized Data
d = 14.8, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

163
Chapter E. Mathematical models

1.2

0.8

0.6

d = 6.5, Normalized Data


d = 7.9, Normalized Data
0.4 d = 9.3, Normalized Data
d = 10.7, Normalized Data
d = 12.1, Normalized Data
d = 13.6, Normalized Data
0.2 d = 15, Normalized Data
d = 16.4, Normalized Data
Mathematical model

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4

1.2

0.8

0.6

d = 6.5, Normalized Data


d = 7.9, Normalized Data
0.4 d = 9.3, Normalized Data
d = 10.7, Normalized Data
d = 12.1, Normalized Data
d = 13.6, Normalized Data
0.2 d = 15, Normalized Data
d = 16.4, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5

164
E.1 Drained tension

1.2

0.8

0.6

d = 6.5, Normalized Data


d = 7.9, Normalized Data
0.4 d = 9.3, Normalized Data
d = 10.7, Normalized Data
d = 12.1, Normalized Data
d = 13.6, Normalized Data
0.2 d = 15, Normalized Data
d = 16.4, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3

1.2

0.8

0.6

d = 6.5, Normalized Data


d = 7.9, Normalized Data
0.4 d = 9.3, Normalized Data
d = 10.7, Normalized Data
d = 12.1, Normalized Data
d = 13.6, Normalized Data
0.2 d = 15, Normalized Data
d = 16.4, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3 3.5 4

165
Chapter E. Mathematical models

1.2

0.8

0.6

d = 6.5, Normalized Data


d = 7.9, Normalized Data
0.4 d = 9.3, Normalized Data
d = 10.7, Normalized Data
d = 12.1, Normalized Data
d = 13.6, Normalized Data
0.2 d = 15, Normalized Data
d = 16.4, Normalized Data
Mathematical model

0
0 0.5 1 1.5 2 2.5 3 3.5 4

166
E.2 Partially drained tension

E.2 Partially drained tension

E.2.1 Investigation of post-peak behaviour

1. Isolation of data. Data is separated into pre-peak and post-peak parts. The part of interest is
marked with green in the figure.

2. Curve fitting. Each curve in each model is fitted with a function of the form f (x) = p2 /x2 +
p1 x + p0 .

3. Study of fitting coefficients. Since each curve corresponds to a specific depth, and hence a
value of vertical stress σvo
0 , the dependency of p , p and p on the stress level is established. All
2 1 0
coefficients linked to most FE models seem to have a distribution that can be approximated with a
power function f (x) = axb . The concept and typical distribution is illustrated below.

167
Chapter E. Mathematical models

4. Study of coefficients a and b. The link to bucket diameter D and friction angle ϕ is introduced
through coefficients a and b. At this stage there are 3 sets of a and b: one set per p-coefficient.
Each set comprises 20 values that are linked to the 20 FE models involved in the study (those with
ϕ = 30◦ are disregarded). Every FE model has a unique combination of D and ϕ, therefore a and b
are established as functions of two variables.

The following expressions are intended to shorten the equations presented in this section:
 
T D
{α}{Pα } = α1 + α2 + α3 tan ϕ
Dre f
   
T D D
{β }{Pβ } = β1 + β2 + β3 tan ϕ + β4 tan ϕ + β5 tan2 ϕ
Dre f Dre f
   2  
T D D D
{γ}{Pγ } = γ1 + γ2 + γ3 tan ϕ + γ4 + γ5 tan ϕ
Dre f Dre f Dre f

Dre f = 15 m

σre f = 100 kPa

The p-coefficients are found with respect to their affiliated a- and b-coefficients:
 0 b p2
σvo
p2 = a p2
σre f

a p2 = {β }{Pβ }T

b p2 = {α}{Pα }T

β1 β2 β3 β4 β5
0.0496 0.2035 0.0145 -0.2442 -0.0340
α1 α2 α3
9.6825 0.6471 -16.1763

 b p1
σvo
0
p1 = a p1
σre f

a p1 = {α}{Pα }T

b p1 = {γ}{Pγ }T

α1 α2 α3
0.6151 -0.1730 -0.4782
γ1 γ2 γ3 γ4 γ5
0.6915 4.6587 -1.4662 0.8080 -5.5509

