Soil Variability and Its Consequences in Geotechnical Engineering
Soil Variability and Its Consequences in Geotechnical Engineering
in geotechnical engineering
vorgelegt von
2013
[email protected]
Mitteilung 69
des Instituts für Geotechnik
Universität Stuttgart, Germany, 2013
Editor:
Univ.-Prof. Dr.-Ing. habil. Christian Moormann
c
Maximilian Huber ([email protected])
Institut für Geotechnik
Universität Stuttgart
Pfaffenwaldring 35
70569 Stuttgart
All rights reserved. No part of this publication may be reproduced, stored in a retrieval
system, or transmitted, in any form or by any means, electronic, mechanical,
photocopying, recording, scanning or otherwise, without the permission in writing of
the author.
Printed by e.kurz + co, druck und medientechnik GmbH, Stuttgart, Germany, 2013
ISBN 978-3-921837-69-6
(D93 - Dissertation, Universität Stuttgart)
[email protected]
Preface of the editor
With issue no. 69 of the proceedings the Institute of Geotechnical Engineering at the
University of Stuttgart (IGS) the dissertation study of Dr.-Ing. Maximilian Huber is pub-
lished. Dr. Huber’s thesis is related to soil and rock heterogeneity, to the mathematical
approaches for describing soil variability and to its consequences in geotechnical engi-
neering.
Soils and rocks as natural geological materials exhibit spatial variability of material
properties. The spatial variability, frequently referred to as heterogeneity, is anisotropic,
often depth-dependent and occurs at multiple scales: at the very small scale, also called
grain scale, as seen in the arrangement of solid particles of sand or in the arrangement of
pellets in clay, at the decimeter to meter scale, as observed in specimen testes in labora-
tory or in CPT/SPT-soundings in soil layers; and at larger scales too, like the geotechni-
cal scale relevant for geotechnical structures like foundation, excavations or tunnels, or
at the geological scale of several hundreds of meters, which is affected by the layering
of soils of different types. Spatial variability of soils and rocks influences the material
behaviour in mechanical and hydraulic sense, the flow of groundwater water and the
performance of geotechnical structures. Traditional deterministic analyses based on sin-
gle "representative" soil property values lead to partial or global factors of safety, which
provide no information regarding probability of failure and which does not consider the
uncertainty that arises through having incomplete information about the subsoil con-
ditions. In contrast probabilistic approaches allow defining soil properties in statistical
terms and simulating geotechnical performance probabilistically for example on terms
of reliability. The reliability, defined as the complement of the failure probability, is a
rational measure of safety.
The scientific work of Dr. Huber focuses on the evaluation of the effects of soil vari-
ability by using advanced mathematical models within the framework of probabilistic
methods. The evaluation of stochastic soil properties including spatially correlated soil
properties is investigated on a wide theoretical basis. These results are used in case
studies on the evaluation of the spatial correlation of different measurement data sets.
Considering these outcomes finally the effects of soil variability are evaluated in typi-
cal problems in tunnelling and foundation engineering demonstrating the effect of soil
heterogeneity and spatial variability at different scales.
The thesis of Dr. Huber shows that probabilistic analyses are powerful and versatile
tools for investigating the influence of uncertainties on a given geotechnical problem, but
indicates the meaning of the given input data and limitations of probabilistic analyses
also.
[email protected]
Safety, reliability and risk are key issues in situations with continuously increasing
complexity especially in geotechnical engineering. In this regard the doctoral thesis of
Dr. Huber proves as a valuable contribution.
Christian Moormann
Stuttgart, July 2013
[email protected]
Preface of the supervisor of the PhD thesis
As yet probabilistic analyses are not common in geotechnical engineering and will prob-
ably never be introduced for regular structures and foundations, but it is expected that it
will become more and more used for major engineering projects in difficult ground. As
usual in most branches of engineering, loads on structures are stochastic, but the natu-
ral variability of soil properties exceeds by far the variability of man-made engineering
materials. As a consequence, a special need of probabilistic design exist in geotechni-
cal engineering. For this reason, I have encouraged Maximilian Huber to do a doctoral
study on the application of probabilistic methods in geotechnics, leaving the choice of a
more precise topic entirely to him.
The first idea he came up with was to perform borehole jacking tests in a particular
layer of mudstone to measure its stiffness and to assess the corresponding correlation
length. I was very happy to learn that my colleague Prof. Andras Bardossy was will-
ing to support these field measurements even financially. These experiments had to be
done in the Fasanenhof tunnel, being at that time under construction at Stuttgart by the
companies Weiss & Freitag and Max BÃűgl. I am not only indebted to the support from
these companies, but also for the support of the owner of the tunnel, i.e. the Stuttgarter
Strassenbahnen AG.
The candidate has chosen to focus this study on soil variability, being usually taken
into account by distinguishing between different soil layers. In probabilistic studies soil
properties within a layer are based on a probabilistic density function, as defined by
a mean value, a standard deviation and possibly skewness. In such an approach the
spatial variability within a soil layer may be disregarded on assuming homogeneity. On
approaching reality more closely, spatial variability within soil layers may be modelled
in combination with a correlation length, i.e. "a scale of fluctuation" for the soil property
considered. I consider such developments as fascinating and would like to know to
which extent such approaches are already applicable in engineering.
This dissertation study convinced me of the matureness of probabilistic design in
geotechnical engineering. On the other, it made me realise that variability within soil
layers is still a topic of research, as the assessment of correlation lengths is not straight
forward. The assessment of this length is probably the most important scientific achieve-
ment of this dissertation study on soil variability. Many of the case studies may also
serve as a manual for the application of probabilistic design in advanced geotechnical
engineering. As a consequence, his research has already attracted good attention in the
international research community.
[email protected]
I got to know Maximilian Huber not only as a young researcher and teacher before
my retirement from Stuttgart University, but also later in Delft when he worked with his
external advisor Prof. M. Hicks. These visits to Delft also provided the opportunity of
additional social contacts, which I enjoyed very much. So I had the pleasure of learning
to know him not only as a talented researcher with creative ideas, but also as gifted
musician with a wide field of interests.
Pieter A. Vermeer
Nederhorst den Berg, Netherlands, July 2013
[email protected]
Acknowledgments
The research presented in this thesis is the result of the work carried out during the
years 2007 - 2013 at the Institute of Geotechnical Engineering of Stuttgart University. Be-
hind this work there are important contributions and the significant support of a certain
number of persons, whom I would like to thank sincerely.
I wish to express first of all my deep gratitude to Prof. Pieter Vermeer for his extensive,
invaluable and steadfast guidance, for his support and patience, sharing his knowledge
and experience, throughout my PhD studies.
Likewise, I thank Prof. Andras Bardossy and Prof. Michael Hicks for their expert
advise and feedback as my co-supervisor throughout the time of my studies.
Throughout the years, I benefited from the cooperative and friendly support, advise
and guidance of many others. With this regard, I would like to express my gratitude
to Patrick Arnold, Dr. Robby Caspeele, Ed Calle, Anneke Hommels, Dr. Jon Nuttall,
Prof. Christian Moormann, Dr. Massimiliano Moscatelli, Dr. Thomas Most, Dr. Dirk
Proske and to longtime colleagues at the University of Stuttgart. Amongst them are Prof.
Thomas Benz, Dr. Lars Beuth, Fursan Hamad, Dr. Issam Jassim, Dr. Annette Lächler,
Dr. Axel Möllmann, Dr. Marcus Schneider, Peter Ströhle, Prof. Bernhard Westrich and
Bernd Zweschper.
Of those mentioned here and many others whom I met during the exciting years of
my PhD studies I maintain invaluable good memories. I am very grateful to my parents,
to my sister and to friends, who gave me indispensable support through their company
and advise.
Maximilian Huber
Weimar, July 2013
[email protected]
[email protected]
Contents
Summary xv
Zusammenfassung xix
1 Introduction 1
1.1 Background and rationale . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Research aim . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Thesis scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
i
Contents
ii
Contents
Bibliography 160
iii
Contents
iv
List of Figures
1.1 Data on the frequency of natural and industrial disasters against num-
ber of lost humans lives. In colour shading data from Dutch Govern-
ment Group Risk Criteria [395] for floods, data on other natural disaster
as reported by the N ATURAL D ISASTERS D ATABASE NATURAL DISASTERS
DATABASE and Christian & Baecher [23]. . . . . . . . . . . . . . . . . . . . . 2
1.2 Average annual damages ($US billion) caused by reported natural disas-
ters 1990 - 2011 by the I NTERNATIONAL D ISASTERS D ATABASE [378]. . . . 2
2.1 Mapping of the same area of Canada in 1958 (left) and in 1923 (right) using
the same data [22]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Uncertainties in Reliability Based Design modified from Baecher & Chris-
tian [23], Honjo [166], Phoon[286] and Phoon & Kulhawy [282, 283]. . . . . 9
2.3 Illustration of the multi-scale nature of soil after Borja [48], Chen et al.
[73], Christakos [82] and Wackernagel [401]. . . . . . . . . . . . . . . . . . . 10
2.4 Nested variogram from Jaksa [184]. . . . . . . . . . . . . . . . . . . . . . . . 22
2.5 Comparison for computing a variogram from regular sampling on a tran-
sect:
(a) with a complete set of data, indicated with • and
(b) with missing values indicated by ◦ from [403]. . . . . . . . . . . . . . . 23
2.6 The geometry for discretizing the lag into bins by distance and direction
in two dimensions from [403]. . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.7 Theoretical variogram functions (a) and comparison of semivariance func-
tion and autocorrelation function (b). . . . . . . . . . . . . . . . . . . . . . . 25
2.8 Dependency of the correlation length on the sample size, taken from De-
Groot & Baecher [90]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1
Lognormal distribution with a mean value of Ȳ = 1 and different coeffi-
cients of variation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2 Quantile-quantile transformation of normally N (µ = 0, σ = 1) into log-
normally distributed data log-N (µ = 1.5, σ = 10). . . . . . . . . . . . . . . 35
3.3 Evaluated correlation lengths using MoMvar , MoMLA and ML approach
(a,c) and resulting nugget/sill ratios (b,d) using MoMvar and ML approach
for different error levels ε = 0.01 (a,b) and ε = 0.0001 (c,d). . . . . . . . . . 36
3.4 Case study A: Analysis of the indicator correlation lengths for the different
quantiles of the CDF (log-N (µ = 1.5, σ = 10)) and a measurement noise
of N (µ = 0, σ = 0.01). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.5 Case study A: distributions of the relative errors εtarget of the MoMvar ,
[email protected]
MoMLA and ML- approaches for COV = 1 % (a) and COV = 100 % (b). . . 39
v
List of Figures
3.6 Case study B: plan view of the 138 CPT measurement locations. . . . . . . 40
3.7 Case study B: Average, minimum and maximum measurement values of
the CPT data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.8 Steps of the stochastic site characterisation and involved methods. . . . . . 42
3.9 Histogram and fitted normal distribution function of the detrended cone
resistances in sand of CPT 4 N (µ = 0.10 kN/m2 |σ = 2.51 kN/m) with a
skewness γ1 = 0.40( kN/m2 )3 and an excess kurtosis γ2,excess = −0.51( kN/m2 )4 . 44
3.10 (a) Cone resistance qc (z) measured with depth z and soil layer
identification of CPT 4 data using RI and Bartlett statistics and
(b) semivariogram on the raw and detrended data of the silty clay layer
at 10 m depth. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.11 Case study B: probability density function of the MoMvar , MoMLA and ML
approach and the combined probability density functions using Bayesian
Model Averaging of the silty clay layer. . . . . . . . . . . . . . . . . . . . . 48
3.12 Casestudy B: Analysis of the correlation lengths for indicators of the CDF
using MoMvar , MoMLA and ML with fitted lognormal distribution func-
tions for silty clay layer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.13 Plan view and cross section of the experiments in the Fasanenhof tunnel-
ing project. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.14 Experimental equipment of the borehole jacking probe according to the
DIN 4094-5 [104]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.15 Loading, unloading and reloading cycles performed in the boreholejack-
ing tests according to DIN 4094 − 5 [104]. . . . . . . . . . . . . . . . . . . . 53
3.16 Measurement results within the Fasanenhoftunnel. . . . . . . . . . . . . . 53
3.17 Casestudy C: Analysis of the indicator correlation lengths using MoMvar ,
MoMLA and ML of the EU,3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.18 Case study C: combination of the results using BMA. . . . . . . . . . . . . 55
3.19 Case study D: Plan view of the four sites at the Sheikh Zayed road in Dubai. 57
3.20 (a) Observations of the uniaxial compressive strength,
(b) histogram and fitted lognormal probability distribution function of
the UCS at the Sheikh Zayed road in Dubai. . . . . . . . . . . . . . . . 58
3.21 Case study D: Combination of the MoMvar , the MoMLA and the ML ap-
proach using BMA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.22 Case study D: Analysis of the indicator correlation lengths using using the
[email protected]
MoMvar , the MoMLA and the ML approach. . . . . . . . . . . . . . . . . . . 59
vi
List of Figures
3.23 Soil classification into 9 soil behaviour types according to Robertson [308]:
zone 1: sensitive fine grained
zone 2: organic soil-peat
zone 3: clay-silty clay
zone 4: clayey silt - silty clay
zone 5: silty sand - sandy silt
zone 6: clean sand to silty sand
zone 7: gravelly sand to sand
zone 8: very stiff sand to clayey sand and
zone 9: very stiff to fine grained soil. . . . . . . . . . . . . . . . . . . . . . . 63
3.24 Combination of measurement data of silty clay, results of the literature
database and results of the CPT databases using Bayesian Model Averaging. 65
vii
List of Figures
5.11 Local sensitivities of the cohesion c′ and the friction angle ϕ′ with respect
to the deterministic footing pressure in parametric study 1. . . . . . . . . . 108
5.12 Local (a) and global (b) sensitivity factors of the full probabilistic analysis
in parametric study 2 for COVϕ′ = 20%, COVc′ = 40% and COVq = 20%. . 108
5.13 Geometry of the tunnel. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.14 Flow area at collapse (a) and typical pressure displacement curve (b). . . . 111
5.15 Variability of the stability numbers Nc (a) and Nγ (b) for different consti-
tutive failure criteria (Mohr-Coulomb, Matsuoka-Nakai, Lade-Duncan) in
comparison to experimental results from literature [196]. . . . . . . . . . . 112
5.16 Influence of the COV of the cohesion c′ and of the friction angle ϕ′ on face
pressure versus probability of failure using the Mohr-Coulomb criterion. . 114
5.17 Limit state surfaces for different levels of the face pressure qt in (a) and
for probability of failure versus correlation coefficient between cohesion
c′ and friction angle ϕ′ in (b) for the MC- criterion. . . . . . . . . . . . . . . 115
5.18 (a) Variation of the COVs for the cohesion c′ , the friction angle ϕ′ and
the tunnel face pressure qt together with the corresponding probabili-
ties
of failure pf .
(b) Variation of the coefficients of variation for the cohesion c′ and the
friction angle ϕ′ for a given coefficient of variation COVqt = 10%
the tunnel face pressure qt , together with the corresponding probabil-
ities
of failure pf . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.19 Reliability indices βi and local sensitivity αi (a) as well as the global sen-
sitivity values δiP C values (b) for parametric study 2 on the tunnel face
stability using different constitutive failure criteria MC, MN, LADE. . . . 116
6.1
(a) Evaluation scheme of the probability of damage pdamage due to
differential settlements,
(b) cumulative distribution function and
(c) histogram with fitted probability density function of the normally
distributed limit state function g, based on an underlying random field
(B/D = 2, H/D = 1, µ = 60 MN/m2 , COV = 50 %, θh = θv = 2D). . . . 123
6.2 (a) Influence of the width B of the building,
(b) influence of the coefficient of variation,
(c) correlation lengths and
(d) influence of the anisotropy of the correlation structure on the surface
settlements due to tunnel excavation. . . . . . . . . . . . . . . . . . . . . 124
6.3 Effects of the ratio of anisotropy θh /θv on the maximum pdamage . . . . . . . 125
6.4 FEM mesh and location of the boreholes (B1,B2 B3). . . . . . . . . . . . . . 126
6.5 Results of parametric study on tunnelling in a layered soil. . . . . . . . . . 127
6.6 (a) Strength reduction factor versus maximum settlement of a 2D
homogeneous slope (c′ = 50 kN/m2 , ϕ′ = 0) and
(b) FEM mesh of the investigated slopes with 8-noded,
[email protected]
quadrilateral elements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
viii
List of Figures
6.7 Effect of varying COVc on the reliability index β and on the probability
of failure pf using lognormally distributed random variables for µc = 50
kN/m2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
6.8 (a) Effects of varying COVc on the probability of failure using different
random field generators with a Θ = θ/H = 2 and random variables
for µc = 50 kN/m2 ;
(b) effects of varying Θ = θ/H on the probability of failure for different
random field generators in comparison to random variables with
a COVc’ = 20%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
6.9 Influence of the spherical, exponential and Gaussian correlation function
on the probability of failure of a slope. . . . . . . . . . . . . . . . . . . . . . 134
6.10 Different measures for the accuracy of the PCE to the system response of a
2D slope stability analysis (a,b) and (c) global sensitivity measures of the
input variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
6.11 Different failure modes derived in Hicks & Spencer [158]. . . . . . . . . . . 137
6.12 Geometry of the layered slope including one borehole using random vari-
ables (a) and random fields (b). . . . . . . . . . . . . . . . . . . . . . . . . . 138
6.13 C ASE STUDY I: Reliability evaluation of a two layered slope stability prob-
lem (a) and the corresponding global sensitivity measures (b). . . . . . . . 140
6.14 C ASE STUDY II: Normalized reliability index of different RFEM approaches
for a two layered slope with respect to a varying Θboundary = θboundary /H. . 141
6.15 C ASE STUDY III: Different determination coefficients for the PCE accuracy. 142
6.16 C ASE STUDY III: Global sensitivity measures δiP C for the isotropic correla-
tion length of the upper layer (1) and the lower layer (2) as well as for the
layer boundary (3) for the PCE expansion order M = 5. . . . . . . . . . . . 142
6.17 Principles of the Pluri-Gaussian Simulation approach, [14]:
Two uncorrelated Gaussian random fields Y1 (a) and Y2 (b) with different
anisotropies are combined via the lithotype rule (c); the red square high-
lights the way the lithotype rule is used to construct the Pluri-Gaussian
random field simulation (d). . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
6.18 Part of the realisation of the Pluri-Gaussian random field (200/200/75m)
including all soil types. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
6.19 2D FEM model used in this case study. . . . . . . . . . . . . . . . . . . . . . 148
6.20 Sample fragility curve for ultimate differential settlements
αultimate = 1/1, 000 and the fitted lognormal cumulative distribution func-
tion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
6.21 Map of allowable loads pSLS on a footing with αulitmate = 1/300 (a) and
αulitmate = 1/500 (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
6.22 Map of allowable loads pSLS on a footing with αulitmate = 1/600 (a) and
αulitmate = 1/1, 000 (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
ix
List of Figures
A.3 Histogram of vertical correlation lengths in cohesive soils for all scales (a),
for small(b), medium (c) and large scales (d). . . . . . . . . . . . . . . . . . 204
A.4 Histogram of horizontal correlation lengths in cohesive soils for all scales
(a), medium (b) and large scales (c). . . . . . . . . . . . . . . . . . . . . . . 205
A.5 Histogram of vertical correlation lengths in frictional soils for all scales
(a), medium (b) and large scales (c). . . . . . . . . . . . . . . . . . . . . . . 206
A.6 Histogram of horizontal correlation lengths in frictional soils for all scales
(a), medium (b) and large scales (c). . . . . . . . . . . . . . . . . . . . . . . 206
B.1 Plan view and typical CPT profile of NGES site A LAMEDA. . . . . . . . . . 208
B.2 Plan view and typical CPT profile of NGES site E VANSVILLE A REA. . . . . 209
B.3 Plan view and typical CPT profile of NGES site L ANCESTER. . . . . . . . . 210
B.4 Plan view and typical CPT profile of NGES site S AN B ERNADINO C OUNTY. 211
B.5 Plan view and typical CPT profile of NGES site S AN L UIS O BISPO C OUNTY.212
B.6 Plan view and typical CPT profile of NGES site S ANTA C LARA C OUNTY. . 213
B.7 Plan view and typical CPT profile of NGES site S OLANO C OUNTY. . . . . 214
B.8 Plan view and typical CPT profile of PEER site A NSSALL. . . . . . . . . . . 215
B.9 Plan view and typical CPT profile of PEER site B ERKELEY. . . . . . . . . . 216
E.1 Random field realisations of the SGSIM (a) and the SISIM (b) algorithm
with an isotropic, spatial correlation (θver = θhor ). . . . . . . . . . . . . . . . 232
E.2 Cumulative probability distribution (a) and variograms (b) of the SGSIM
and SISIM random field in figure E.1. . . . . . . . . . . . . . . . . . . . . . 233
E.3 Indicator correlation lengths for each threshold of the SGSIM and SISIM
algorithm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
E.4 Main characteristics of various random field simulation methods from
Chiles & Delfiner [79]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
E.5 Main characteristics of various random field simulation methods from
Chiles & Delfiner [79]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
G.1 Input variables and system response of an PCE with order M = 5 and a
face pressure of qt = 40 kN/m2 . . . . . . . . . . . . . . . . . . . . . . . . . . 240
G.2 Comparison of non-intrusive SFEM and FORM in terms of the probability
of failure for a COVϕ′ = 5 % and COVϕ′ = 10 % . . . . . . . . . . . . . . . . 241
G.3 Acccuracy of the PCE fitting for a face pressure qt = 40 kN/m2 . . . . . . . . 241
G.4 Collocation points in Gaussian space and physical space for several PCE
expansion orders M. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
G.5 Statistical moments of the approximated system response for a face pres-
sure qt = 40 kN/m2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
[email protected]
G.6 Probability of failure as a function of the PCE expansion order. . . . . . . . 243
x
List of Figures
G.7 Sobol indices for several PCE expansion orders for the face pressure qt =
40 kN/m 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
xi
[email protected]
List of Tables
xiii
List of Tables
xiv
Summary
The core competences of civil engineers are designing, building and maintaining struc-
tures and buildings to enable life and business for society. This includes the prevention
against natural hazards such as climatological, hydrological, meteorological and geo-
physical disasters. Such complex hazards asks for sophisticated techniques to ensure
appropriate safety standards for society. Too low safety standards can result in many
casualties and much economic damage, whereas too high standards results in overly ex-
pansive systems. Therefore, these phenomena ask for sophisticated methods to consider
their impacts on structures. Especially geotechnical engineers are asked for integrated
concepts to design structures withstanding the above mentioned hazards.