168
E.2 Partially drained tension

 b p0
σvo
0
p0 = a p0
σre f

a p0 = {γ}{Pγ }T

b p0 = {α}{Pα }T

γ1 γ2 γ3 γ4 γ5
0.3310 -0.9759 1.7063 0.8021 -0.7690
α1 α2 α3
-0.4111 -0.4759 1.2182

E.2.2 Peak shear stress and peak displacement

!b
σvo
0
τp = a D
L tan ϕ

   
D D
a = −0.4675 − 0.0753 + 1.3639 tan ϕ + 0.0806 tan ϕ − 0.8737 tan2 ϕ
Dre f Dre f
   
D D
b = 2.9391 + 0.2263 − 4.854 tan ϕ − 0.2933 tan ϕ + 3.3032 tan2 ϕ
Dre f Dre f

 d
σvo
0
zp = c D
σre f
   2  
D D D tan ϕ
c = 0.0048 − 0.0037 + 0.0099 tan ϕ + 0.0096 − 0.0204
Dre f Dre f Dre f

d = 0.5913

169
Chapter E. Mathematical models

E.2.3 Normalized curves vs t-z formulation

1.6

1.4

1.2

1
d = 3.6, Normalized Data
d = 4.2, Normalized Data
0.8 d = 4.8, Normalized Data
d = 5.3, Normalized Data
d = 5.8, Normalized Data
d = 6.4, Normalized Data
0.6 d = 7, Normalized Data
d = 7.5, Normalized Data
d = 3.6, Mathematical Model
d = 4.2, Mathematical Model
0.4 d = 4.8, Mathematical Model
d = 5.3, Mathematical Model
d = 5.8, Mathematical Model
0.2 d = 6.4, Mathematical Model
d = 7, Mathematical Model
d = 7.5, Mathematical Model

0
0 1 2 3 4 5 6

1.4

1.2

0.8 d = 3.6, Normalized Data


d = 4.1, Normalized Data
d = 4.6, Normalized Data
d = 5.3, Normalized Data
0.6 d = 5.9, Normalized Data
d = 6.4, Normalized Data
d = 6.9, Normalized Data
d = 7.5, Normalized Data
d = 3.6, Mathematical Model
0.4 d = 4.1, Mathematical Model
d = 4.6, Mathematical Model
d = 5.3, Mathematical Model
d = 5.9, Mathematical Model
0.2 d = 6.4, Mathematical Model
d = 6.9, Mathematical Model
d = 7.5, Mathematical Model

0
0 1 2 3 4 5 6

170
E.2 Partially drained tension

1.8

1.6

1.4

1.2

d = 3.6, Normalized Data


1 d = 4.1, Normalized Data
d = 4.6, Normalized Data
d = 5.3, Normalized Data
0.8 d = 5.9, Normalized Data
d = 6.4, Normalized Data
d = 6.9, Normalized Data
0.6 d = 7.5, Normalized Data
d = 3.6, Mathematical Model
d = 4.1, Mathematical Model
0.4 d = 4.6, Mathematical Model
d = 5.3, Mathematical Model
d = 5.9, Mathematical Model
d = 6.4, Mathematical Model
0.2 d = 6.9, Mathematical Model
d = 7.5, Mathematical Model

0
0 1 2 3 4 5 6 7

2.5

1.5
d = 3.6, Normalized Data
d = 4.1, Normalized Data
d = 4.6, Normalized Data
d = 5.3, Normalized Data
d = 5.9, Normalized Data
1 d = 6.4, Normalized Data
d = 6.9, Normalized Data
d = 7.5, Normalized Data
d = 3.6, Mathematical Model
d = 4.1, Mathematical Model
d = 4.6, Mathematical Model
0.5 d = 5.3, Mathematical Model
d = 5.9, Mathematical Model
d = 6.4, Mathematical Model
d = 6.9, Mathematical Model
d = 7.5, Mathematical Model