This research is focusing on the evaluation of the effects of soil variability within the
framework of probabilistic methods. The evaluation of stochastic soil properties includ-
ing spatially correlated soil properties is deeply investigate on a theoretical basis. These
results are used in case studies on the evaluation of the spatial correlation of different
measurement data sets. These outcomes are also used in different case studies, which
are focusing on the effects of soil variability for typical geotechnical problems.
Chapter 2: The basics of probabilistic site characterization are described within this
chapter including a comparison to the state-of-the-art. This involves the description of
the basics of statistics and geostatistics in order to describe the variability and the spatial
correlation of soil properties. Besides this, the main sources of error in geotechnical
engineering are summarized, which influence probabilistic site characterization. Also an
overview of sampling schemes is provided. Three main methods for the mathematical
characterization of spatial variability are derived. These methods are used to analyse
and quantify spatial variability of soil properties.
Chapter 3 The possibilities and limitations of the three different approaches on the
evaluation of spatial variability can be deduced from the results of four different case
studies. At first, an analytically defined random process is used to show the capabilities
of the three different methods. On basis of this, three case studies are presented on the
evaluation of spatial variability of equally spaced and irregular spaced data sets. These
datasets are used to explain the concepts of probabilistic site characterization, which
includes the identification of trends and layers as well as a novel scheme for the combi-
nation of different models for spatial variability by means of statistical approaches.
The results of these case studies are compared to a large literature study on the spa-
[email protected]
tial correlation of soil properties, which is enriched by the results of an extensive study
xv
Summary
on cone penetration tests in different soil types. Finally, these different sources of in-
formation on the correlation length are combined via the Bayesian Model Averaging
methodology.
Chapter 4: This chapter focuses on the basics of safety and reliability in engineering. It
provides a description of common approaches for dealing with uncertainties and safety
in geotechnical engineering. Global and partial safety factors as well as the basics of
uncertainty quantification and reliability based design are described and compared with
a special focus on the basics of the generation of random numbers and random fields as
well as on the computation of failure probabilities.
Moreover, an introduction into local and global sensitivity analyses is given.
Chapter 5: Typical geotechnical design problems are investigated within the frame-
work of uncertainty quantification.
At first, an analytical limit state equation (LSE) is used to investigate the effects of soil
variability on the design of a tunnel lining.
In another case study the bearing capacity of a strip footing is investigated by means
of probabilistic methods. Within this, semi-analytical LSEs are derived from 2D FEM
simulations, which considers the Mohr-Coulomb, Matsuoka-Nakai, Lade-Duncan fail-
ure criteria to model the soil behaviour. These LSEs are used to investigate the effects
of the soil variability, of the geometry of the problem and of the load uncertainty. Sen-
sitivity analyses evaluate the contribution of the random properties to the probability of
failure.
The tunnel face stability problem is investigated in a similar way. The semi-analytical
LSEs are derived from parametric 3D FEM simulations using the Mohr-Coulomb, Mat-
suoka-Nakai, Lade-Duncan failure criteria. The effects of soil variability, geometry and
construction processes on the probability of failure are quantified by means of uncer-
tainty quantification and sensitivity analyses.
Chapter 6: The effects of spatial soil variability are evaluated in typical problems in
tunnelling and foundation engineering by using the framework of uncertainty quantifi-
cation.
In the first case study, the effects of spatially correlated soil properties on surface set-
tlements, which are induced by the construction of a tunnel, are investigated. Starting
from traditional approaches of a single-layered subsoil, a novel concept for considering
the spatial variability at multiple scales is presented.
The effects of soil variability at different scales are also investigated within slope sta-
bility problems. These studies also include sensitivity analyses to investigate the contri-
bution of spatial variability to the probability of failure.
Finally, the effects of large scale spatial variability is focused in the case study on the
risk-based characterisation of an urban site. The macro-scale variability of the subsoil is
simulated via the Pluri-Gaussian simulation approach. This approach captures the un-
[email protected]
certainty of the boundaries of the spatially distributed soil types, incorporating expert
xvi
judgements, soil investigations and stochastic properties of the soil types. These simula-
tion results analysed using fragility curves. Herein, fragility curves suggest admissible
footing pressure due to differential settlements. These results are used for the generation
of risk maps of the investigated urban site.
Chapter 7: This chapter comprises a summary of the objectives of this thesis and a
summary of the conclusions. Further research topics are announced in the outlook.
[email protected]
[email protected]
Zusammenfassung
Die Kernkompetenzen des Bauingenieurwesens sind das Planen, Erstellen und Erhal-
ten von Bauwerken, um die Existenz und Wirtschaftsleben einer Gesellschaft zu er-
möglichen. Dies umfasst auch Maßnahmen gegen Naturkatastrophen wie z.B. Klima-
wandel, Hochwasser oder Massenbewegungen. Diese komplexen Gefahren erfordern
hochentwickelte Technologien, um ein angemessenes Level an Sicherheit für die Gesell-
schaft zu sichern. Zu niedrige Sicherheiten können zu großen Verlusten und hohen
ökonomischen Schäden führen, wohingegen zu hohe Sicherheitsanforderungen in einem
teuren und einer nicht wirtschaftlichen Bemessung von Bauwerken enden. Daraus kann
abge-leitet werden, dass hochentwickelte Methoden notwendig sind, um die Einwirkun-
gen auf Bauwerke zu simulieren. Speziell in der Geotechnik ist es in Folge der hohen
Variabilität des Untergrundes erforderlich, integrierte Konzepte für die Bemessung von
Bauwerken gegen die zuvor genannten Einwirkungen anzuwenden.
Im Rahmen dieser Dissertation wird die mathematische Beschreibung von Boden-
variabilität und die Bestimmung der Konsequenzen dieser Variabilität in verschiede-
nen geotechnischen Fragestellungen untersucht. Hierbei werden probabilistische Meth-
oden angewendet, um die Folgen von unsicheren und räumlich streuenden Bodenken-
ngrößen zu bestimmen. Neben der Zusammenstellung der theoretischen Grundlagen
der Quantifizierung von räumlichen Varaibliät wird die Anwendung exemplarische in
mehreren Fallbeispielen aufgezeigt. Diese Resultate werden in weiterführenden Unter-
suchungen verwendet, in welchen die Folgen von Bodenvariabilität in geotechnischen
Problembestellungen beispielhaft analysiert werden.
Kapitel 2: In diesem Kapitel werden die Grundlagen für eine probabilistische Charak-
terisierung des Untergrundes zusammengestellt. Dies umfaßt die theoretischen Grund-
lagen der Statistik und Geostatistik, welche für die Beschreibung von Variabilität und
räumlicher Korrelation von Bodeneigenschaften erforderlich sind. Drei gängige Meth-
oden für die Analyse von räumlicher Variabilität in der Geotechnik werden hergeleitet
und einander gegenübergestellt. Diese ist notwendig, um die Ergebnisse der Analyse
von Messdaten in Hinblick auf die räumliche Variabilität zu verstehen und zu inter-
pretieren.
xix
Zusammenfassung
Auf Basis dieser Ergebnisse wird in der zweiten Fallstudie eine Datenbank von CPT
Feldversuchen analysiert, welche im Rahmen einer Großbaustelle durchgeführt wur-
den. Exemplarisch wird hier das Erstellen eines stochastischen Modells des Baugrundes
gezeigt. Hierbei wird mittels empirischer und statistischer Methoden die Variabilität des
Untergrundes auf verschiedenen Skalen charakterisiert. Exemplarisch wird das Kom-
binieren verschiedener Analysemodelle von räumlicher Variabilität aufgezeigt. Dies
Herangehensweise beruht auf dem Satz von Bayes.
In einer weiteren Fallstudie wird die räumliche Variabilität von Steifigkeiten in einer
homogenen Bodenschicht untersucht. Anhand dieser Analyse wird eine Methodik zur
Bestimmung der Unsicherheit von Modellen der räumlichen Variabilität abgeleitet.
In der vierten Fallstudie wird die räumliche Variabilität einer sehr großen Anzahl von
Festigkeitsmessungen analysiert, welche auf verschiedenen Baustellen durchgeführt wur-
den. Die Ergebnisse werden mit den Daten einer umfangreichen Literaturstudie ver-
glichen und verifiziert.
Abschließend wird eine Methodik für die Kombination von gemessener räumlicher
Varaibität mit einer Expertenmeinung vorgestellte. Das Bayes’sche Prinzip wird ver-
wendet, um verschiedene Vorinformationen, Expertenwissen, Literaturergebnisse und
Messdaten miteinander zu kombinieren. Dieses Vorgehen wird exemplarisch an einem
Datensatz aufgezeigt.
Kapitel 4: In diesem Kapitel sind die Grundlagen für Sicherheit und Zuverlässigkeit
zusammengestellt. Ausgehend vom Stand der Technik, bei dem globale und partielle
Sicherheiten verwendet werden, werden die Grundlagen der Zuverlässigkeitsanalyse
von Bauwerken erklärt. Hierbei wird besonders auf die in der Geotechnik gängigen
Näherungsverfahren zur Ermittelung der Versagenswahrscheinlichkeit eingegangen, wo-
bei auch das Generieren von Zufallszahlen und -feldern beschrieben wird.
Die mathematischen Grundlagen für die Konzept zur lokalen und globalen Sensitiv-
itätsanalyse werden am Ende dieses Kapitels erklärt.
Kapitel 5: In diesem Kapitel wird das Konzept der Quantifizierung von Unsicherheit
anhand von verschiedenen Fallstudien in der Geotechnik erläutert. Hierbei wird die
Bodenvariabilität durch Zufallszahlen simuliert.
In der ersten Fallstudie im Tunnelbau werden die Folgen von streuenden Bodenken-
ngrößen in der Bemessung einer Tunnelschale untersucht. Hierfür wird eine analytische
Versagenszustandsgleichung für die Bestimmung der Änderung der Wahrscheinlichkeit
eines Versagens in Folge von streuenden Bodenkenngrößen untersucht. Lokale bzw.
globale Sensitivitäten quantifizieren den Einfluss der streuenden Eigenschaften mit Hilfe
statistischer Methoden.
Die Folgen von Bodenvariabilität auf die Tragfähigkeit von vertikal belasteten Streifen-
fundaments mittels probabilistischen Methoden untersucht. Darüber hinaus werden
neben der Variabilität der Festigkeitseigenschaften des Untergrundes auch die Geome-
trie und die Belastung des Streifenfundamentes analysiert. Dies wird mit verschiedenen
[email protected]
Versagenszustandsgleichungen beschrieben, welche aus 2D FEM Untersuchungen mit
xx
verschiedenen stofflichen Versagenskriterien (Mohr-Coulomb, Matsuoka-Nakai, Lade-
Duncan) abgeleitet werden. Für die Analyse dieser Fragestellung werden Fragilitäts-
graphen verwendet. Fragilitätsgraphen beschreiben in diesem Zusammenhang zwis-
chen aufgebrachter Last und Versagenswahrscheinlichkeit.
In einer weiteren Fallstudie wird Stabilität der Ortsbrust im Tunnelbau untersucht.
Das Versagen der Ortsbrust wird mit verschiedenen stofflichen Versagenskriterien (Mohr-
Coulomb, Matsuoka-Nakai, Lade-Duncan) in Kombination mit Methoden der Zuverläs-
sigkeitsanalyse untersucht, wobei in diesem Zusammenhang neben der Untergrund-
varibilität auch die Geometrie und der Herstellungsvorgang mittels Zufallsvariablen
berücksichtigt werden. Der Vergleich der verschiedenen Ergebnisse wird mit einem Ver-
gleich der Sensitivitäten der verwendeten Zufallsvariablen abgerundet.
[email protected]
[email protected]
Chapter 1
Introduction
The core competence of civil engineers are designing, building and maintaining struc-
tures and buildings to enable life and business for society. This includes the prevention
against natural hazards. Natural hazards can be subdivided into climatological, hydro-
logical, meteorological and geophysical disasters. As defined in the I NTERNATIONAL
D ISASTERS D ATABASE, climatological disasters are (bush, forest, scrub and grassland)
fires, meteorological disasters are local storms, extra-tropical and tropical cyclones. Hy-
drological disasters are general floods, flash floods, mudslides, storm surges and coastal
floods. The complexity of natural hazards becomes more evident in figure 1.1. Herein,
the annual frequency of floods, mass movements, seismic activities, storms, vulcanos,
and wildfires is plotted with respect to the lost lives. These results from the I NTERNA -
TIONAL D ISASTERS D ATABASE [378] show good agreement with data from Christian &
Baecher [23] on nuclear power plants, dam failures, explosions air crashes and man-
caused fatalities. The Dutch Government Group Risk Criteria [395] indicate that these
natural hazards are not acceptable risks and have to be mitigated with advanced con-
cepts. This asks for sophisticated techniques to ensure appropriate safety standards for
society.
Too low safety standards can result in many casualties and much economic damage,
whereas too high standards results in overly expansive systems. Therefore, it is impor-
tant to evaluate the safety of structures with appropriate approaches, which allow to
consider variability and uncertainty in a proper way. These complex phenomena cause
tremendous economical damage as shown in figure 1.2. One can deduce that sophisti-
cated methods to consider their impacts on structures are urgently needed. Especially
geotechnical engineers are asked for integrated concepts to design structures withstand-
ing the above mentioned hazards.
In international conferences and workshops the state-of-the-art safety concepts are
continuously improved. The global and partial safety factor concepts are mainly driven
by experience and now enriched and extended by the results of probabilistic analy-
ses. Theses new developments help to contribute to more economic and safe design
[email protected]
approaches.
1
Chapter 1 Introduction
annual frequency
1
10
floods
0
mass movements
10 seismic activities
exp storms
l osi
-1 on vulcano
10
wildfire
10
-2 acceptable risk
reduction desired risk
-3
not acceptable risk
10
da
m
f a il
-4
10 ai
ure
nuc
lea
rc
-5 rp
ra
10 ow
s
he
er
st
pl
a
ot
-6
10
al
nt
s
-7
10 1 2 3 4 5 6
10 10 10 10 10 10
lives lost
Figure 1.1: Data on the frequency of natural and industrial disasters against number of
lost humans lives. In colour shading data from Dutch Government Group
Risk Criteria [395] for floods, data on other natural disaster as reported by the
N ATURAL D ISASTERS D ATABASE NATURAL DISASTERS DATABASE and Chris-
tian & Baecher [23].
1.0
Average annual estimated damages
geophysical
Proportion of average annual
damages per disaster group
($US billion) per disaster group
hydrological
40 meteorological 0.8
climatological
30 0.6
20 0.4
10 0.2
0 0.0
A s
A s
a
a
ia
ia
a
a
a
a
pe
pe
ic
ic
si
si
an
an
ic
ic
fr
fr
ro
ro
er
er
ce
ce
A
A
Eu
Eu
m
m
O
O
A
Figure 1.2: Average annual damages ($US billion) caused by reported natural disasters
1990 - 2011 by the I NTERNATIONAL D ISASTERS D ATABASE [378].
[email protected]
2
1.2 Research aim
3
Chapter 1 Introduction
the estimation of tunnelling induced settlements are investigated for a single- and two
layered subsoil using probabilistic methods. In addition to this, the effects of multi-scale
soil variability are also investigated for soil slopes. The quantification of these effects
by means of global sensitivity measures offers insight into the contribution of uncertain
properties of single- and two-layered soil slopes. In the case study on the risk-based site
characterisation the effects of macro-scale soil variability are investigated by means of
the Pluri-Gaussian simulation method, expert judgement and the framework of uncer-
tainty quantification.
Chapter 7 is a summary of the most relevant findings of this research, drawing con-
clusions and including recommendations for further research.
4
Chapter 2
Characterizing soil variability at different scales
2.1 Introduction
Interpretation of site exploration data is illustrated in figure 2.1 [22]. The presented geo-
logical maps were drawn 30 years apart and are quite different from each other, although
the sample data are the same. According to Baecher [22], the theory of geology was re-
evaluated in the profession and this led to a different reinterpretation of the data.
One can conclude from [22] that for site characterization only parts of exploration can
be statistically modelled. Statistical analysis of data can at most indicate how to logically
modify what was thought before to what should be thought after.
For this reason, it is useful to use different statistical approaches and schemes to de-
scribe various phenomena in detail. Within this chapter, the basics of probabilistic site
characterization are described, which involves a summary of uncertainties and main
sources of error in geotechnical engineering. The focus of the author is to enlighten
the mathematical framework to describe spatial variability at different scales. Herein, a
basic description of the uncertainty of the spatial correlation is offered together with a
summary of the anisotropy ratios of the spatial correlation.
Figure 2.1: Mapping of the same area of Canada in 1958 (left) and in 1923 (right) using
[email protected]
the same data [22].
5
Chapter 2 Characterizing soil variability at different scales
• Small and relative simple structures, for which basic stability and performance
requirements can be fulfilled by experience and qualitative ground investigation
and for which risks are negligible. For this category, it is assumed that ground
conditions are known by experience.
• This category refers to major civil engineering projects in difficult ground where
both large sampling and large modelling is justified.
6
2.3 Sources of soil variability
the definition of uncertain regions of a site and to optimize any additional characteriza-
tion efforts with the objective of reducing the amount of uncertainty. Journel and Alabert
[189] report that the amount of material that is sampled in site characterization boreholes
is typically only 10−6 to 10−9 of the total site volume, which creates a need for statistical
techniques for characterizing uncertainty.
Several authors [320, 337, 341] used methods of artificial intelligence, e.g. artificial
neural networks or support vector machines to characterize soil properties within a sta-
tistically homogeneous soil layer. Also geostatistical interpolation simulation techniques
can be used to evaluate the uncertainty of a conceptual geological-geotechnical model
[79, 250].
7
Chapter 2 Characterizing soil variability at different scales
Table 2.1: Approximate guidelines for coefficients of variation of some design soil pa-
rameters taken from [183, 282, 283].
8
2.3 Sources of soil variability
shall take this into account; the transformation error is introduced through the design
processes [166]. Uncertainties of the load, of the (mechanical) model and of the decision
model are finally ’hidden’ in the design result, as described in detail in Honjo & Kuroda
[166]. As stated in the introduction, the focus of this chapter is in the description of
the uncertainty estimation, which is introduced through the setting up of the conceptual
geological ground model.
After doing site investigations, the engineer’s aim is to estimate the values at unsam-
pled locations. One would intuitively choose a procedure like interpolation in order to
take the sampled properties in the neighbourhood of an unsampled location. A vast va-
riety of methods for interpolation can be found in different textbooks, which are very
present in fields like hydrology, statistics or geostatistics. These interpolation methods
are based on the concept of spatial correlation. In contrast to the interpolation approach,
there are the geostatistical simulation approaches, which are described in chapter 4.
load modelling
uncertainty uncertainty
parameters spatial
values
time preferences
SPT values SPT values j’-values
z z z
m±s m±s
Figure 2.2: Uncertainties in Reliability Based Design modified from Baecher & Christian
[23], Honjo [166], Phoon[286] and Phoon & Kulhawy [282, 283].
9
Chapter 2 Characterizing soil variability at different scales
sa
failure compacted
surface P grain
zone
sf
dilative
“homogeneous” shear zone void
soil band
Figure 2.3: Illustration of the multi-scale nature of soil after Borja [48], Chen et al. [73],
[email protected]
Christakos [82] and Wackernagel [401].
10
2.4 Describing spatial variability
The geotechnical level is between the specimen scale and the geological scale; there-
fore, it is important to keep in mind that there is not a single spatial scale, but multiple
spatial scales contributing to soil variability. Of course, this plays a role in the evalua-
tion of spatial variability of soil properties as well in the evaluation of the effects of soil
variability.
11
12
Scale name: Basin Depositional envi- Channels Stratigraphical Flow regime fea- Pores
ronments features tures
Approximate 3 km– 100 km 80 m–3 km 5 m–80 m 0.1 m–5 m 2 mm–0.1 m < 2 mm
length scale
Geologic features Basin geometry, Multiple facies, Channel geome- Abundance of Primary sedimen- Grain size, shape,
strata geometries, facies relations, try, bedding type sedimentary tary structures: sorting, packing, ori-
[email protected]
structural fea- morphologic and extent, lithol- structures, strat- ripples, cross- entation, composi-
tures, lithofacies, features ogy, fossil content ifcation type, bedding, parting tion, cements, inter-
regional facies upward fning/ or lineation, lamina- stitial clays
trends coarsening tion, soft sediment
deformation
Heterogeneity Faults (sealing) Fractures (open Frequency of Bed boundaries, Uneven diage- Provenance, dia-
by Koltermann & Gorelick [204].
affected by folding, external or tight), intra- shale beds, minor channels, netic processes, genesis, sediment
controls (tec- basinal controls sand and shale bars, dunes sediment trans- transport mecha-
tonic, sea level, (on fuid dynamics body geometries, port mechanisms, nisms
climatic history), and depositional sediment load bioturbation
thickness trends, mechanism) composition
unconformities
Observations/ Maps, seismic Maps, cross- Outcrop, cross- Outcrop, lithologic Core plug, hand Thin section, hand
measurement profles, cross- sections, litho- well tomography, and geophysical sample, outcrop lens, individual
Chapter 2 Characterizing soil variability at different scales
techniques sections logic and geo- lithologic and logs clast, aggregate
physical logs, geophysical logs analysis
seismic profles
Support volume of Shallow crustal Regional (long Local (short term Near-well Core plug anal- Several pores
hydraulic measure- properties term pumping or pumping or tracer (non-pumping ysis (permeame- (mini-permeameter)
ments tracer tests) tests) tests-height of ter)
screened interval)
Table 2.2: Different scales of spatial variability of soil properties of sedimentary deposits
2.5 Quantifying spatial dependence
correlation. In the case of a multivariate distribution, all random variables are linked
through a covariance matrix, as defined in equation C.1 in appendix C.
The basic assumptions for the description of spatial variability are as follows:
• S TATIONARITY is defined as when the mean value µ and the standard deviation σ
are constant over the whole domain. Moreover, the covariance C(τ ) is only depen-
dent on the separation τ and not on the absolute position [403].
13
Chapter 2 Characterizing soil variability at different scales
The variogram does not require the knowledge of the mean of the random function Z(X)
because the squared difference in equation 2.2 eliminates the mean value. Moreover,
small variations are filtered out [79].
It has to be pointed out that the variogram approach describes the spatial dependence
as an integral of the whole distribution of parameter values. The spatial correlation of
the extreme values of the random function Z(X) cannot be investigated separately. In
contrast to this is the indicator approach. In equation 2.3 a threshold cutk is used for
truncating the random function Z(X), which is investigated by the variogram approach.
The truncation of the random function introduces difficulties in the estimation of the
spatial correlation because the truncated random function suffers from a sometimes very
skewed distribution. As shown in the following case studies in chapter 3, this causes
additional difficulties in the estimation of the spatial correlation.
Choosing different percentile values (appendix C) of the cumulative distribution func-
tion of Z(X), one can analyse the different spatial correlation of extreme values. In fact, it
has long been recognized [29, 189] that different percentile values for example extremes
can have a different spatial dependence structure from the central values. Indicator vari-
ograms can be used to express the difference in dependence as a function of the observed
values. This requires the specification of the random function Z(X) and threshold cutk
to create the indicator transform I(X; cutk ) as defined in equation 2.3. Different authors
[29, 176, 189] noted that with the help of indicator variograms in many cases the depen-
dence between variables departs considerably from the variogram approach.
(
1 if Z(X) ≤ cutk
I(Z(X; cutk ) = (2.3)
0 otherwise
The expected value of the indicator random function I(X; cutk ) identifies the cumulative
probability, i.e. the proportion of the property no greater than cutk [29]:
Similarly, the indicator cross-covariances for two thresholds cutk and cut′k and for a sep-
[email protected]
aration distance τ identifies the bivariate (two-point) cumulative distribution function
14
2.5 Quantifying spatial dependence
F (·) [144].
The indicator cross-covariance describes the covariance between cutk and cut′k and is
written as:
For cutk = cut′k the indicator cross-covariance becomes the indicator covariance, which
can be interpreted as a two point connectivity function [144].
It has to be stressed that connectivity is here used in a probabilistic sense; a high value of
the indicator covariance indicates that there is a high probability that any two locations
separated by the vector τ be jointly below the threshold cutk , but this is not enough to en-
sure the existences of a continuous path between two points. The indicator covariances
are symmetric with respect to the median threshold as derived in [144].
Illustrative results of an indicator analysis of an analytically generated random func-
tion can be found in chapter 3.1. The indicator varigoram analysis of measurement data
and the involved difficulties are also shown in the case studies in chapter 3.
15
Chapter 2 Characterizing soil variability at different scales
bn2
σ
Γ2 (Zn ) = (2.10)
b2
σ
Herein, σbn2 is the variance of the derived moving average series of degree n, and σ b2 is
the variance of the original data. If the spacing of the data is d, Zn in equation 2.10 will
be equal to (n − 1)d. The variance function Γ2 (Zn ) in equation 2.10 can be determined
for different sizes of the averaging window, which is used for the caluculation of σ bn .
Wickremesinghe & Campanella [404] derive the scale of fluctuation for large values of Z
and very large values of n.
θ = max( Γ2 (Zn ) Z ) (2.11)
Wickremesinghe & Campanella [404] recommend to pick the value of Γ2 (Z) from the
curve of Γ2 (Zn ) Z vs. n at a reasonably high value of Z, where there is a distinct change
in the curve. MoMLA offers nearly the same results like other methods of the state-of-
the-art for equally spaced and normally distributed data. It can be deduced from the
case studies B, C and D in chapter 3 that the MoMLA approach does not offer reliable
results in the case of non equally spaced measurement data. This can be related to the
different number of measurement data, which have been used to calculate the variance
of different distance classes. This results in a not clearly detectable maximum of equation
[email protected]
2.11.
16
2.5 Quantifying spatial dependence
The covariance matrix CZZ contains the values of the auto-covariance function C(Zi , Zj )
of each possible pair of measurements. Selecting the unknown parameters in a vector
Θ = [Z̄, σr , θh , θv ]T the log-likelihood for Θ is given in equation 2.13.
n 1 1
L(Θ|z) = − ln(2π) − ln |CZZ | − (z − Z̄)T C−1
ZZ (z − Z̄) (2.13)
2 2 2
By maximizing the likelihood, the optimal parameter set Θ can be obtained by stan-
dard optimization strategies, for example the simplex method. The advantage of the
simplex algorithm is that the results are independent of the initial parameters, hence
only depending on data.
De Groot & Baecher [90] state that the maximum likelihood estimators Θ b are asymp-
totically jointly normally distributed:
b ∼ N (Θ, B−1 )
Θ (2.14)
B = diag(BZ̄ , BΘ ) (2.15)
BZ̄ = 1T C−1 1
ZZ
1 −1 ∂CZZ −1 ∂CZZ
(2.16)
BΘij = tr CZZ CZZ
2 ∂Θi ∂Θj
where 1 is a unit vector of length n. Using the information matrix B, the accuracy of the
[email protected]
obtained parameters is estimated.
17
Chapter 2 Characterizing soil variability at different scales
where p(x) is the basis vector of the regression model and β contains the regression
coefficients. If the correlation structure of the measurements is known in advance, the
regression coefficients could be estimated.
β̂ = (PT R−1
ZZ P)
−1
PT R−1
ZZ z (2.18)
where P is the so-called level matrix in the regression model containing the basis vector
terms for the measurement positions, and RZZ is the correlation matrix, which is related
to the covariance matrix by the residual variance CZZ = σr2 RZZ .
The correlation matrix, which can be directly calculated from the correlation lengths
and the measurement positions, is generally not known in advance. Thus, the regression
can be done only by assuming an initial guess and updating the required parameters
iteratively by using either the moment estimator or the maximum likelihood formulation
from equation (2.13). Another possibility is to incorporate the regression coefficients
directly in the maximum likelihood approach as proposed in [90].
n 1 1
L(Θ|z) = − ln(2π) − ln |CZZ | − (z − P β)T C−1
ZZ (z − P β). (2.19)
2 2 2
The accuracy of the estimated parameters can be estimated again using the inverse of the
information matrix B where only a slight modification of equation (2.16) is necessary:
B = diag(Bβ , BΘ )
Bβ = PT C−1 P
ZZ (2.20)
1 −1 ∂CZZ −1 ∂CZZ
BΘij = tr CZZ CZZ
2 ∂Θi ∂Θj
z̄ ≡ Λ z (2.21)
18
2.5 Quantifying spatial dependence
z̄ = P̂ z (2.23)
p increments are linearly dependent on the others [197]. The matrix P̂ has the property
that:
P̂ P = 0 (2.24)
then
P̄ z̄ = P̂ P β + P̂e = P̂ ε (2.25)
and the drift is filtered out, whatever the coefficients β are. Here, the increments z are
assumed to be normally distributed N (µ = 0, σ = 1) and the covariance parameters are
estimated by the minimization of the negative log-likelihood function similar to equation
2.19. Herein, z is substituted by z̄, CZZ by Λ CZZ ΛT and P β is dropped [270].
m m m
σ 2 , θ|z̄) =
L(b ln(2 π) + − ln(m) +
2 2 2
1 1
+ ln(Λ CZZ ΛT ) + ln z̄T (Λ CZZ ΛT )−1 z̄ (2.26)
2 2
where m = n − p. The same method of minimization can be used to obtain the Residual
Maximum Likelihood estimates of Θ.
The accuracy of the estimated parameters can be estimated by using equation 2.27 as
shown in [403].
1 −1
σ= zT Λ CZZ ΛT z (2.27)
n−p
19
Chapter 2 Characterizing soil variability at different scales
where the symbol ”|” is the symbol for conditional probability. In one dimensional prob-
lems a Markov chain is described by a single transition probability matrix. Transition
probabilities correspond to relative frequencies of transitions from a certain state to an-
other state. Theses transition probabilities can be arranged in a square matrix form.
p11 p12 . . . . . . p1n
. .
p11 . . . . . . . . ..
p= ... . . . p . . . ...
(2.29)
lk
.
pn1 . . . . . . . . pnn
where plk denotes the probability of transition from state Sl to state Sk ; n is the number
of states in the system. Thus the probability of a transition from S1 to S1 , S2 , . . . Sn is
given by p1l , l = 1, 2, . . . n in the first row and so on. The matrix p has to fulfil specific
properties. Its elements are non-negative, plk ≥ 0 and the elements of each row sum up
to one. For a more detailed explanation and extension of this approach (e.g. to more
dimensions, to multiple steps or to conditioning to measurements) and application of
this concept the reader is referred to Elfeki [115] and Caers [64].
Copula approach: Geostatistical literature [4, 29, 31, 151] offers via the copulas statis-
tics another means to describe spatial dependence. Copula describe the dependence
structure between random variables in a more general but also complex way than the
variogram. Therefore, copulas are useful tools to describe the spatial dependence. A de-
tailed explanation of copulas and their relationship to indicator variograms is presented
by Bardossy [29].
Table 2.3: Types of sampling strategies defined by two sources of randomness according
to Brus & de Gruijter [56].
20
2.5 Quantifying spatial dependence
In the design based approach, stochasticity is introduced at the stage of sampling [56];
the sample locations are selected by predefined procedures such as pure random sam-
pling. Stratified sampling, clustered sampling or nested sampling offer a smooth way of
describing a heterogeneous population [21, 23, 56, 84]. Basically spoken, homogeneous
groups are set up inside the heterogeneous population, in which properties can be linked
to the variance of the population. This approach is widely used in geostatistics and will
be described in detail later.
In the model-based approach the soil forming process, which has led to the field of val-
ues of a property in the study area, is modelled as a stochastic process. The difference
of these two approaches is how the sample data are weighted. In the design-based ap-
proach the weights are derived from the sampling design; in the model-based approach
the weights are derived from the chosen model and the actual configuration of the sam-
ple locations. According to [56], the model based approach is used in the field of geo-
environmental engineering in the context of contaminated soils.
Chiles & Delfiner [79] state that the choice of the sampling concept depends on the
objective of the sampling (e.g. exploration or placement of a new well in petroleum
exploration). Of course, one has also to take practical constraints such as accessibility
and costs into account. They emphasize that the randomness as introduced by sampling
on a regular grid is smaller compared to stratified random sampling and pure random
sampling.
In geotechnical literature, Baecher [21] provides an overview of search theory and its
implications in probabilistic site description. Later, Tang [370] and Halim & Tang [150]
worked on geometric models of anomalies using the Bayesian principle. Herein, single
stage search and grid search concepts are used in a Bayesian context leading finally to a
sequential search approach.
Recently, Schweckendiek & Calle [331] used the Bayesian approach in a risk based con-
cept. Li et al. [220] proposed an alternative approach using copula statistics in this con-
text. In the field of geography [146] an alternative approach based on the fuzzy method
is presented. Sampling schemes, which are optimized by adaptive sampling [237] pro-
cedures or simulated annealing [359] are presented in the soil sciences as promising new
developments.
21
Chapter 2 Characterizing soil variability at different scales
100
0
0.20.5 3 6 25
mutual distance t in meter
n elementary variograms (figure 2.4). This approach is not very common in geotechnical
literature [184, 384].
Xn
γ(τ ) = γi (τ ) (2.30)
i=1
An alternative way is offered by the so called fractal approach. Herein, fractals are used
to model the self-similarity of natural phenomena. This offers the opportunity to de-
scribe the phenomena at different scales and fractals can portray an exact self-similarity.
Fenton [126] suggested that such fractal or long memory behaviour is likely to be present
due to the large scale mixing processes (e.g. erosion, transportations, deposition and
weathering) that are involved in the formation of soils.
According to Webster & Oliver [403], Bellehumeur & Legendre [33], Cheng [74] and
Burrough [59–61], the fractal dimension of transects and surfaces can be related to the
variogram. One big advantage of this approach is the modelling of irregular spaced
data, but a vast amount of data are needed for this method.
In environmental and agricultural sciences, the nested sampling approach is used to
characterize multi-scale phenomena. Nested sampling refers to a form of multi-stage
sampling, because the higher stage units are ”nested” within the lower stage units [281,
403].
Apart from these nested structures and the nested sampling approach, various authors
offer hybrid approaches combining the different approaches mentioned above or adding
new promising schemes like adaptive schemes [237], sequential schemes [187], Bayesian
approaches [150] or copula statistics [220].
22
2.6 Variogram calculation
1 X
γ̂(τ ) = [Z(Xi ) − Z(Xj )]2 (2.31)
2N (τ )
(ij)∈R(τ )
where
R(τ ) = {τ − w/2 ≤ ds (ui , uj ) ≤ τ + w/2} (2.32)
dS (Xi , Xj ) is the spatial distance between the spatial point sets Xi , Xj , N (τ ) is the num-
ber of pairs in R(τ ) and w is the width of the spatial distance class as shown in figure
2.6. The bigger the w becomes, the smoother is the semi-variogram because the ε filters
out the very high and low values.