0
0 1 2 3 4 5 6 7

171
Chapter E. Mathematical models

1.4

1.2

0.8 d = 4.7, Normalized Data


d = 5.4, Normalized Data
d = 6.1, Normalized Data
d = 6.8, Normalized Data
0.6 d = 7.6, Normalized Data
d = 8.3, Normalized Data
d = 9, Normalized Data
d = 9.8, Normalized Data
d = 4.7, Mathematical Model
0.4 d = 5.4, Mathematical Model
d = 6.1, Mathematical Model
d = 6.8, Mathematical Model
d = 7.6, Mathematical Model
0.2 d = 8.3, Mathematical Model
d = 9, Mathematical Model
d = 9.8, Mathematical Model

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

1.4

1.2

0.8 d = 4.7, Normalized Data


d = 5.4, Normalized Data
d = 6.1, Normalized Data
d = 6.8, Normalized Data
0.6 d = 7.6, Normalized Data
d = 8.3, Normalized Data
d = 9, Normalized Data
d = 9.8, Normalized Data
d = 4.7, Mathematical Model
0.4 d = 5.4, Mathematical Model
d = 6.1, Mathematical Model
d = 6.8, Mathematical Model
d = 7.6, Mathematical Model
0.2 d = 8.3, Mathematical Model
d = 9, Mathematical Model
d = 9.8, Mathematical Model

0
0 1 2 3 4 5 6

172
E.2 Partially drained tension

1.4

1.2

0.8 d = 4.7, Normalized Data


d = 5.4, Normalized Data
d = 6.1, Normalized Data
d = 6.8, Normalized Data
0.6 d = 7.6, Normalized Data
d = 8.3, Normalized Data
d = 9, Normalized Data
d = 9.8, Normalized Data
d = 4.7, Mathematical Model
0.4 d = 5.4, Mathematical Model
d = 6.1, Mathematical Model
d = 6.8, Mathematical Model
d = 7.6, Mathematical Model
0.2 d = 8.3, Mathematical Model
d = 9, Mathematical Model
d = 9.8, Mathematical Model

0
0 1 2 3 4 5 6

1.4

1.2

0.8 d = 4.7, Normalized Data


d = 5.4, Normalized Data
d = 6.1, Normalized Data
d = 6.8, Normalized Data
0.6 d = 7.6, Normalized Data
d = 8.3, Normalized Data
d = 9.1, Normalized Data
d = 9.8, Normalized Data
d = 4.7, Mathematical Model
0.4 d = 5.4, Mathematical Model
d = 6.1, Mathematical Model
d = 6.8, Mathematical Model
d = 7.6, Mathematical Model
0.2 d = 8.3, Mathematical Model
d = 9.1, Mathematical Model
d = 9.8, Mathematical Model

0
0 1 2 3 4 5 6 7

173
Chapter E. Mathematical models

1.2

0.8

d = 5.4, Normalized Data


d = 6.3, Normalized Data
0.6 d = 7.1, Normalized Data
d = 8, Normalized Data
d = 8.8, Normalized Data
d = 9.6, Normalized Data
d = 10.4, Normalized Data
0.4 d = 11.2, Normalized Data
d = 5.4, Mathematical Model
d = 6.3, Mathematical Model
d = 7.1, Mathematical Model
d = 8, Mathematical Model
0.2 d = 8.8, Mathematical Model
d = 9.6, Mathematical Model
d = 10.4, Mathematical Model
d = 11.2, Mathematical Model