The variograms of second-order stationary processes reach upper bounds, at which
they remain constant after their initial increases as shown in figure 2.7. A variogram
may reach its sill at a finite lag distance, in which case it has a range, also known as the
correlation length; since this is the range at which the autocorrelation becomes 0 (figure
2.7). This separation marks the limit of spatial dependence. Places further apart than
this are spatially independent. For practical purposes their effective ranges are usually
taken as the lag distances at which they reach 95% of their sills [79, 97, 190, 401].
Some semi-variograms may approach their sills asymptotically, and so they have no
strict ranges. This can indicate a trend in the data. In some instances the variogram
decreases from its maximum to a local minimum and then increases again, figure 2.7.
This maximum is equivalent to a minimum in the covariance function, which appears
as a hole. This form arises from fairly regular repetition in the process. A variogram
(a) (b)
lag 1 lag 1
lag 2 lag 2
lag 3 lag 3
Figure 2.5: Comparison for computing a variogram from regular sampling on a transect:
(a) with a complete set of data, indicated with • and
[email protected]
(b) with missing values indicated by ◦ from [403].
23
Chapter 2 Characterizing soil variability at different scales
Figure 2.6: The geometry for discretizing the lag into bins by distance and direction in
two dimensions from [403].
that continues to fluctuate with a wave-like form with increasing lag distance signifies
greater regularity.
If there is only a neglectable or just a little range in the semivariogram, this is called
a nugget effect. This discontinuity at the origin of the semivariogram is used to charac-
terize the residual influence of all variabilities, which have a range much smaller than
the available distances of observation [190]. The nugget effect is equivalent to the well
known phenomenon of white noise in physics.
A semivariogram is said to display a hole effect when its growth is not monotonic
and shows bumps, which reflects the tendency for high values to be systematically sur-
rounded by low values and vice versa [79]. Journel & Huijbregts [190] attribute the hole
effect to different reasons. They recommend in case of periodic sampling distances that
the hole effect may result in a refined investigation scheme.
24
2.6 Variogram calculation
Gaussian variogram
spherical variogram
nugget
bilinear variogram r(t)
effect
range lag t
lag t
Figure 2.7: Theoretical variogram functions (a) and comparison of semivariance function
and autocorrelation function (b).
behaviour at the origin, equation 2.34), exponential variogram (linear behaviour at the
origin, 2.35), Gaussian variogram (parabolic behaviour at the origin, equation 2.36), models
without a sill (power functions, fractal model, logarithmic variogram). Other models
like the cubic model, generalized Cauchy models, K-Bessel model, power-law model,
pentaspherical model, Matern model or logarithmic model can be found in standard
geostatistical textbooks [79, 190, 401, 403].
(
τ /a for τ ≤ a,
γ(τ ) = (2.33)
a for τ > a
(
τ 3
c{ 23τb − 1
} for τ ≤ b ,
γ(τ ) = 2 b
(2.34)
c for τ > b
τ
γ(τ ) = 1 − exp − (2.35)
c
τ2
γ(τ ) = 1 − exp − 2 (2.36)
d
25
Chapter 2 Characterizing soil variability at different scales
Herein, γ
b(τj ) is the vector of the empirical variogram, V is the variance-covariance
matrix of γ
b(τ ). The calculation of the variance covariance matrix is rather compli-
cated as highlighted by Ortiz & Deutsch [268].
V(τ ) = E [ Z(Xi ) − Z(Xi + τ ) ]2 · [ Z(xj ) − Z(Xj + τ ) ]2 − [ 2 γ
b(τ ) ]2 (2.39)
X
n
Q(b) = γ (τ ) − γ(τj ; b))]2
wj2 [b (2.40)
j=1
Herein, w is the weight, which can be the reciprocal of the number of pairs at each
lag, as proposed by Matheron [240] or also the variance at each point [79, 268]. The
variance-covariance matrix in equation 2.39 can be used to calculate the variance
of the variogram. The expression 2.41 tells us that the uncertainty in the variogram
at a distance τ is the average covariance between the pairs of the pairs used to
calculate the variogram for that particular lag assuming a multivariate Gaussian
distribution of the variables.
n(τ ) n(τ )
1 XX
w= 2
= n(τ )/ Vij (τ ) (2.41)
σ2γ(τ ) i=1 j=1
There are different approaches for choosing the weights w. Cressie [87] shows
that for equally spaced Gaussian variables, the variance of the estimates can be
approximated by equation 2.42:
N (τ )
w≈ (2.42)
b(τ ) ]2
2[γ
26
2.6 Variogram calculation
27
Chapter 2 Characterizing soil variability at different scales
Table 2.4: Relationship between the scale of fluctuation δ and the correlation distance θ
for various autocorrelation functions from Vanmarcke [384].
Autocorrelation function δ θ
δ
(
1 − |τ | /a for |τ | < a
ρ(τ ) = bilinear model a 1
0 for |τ | ≥ a
ρ(τ ) = e −|τ |/b
Markov model 2b 1/2
2 √ √
ρ(τ ) = e−(|τ |/c) Gaussian model πc 1/ π
The correlation length is defined without the factor of 2 shown to the right-hand side
of equation 2.47 especially in the geostatistical literature e.g. Journel & Huijbregts [190].
Equation 2.47 implies that θ has to be finite; otherwise, alternative concepts like fractal
processes have to be used [127]. Another consequence for the application in engineering
sciences as well as in earth sciences is that the correlation function is only meaningful
for strictly non-negative correlation functions.
The correlation length can also be defined in terms of the variance function in the local
averaging context as a limit [127, 385], whereas the correlation length θ is assumed to be
finite.
θ = lim T γ(T ) (2.48)
T →∞
DeGroot & Baecher [90] state that the Method-of-Moments (MoM) is unbiased in the
case of infinite samples. Otherwise, the correlation length is dependent on the sample
size. It can be clearly seen in figure 2.8 that for a finite record length L of the data, which
are sampled with a mutual separation distance τ , the estimation of the correlation length
is strongly biased. These findings coincide with the suggestions of Journel & Huijbregts
[190] or Chiles & Delfiner [79]. They recommend to sample with a distance between the
[email protected]
measurement points that is at least smaller than 1/5 to 1/4 of the correlation length.
28
2.7 On the correlation length
1.0
0.4
0.2
L/t →∞
0.0 20
10
5
-0.2 3
1
-0.4
0 1 2 3 4
Normalized lag distance t/t0
Figure 2.8: Dependency of the correlation length on the sample size, taken from DeGroot
& Baecher [90].
29
Chapter 2 Characterizing soil variability at different scales
Table 2.5: Typical ratios of the horizontal θhor and vertical correlation lengths θver col-
lected from literature [27, 79, 96, 97, 190, 204, 401, 408].
point bars
fluvial
eolian
estuariane
deepwater
deltaic
carbonates platform
is governing the whole evaluation of the correlation length. The estimation of the uncer-
tainty of the evaluated spatial correlation is a by-product of the ML procedure as shown
in equation 2.16 and for data with a linear trend equation 2.19. Via these equations it is
possible to estimate the error of the estimated variogram model-parameters and conse-
quently also of the correlation length.
30
2.8 Synopsis
2.8 Synopsis
Within this chapter different approaches to quantify spatial variability are presented,
which are used to quantify the different scales of soil variability. The literature review
on this showed that the most used approaches are the variogram approach, the local
average approach and the Maximum Likelihood approach. Therefore, the author is con-
centrating on these methods to evaluate purely the spatial variability of soil properties
without considering a trend.
31
[email protected]
Chapter 3
Case studies on the evaluation of spatial
variability
This chapter provides a comparison of the Method-of-Moments (MoM) and the Maxi-
mum Likelihood (ML) approaches by applying them to four different case studies. The
Method-of-Moments methods include the variogram (MoMvar ) in chapter 2.5.1.1 and the
local average approach (MoMLA ) in chapter 2.5.1.3. These approaches are used to anal-
yse data fulfilling the basic assumptions of stationarity, homogeneity and ergodicity. Via
this, the possibilities and limitations of these approaches are discussed while applying
them to analytically generated random sequence as well as to equally and non equally
spaced data.
Moreover, the indicator approach is used to analyse the correlation structures of mea-
surement data to compare theses results to the indicator correlation lengths of an analyt-
ically defined random process.
The results of these case studies are compared to the literature database, which is cov-
ering the author’s knowledge of publications on the spatial variability of soil properties.
On top of this, the results of a study on the spatial correlation of different soil types
is presented, which is based on the analysis of CPT databases. Finally, these different
sources of information on spatial variability are merged via the Bayesian Model Averag-
ing approach.
33
Chapter 3 Case studies on the evaluation of spatial variability
3.1.2 Analysis
The MoMvar and the MoMLA approach as well as the ML approach are used to anal-
[email protected]
yse a set of 300 independent, identically distributed (iid) random functions Ẑ. MoMvar
34
3.1 Case study A – Random function
2.0
probability density
COV = 1 %
COV = 50 %
function
1.5
COV = 100 %
1.0
0.5
0.0
0 1.0 2.0 3.0 4.0
generated data in [m]
Figure 3.1: Lognormal distribution with a mean value of Ȳ = 1 and different coefficients
of variation.
Cumulative probability
0.6 0.6
0.3 0.3
0.0 0.0
-2 -1 0 1 2 1 2 3 4 5
Data Data
35
Chapter 3 Case studies on the evaluation of spatial variability
(a) (b)
2
1.5
15 1
0.5
0 0 1 2
0 0 1 2
10 10 10 COV in % 10 10 10 COV in %
0 3 30 g1 0 3 30 g1
ML MoMLA MoMvar
(c) (d)
nugget effect / sill in %
40 ε = 0.01
|1-q/qtarget| in %
4
20
2
0 0 0
1 2
10 10 10 COV in % 10
0
10
1
10
2
COV in %
0 3 30 g1 0 3 30 g1
Figure 3.3: Evaluated correlation lengths using MoMvar , MoMLA and ML approach (a,c)
and resulting nugget/sill ratios (b,d) using MoMvar and ML approach for dif-
ferent error levels ε = 0.01 (a,b) and ε = 0.0001 (c,d).
and MoMLA have been implemented into a M ATLAB program to analyse the data in an
efficient way for all case studies in this thesis. The program of Pardo-Iguzquiza [270]
has been modified and applied for the ML approach. The ML approach was chosen for
calculating the correlation length of data without a trend in all following case studies;
otherwise, the Residual Maximum Likelihood method (chapter 2.5.2.2) would be more
difficult to compare because the trend of data is eliminated also. The bilinear, spheri-
cal, exponential and Gaussian variogram functions have been fitted to the experimental
variogram values using the weighted least squares method using the weight defined by
McBratney & Webster [247]. The same theoretical variograms have been used within the
ML approach to evaluate the range of noisy random sequences. After this, the Akaike
Information Criterion approach has been applied to select the best fitting correlation
length of the variogram and the ML approach.
The evaluated correlation lengths of the iid random functions ẑ are fitted to a lognor-
mal distribution function using the maximum likelihood method. In figure 3.3 one can
see the influence of the skewness of the distribution as well as consequences of the noise.
[email protected]
The deviation |1 − θ/θtarget | of the evaluated correlation length from the target correla-
36
3.1 Case study A – Random function
MoMVar MoMLA ML
100 %
60 %
40 %
20 %
0 % 20 % 40 % 60 % 80 % 100 %
Figure 3.4: Case study A: Analysis of the indicator correlation lengths for the different
quantiles of the CDF (log-N (µ = 1.5, σ = 10)) and a measurement noise of
N (µ = 0, σ = 0.01).
tion length θtarget = 10 m is shown as a function of the skewness and asymmetry of the
b
underlying distribution of Z(X). Moreover, the ratio of the nugget effect and the sill is
shown as a function of the COV in figure 3.3.
The bigger the skewness of the underlying distribution and the bigger the measure-
ment noises, the worse is the evaluated spatial correlation. It was also observed by Kerry
& Oliver [193, 194] that the variogram approach becomes unreliable when the data are
strongly asymmetric or skewed as well as in presence of outliers or extreme values. Sim-
ilar to the findings of Webster & Oliver [403], it can be seen in figure 3.3 that as COV and
the skewness increase the nugget and the sill variances also increase. It is also shown
that the ratios of the nugget to sill increase as the skewness increases even though this
is not considered in the generation of the random process. Webster & Oliver [403] also
observed this and point out that the ratio of nugget to sill is a combination of the degree
of asymmetry and the spatial distribution of data points of the tail of the distribution.
By comparing the three approaches in figure 3.3, MoMvar and MoMLA show more sta-
ble results in comparison to the ML approach, which can be traced back to the basic
assumptions of normally distributed data. All three approaches become more unreliable
the more noisy the data are.
The indicator approach is employed to evaluate the correlation lengths of 300 iid ran-
dom processes ẑ. The different percentiles of the cumulative distribution function of
[email protected]
ˆ are used as thresholds cutk , as described in section 2.5.1.1. The random process
Z(X)
37
Chapter 3 Case studies on the evaluation of spatial variability
Z(X) is truncated by the thresholds cutk . MoMvar , MoMLA and ML approaches are used
to calculate the indicator correlation lengths θind . One would expect that theses indicator
correlation lengths would be the same, but this is not the case as shown in figure 3.4.
This becomes clear, if one recalls the generation of the random process: only one correla-
tion length is used to define the spatial correlation. The indicator correlation lengths of
the different thresholds are symmetric towards the median value, as reported in [143].
Moreover, the extreme values have significantly lower correlation lengths than the me-
dian value, as shown in in figure 3.4.
3.1.3 Remarks
The non-parametric MoMvar technique has proven its strengths in comparison to the
MoMLA and ML approach. It can be seen in figure 3.3 that the variogram technique
is more robust than the MoMLA and ML approaches in the presence of a high noise
as well as when the underlying distribution is highly skewed and asymmetric, which
coincides with the findings in [193, 194]. The MoMLA and ML approaches use the q-q
transformation to convert the underlying distribution into standard normal distributed
variables N (µ = 0, σ = 1). This transformation causes the inaccuracies shown in figure
3.3, especially in the presence of an asymmetric distribution.
It has to be pointed out that the variogram methodology is working in the best way
for normally distributed values N (µ = 0, σ = 1). If the underlying distribution becomes
skewed the outcome is less reliable, which is also true for the MoMLA and the ML ap-
proach. For this reason, robust techniques are offered in a geostatistical framework as
described in detail amongst other by Chiles & Delfiner [79], Bardossy & Kundzewicz
[30] and Marchant & Lark [236]. Following publications summarized in [403], differ-
ent estimators like the Cressie & Hawkins estimator, median variogram estimators or
Genton’s estimator can overcome this problem in a complicated way. The form of these
estimators does not allow explicit computing of their correlation structure as stated by
Genton [137].
There are differences in the evaluated correlation length using the non-parametric
MoMvar and MoMLA approaches as well as the parametric ML approach. Moreover, one
has to keep in mind that there has been performed a q-q transformation for the MoMLA
and the ML. The difference for the ML approach can be related to the smoothing due to
the definition of lag classes in the variogram approach. In this context Webster & Oliver
[403] stress the benefit of the ML approach not to smooth the spatial structure because
there is no ad hoc definition of lag classes; the model parameters are calculated directly
from the variance-covariance matrix of the full data. The ML and MoMvar approach as-
sume that the data follow a multivariate Gaussian distribution, which is a simplification
of the data and very difficult to verify in practice [403].
The relative error εtarget is shown for the MoMvar , MoMLA and ML approaches for a
lognormally distributed random function with a COV= 1 % in figure 3.5 (a) and COV=
100 % in figure 3.5 (b). A measurement noise of N (µ = 0, σ = 0.01) was added to show
the influence of the asymmetry of the underlying lognormal distribution function on the
[email protected]
evaluated correlation length. The shifted, lognormal distribution functions show in the
38
3.1 Case study A – Random function
case of a nearly symmetric distribution with COV = 1 % comparable results. The MoMvar
and ML approaches offer more or less the same result for the correlation length. The
mean value of the distribution function is below εtarget < 10 %; the results of the MoMLA
approach offer a slightly higher relative error εtarget . In the case of a COV= 100 % of the
underlying distribution, the robustness of the MoMvar approach is strengthened again
in comparison to the ML and MoMLA methods, which are very sensitive to asymmetric
and skewed distributions.
For the correlation lengths in general, it can be concluded for that one has to pay atten-
tion to the univariate distribution of the data: the higher the skewness and asymmetry,
the more attention has to be paid in evaluating the correlation length. One has to pay
more attention when analysing skewed data because then the results of the three meth-
ods scatter significantly.
0.14
probability density funciton
probability density funciton
0.1
0.12
0.08
0.1
0.06 0.08
0.06
0.04
0.04
0.02
0.02
0.00 0.00
-50 % 0 % 50 %100 % -50 % 0 % 50 %100 %
εtarget = (θ - θtarget )/θtarget εtarget = (θ - θtarget )/θtarget
Figure 3.5: Case study A: distributions of the relative errors εtarget of the MoMvar , MoMLA
[email protected]
and ML- approaches for COV = 1 % (a) and COV = 100 % (b).
39
Chapter 3 Case studies on the evaluation of spatial variability
Figure 3.6: Case study B: plan view of the 138 CPT measurement locations.
40
3.2 Case study B – CPT data evaluation
15
20
25
30
Figure 3.7: Case study B: Average, minimum and maximum measurement values of the
CPT data.
depth of the field investigations sand can be found again with different properties: fine
to coarse sand, sometimes even sand with a silty part. The density has been detected
over a range from loose to dense. The groundwater level has been measured at a depth
of 0.6 to 1.4 m below ground level. Due to the high precipitation, the groundwater level
is expected to be at the surface of the area.
The summary of the measurement results is shown in figure 3.7. The upper and lower
bounds of the measurement data are shown together with the mean value of the CPT
measurements.
41
Chapter 3 Case studies on the evaluation of spatial variability
Figure 3.8: Steps of the stochastic site characterisation and involved methods.
Using this scheme, one can separate different scales of spatial variability (figure 2.3):
the large geologically based spatial variability are separated from the meso-scale phe-
nomena, which can be investigated without injuring the basic assumptions of the theory
pointed out in section 2.4.2. Via this separation of the different scales of spatial variabil-
ity, it is possible to reduce the noise significantly, which was simulated in section 3.1 via
a normally distributed random noise ε because the homogeneity and stationarity criteria
are fulfilled.
By looking at the measurement data CPT(z) in figure 3.7, one can clearly indicate a
trend of the measurements with depth, which can be described by equation 3.4 where
m(z) is a deterministic function giving the mean measurement value at a depth z below
the surface level; and ε(z) are the random residuals.
To apply the theory to evaluate the spatial variability, one has to check the basic assump-
tions of this theory in order to use the MoMvar MoMLA and ML approaches. If a trend
is not removed from the measurement data, the evaluation of the correlation lengths is
more difficult or even impossible as shown in figure 3.10 (b). Therefore, a selection of
techniques to identify and test measurement data on stationarity and homogeneity are
presented in this section.
S TATIONARITY can be defined by a constant mean value and variance within the test
data. Various authors have proposed techniques for detecting the stationarity of the
data. Only some are enlightened in this section. Amongst others, Jacksa [184] as well as
Bennett [36] propose different methods to detect a trend.