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

1.2

0.8

d = 5.4, Normalized Data


d = 6.3, Normalized Data
0.6 d = 7.1, Normalized Data
d = 8, Normalized Data
d = 8.8, Normalized Data
d = 9.6, Normalized Data
d = 10.4, Normalized Data
0.4 d = 11.2, Normalized Data
d = 5.4, Mathematical Model
d = 6.3, Mathematical Model
d = 7.1, Mathematical Model
d = 8, Mathematical Model
0.2 d = 8.8, Mathematical Model
d = 9.6, Mathematical Model
d = 10.4, Mathematical Model
d = 11.2, Mathematical Model

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

174
E.2 Partially drained tension

1.2

0.8

d = 5.4, Normalized Data


d = 6.3, Normalized Data
0.6 d = 7.1, Normalized Data
d = 8, Normalized Data
d = 8.8, Normalized Data
d = 9.6, Normalized Data
d = 10.4, Normalized Data
0.4 d = 11.2, Normalized Data
d = 5.4, Mathematical Model
d = 6.3, Mathematical Model
d = 7.1, Mathematical Model
d = 8, Mathematical Model
0.2 d = 8.8, Mathematical Model
d = 9.6, Mathematical Model
d = 10.4, Mathematical Model
d = 11.2, Mathematical Model

0
0 1 2 3 4 5 6

1.2

0.8

d = 5.4, Normalized Data


d = 6.3, Normalized Data
0.6 d = 7.1, Normalized Data
d = 8, Normalized Data
d = 8.8, Normalized Data
d = 9.6, Normalized Data
d = 10.4, Normalized Data
0.4 d = 11.2, Normalized Data
d = 5.4, Mathematical Model
d = 6.3, Mathematical Model
d = 7.1, Mathematical Model
d = 8, Mathematical Model
0.2 d = 8.8, Mathematical Model
d = 9.6, Mathematical Model
d = 10.4, Mathematical Model
d = 11.2, Mathematical Model

0
0 1 2 3 4 5 6 7

175
Chapter E. Mathematical models

1.2

0.8

d = 6.5, Normalized Data


d = 7.5, Normalized Data
0.6 d = 8.5, Normalized Data
d = 9.4, Normalized Data
d = 10.5, Normalized Data
d = 11.5, Normalized Data
d = 12.6, Normalized Data
0.4 d = 13.5, Normalized Data
d = 6.5, Mathematical Model
d = 7.5, Mathematical Model
d = 8.5, Mathematical Model
d = 9.4, Mathematical Model
0.2 d = 10.5, Mathematical Model
d = 11.5, Mathematical Model
d = 12.6, Mathematical Model
d = 13.5, Mathematical Model

0
0 0.5 1 1.5 2 2.5 3

1.2

0.8

d = 6.5, Normalized Data


d = 7.5, Normalized Data
0.6 d = 8.5, Normalized Data
d = 9.4, Normalized Data
d = 10.5, Normalized Data
d = 11.5, Normalized Data
d = 12.6, Normalized Data
0.4 d = 13.5, Normalized Data
d = 6.5, Mathematical Model
d = 7.5, Mathematical Model
d = 8.5, Mathematical Model
d = 9.4, Mathematical Model
0.2 d = 10.5, Mathematical Model
d = 11.5, Mathematical Model
d = 12.6, Mathematical Model
d = 13.5, Mathematical Model

0
0 0.5 1 1.5 2 2.5 3 3.5

176
E.2 Partially drained tension

1.2

0.8

d = 6.5, Normalized Data


d = 7.5, Normalized Data
0.6 d = 8.5, Normalized Data
d = 9.4, Normalized Data
d = 10.5, Normalized Data
d = 11.5, Normalized Data
d = 12.6, Normalized Data
0.4 d = 13.5, Normalized Data
d = 6.5, Mathematical Model
d = 7.5, Mathematical Model
d = 8.5, Mathematical Model
d = 9.4, Mathematical Model
0.2 d = 10.5, Mathematical Model
d = 11.5, Mathematical Model
d = 12.6, Mathematical Model
d = 13.5, Mathematical Model

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

1.2

0.8

d = 6.5, Normalized Data


d = 7.5, Normalized Data
0.6 d = 8.5, Normalized Data
d = 9.4, Normalized Data
d = 10.5, Normalized Data
d = 11.5, Normalized Data
d = 12.6, Normalized Data
0.4 d = 13.5, Normalized Data
d = 6.5, Mathematical Model
d = 7.5, Mathematical Model
d = 8.5, Mathematical Model
d = 9.4, Mathematical Model
0.2 d = 10.5, Mathematical Model
d = 11.5, Mathematical Model
d = 12.6, Mathematical Model
d = 13.5, Mathematical Model

0
0 1 2 3 4 5 6

177
Chapter E. Mathematical models

1.2

0.8

d = 7.2, Normalized Data


d = 8.2, Normalized Data
0.6 d = 9.3, Normalized Data
d = 10.6, Normalized Data
d = 11.8, Normalized Data
d = 12.9, Normalized Data
d = 14, Normalized Data
0.4 d = 15, Normalized Data
d = 7.2, Mathematical Model
d = 8.2, Mathematical Model
d = 9.3, Mathematical Model
d = 10.6, Mathematical Model
0.2 d = 11.8, Mathematical Model
d = 12.9, Mathematical Model
d = 14, Mathematical Model
d = 15, Mathematical Model

0
0 0.5 1 1.5 2 2.5

1.2

0.8

d = 7.2, Normalized Data


d = 8.2, Normalized Data
0.6 d = 9.3, Normalized Data
d = 10.6, Normalized Data
d = 11.8, Normalized Data
d = 12.9, Normalized Data
d = 14, Normalized Data
0.4 d = 15, Normalized Data
d = 7.2, Mathematical Model
d = 8.2, Mathematical Model
d = 9.3, Mathematical Model
d = 10.6, Mathematical Model
0.2 d = 11.8, Mathematical Model
d = 12.9, Mathematical Model
d = 14, Mathematical Model
d = 15, Mathematical Model

0
0 0.5 1 1.5 2 2.5

178
E.2 Partially drained tension

1.2

0.8

d = 7.2, Normalized Data


d = 8.2, Normalized Data
0.6 d = 9.3, Normalized Data
d = 10.6, Normalized Data
d = 11.8, Normalized Data
d = 12.9, Normalized Data
d = 14, Normalized Data
0.4 d = 15, Normalized Data
d = 7.2, Mathematical Model
d = 8.2, Mathematical Model
d = 9.3, Mathematical Model
d = 10.6, Mathematical Model
0.2 d = 11.8, Mathematical Model
d = 12.9, Mathematical Model
d = 14, Mathematical Model
d = 15, Mathematical Model

0
0 0.5 1 1.5 2 2.5 3 3.5

1.2

0.8

d = 7.2, Normalized Data


d = 8.2, Normalized Data
0.6 d = 9.3, Normalized Data
d = 10.6, Normalized Data
d = 11.8, Normalized Data
d = 12.9, Normalized Data
d = 14, Normalized Data
0.4 d = 15, Normalized Data
d = 7.2, Mathematical Model
d = 8.2, Mathematical Model
d = 9.3, Mathematical Model
d = 10.6, Mathematical Model
0.2 d = 11.8, Mathematical Model
d = 12.9, Mathematical Model
d = 14, Mathematical Model
d = 15, Mathematical Model

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

179
Chapter E. Mathematical models

E.3 Drained compression

ϕ
33◦ 35◦ 38◦ 40◦ 33◦ 35◦ 38◦ 40◦
D
10 m 19.9289 22.9409 32.4527 34.5442 0.5254 0.5623 0.6223 0.6136
13 m 25.0058 25.2997 32.0182 40.6087 0.5111 0.5311 0.5803 0.6199
15 m 30.7947 29.2518 34.2847 40.6333 0.5074 0.5242 0.5666 0.5960
18 m 45.7884 37.4970 38.9360 44.9747 0.5110 0.5133 0.5466 0.5793
20 m 61.5876 47.3954 45.3220 48.0943 0.5129 0.5166 0.5452 0.5654

Table E.1: Values of A-coefficients (left section) and B-coefficients (right section) used for regression
analysis in the context of formulating the mathematical model for drained compression.