• V ISUAL INSPECTION: Mere inspection of the raw data is often sufficient to detect
non-stationarity. Visual inspection, however, is not sufficient to detect the form of
the trend, nor most cases of non-stationary variance.
• H ISTOGRAM PLOTS: A simple, though crude technique, is to split the random field
into a number of subsections and to plot each of their histograms. Comparison of
these histograms enables shifts in the means and variances to be detected.
• I NSPECTION OF THE EMPIRICAL VARIOGRAM: Box and Jenkins [51] as well as Ben-
[email protected]
nett [36] investigated this approach amongst others. It can be deduced from look-
42
3.2 Case study B – CPT data evaluation
ing at equation (2.2) that non-stationary data will impose also a trend in the em-
pirical variogram. The range cannot be determined. This feature can be used to
indicate a trend of measurement data as shown in figure 3.10.
• S IGNIFICANCE TESTS ON TRENDS: Another smooth way to detect stationarity of
measurements is offered by the non-parametric method of Mann-Kendall’s τ [234].
Alternatively, Sachs [316] offers a different test for testing the significance of trends,
namely the Cox-Stuart test or Neumann test. The Mann-Kendall test is based on
the statistics S. Each pair of observed values yi , yj (i > j) is inspected to find out
whether yi > yj or yi < yj . Let the number of the former type of pairs be P , and the
number of the latter type of pairs be M . Then S is defined as
S =P −M (3.5)
For n > 10, the sampling distribution of S is as follows. Z follows the standard
normal distribution, where
(S − 1)σs if S > 0
Z= 0 if S = 0 (3.6)
(S + 1)σs if S < 0
r
n(n − 1)(2n + 5)
σs =
18
The null hypothesis that there is no trend is rejected if the computed Z value is
greater than Zα/2 in absolute value. Herein, the significance α = 1% is chosen
according to [287].
After identifying the trend via the described approaches, the least squares method [34]
is used to fit a linear trend to the data as recommended by [23, 286]. Alternatively, one
could also use different approaches based on the Bayesian principle or on a Maximum
Likelihood method [23, 286].
After removing the trend of the measurement data by fitting a linear function by least
squares, Maximum-Likelihood or Bayesian approach to the measurement data, one has
to check the data on homogeneity. This can be done be engineering judgement; but also
statistical can support this judgement on a mathematical basis.
Homogeneity can be defined as stationarity of the variance as presented in [284, 316].
The intra-class correlation coefficient and the Bartlett statistics are used herein for check-
ing the homogeneity of the measurements.
43
Chapter 3 Case studies on the evaluation of spatial variability
0.15 histogram
fitted pdf
0.10
0.05
0.00
-10 -5 0 5 10 15
residual cone resistance in MN/m²
Figure 3.9: Histogram and fitted normal distribution function of the detrended cone re-
sistances in sand of CPT 4 N (µ = 0.10 kN/m2 |σ = 2.51 kN/m) with a skew-
ness γ1 = 0.40( kN/m2 )3 and an excess kurtosis γ2,excess = −0.51( kN/m2 )4 .
Basically, ergodicity of the mean value and the variance within the moving window
is assumed. This is only valid for symmetric distributions according to various au-
thors [154, 417]. The critical RI value RIcrit is estimated according to Hegazy et al.
[154]. The boundaries are identified quantitatively at locations where RI exceeds the
empirical relationship of the mean µRI and standard deviation σRI of the RI profile:
RIcrit = µRI + 1.65σRI ; it is recommended to check the computed results visually and to
judge the evaluated soil layer. The critical RI according to Hegazy et al. [154] is slightly
higher than the recommendation of Zhang & Tumay [417] RIcrit = 0.7, which is also an
empirical rule. Others [154, 417] also point out that the choice of RIcrit does not seem to
depend on the underlying correlation structures of the profile, which is also discussed
by Phoon et al. [284].
Bartlett statistics: This classical test is used to test the equality of multiple sample vari-
ances for independent data sets. This has not to be taken into account in this case study
because a normal distribution function can be fitted to the residuals of the CPT measure-
ments as depicted in figure 3.9.
For the case of two sample variances, s21 and s22 , the Bartlett test statistic reduces to:
2.30295(m − 1)
Bstat = [2 log s2 − (log s21 + log s22 )] (3.8)
1 + 2/(1(m − 1))
where m is the number of data points used to evaluate s21 or s22 . The total variance s22 is
defined as:
s2 + s22
s= 1
2
While using the Bartlett statistics, one has to keep in mind that this procedure is very
sensitive to non-normally distributed and skewed variables. According to Sachs [316],
[email protected]
he implies that in the case of a small deviation from the symmetric normal distribution,
44
3.2 Case study B – CPT data evaluation
the procedure will not offer reliable results. Especially in the presence of a skewness
γ1 6= 0( kN/m2 )3 and an excess kurtosis γ2,excess 6= 0( kN/m2 )4 , which can be observed
very often in the case of measurement data.
A continuous Bartlett statistic profile can be easily generated by moving a sampling
window over the simulated soil profile. Campanella et al. [68] as well as Wickremesinghe
[405] recommend a window width of approximately the scale of fluctuation in the layer.
This is also pointed out by Phoon et al. [284]. This implies an iterative approach. The
sampling window is divided into two equal segments and the sample variance s21 and s22
is calculated from data points lying within each segment. The Bartlett statistic basically
indicates the difference between the sample variances in these two adjacent segments.
As shown in equation 3.8, the Bartlett statistic is zero if s21 and s22 are equal. Phoon et al.
[284] offer a critical value Bcrit under the framework of the MODIFIED B ARTLETT STATIS -
TICS taking into account the spatial correlation using an exponential model. Herein,
I1 = n/k ranges between 5 and 50 and I2 = m/k where k is the number of points in one
scale of fluctuation; n is the total number of points in the entire soil record and m is the
number of points in one sampling window.
This critical value Bcrit is calculated for every layer to take the different correlation
lengths into account. The lowest critical value Bcrit is used for the whole CPT profile.
In figure 3.10 the detrended measurement data are used to evaluate the RI as well as
the Bartlett statistics. The width of the sampling window is chosen as big as the correla-
tion length. The critical values RIcrit and Bcrit indicate the boundary between both soil
layers.
Data processing: The CPT measurement data are detrended by the a least square fitting
[email protected]
of a linear function within one layer.
45
Chapter 3 Case studies on the evaluation of spatial variability
5 clayey sand 5
10 silty clay 10
Depth
15 15
silty sand
20 20
silty clay
25 25
silty sand
30 30
(b)
g(t)
raw data
20
10
detrended data
0
0 2 4 6 8
lag distance t
Figure 3.10: (a) Cone resistance qc (z) measured with depth z and soil layer
identification of CPT 4 data using RI and Bartlett statistics and
(b) semivariogram on the raw and detrended data of the silty clay layer
at 10 m depth.
46
3.2 Case study B – CPT data evaluation
After identifying the homogeneous section of the CPTs and the detrending by least
square fitting of a linear function to these homogeneous sections, the MoMvar , MoMLAS
and ML approaches are used to evaluate the spatial variability of the site for each layer.
Different theoretical variogram functions (namely spherical, exponential bilinear and
Gaussian functions) are fitted to the experimental variogram values using the weighted
least squares and Akaike Information criterion to identify the best fitted theoretical var-
iogram function. Also within the ML approach the Akaike Information criterion was
used to identify the best suitable theoretical variogram model.
The analysis of the correlation lengths and of the indicator correlation lengths is car-
ried out for the different soil types. For this reason, the charts of Robertson [308] were
used to evaluate the the soil types from the normalized cone resistance and the friction
ratio of a CPT test as shown in figure 3.23. For theses soil layers the the cumulative
distribution function was used for the evaluation of the different thresholds cutk,i . This
has been used for each of the 138 CPT measurements to evaluate the indicator correla-
tion length of each threshold cutk,i . It is found that these indicator correlation lengths
follow a lognormal distribution function. The mean values θind of each threshold cutk,i
are shown in figure 3.12, which is scaled by the correlation length evaluated by the vari-
ogram approach θ.
3.2.3 Remarks
Combination of different models: One can clearly see in figure 3.11 that the MoMvar ,
the MoMLAS and the ML approaches offer comparable results. This can be deduced from
the nearly symmetric distribution of the residual values shown in figure 3.9, because
there is only a very small skewness of the measurement data. These three probabil-
ity density functions of the MoM and ML approaches are equally probable and can be
merged by using the so called B AYESIAN MODEL AVERAGING (BMA) scheme [162, 300],
which is based on the Bayes’ theorem.
The Bayes’ theorem updates a subjective, prior probability distribution f (θ) with a
likelihood function L(θ|z1 , z2 , ..., zn ), which is the conditional probability function of z1 , z2 , ..., zn
e.g. measurement values.
f (θ|z1 , z2 , ..., zn ) ∝ f (θ) · L(θ|z1 , z2 , ..., zn ) (3.10)
The resulting posterior pdf f (θ|z1 , z2 , ..., zn ) of the variable of interest θ is conditioned on
the prior probability f (θ) and on the Likelihood function L(θ|z1 , z2 , ..., zn ), [70, 136, 382].
If θ is the quantity of interest, then its posterior distribution given data D is
K
X
Prob(θ|D) = Prob(θ|Mk , D) · Prob(Mk |D) (3.11)
k=1
This is an average of the posterior distributions under each of the models considered,
weighted by their posterior model probability. In equation 3.11 M1 , M2 , ..., Mk are the
models considered. The posterior probability for the model Mk is given by
Prob(D|Mk ) · Prob(Mk )
[email protected](Mk |D) = PK (3.12)
l=1 Prob(D|Ml ) Prob(Ml )
47
Chapter 3 Case studies on the evaluation of spatial variability
Figure 3.11: Case study B: probability density function of the MoMvar , MoMLA and ML
approach and the combined probability density functions using Bayesian
Model Averaging of the silty clay layer.
where Z
Prob(D|Ml ) = Prob(D|θk , Mk ) Prob(θk |Mk ) dθk (3.13)
48
3.2 Case study B – CPT data evaluation
40 %
20 %
Figure 3.12: Casestudy B: Analysis of the correlation lengths for indicators of the CDF
using MoMvar , MoMLA and ML with fitted lognormal distribution functions
for silty clay layer.
used to interpolate or simulate spatial variable data under the assumption of one single,
known correlation length. On basis of the presented results, the author recommends to
take the most probable value e.g. the mean value from the distribution in figure 3.11 for
this issues. Moreover, the effects of a distributed correlation lengths is investigated in
chapter 6.2.
Indicator correlation lengths analysis: The results of the indicator correlation length
analyses are shown in figure 3.12. Around the median value the indicator correlation
lengths are slightly higher in comparison to the other thresholds, but have nearly the
same indicator correlation lengths for all quantiles of the CDF. Comparing these re-
sults to figures 3.4 and 3.12, one can conclude that the correlation structure of these
CPT measurements cannot be fully characterized by the mean value, standard devia-
tion and covariance function; it can be concluded that this measured structure of spa-
tial variability cannot be described by only one correlation function as it is done in the
well known multi-Gaussian model ,[143]. This was also found out by other researchers
[143, 144, 189]. The extreme low and high values have indicator correlation lengths. The
results of the MoMvar , MoMLA and ML approaches show a comparable behaviour: the
indicator correlation lengths are nearly the same for all thresholds. The scattering of the
indicator correlation lengths can be related to the indicator approach: The symmetrically
distributed data shown in figure 3.9 are transformed through the indicator approach in
equation 2.3 into skewed and asymmetric datasets. As shown before, this asymmetric
distribution causes less reliable results for the correlation length analysis. The MoMvar ,
MoMLA and ML methods are very sensitive to deviations from the normal distribution
(e.g. skewness γ1 > 1, excess kurtosis γ2,excess 6= 0). Finally, it has to be emphasised that
[email protected]
the indicator correlation length analyses clearly show a non-multi-Gaussian behaviour,
49
Chapter 3 Case studies on the evaluation of spatial variability
Figure 3.13: Plan view and cross section of the experiments in the Fasanenhof tunneling
project.
50
3.3 Case study C – Fasanenhof-tunnel
∆p
E = f (ν) d (3.14)
∆d
51
Chapter 3 Case studies on the evaluation of spatial variability
pressure
data acquisition
guide rod control
hydraulic
pump
pressure hose
Figure 3.14: Experimental equipment of the borehole jacking probe according to the DIN
4094-5 [104].
The MoMvar , MoMLA and ML approach have been used to analyse the spatial variabil-
ity of the measurements. The Akaike-Information-Criterion has been used to identify the
best–fit in the variogram and ML approaches. The results are summarized in table 3.2.
The different approaches offer different results: the correlation lengths for loading and
reloading are different for the loading and reloading data within the variogram method,
which is not present in the other approaches; for the initial loading measurements, the
MoMvar and the MoMLA approach have more or less the same results, whereas the ML
method has longer correlation lengths.
The ML approach offers in contrast to MoMvar and MoMLA methods a possibility to
make a statement on the accuracy of the estimated correlation lengths as described in
equation 2.16. To overcome this drawback of the MoMvar and MoMLA methods, the non-
parametric J ACKKNIFE method is introduced to evaluate the reliability of the correlation
length. The basic idea of the J ACKKNIFE approach lies in systematically recomputing the
statistic estimate, leaving out one or more observations at a time from the sample set as
Table 3.1: Lognormally distributed results of the experiments in the Fasanenhof tunnel.
52
3.3 Case study C – Fasanenhof-tunnel
Figure 3.15: Loading, unloading and reloading cycles performed in the boreholejacking
tests according to DIN 4094 − 5 [104].
EL,1
6
EL,1 , EL,2 , EL,3
EL,2
in MN/m²
EL,3
4
2
0 20 40 60 80 100
location in the tunnel in m
ER,1
EU,1 , EU,2 , EU,3
8
in MN/m²
ER,2
ER,3
6
4
0 20 40 60 80 100
location in the tunnel in m
53
Chapter 3 Case studies on the evaluation of spatial variability
ML
60 %
40 %
20 %
0%
0.85 0.90 0.95 1.0 1.05
normalised correlation length qind / q
Figure 3.17: Casestudy C: Analysis of the indicator correlation lengths using MoMvar ,
MoMLA and ML of the EU,3 .
54
3.3 Case study C – Fasanenhof-tunnel
0.01
0.00
0 20 40 60 80 100 120 140 160
correlation length q in meter
Indicator correlation length analyses: The J ACKKNIFE method is used in a similar way
for the analysis of the indicator correlation lengths. The thresholds are the percentile
values of the measurement data. MoMvar , MoMLA and ML methods offer comparable
results in figure 3.17. The indicator correlation lengths are nearly the same for all thresh-
olds; comparing theses results with the ones of case study A, one can clearly see that
the extreme high and low values show indicator correlation lengths that are more or less
the same as for the median value, as shown for EU,3 in figure 3.17. Although there is
a scattering, the distribution of the indicator correlation length at different thresholds
does not show the same distribution as in the analytical case shown in figure 3.4. This
can be attributed to a non Multi-Gaussianity of the measurement data, which cannot be
described by knowing solely the mean valued, the standard deviation and the spatial
correlation of measurements.
Table 3.3: Correlation lengths using the J ACKKNIFE approach for the MoMvar approach
in comparison to the mean value and the standard error of the ML approach.
MoMvar ML BMA
µ ± σ [MN/m2 ] µ ± σ [MN/m ] µ ± σ [MN/m2 ]
2
55
Chapter 3 Case studies on the evaluation of spatial variability
3.3.3 Remarks
A main concern in this measurement setting has been the different sources of error,
which are summarized in chapter 2.3. To keep the measurement error as small and
constant as possible, the same people carried out all the tests with the same equipment.
The knowledge error and description error were kept minimal by interviewing different
experts to judge the geological and geotechnical circumstances in order to identify the
homogeneous layer as shown in figure 3.13. After this, the spatial variability inside a
homogeneous layer was assumed to be the main source of uncertainty. But one has to be
aware that the mutual distance between the samples itself is also a source of error.
In tables 3.2 and 3.3 the results of the correlation length are presented; the results for
each loading and reloading cycle show an almost similar spatial correlation within the
MoMvar , MoMLA and ML approaches. The MoMvar , MoMLA and ML approaches show
different results. The differences between the three approaches can be related to the dif-
ferent evaluation schemes of the spatial correlation. By looking at the results, one can
conclude that the results of the MoMLA approach are not reliable. If one compares the
results of the MoMLA approach to the sampling scheme of the test results every 2 m, the
results of the MoMLA approach are not reliable because there are just 2 measurements
within the correlation length. Therefore, MoMLA is not considered for J ACKKNIFE ap-
proach.
The J ACKKNIFE approach is employed to evaluate the uncertainty of the spatial cor-
relation, which is new to the reader’s knowledge. Moreover, this offers the possibility
to combine via the Bayesian Model Averaging Scheme the results of the MoMvar and
the ML approach. Different assumptions and different approaches are merged in the
Bayesian Model Averaging scheme, which enables a more precise characterization of the
spatial correlation length with a lower COV = 49% in figure 3.18.
When looking at the results of the indicator correlation length analysis, one can clearly
see that the normalised correlation lengths is nearly the same for all investigated thresh-
olds. The normalised correlation lengths differ significantly from the analytically de-
fined case in case study A (section 3.1) From these new findings, one can conclude that
the simple methods for the simulation of spatial variability (by using a mean value, a
standard deviation and only one single covariance function) do not fully acknowledge
the measured spatial correlation structure.
56
3.4 Case study D – Sheikh Zayed road in Dubai
D
B
C
[email protected]
Figure 3.19: Case study D: Plan view of the four sites at the Sheikh Zayed road in Dubai.
57
Chapter 3 Case studies on the evaluation of spatial variability
probability density
0.30 fitted pdf
10
depth in m
15
0.20
20
25
0.10
30 measurements
35 mean value
trend 0.00
0 2 4 6 8
UCS in MN/m²
3.4.3 Remarks
The data, which have been evaluated in this section are non equally spaced, as shown in
figure 3.20. Therefore, the evaluation of the variogram has to follow equation 2.31 and
figure 2.6.
The strength of the MoMvar approach can be clearly seen in figure 3.21. It has by far
the lowest variations of the resulting correlation length, which is not the case for the
ML and MoMLA approaches. This can related to the subdivision of the non-equidistant
measurements into distance classes. If there are outliers in these distance classes, they
can influence the correlation length significantly. The problem of outliers is also cap-
tured by the J ACKKNIFE approach as a side effect; via this algorithm the influence of
these outliers on the correlation length is evaluated through the variance of the correla-
tion length as shown in figure 3.21. This challenge of non-equidistant data is not really
present for the ML approach because all data and their mutual correlation are captured
trough the correlation matrix. The ML approach is also just slightly influenced by the
q-q transformation.
In comparison to the variogram approach, the ML and MoMLA approaches show a
high COV of the correlation length. This can be related to the q-q transformation and
to the relative high skewness of the distribution, which is influencing the reliability of
these methods, as shown in case study A.
When looking at the results of the indicator correlation length analysis in figure 3.22,
[email protected]
one has to keep in mind the high skewness and asymmetry introduced by the indica-
58
3.4 Case study D – Sheikh Zayed road in Dubai
Figure 3.21: Case study D: Combination of the MoMvar , the MoMLA and the ML approach
using BMA.
cumulative distribution function
80 %
MoMvar
percentile of the
60 % MoMLA
ML
40 %
20 %
0%
0.95 1.0 1.05 1.10
normalised correlation length qind / q
Figure 3.22: Case study D: Analysis of the indicator correlation lengths using using the
MoMvar , the MoMLA and the ML approach.
59
Chapter 3 Case studies on the evaluation of spatial variability
tor approach. As in case studies B and C, the results are different from the findings of
the generated random process following a multi-variate Gaussian distribution in case
study A. The lower extreme values show a significantly higher correlation length than
the lower extreme values and the median value. Similar findings for permabaility mea-
surement data are presented in literature [144]. In this context, the indicator correlation
lengths are more or less the same for all thresholds [144].
60
3.5 Literature database on the spatial correlation of soil properties
61
Chapter 3 Case studies on the evaluation of spatial variability
Table 3.4: Properties of the lognormally distributed horizontal and vertical correlation
lengths of CPT data for frictional and cohesive soils.
into correlation lengths of different soil. The presented database is used to derived upper
and lower limits of vertical and horizontal correlation lengths of different soil types. This
can be used to set up efficient sampling schemes because the spatial correlation length
is essential to investigate the micro-, meso- or macro scale soil variability. Moreover, the
presented database links soil types and spatial variability presented in literature.
62
3.6 Evaluation of the vertical correlation length using CPT databases of different soil types
1000 1000
7 8 sv0
qt
7
no u
rm 9
all
100 100 6
6
yc
on
Qt Qt 5
sol
ida
5 4
dte
10 4 10 3
1 3
1
2
1 1
0.1 1 10 -0.4 0.0 0.4 0.8 1.2
Fr [%] Bq
q t -σ v0 u z -u 0 fs
Qt = Bq = Ft = × 100%
σ'v0 q t -σ v0 q t -σ v0
Figure 3.23: Soil classification into 9 soil behaviour types according to Robertson [308]:
zone 1: sensitive fine grained
zone 2: organic soil-peat
zone 3: clay-silty clay
zone 4: clayey silt - silty clay
zone 5: silty sand - sandy silt
zone 6: clean sand to silty sand
zone 7: gravelly sand to sand
zone 8: very stiff sand to clayey sand and
zone 9: very stiff to fine grained soil.
The vertical correlation lengths for the soil types are statistically analysed and the
mean values and coefficients of variation of the fitted lognormal distribution functions
are summarized in table 3.6. This table also contains the results of the BMA scheme,
which were used to combine the results of the MoMvar , the MoMLA and the ML approach.
The above cited databases do not cover the investigated CPT measurements of the soil
type 7 "gravelly sand to sand", soil type 8 "very stiff sands to clayey sand" and soil type
9 "very stiff, fine grained sand".
One can clearly see that the difference between "sensitive fine grained soils" and "clean
sands to silty sands" is not that big. For the MoMvar , the MoMLA and the ML approach,
the mean values of the vertical correlation length ranges between θvert = 0.4 m and
[email protected]
θvert = 0.8 m. Also the COV values do not show a big scatter. Maybe this can be linked
63
Chapter 3 Case studies on the evaluation of spatial variability
Table 3.5: Summary of the different CPT databases [50, 129, 275, 380].
Table 3.6: Mean value and COV of the vertical correlation lengths of the different CPT
databases and the BMA combination results.
to the genesis of the tested soils because all of the tested soils can belong to sediment
soils. It is very difficult to compare the presented results to other works because no
comparable study is available to the author’s knowledge.
The presented results are an extension of the literature database in section 3.5 using
measurement data. On basis of CPT measurements, the vertical correlation lengths of
various soil types are evaluated. These results offer a detailed insight into the spatial
correlation of CPT measurements, which have been performed in 6 different soil types.
These results are contributing to a general description of stochastic soil properties of 6
different soil types.
64
3.7 Combining expert knowledge and measurements
0.5
0.0
0 0.5 1 1.5 2
vertical correlation length q in meter
Figure 3.24: Combination of measurement data of silty clay, results of the literature
database and results of the CPT databases using Bayesian Model Averaging.
65
Chapter 3 Case studies on the evaluation of spatial variability
Case study A: The simplest boundary conditions are used in CASE STUDY A; artificially
generated, homogeneous, stationary data with a known correlation length θtarget and
an added error are analysed by the MoMvar , the MoMLA and the Maximum Likelihood
approaches. The data are quasi continuously defined in order to exclude a sampling
induced errors. It is shown that all three approaches offer nearly the same results under
ideal conditions of a symmetric probability density function.
In addition to this, the indicator approach is employed to analyse the spatial corre-
lation of extreme low and high values. The indicator correlation length of the extreme
low and high values of the random process show a significantly lower correlation than
the median values. Via this, the indicator correlation length of a theoretical multivariate-
Gaussian random process is analysed and shown, which help to understand the results
of the case studies in the latter.
Case study B: A scheme for stochastic site characterization is used to describe the soil
tested by 138 vertical CPT measurements in CASE STUDY B. This procedure combines
engineering judgement and these statistical techniques to identify soil layering. More-
over, the spatial variability of each soil layer is identified and statistically described by a
lognormal distribution function. The presented scheme enables the engineer to separate
different scales of variability, e.g. the geological macro-scale and and smaller geotechni-
cal meso-scale inside each layer. By separating the macro- and the meso-scale, the evalu-
ation of the correlation lengths becomes more easy because measurement data from the
observed layers do contain less erroneous parts of e.g. large scale fluctuations.
The results of the the MoMvar , the MoMLA and the ML approaches offer independent
interpretation of the measurement data. Due to the big number of CPT measurements
it is possible to identify a lognormally distributed vertical correlation length, which is a
novel insight into this context.
The results of the MoMvar , the MoMLA and the ML approaches are combined using
the B AYESIAN M ODEL AVERAGING scheme, which combines the information of differ-
ent models using the B AYES theorem. This procedure results in a probability density
function of the correlation length, which contains different models and assumptions in
one results. The resulting probability density function of the correlation length is a new
insight into spatial variability.
The indicator correlation length analyses clearly show that the measurement data have
a particular correlation of the extreme low and high values. This implies that these spa-
tial correlations cannot be modelled by conventional simulation approaches using one
[email protected]
single correlation function.
66
3.8 Summary and consequences of the case studies
Case study C: Equi-distant in-situ measurement of stiffnesses are analysed within this
case study. In order to analyse the uncertainty of the correlation length, the well known
the MoMvar and the MoMLA approach approach is extended by using the J ACKKNIFE
procedure. Via this novel approach, a comparison with the uncertainty measures of the
ML approach is possible.
Furthermore, the results of the the MoMvar , the MoMLA and the ML approaches can be
combined with the introduced Bayesian Model Averaging scheme.
In addition to this, the indicator correlation length analyses indicate a spatial structure
of the measurement data, which cannot be fully described by standard approaches using
one single correlation length used in case study A.
Case study D: In case study D the non-equidistant data are investigated. These mea-
surement results of the soil strength follow a skewed, lognormal distribution. Apart
from the challenges of analysing these data, the B AYESIAN M ODEL AVERAGING ap-
proach offers an elegant tool to combine the results of the the MoMvar , the MoMLA and
the ML methods.
The results of the presented case studies show that the presented methodologies can
evaluate the correlation length of equally and irregularly spaced measurement data. The
results in the case studies show different ratios of the nugget effects to sill values. This
implies that different scales of variability are involved. But one has to keep in mind that
the the MoMvar , the MoMLA and the ML approaches can only identify the spatial corre-
lation of one scale. Therefore, the sampling concept of the measurement data is essential
and has a major impact on the results. One has to know in advance the scale in order to
measure it correctly. But there are no satisfying solutions offered in literature, as pointed
out in section 2.5.4. The big drawback of the presented design-based sampling concepts
of geostatistics is the high sampling effort, which implies high costs. Model based design
approaches like nested-sampling design [403] or adaptive sampling approaches offer a
promising alternative; but this is suffering from the assumption of normally distributed
variables.
Due to the multi-scale nature of spatial soil variability no general answer for the min-
imum sampling effort to describe spatial variability can be given. Barnes [32] states that
the upper bound of required sampling for spatially dependent samples has to be lower
than for identically, independently distributed samples. Different authors focused on
this but only some of them like Journel & Huijbregts [190] or Lark [214] offer rules of
thumb for a one dimensional spatial correlation. Here, the number of pairs should be at
least 30 to 50 pairs. In the field of ecology and environmental engineering, Marchant &
Lark [236] recommend to have at least 100 to 150 samples for a two dimensional spatial
analysis, whereas Webster & Oliver [402] recommend at least 200 samples. Other publi-
cations [90, 197–200, 270, 271] refer to a minimum sampling size of 100 − 150 samples in
a two dimensional analysis of the spatial correlation using the ML approach. DeGroot &
Baecher [90] provide in figure 2.8 a relationship between the correlation length and the
sample size, which also gives comparable suggestions for a minimum required sample
size.
[email protected]
If the number of pairs is lower, the uncertainty of the correlation length becomes very
67
Chapter 3 Case studies on the evaluation of spatial variability
dominant and complicated for the ML approach as well for the presented J ACKKNIFE
method.
68
Chapter 4
Safety and uncertainties
Within this chapter the basics of safety and reliability are highlighted. It provides a
description of common approaches for dealing with uncertainties and safety in geotech-
nical engineering. Global and partial safety factors, as well as the basics of uncertainty
quantification and reliability based design, are described and compared with a special
focus on the basics of the generation of random numbers and random fields, as well as
on the computation of failure probabilities. Moreover, an introduction to the local and
global sensitivity analysis of systems is presented.
• Safety is a state in which no disturbance of the mind exists, based on the assump-
tion that no disasters or accidents are impending.
• Safety is a feeling based on the experience that one is not exposed to certain hazards
or dangers.
Safety requirements and safety concepts have a long history in some technical fields
[295]. Nearly 4,000 years ago, this can be seen in the code of H AMMURABI, in which
strong penalties in the case of collapsing structures were fixed. Probably the first ap-
plication of a global safety factor in structural engineering dates back to P HILO from
Byzantium [338] in 300 B.C., who introduced the global safety factor η in terms of:
[email protected] resistance
η= (4.1)
load
69
Chapter 4 Safety and uncertainties
Table 4.2: Factors of global safety in geotechnical engineering after Terzaghi & Peck [373].
Only in the last few centuries have the application of global safety factors become widespread.
Over time many different values were developed for different materials [295]. In most
cases, the values dropped significantly during the last century. Proske et al. [297] report
that in 1880 in brick masonry, a factor of η = 10 was required, whereas 10 years later, a
factor between η = 7 − 8 was required. In the 20th century, the values have changed from
a factor of η = 5, then to η = 4, and now for the recalculation of historical structures a
factor of η = 3 is chosen [297].
Especially with the development of new materials, an increase in concern over the
safe application of these materials has arisen. At the beginning of the 20th century the
development of safety factors for different materials led to initial efforts in developing
material-independent factors, such as those shown in table 4.1. Visodic [397] shows
that the knowledge of load, material and environment have an influence on the global
safety factor depending on the state of knowledge: if there is more knowledge on the
load, material and environment, the factor of global safety can be reduced to a certain
limit [297]. A similar tendency can also be observed for other materials like steel. In
geotechnical engineering, Meyerhof [251] reports that the factor of global safety for the
stability of a retaining wall remained the same, since being introduced in the 18th century
by Belidor and Coulomb. The global factors of safety for different geotechnical problems
are summarized in table 4.2. These values by Terzaghi & Peck [373] do not take into
[email protected]
account the variability of the soil properties or additional knowledge on the soil.
70
4.1 Safety in engineering
E d ≤ Rd (4.2)
Subsequently, the load event can be built from several single elements, such as the char-
acteristic dead load Gk,j connected with a special safety factor γG,i only for dead load,
and the characteristic life load Qk,j connected with a special safety factor γQ,i and ψ0,i for
the combination of different life loads:
X X
Ed = γG,i Gk,i + γQ,i Qk,i γQ,i ψ0,i Qk,i ≤ Rd (4.3)
According to Proske [297], although the partial safety factor concept was first introduced
in structural engineering after World War II, it took quite some time to become practi-
cally applicable. In geotechnical engineering Talyor [372] was one of the first amongst
others to introduce separate factors of safety on the components of shear strength in
slope stability estimation. It is reported in [251] that Brinch-Hansen generalized this
approach and initiated partial safety factors in geotechnical engineering. In civil engi-
neering, these partial factors were chosen to give about the same design estimate as con-
ventional total factors of safety as stated by Schuppener [325, 330]. These factors have
been refined subsequently by semi-probabilistic methods on the basis of the variability
of the loads, soil strength parameters and other design data in practice [251]. Therefore,
the development of partial safety factors is strongly connected to the development of the
probabilistic safety concept in structural engineering [297].
The calibration of partial safety factors is described in E UROCODE [71] as shown in
figure 4.1. The deterministic as well as the probabilistic approach can be used for cal-
ibration of the partial safety factors. The deterministic approach includes historical as
well as empirical methods in civil engineering, which have shown their strengths over
several years or even decades [297], for the calibration of partial safety factors. The
probabilistic approach can use reliability methods and fully probabilistic methods (e.g.
Monte Carlo approach) for the calibration of the partial safety factors. The probabilistic
methods offer the basis for the calibration of the partial safety factors, which are used
in the semi-probabilistic approach. In the semi-probabilistic approach, partial safety fac-
tors are used to consider the uncertainties in load, resistances and model uncertainties,
[297].
The first proposals about probabilistic-based safety concepts were found in the 1930s
in Germany and in the Soviet Union. The development of probabilistic safety concepts, in
general, experienced a strong impulse during and after World War II, not only in the
[email protected]
field of structures but also in the field of aeronautics. The Joint Committee of Structural
71
Chapter 4 Safety and uncertainties
semi-probabilistic
methods
method C
design using
method A partial safety factors method B
Figure 4.1: Calibration of the partial safety factors according to E UROCODE [71].
Safety (JCSS) introduced the probabilistic model code [121] on the basis of the proba-
bilistic safety concept of structures. The probabilistic model code includes a detailed
introduction to model the load and resistance parameters as random variables in order
to calculate the probability of failure of structures.
In E UROCODE [71] the safety of structures is defined as the capability of structures to
resist loads. Due to the fact that no building can resist all theoretically possible loads, the
resistance has to reach only a sufficient level [71]. Only by using a quantitative measure
can one offer a basis on decision on whether a structure is reliably or not. The reliability
is interpreted as a probability of failure not occurring, which is explained in detail later
in this chapter.
The design process including reliability is called reliability based design (RBD) [286].
Herein, the uncertain components of a system are simulated as random variables within
the reliability framework to evaluate the probability of failure. The E UROCODE, as well
other design codes [121, 127], probabilistic methods are used with the reliability based
design concept calibration of partial safety factors.
Proske [297] offers in figure 4.2 an overview of most of the safety concepts in struc-
tural engineering. Starting from basic empirical rules, Proske [297] adds concepts with
increasing complexity: As pointed out above, the global and semi-probabilistic safety
concepts are less complex than simplified and exact probability safety concepts. As a
next step, the reliability index and the probability of failure of a system are compared to
the a target reliability and probability of failure. Probabilistic methods like First Order
Reliability Method or Monte Carlo Method can be used for this, as described in the E U -
ROCODE [71]. Relatively new concepts like fuzzy-probabilistic procedures [252] and risk
based design concepts are also included in figure 4.2. One can deduce form this figure
that different safety concepts can cover different aspects. Proske [297] states that only
[email protected]
complex safety concepts fulfil the basic requirements indicated by basic human rights
72
4.1 Safety in engineering
R=C ·P (4.4)
73
74
human right “safety”
[email protected]
risk-based safety concept
Chapter 4 Safety and uncertainties
semi-probability concept
Figure 4.2: Safety concepts in the context of structural engineering from Proske [295].
4.1 Safety in engineering
where R is the risk, p the probability of this event, A0 the value of an object, ν0 the
vulnerability of the object during the event and p0 the probability of exposure of the
object during the event, e.g. flood or landslide amongst other natural hazards [133].
Equation 4.5 can be further extended to several event scenarios and objects. However,
the term vulnerability remains to be defined. Again, as with many other terms, the
variety of definitions of vulnerability is virtually unmanageable according to [133, 295].
Deeper discussion and examples can be found amongst others in [133, 295].
At the basis of risk based design, Baecher & Christian [23] summarized the annual failure
probability and expected losses for a variety of common civil facilities and other large
structures or projects in figure 4.3. Baecher & Christian [23] and Meyerhof [251] are the
only two sources of FN-curves in geotechnical engineering according to the knowledge
in geotechnical engineering. In figure 4.4 Meyerhof [251] shows the global safety factor
in comparison to the lifetime probability of stability failure using different coefficients of
variability (COV). On top of this, observed and theoretical probabilities of failure in soil
engineering is compared to lifetime fatality risk per person. From figures 4.3 and 4.4 one
can derive that there are different reliability levels for different structures.
Table 4.3 shows the classification of target reliability levels provided in the E UROCODE
[71]. Reliability indices are given for two reference periods T ( 1 and 50 years) but with-
out any explicit link to the design working life Td . The reliability index β is a scalar
measure of safety equivalent to the probability of failure pf , but measured on another
scale. A detailed description of the reliability index is given in section 4.3.1.
The values are based on calibration and optimization and reflect results from several
studies. It is noted that similar β values as in table 4.3 are given in other national and
international guidelines. Examples of buildings and civil engineering works for RC 3
are bridges and public buildings, for RC 2 residential and office buildings and for RC 1
agricultural buildings and greenhouses.
In addition to this, reliability indices with respect to consequences and to relative cost
of safety measures are presented in JCSS [121] and ISO 2394 [183] offers more detail. The
bigger the reliability index β becomes, the more unreliable is the occurrence of failure,
which will be covered in section 4.3.1.
Calgaro [66] states that different criteria may be taken into account when choosing a
target reliability index. He emphasis that the combination of economic, risk acceptance,
psychological and legal criteria have to be taken into account.
The Life Quality Index (LQI) is a recently developed concept for determining accept-
[email protected]
ability of decisions involving life safety risks in engineering. It provides a rational basis
75
Chapter 4 Safety and uncertainties
0
10
marginally accepted
accepted
annual probability of failure
-1 mine
10 pit
slopes merchant shipping
geisers
10
-2 mobile
MOBIL
drill rigs
DRILL RIGS
foundations super
-3
fixed tankers
10 FIXED
drill rigs
canvey refi
neries
-4
10 super
dams tankers
-5 estim
10 ated
US
commercial d am
-6 aviation s
10
-2 -1 0 1 2 3 4
10 10 10 10 10 10 10 lives lost
Figure 4.3: Chart showing average annual risks posed by a variety of traditional civil fa-
cilities and other large structures or projects proposed by Baecher & Christian
[23].
lifetime fatality
global factor of safety risk per person
1.0 1.5 2.0 2.5 3.0
10
0
human life
lifetime probability of failure
natural disasters
high
COV mining
theor
y = 30 %
10
-2
FO motor vehicles
medium
CO
E CO ships, fires
V
Vt
R the
= 20 railways
he
-4 o ry
or
10 %
y
aircraft
=1
low
0%
gas piplines
-6
nuclear reactors
10
E FO , R FL
Figure 4.4: Lifetime probabilities of stabilitiy failures and comparative human risks from
Meyerhof [251].