 
D
A = −21.1 + 34.1 + 31.6 tan ϕ
Dre f
 
D
B = 0.2932 − 0.0694 + 0.4439 tan ϕ
Dre f

NRMSE = 14.25 %
70

60

50

40

30

20

10
0.9
1.4
0.8
1.2
0.7 1
0.8
0.6 0.6

Figure E.1: Fitting A-coefficients with a plane.

180
E.3 Drained compression

NRMSE = 10.73 %
0.65

0.6

0.55

0.5

0.45
0.9
1.4
0.8
1.2
0.7 1
0.8
0.6 0.6

Figure E.2: Fitting B-coefficients with a plane.

4
Normalized Data
3.5 Power fit
Mathematical Model: t-z formulation

2.5

1.5

0.5

0
0 0.005 0.01 0.015 0.02 0.025 0.03

181
Chapter E. Mathematical models

4
Normalized Data
3.5 Power fit
Mathematical Model: t-z formulation

2.5

1.5

0.5

0
0 0.005 0.01 0.015 0.02 0.025 0.03

5
Normalized Data
4.5 Power fit
Mathematical Model: t-z formulation
4

3.5

2.5

1.5

0.5

0
0 0.005 0.01 0.015 0.02 0.025 0.03

182
E.3 Drained compression

4.5
Normalized Data
4 Power fit
Mathematical Model: t-z formulation
3.5

2.5

1.5

0.5

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

4.5
Normalized Data
4 Power fit
Mathematical Model: t-z formulation
3.5

2.5

1.5

0.5

0
0 0.005 0.01 0.015 0.02 0.025

183
Chapter E. Mathematical models

4.5
Normalized Data
4 Power fit
Mathematical Model: t-z formulation
3.5

2.5

1.5

0.5

0
0 0.005 0.01 0.015 0.02 0.025

4.5
Normalized Data
4 Power fit
Mathematical Model: t-z formulation
3.5

2.5

1.5

0.5

0
0 0.005 0.01 0.015 0.02 0.025

184
E.3 Drained compression

5
Normalized Data
4.5 Power fit
Mathematical Model: t-z formulation
4

3.5

2.5

1.5

0.5

0
0 0.005 0.01 0.015 0.02 0.025

5
Normalized Data
4.5 Power fit
Mathematical Model: t-z formulation
4

3.5

2.5

1.5

0.5

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

185
Chapter E. Mathematical models

4.5
Normalized Data
4 Power fit
Mathematical Model: t-z formulation
3.5

2.5

1.5

0.5

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

5
Normalized Data
4.5 Power fit
Mathematical Model: t-z formulation
4

3.5

2.5

1.5

0.5

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

186
E.3 Drained compression

5
Normalized Data
4.5 Power fit
Mathematical Model: t-z formulation
4

3.5

2.5

1.5

0.5

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

6
Normalized Data
Power fit
5 Mathematical Model: t-z formulation

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

187
Chapter E. Mathematical models

6
Normalized Data
Power fit
5 Mathematical Model: t-z formulation

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

5
Normalized Data
4.5 Power fit
Mathematical Model: t-z formulation
4

3.5

2.5

1.5

0.5

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

188
E.3 Drained compression

6
Normalized Data
Power fit
5 Mathematical Model: t-z formulation

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

8
Normalized Data
7 Power fit
Mathematical Model: t-z formulation

0
0 0.005 0.01 0.015

189
Chapter E. Mathematical models

6
Normalized Data
Power fit
5 Mathematical Model: t-z formulation

0
0 0.005 0.01 0.015

6
Normalized Data
Power fit
5 Mathematical Model: t-z formulation

0
0 0.005 0.01 0.015

190
E.4 Partially drained compression

6
Normalized Data
Power fit
5 Mathematical Model: t-z formulation

0
0 0.005 0.01 0.015

E.4 Partially drained compression

ϕ
33◦ 35◦ 38◦ 40◦ 33◦ 35◦ 38◦ 40◦
D
10 m 1.6505 1.8561 2.5583 3.5440 0.0991 0.1084 0.1710 0.2230