[email protected]
76
4.2 Basics of uncertainty quantification
Table 4.4: Reliability based design: different levels of accuracy from Honjo et al. [168].
for establishing target reliabilities for civil engineering systems [218, 299, 361]. The LQI
is a socio-economic utility function that depends on the wealth and life expectancy of a
society. Any decision that increases the value of the LQI is deemed acceptable. This in-
crease can be due to an increase in life expectancy (reduction of fatalities) or an increase
in societal wealth (reduced use of resources). In this way the LQI establishes a relation
between the resources invested in improving the safety of an engineering facility and
potential fatalities and injuries that are avoided by the investment. Hence it provides a
means to quantify the optimal trade-off between safety and cost [299, 361].
77
Chapter 4 Safety and uncertainties
this approach, it is possible to evaluate the probability of failure more precisely because
more information is available in comparison to the other levels of RBD and uncertainty
quantification.
78
4.2 Basics of uncertainty quantification
The set of points {x : g(x) = 0} defines the limit state surface [364, 366]. Denoting by
fX (x) the joint probability density function of a random vector X, the probability of
failure pf of the system is:
Z
pf = Prob [g (X1 , X2 , . . . , Xn ) ≤ 0] = fX (x)dx (4.8)
Df
Sudret [366] emphasis that in all but academic cases, this integral cannot be computed
analytically. Indeed, the failure domain depends on response quantities (e.g. displace-
ments, strains, stresses, etc.), which are usually computed by simplified equations [44,
57, 176, 286] or by means of numerical methods e.g. the finite element method [57, 176,
364, 366]. In other words, the failure domain is implicitly defined as a function of X.
Thus, numerical methods have to be employed to evaluate the probability of a failure.
Here, L is the lower triangular matrix, which results from a Cholesky decomposi-
[email protected]
tion of the matrix of correlation coefficients ρ.
79
Chapter 4 Safety and uncertainties
Conditional random fields: All the above mentioned approaches simulate uncondi-
tional random fields, which are not taking data points into account from e.g. measure-
ments, pre-knowledge, etc. Therefore, unconditional random fields are not spatially
consistent in the presence of data points. By studying the explanation of the SGSIM
in appendix E, one can derive the scheme of the direct conditional simulation within this
sequential scheme. In figures E.4 and E.5 in appendix E an overview of the main charac-
teristics of various random field simulators are shown.
In addition to this, there are also other approaches like the Kriging approach or simu-
lated annealing, which can be used for conditioning random fields [79, 97].
Li & Der Kiureghian [219] state that many applications in civil engineering call for a
combination of continuum mechanics and representation of uncertain media as random
fields. Therefore, it is necessary to map a random field onto a grid of a Finite Element
mesh. Several methods for discretisation of random fields have been proposed in the past
amongst others by [219, 245, 364, 366] in the framework of stochastic Finite Elements (see
section 4.3.3). These include the midpoint method (MP), the spatial averaging method
(SA), weighted integral method, the shape function method (SF) and the series expan-
sion method (SE), the Karhunen Loeve expansion (KL), the orthogonal series expansion
and the expansion optimal linear exstimation method (EOLE).
The simplest method of discretisation of a random field within the domain Ω is the
midpoint method. In this method, the field within the domain Ωe of an element is de-
scribed by a single random variable representing the value of the field at a central point
[email protected]
of the element, e.g. the centroid xc , as shown in figure 4.6. The field value within the
80
4.3 Computation of failure probabilities
The average values vbe now form the vector v. A realization of the field so defined is also
a stepwise function with discontinuities along the element boundaries. Li & Kiureghian
[219] state that the variance of the spatial average variable over an element is smaller
than the local variance of the random field, in general. Moreover, this method can map
a random field on structured as well as on non-structured finite element meshes.
The other above mentioned discretisation methods offer a continuous function of the
discretized field, but ask for a more complicated mathematical background as explained
in detail in literature [219, 245, 364, 366].
xi
We
xc
x
[email protected]
Figure 4.6: Random field discretisation after Li & Kiureghian [219].
81
Chapter 4 Safety and uncertainties
278], grey number theory [296], imprecise probability method based on p-box repre-
sentation [128, 415] and random set approaches [279, 290, 332], which are summarized
briefly in [262]. The probabilistic approaches aim for a computation of the probability of
failure, which is faster than the computationally time consuming Monte Carlo (MC) sam-
pling approach. Here it should be realised that each alternative to the robust MC method
implies some loss of accuracy. Therefore, the MC approach is used for verification and
calibration of these approaches. The Bayesian approach in uncertainty quantification is
described in various publications [23, 416] as well as the standard reliability methods
(e.g. FOSM, FORM, SORM) in [57, 326] and iterative random point sampling methods
in [57, 326, 327]. Prefixed point sampling methods like Taylor series, finite difference
methods or the Point Estimate method can be found in recent publications [13, 374].
Fenton has worked in various publications [127] and different applications in geotech-
nical engineering on the simulation of spatial variability using random fields within the
non-deterministic approaches
interval analysis
Bayesian approach Standard reliability
methods
Random Finite fuzzy apporaches
Element Method FOSM
Iterative random grey number
Stochastic Finite FORM point sampling theory
Element Method
SORM directional samling,
Hasover-Lind method; imprecise
response surface methods adaptive sampling, probability
MCMC
random set
interpolation regression prefixed point approach
methods methods
sampling methods
Kriging
radial basis Taylor series finite
polynomial chaos
funciton difference method
expansion method
82
4.3 Computation of failure probabilities
Gaussian space
xj’
xc’
pdf
aj’
ac’
xj’ mj’
g>0
tan
(c’*|j’*)
ge
g>0 g<0
nt
g<0
b g<0
xc’ mc’
g<0 cohesion c’ [kN/m²]
Figure 4.8: First order reliability method in Honjo [168] and Sudret [364].
In the case of correlated variables Y the transformation from correlated Gaussian space
to standard Gaussian space can be done by means of
Ξ = L−1 Y (4.14)
CZZ = L LT (4.15)
Then a linearisation of the limit state function is performed in Gaussian space (ξc′ , ξϕ′ ).
The expansion point ξ ∗ is chosen so as to maximize the pdf within the failure domain Df .
Geometrically, this coincides with the point in the failure domain, having the minimum
[email protected]
distance β from the origin. From a safety engineering point of view, the point x∗ =
83
Chapter 4 Safety and uncertainties
[c′∗ , ϕ′∗ ] corresponding to ξ ∗ = [ξc′∗ , ξϕ′∗ ] is called the design point or most probable point
(MPP).
From the geometrical interpretation of the expansion point ξ ∗ in standard Gaussian
space it becomes quite clear that the calculation of the design point can be reduced to an
optimization problem.
∗ 1 T
ξ = argmin ξ ξ ; subjected to: g [z(ξ)] = 0 (4.16)
2
1
L = ξ T ξ + λ g(ξ) → Min. (4.17)
2
Standard optimization procedures can be utilized to solve for the location of ξ like
e.g. Rackwitz-Fiessler algorithm, particle swarm algorithm amongst other methods de-
scribed in e.g. [57].
The equation of this hyperplane may be cast as:
β−αξ =0 (4.18)
where the unit vector α = ξ/β is also normal to the limit state surface at the design point
ξ∗ :
∇g ( T −1 (ξ ∗ ) )
α=− (4.19)
k∇g ( T −1 (ξ ∗ ) )k
Herein, T −1 is the q-q transformation from the Gaussian space into the physical space
shown in figure 4.8. The vector α describes the contributions of the random variables ξi
to the probability of failure pf .
A linear approximation of the failure surface at the design point will be accurate if
the failure function is linear or weakly non-linear (relatively flat). For heavily non-linear
failure functions, the FORM methods may not always be adequate to find a reasonably
correct failure probability. In such cases, a better approximation of the failure surface at
the design point is required. For this purpose, a second order (parabolic) failure surface
is fitted to the non-linear failure function at the design point [53, 93, 131, 376] to give
the Second Order Reliability Methods (SORM). It is a relatively complicated process and
computationally more time consuming as well. A detailed description of this method
can be found in Bucher [57] amongst others [53, 131, 191, 286, 366].
84
4.3 Computation of failure probabilities
Importance sampling: Bucher [57] states that in order to reduce the standard deviation
σp2f of the estimator to the order of magnitude of the probability of failure itself m must
be in the range of m = 1/pf . For values of pf in the range of 10−6 this cannot be achieved
if each evaluation of the limit state function requires a complex analysis. Alternatively,
strategies are employed, which increase the "hit-rate" by artificially producing more sam-
ples in the failure domain than should occur according to the distribution functions. One
way to approach this solution is the introduction of a positive weighting function hY (z)
which can be interpreted as a density function of a random vector Y. Samples are taken
according to hY (z). The probability of failure is then estimated from
m
1 X fY (z) fY (x)
pf = Ig (z) = E Ig (z) (4.23)
m k=1 hY (z) hY (z)
From the estimation procedure it can be seen that the variance of the estimator pf be-
comes
2 1 fY (x)2
σp f = E Ig (x) (4.24)
m hY (x)2
A useful choice of hY should be based on minimizing σp2f . Ideally, the weighting function
should reduce the sampling error to zero. However, this cannot be achieved in reality
since such a function must have the property
(
1
f (z) for g(z) ≤ 0,
pf Z
hY (z) = (4.25)
0 for g(z) > 0
85
Chapter 4 Safety and uncertainties
obtain a general importance sampling concept. The efficiency of this concept depends
on the geometrical shape of the limit state function. In particular, limit state functions
with high curvatures or almost circular shapes cannot be covered very well.
Apart form importance sampling there are several other techniques like line sampling,
directional sampling, adaptive sampling, asymptotic sampling, line sampling or subset
simulation, which are explained in detail in several sources [57, 326, 364].
Intrusive SFEM: Within SFEM, different authors [286, 364, 366] distinguish between
the so called intrusive and the non-intrusive approach.
Within the intrusive SFEM, named after the pioneering work of Ghanem & Spanos
[138], the aim is to represent the complete response PDF in an intrinsic way. The imple-
mentation of the intrusive SFEM has to be carried out for each class of problem. Herein,
the stiffness matrix as well as the boundary conditions consist of a mean (deterministic)
part and of stochastic parts, which can be solved by using various methods such as the
"weighted integral method", the "Neumann series expansion" method or the "Taylor se-
ries expansion" method, as described in detail in [245, 286, 366]. The response of the sys-
tem (which, after proper discretisation of the problem, is a random vector of unknown
joint probability density function) is expanded onto a particular basis of the space of
random vectors of finite variance called "polynomial chaos".
There are two main variants of SFEM in the literature: i) the perturbation approach,
which is based on a Taylor series expansion of the response vector [203] and ii) the spec-
tral stochastic finite element method (SSFEM) [138], where each response quantity is
represented using a series of random Hermite polynomials. A detailed description can
be found in [336, 358, 366]. Amongst others, Sett & Jeremic [336] also applied the SSFEM
framework to highly non-linear and dynamical geomechanical problems.
Sudret & Der Kiureghian [367] state as an overall conclusion that SSFEM has limited
applicability to reliability problems involving small failure probabilities. The polynomial
[email protected]
chaos expansion provides a global fit to each response quantity, which may be good in
86
4.3 Computation of failure probabilities
the central region of the respective distribution, but poor in the tail regions. Since small
probability events are influenced by the tail regions of these probability distributions,
accurate results from SSFEM cannot be expected for such problems. This limitation is
more severe for problems involving random fields with short correlation lengths or large
coefficients of variation.
Non-intrusive SFEM: In order to apply this approach in a similar way to a wider range
of mechanical problems, the so called non-intrusive SFEM has been developed [286, 364].
Within the non-intrusive SFEM, the scalar response quantities S of the system e.g.
nodal displacements, strain or stress components are directly expanded onto the poly-
nomial chaos, which is truncated after the P term.
∞
X P
X −1
S = h(X) = Sj Ψj = Sj Ψj (4.26)
j=0 j=0
The big advantage of this approach is the fact that - basically - it is just the post-processing
of simulation results. Therefore, any FEM-code can be used to calculate the system re-
sponse in contrast to the intrusive SFEM [286, 364].
Sudret [364] proposes two methods to compute the coefficients in this expansion from
a series of deterministic finite element analyses, namely the projection method and the
regression method.
The PROJECTION METHOD is based on the orthogonality of the polynomial chaos [3,
286]. By premulitplying equation 4.26 with Ψi and taking the expectation of both mem-
bers, it becomes: "∞ #
X
E [S Ψ] ≈ E Sj Ψi Ψj (4.27)
j=0
Due to the orthogonality of the basis E [Ψi Ψj ] for any i 6= j, one can reformulate the
following equation:
E [S Ψj ]
S = 2 (4.28)
E Ψj
In equation 4.28 the denominator is known analytically, as derived in Appendix D, and
the numerator may be cast as a multi-dimensional integral:
Z
E [SΨj ] = h( X(ξ) ) Ψj (ξ) ϕm (ξ) dξ (4.29)
RM
87
Chapter 4 Safety and uncertainties
the integral [43, 244] as a weighted summation of the integrand evaluated at selected
points (the so-called integration points).
The Smolyak sparse grids are also quoted in literature [3, 286] as a promising approach
for computing the integral of equation 4.29.
The REGRESSION METHOD is another approach for computing the response expansion
coefficients. It is the regression of the exact solution S with respect to the polynomial
chaos basis Ψi (ξ). The scalar response quantity S consists of a residual ǫ (a zero mean
random variable) and unknown coefficients S. e
p−1
X
S = h(X) = Sj Ψj + ǫ (4.30)
j=0
The minimization of the variance of the residual with respect to the unknown coefficients
leads to equation 4.31 by using a set of Q regression points in the standard normal space
ξ i and their isoprobabilistic transform xi
Q
( p−1
)2
1 X X
e = Argmin
S i
h(x ) − i
Sj Ψj (ξ ) (4.31)
Q i=1 j=0
Sudret [364] solves this minimization problem in the following way: Denoting by Ψ the
matrix whose coefficients are given by Ψij = Ψj (ξ i ), i = 1, . . . , Q; j = 0, . . . , p − 1 and
by Sex the vector containing the exact response valuse computed by the model Sex =
h(xi ), i = 1, . . . , Q, the solution to equation 4.31 reads:
−1 T
S = ΨT Ψ Ψ Sex (4.32)
This approach is comparable to the so called response surface method used in many
domains of natural sciences and engineering. Within this context, the set of x1 , . . . , xQ
is the so called experimental design. In equation 4.32, ΨT · Ψ is the information matrix.
Sudret [364] shows that an efficient design can be built from the roots of the Hermite
polynomials as follows:
• If p denotes the maximal degree of the polynomials in the truncated PC expansion,
then the p + 1 roots of the Hermite polynomial of degree p + 1 (denoted by Hep+1 )
are computed, say r1 , ..., rp+1 .
• From this set, M -tuplets are built using all possible combinations of the roots: rk =
(ri1 , . . . , riM ), 1 ≤ i1 ≤ . . . ≤ iM ≤ p + 1, k = 1, . . . , (p + 1)M .
• The Q points in the experimental design ξ1 , . . . , ξQ are selected among the rj by re-
taining those which are closest to the origin of the space, i.e. those with the smallest
norm, or equivalently those leading to the largest values of the PDF ϕM (ξ j ).
To choose the size of Q of the experimental design, the following empirical rule was
proposed by Berveiller et al. [44] based on a large number of numerical experiments.
[email protected] Q = (M − 1) P (4.33)
88
4.3 Computation of failure probabilities
Herein, P is the number of unknown coefficients defined by the following equation com-
bining the PCE order M and the degree of the Hermite polynomial p.
M +p (M + p)!
P = = (4.34)
p M ! p!
P −1 P −1 P −1
1 3
1 XXX
δS ≡ 3 E (S − E[S]) = 3 E[Ψi Ψj Ψk ] Si Sj Sk (4.37)
σS σS i=1 j=1 k=1
P −1 P −1 P −1 P −1
1 4
1 XXXX
κS ≡ E (S − E[S]) = E[Ψi Ψj Ψk Ψl ] Si Sj Sk Sl (4.38)
σS4 σS3 4 i=1 j=1 k=1 l=1
where the expectation of Ψ2j is given in Appendix D. Various authors [286, 364, 366] de-
scribe that the moments of higher order are obtained in a similar manner.
Reliability analysis of the SFEM results: This metamodel can be used to construct the
response PDF of the system as well as to compute the reliability of the observed sys-
tem, which is approximated by a polynomial chaos expansion. Surprisingly, this link
between structural reliability and the non-intrusive SFEM based on PC expansions is
relatively new [43, 364, 366]. The PC expansion can be used as a meta-model within the
framework of non deterministic approaches for uncertainty quantification, which offers
the engineer a first insight into the reliability of a system. One has to keep in mind that
the non-intrusive SFEM also has limited accuracy in evaluating small probabilities of
failure due to the approximation via PC expansions.
Assessment of the polynomial chaos approximation: It has been shown in the pre-
vious section that polynomial chaos (PC) approximations of the mathematical model
can be obtained using non-intrusive techniques, namely the projection approach or the
regression approach. Both methods provide a stochastic response surface whose perfor-
[email protected]
mance has to be assessed [47, 365]. Blatman & Sudret [47] point out that, in terms of
89
Chapter 4 Safety and uncertainties
statistical learning theory (see e.g. [388] ), the discrepancy between the model response
and the metamodel is measured by means of a risk functional, for instance the commonly
used mean-square error. Such a quantity depends on the PDF of the response, which is
unknown in our context.
The generalized error is defined in equation 4.39:
h i Z 2
2
I MX,p
b = E (M(x) − Mxb,p ) = M(x) − MX,pb fx (x) dx (4.39)
Xb = x(1) , . . . , x(N ) T is the experimental design, the corresponding model evaluations
are Yb = y (1) = M x(1) , . . . , y (N ) = M x(N ) T and M b are the resulting PC approx-
X,p
imations. The notion of generalization error is a basic concept of statistical learning the-
ory as presented in [388]. Computing I[MX,p b ] requires a perfect knowledge of the model
function M, which is not the case in the context of geotechnical engineering in general
since the model is usually not analytical but numerical.
In literature [47, 388] it is proposed to compute the following empirical error or training
error in order to estimate equation 4.39.
h i 1 X
N 2
(i) (i)
IX MX,p
b = M(x ) − M b
X,p (x ) (4.40)
N i=1
where !2
N N
d 1 X (i) 1 X (i)
V ar[Y ] = M(x ) − y
N − 1 i=1 N i=1
However, the use of R2 statistics might be misleading for comparing two different re-
gression base meta models since it automatically increases with the number of P basis
polynomials; furthermore, it is highly biased since it tends to R2 = 1 as P increases.
Blatman & Sudret [47] report that R2 generally underestimates the generalization error.
Therefore, the adjusted determination coefficient Radj
2
is recommended.
h i N −1 h i
2 2
Radj MX,p
b =1− 1 − R MX,p
b (4.42)
N −P −1
The Radj
2
statistic is penalized as P increases. Baltman & Sudret [47] report that Radj
2
still
often overpredicts the true approximation accuracy.