13 m 2.1714 2.4413 3.0033 4.0300 0.1450 0.1509 0.1868 0.2335
15 m 2.5091 2.7429 3.2768 4.0936 0.1705 0.1703 0.1982 0.2350
18 m 2.8286 3.0187 3.4114 3.7670 0.1893 0.1865 0.2022 0.2233
20 m 3.0667 3.1956 3.8761 4.2663 0.2013 0.1962 0.2180 0.2409

Table E.2: Values of A-coefficients (left section) and B-coefficients (right section) used for regression
analysis in the context of formulating the mathematical model for partially drained compression.

 
D
A = −4.3866 + 1.6422 + 7.7951 tan ϕ
Dre f
 
D
B = −0.1833 + 0.0893 + 0.3776 tan ϕ
Dre f

191
Chapter E. Mathematical models

NRMSE = 8.77 %
4.5

3.5

2.5

1.5
0.9
1.4
0.8
1.2
0.7 1
0.8
0.6 0.6

Figure E.3: Fitting A-coefficients with a plane.

NRMSE = 12.44 %

0.3

0.25

0.2

0.15

0.1

0.05
0.9
1.4
0.8
1.2
0.7 1
0.8
0.6 0.6

Figure E.4: Fitting B-coefficients with a plane.

192
E.4 Partially drained compression

1.4

1.2

0.8

0.6

0.4

Normalized Data
0.2
Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015 0.02 0.025 0.03

1.5

0.5

Normalized Data
Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015 0.02 0.025 0.03

193
Chapter E. Mathematical models

1.8

1.6

1.4

1.2

0.8

0.6

0.4
Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015 0.02 0.025 0.03

1.8

1.6

1.4

1.2

0.8

0.6

0.4
Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035

194
E.4 Partially drained compression

1.4

1.2

0.8

0.6

0.4

Normalized Data
0.2
Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015 0.02 0.025

1.6

1.4

1.2

0.8

0.6

0.4

Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015 0.02 0.025

195
Chapter E. Mathematical models

1.8

1.6

1.4

1.2

0.8

0.6

0.4
Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015 0.02 0.025

1.8

1.6

1.4

1.2

0.8

0.6

0.4
Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015 0.02 0.025

196
E.4 Partially drained compression

1.4

1.2

0.8

0.6

0.4

Normalized Data
0.2
Power fit
Mathematical Model: t-z formulation
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

1.6

1.4

1.2

0.8

0.6

0.4

Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

197
Chapter E. Mathematical models

1.8

1.6

1.4

1.2

0.8

0.6

0.4
Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

1.8

1.6

1.4

1.2

0.8

0.6

0.4
Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02

198
E.4 Partially drained compression

1.4

1.2

0.8

0.6

0.4

Normalized Data
0.2
Power fit
Mathematical Model: t-z formulation
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

1.5

0.5

Normalized Data
Power fit
Mathematical Model: t-z formulation
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

199
Chapter E. Mathematical models

1.8

1.6

1.4

1.2

0.8

0.6

0.4
Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

1.8

1.6

1.4

1.2

0.8

0.6

0.4
Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018

200
E.4 Partially drained compression

1.4

1.2

0.8

0.6

0.4

Normalized Data
0.2
Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015

1.5

0.5

Normalized Data
Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015

201
Chapter E. Mathematical models

1.8

1.6

1.4

1.2

0.8

0.6

0.4
Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015

1.8

1.6

1.4

1.2

0.8

0.6

0.4
Normalized Data
0.2 Power fit
Mathematical Model: t-z formulation
0
0 0.005 0.01 0.015

202

You might also like