The cross-validation technique consists of dividing the data sample into two subsam-
ples. A metamodel is built from one subsample, i.e.the training set, and its performance
is assessed by comparing its predictions to the other subset, i.e.the test
set. Let MX\i
[email protected]
be the metamodel that has been built from the experimental design X\ x(i) , e.g when
90
4.3 Computation of failure probabilities
removing the ith observation from the training set X. The predicted residual is defined
as the difference between the model evaluation at x(i) and its prediction based on MX\i
∆(i) = M x(i) − MX\i x(i) (4.43)
The generalization error is then estimated by the mean predicted residual sum of squares
(PRESS), i.e. the following empirical mean square predicted residual, [47].
h i N
1 X (i) 2
∗
IX MX,p
b = ∆ (4.44)
N i=1
h i N !2
1 X M x(i) − Mxb,p x(i)
Ix∗ MX,p
b = (4.46)
N i=1 1 − hi
The above presented metrics for assessing the accuracy of the PC approximation are also
quoted amongst other means by Field & Grigoriu [130].
91
Chapter 4 Safety and uncertainties
In most applications it is quite likely that the exact response function will not be known.
Therefore, it has to be replaced by a sufficiently versatile function, which will express
the relation between the response and the input variables satisfactorily.
Depending on the selected response surface model, support points have to be chosen
to estimate the unknown parameters of the response surface in a sufficient way. A set
of samples of the basic variables is generated for this purpose. In general, this is done
by applying predefined schemes, so called DESIGN OF EXPERIMENTS. Bucher [57] rec-
ommends that it is most helpful to setup an experimental scheme in a dimensionless
space. Within this, Bucher [57] describes saturated designs, which provide a number of
support points that just suffice to represent a certain class of response functions exactly,
and redundant designs, which provide more support points than required to define the
response surface.
92
4.4 Sensitivity analysis
• G LOBAL SENSITIVITY ANALYSIS aims for the quantification of the output uncer-
tainty due to the uncertainty in the input parameters, which are taken singly or in
combination with others, [364].
Saltelli et al. [317] group the different techniques in SA into regression-based methods
and variance-based methods. Within the regression-based methods, the standardized re-
gression coefficients, Pearson correlation coefficients, partial correlation coefficients, and
standardized partial rank correlation coefficients are used to describe the correlation be-
tween inut and output. Sudret [364] points out that in the case of general non linear non
monotonic models, these approaches fail to produce satisfactory sensitivity measures.
The variance-based methods aim at decomposing the variance of the output as a sum
of the contributions of each input variables or combinations thereof. They are sometimes
called ANOVA techniques for "ANalysis Of VAriance", [364]. The correlation ratios in
McKay [249], the Fourier amplitude sensitivity test indices [319] and the Sobol indices
[318, 347] enter this category.
For the sake of completeness, an extensive overview on additional methods of local
and global sensitivity approaches can be found in Cacuci et al. [63] including screening
methods, non-parametric methods, variance based methods and density based methods.
Herein, β is the reliability index and α is the unit vector to the design point. Sudret [364]
considers this linearised limit state function to be a margin function, which quantifies the
distance between a realization of the transformed input random vector and the failure
surface. Its variance straightforwardly reads:
M
X
Var [gFORM (ξ)] = αi2 = 1 (4.48)
i=1
Thus, the coefficients {α2 , i = 1 . . . , M }, which are also called FORM importance factors
i
[email protected]
by Ditlevsen & Madsen [105], correspond to the portion of the variance of the linearised
93
Chapter 4 Safety and uncertainties
margin, which is due to each ξi . When the input random variables X are independent,
there is a one-to-one mapping between Xi and ξi , i = 1 · · · M . Thus, αi2 is interpreted
as the importance of the i-th input parameter in the failure event, [364]. When the in-
put random variables are correlated, other measures of importance should be used as
explained in Haukaas & der Kiureghian [153].
VarXi [ E [S|Xi ] ]
δi = (4.49)
Var [S]
can be evaluated form the coefficients of the PC expansions in equation 4.26 as follows:
X
δiP C = Sα2 E [Ψα ] /σS2 (4.50)
α∈Ii
Herein, σS2 is the variance of the model response computed form the PC coefficients in
equation 4.36 and the summation set
Higher order Sobol’ indices, which correspond to interactions of the input parameters,
can also be computed using this approach as described in Sudret [365] in detail.
By virtue of the knowledge of SA, engineers can rank the input variables by the
amount of their contributions to the output, and thus take measures accordingly to im-
prove the performance of the model, which is a core task in engineering.
4.5 Synopsis
This chapter summarises the concepts of safety and uncertainty. Besides this, the basics
of uncertainty quantification are explained in detail: The mechanical system is repre-
sented via a limit state function and its variables are represented via random variables
and/or random fields. Different methods to compute the failure probability of the me-
chanical system are discussed and followed up by the description of sensitivity analyses.
Global and local sensitivity analyses quantify the importance of each input parameter
within the scheme of uncertainty quantification.
[email protected]
94
Chapter 5
5.1 Introduction
Possibilities and limitations of probabilistic methods in civil engineering have been stud-
ied by Elishakoff [116]. Within a survey Elishakoff asked engineers and scientists on the
applicability of probabilistic methods in their fields all over the world. He concluded
that the engineering community can be divided into enthusiastic supporters of proba-
bilistic approaches in engineering and sceptic opponents. The criticism of the opponents
is mainly based on the complex mathematical and conceptual theory, which is necessary
to apply probabilistic design in engineering .
On top of this, it is stated in various other publications [23, 168, 286] that the broader
geotechnical engineering community is unfamiliar with reliability methods, particularly
pertaining to computational details, practical usefulness, pitfalls, and probabilistic char-
acterization of input parameters. In addition to this, the British Health and Safety Execu-
tive authority published the results of a survey on the application of reliability methods
for offshore engineering in a technical report [410]. Many of the respondents thought
that probabilistic techniques would be a welcome addition to their methods of analysis,
but that the approach was yet to be widely applied. Besides this, it is pointed out that
industry has concerns with the relatively small numbers of qualified and experienced
companies and individuals who can carry out designs and, more importantly, audit and
verify these designs.
Therefore, it is necessary to set up case studies to broaden and disseminate reliability
applications beyond the researcher and enlighten the background of probabilistic design
approaches as proposed by Phoon [286] amongst others. The key objectives of such
case studies are education by examples, demonstration of usefulness of reliability based
design in practical applications, development of user-friendly tools and evaluation of
reliability methods as proposed by the Geotechnical Safety Network (GeoSNet) [360].
Within this chapter, typical geotechnical case studies dealing with tunnelling and foot-
ing analysis are investigated within the framework of uncertainty quantification. Start-
ing from semi-analytical limit state equations (LSE), the reader is introduced to more
[email protected]
complex LSE, which are derived from two and three dimensional FEM models.
95
Chapter 5 Selected case studies in uncertainty quantification using random properties
96
5.2 Tunnel lining
sv sv
sh sh H+R
Econcrete , d
q R
Esoil
gsoil
n
K0
sh sh
sv sv
son’s ratio of the soil, and Econcrete F is the normal stiffness and Econcrete I is the flexural
rigidity of the lining respectively. Esoil is the elasticity modulus of the soil.
1 + K0 β β 1 − K0
N = γsoil H R/ 1 + + + γsoil H R n2 cos (2θ) (5.1)
2 1+ν α 2
1 − K0 2
M = γsoil H R m2 cos (2θ) (5.2)
2
Esoil R3 Esoil R
α= β= (5.3)
Econcrete I Econcrete F
α β
1 + 12(1+ν) + 4(1+ν)
n2 = α(3−2ν) β(5−6ν) αβ
(5.4)
1 + 12(3−4ν)(1+ν) + 4(3−4ν)(1+ν) + 12(3−4ν)(1+ν) 2
β
1+ 2(1+ν)
m2 = α(3−2ν) β(5−6ν) αβ
(5.5)
2+ 6(3−4ν)(1+ν)
+ 2(3−4ν)(1+ν)
+ 6(3−4ν)(1+ν)2
The system can be described with the limit state equation 5.6. This limit state equation
combines the approaches of Erdmann [120] and Sudret [364]. Sudret [364] considers
the interaction of the internal moment and normal force (see figure 5.1) via equation
(5.6). Herein, ultimate internal forces are Nult and Mult , which are dependent on the
compressive strength of the concrete fconcrete , as described by Sudret [364].
(d · Nult )2 + Mult
2
g (N, M ) = −1 (5.6)
[email protected] (d · N (θ))2 + M (θ)2
97
Chapter 5 Selected case studies in uncertainty quantification using random properties
Parametric study 2: In order to quantify the effects of a fully random system, all vari-
ables are introduced with their random properties (table 5.1) in parametric study 2.
Herein, the coefficients of variation are taken from literature [23, 217, 286, 364, 399].
Looking at the results shown in figure 5.2, one can clearly observe that the most influ-
ential parameters have been modelled by stochastic variables: the results of both para-
metric studies are almost similar, but one cannot quantify the influence of each random
[email protected]
variable to the system behaviour.
98
5.2 Tunnel lining
(a) 0 (b)
10
probability of failure pf
14
-10
12 10
reliability index b
10 -20
10
8 parametric study 1
-30
10
6 parametric study 2
-40
4 10
20 30 40 COV in % 20 30 40 COV in %
Esoil Esoil
20 30 40 COVg’ in % 20 30 40 COVg’ in %
15 25 35 COVn in % 15 25 35 COVn in %
2 4 6 COVK0 in % 2 4 6 COVK0 in %
Figure 5.2: Influence of the variability in parametric studies 1 and 2 on the tunnel lining.
99
Chapter 5 Selected case studies in uncertainty quantification using random properties
(a) (b)
determination coefficient
15 19
x 10 x 10
empirical error
2 4 1.000
PRESS
1.5 3
0.995 det. coeff.
1 2
empirical error adj. det. coeff.
0.5 1 0.990
PRESS Q2 coefficient
0 0.985
0 1 2 3 4 1 2 3 4
expansion order M of the PCE expansion order M of the PCE
Figure 5.3: Accuracy estimation of the PCE in for the global sensitivity analysis.
0.6
0.6
0.5
0.4
0.4
0.2
0.3
0.0 0.2
-0.2 0.1
-0.4 0.0
d r EconcreteEsoil g’ n K0 H fc d r EconcreteEsoil g’ n K0 H fc
Figure 5.4: Results of the local (a) and global (b) sensitivity analysis in parametric study
2 of a tunnel lining in a soil with COVϕ’ = 20% and a COVc’ = 10%.
mination coefficient, the adjusted determination coefficient and the Q2 coefficient. One
can derive from this that these measures quantify the PCE-fitting in a similar way: The
expansion order M = 3 can be taken for the evaluation of the global sensitivity factors in
figure 5.4.
By looking at the results of the global and local sensitivity analysis shown in figure
5.4 one can see comparable results for both techniques: Esoil , γsoil , H and fconcrete can
be identified as most influential variables on the system behaviour from the results of
both sensitivity measures. One would also expect the K0 value to play a major role in
this context, but both sensitivity measures are very small. This can be deduced to the
low coefficient of variation in table 5.1. In case of a bigger COV, the sensitivity of the K0
value is expected to increase.
The differences can be deduced to the afore mentioned definitions of these sensitivity
measures. Besides this, it has to be pointed out that negative local sensitivity measures
αi indicate a resistance parameter of the investigated system and vice versa for load
[email protected]
parameters.
100
5.2 Tunnel lining
5.2.4 Conclusions
This first case study is showing exemplarily the application of the uncertainty quantifi-
cation framework in tunnelling. An analytical solution of a continuum problem in tun-
nelling is adapted to quantify the effects of soil variability as well as the performance of a
local and global sensitivity study of the input parameters. The proposed concept allows
one to investigate effects of soil variability precisely without using design approaches
described in the E UROCODE 7 [329].
The soil behaviour is modelled on basis of a linear elastic, perfectly plastic soil model.
Therefore, the results of this case study cannot be fully transferred into applied engi-
neering. But the presented studies offer a probabilistic description of a soil structure
interaction problem in tunnelling. It can be concluded that the reliability based design
methodology can be applied to any geotechnical problem, which can be described via a
limit state equation. This limit state equation can be a closed form solution or an equa-
tion, which is derived from FEM simulations.
On basis of these results, further investigations on the interaction between soil and
tunnel using numerical methods in combination with advanced constitutive models
would offer additional insights into this problem.
101
Chapter 5 Selected case studies in uncertainty quantification using random properties
qf = Nc · c′ + q0 · Nq + γsoil · b · Nb (5.7)
(
2+π for ϕ′ = 0
Nc = (5.8)
(Nq − 1)/ tan ϕ′ for ϕ′ 6= 0
π tan ϕ′ 1 + sin ϕ′
Nq = e Nb = 2 (Nq − 1) tan ϕ′ (5.9)
1 − sin ϕ′
Herein, Nc is the cohesion stability number, c′ the cohesion, Nq the depth stability num-
ber, q0 the surface load, ϕ′ is the effective friction angle, Nb the width stability number
and γsoil the soil unit weight.
The stability number Nc was derived analytically by Prandtl [293]. In 1920, Prandtl
succeeded in finding a solution for the problem of a strip load on a half plane that is
both statically and kinematically admissible. The material beneath the strip load can
be subdivided into three zones, as pictures in figure 5.5: (I) a wedge shapes zone, in
which the major principal stresses are vertical, called active Rankine zone, (II) a radial
shear zone, called Prandtl zone and (III) a passive Rankine zone, [394]. Based on Mohr’s
stress theory and using Airy’s stress function, Prandtl obtained an analytical expression
for the ultimate bearing capacity of a weightless soil.
Veruijt [394] reports that Buisman extended Prandtl’s theory by superimposing over-
[email protected]
burden pressure q0 and the unit weight of soil γsoil , which was further extended with
102
5.3 Bearing capacity of vertically loaded strip footings
q0 q q0
I III I
II II j’ c’
gsoil = 0 kN/m³
Figure 5.5: Failure mechanism of the classic bearing capacity theory from Prandtl [293].
(a) (b)
Nb / Nb, DIN 4017
200 % q
180 %
b gsoil
160 %
j’ c’
140 %
120 % Weißmann Löffler
Vollweiden
DIN 4017
100 %
TGL Pregl
80 % Steenfelt Smoltczyk
60 %
10 15 20 25 30 35 40 45 50
friction angle j’
Figure 5.6: (a) Variability of the theoretical results of the width stability number Nb from
Perau [277] and
[email protected]
(b) geometry of the rigid strip footing.
103
Chapter 5 Selected case studies in uncertainty quantification using random properties
(MC) [243], Matsuoka-Nakai (MN) [243] and Lade-Duncan (LADE)[213]. Herein, the
contact between the rigid footing and the contact with the soil is assumed as rough with
full bounding. These three constitutive failure criteria are shown in figure 5.7. It can be
seen that the MC, MN and LADE criteria are identical for triaxial compression, whereas
MC and MN are identical for triaxial extension. It is clearly shown in the description in
Appendix F that for the MN and LADE criteria no additional input variables are needed.
The results in figure 5.8 (a) show that the width stability number Nb is slightly influ-
enced by the different constitutive failure criteria; the influence of MC, MN and LADE
on the cohesion stability number Nc can be clearly seen in figure 5.8 (b). It can be con-
cluded that the MC-criteria is a conservative lower bound for the estimation of the ulti-
mate shear strength of soil, as reported by others e.g. Schad [321]. Moreover, it has to be
pointed out that the advanced criteria MN and LADE have a larger friction angle under
plane strain conditions.
The bearing capacity problem can be described with the limit state equation 5.10. This
limit state equation compares the actual footing pressure q with the bearing capacity qf .
g (c, ϕ′ ) = q − qf (5.10)
s1’
Mohr-Coulomb
Matsuoka-Nakai
Lade-Duncan
s2’ s3’
104
5.3 Bearing capacity of vertically loaded strip footings
(a) (b)
50 100 Mohr-Coloumb
Matsuoka-Nakai
40 80 Lade-Duncan
30 60
20 40
10 20
0 0
0 10 20 30 40 0 10 20 30 40
effective friction angle j’ effective friction angle j’
Figure 5.8: Stability numbers Nb (a) and Nc (b) for different constitutive failure criteria.
that structure might be exposed. This offers a more general approach for the proba-
bilistic description of a system by incorporating deterministic input parameters into a
probabilistic framework. The deterministic input in this parametric study is the footing
pressure. Furthermore, Ellingwood et al. [117] show that the fragility curves of build-
ings follow a lognormal cumulative distribution function in general. This is a benefit
because the fragility curve represented by a lognormal distribution function is continu-
ously defined over the entire range of loads instead of discrete points, [117]. Therefore,
lognormal cumulative distribution functions are fitted to the calculated failure proba-
bilities as shown in figure 5.9. Furthermore, it can be clearly seen that the larger COV
values of the cohesion and the friction angle cause a higher probability of failure.
In addition to this, the effects of different constitutive failure criteria are shown in
Table 5.2: Stochastic variables of the silty sand in two parametric studies.
105
Chapter 5 Selected case studies in uncertainty quantification using random properties
COVj’ = 5 % COVc’ = 10 %
0.0 1.0
probability of non-failure 1-pf
0.1 0.9
COVj’ = 20 %
probability of failure pf
0.2 0.8
COVc’ = 40 %
0.4 0.7
0.6 0.4
0.7 0.3 MC
0.8 0.2 MN
LADE
0.9 0.1
1.0 4
0 1 2 3 4 5 6 7 8 9 x 10
footing pressure q in kN/m²
Figure 5.9: Fragility curves for different constitutive failure criteria and different COVs
of the cohesion c′ and of the friction angel ϕ’.
figure 5.9. The MC, MN and the LADE criteria have a significant influence on the prob-
ability of failure. For all investigated levels of soil variability, the MC criterion offers a
higher probability of failure in comparison to the MN and LADE criteria. From these
fragility curves it can be concluded that the MC criterion is the most conservative one in
this context.
Parametric study 2: The concept of the fragility curves offers a good insight into the
effects of soil variability. But this concept is based on a deterministic footing pressure.
Therefore, this approach is extended in this second parametric study to a fully proba-
bilistic analysis, which considers all input variables as random. Besides this, different
levels of variability of the load q are investigated. The details of theses lognormal distri-
butions are indicated in table 5.2.
It can be clearly deduced from figure 5.10 that the uncertainty and reliability of this
system are related: the higher the degree of uncertainty is, the higher is the probability of
failure. Again, the probability of failure for different levels of variability is evaluated by
a combination of FORM and IS within the FERUM libraries [49]. In the same fashion is
the influence of the variability of the load q. A high load variability is causing an higher
[email protected]
probability of failure. This effect is more evident for low levels of soil strength properties
106
5.3 Bearing capacity of vertically loaded strip footings
(a) 0 (b)
10
probability of failure pf
14
-10
12 10
reliability index b
10 -20
10
8 parametric study 1
-30
10
6 parametric study 2
-40
4 10
20 30 40 COV in % 20 30 40 COV in %
E soil Esoil
20 30 40 COVg in %
soil
20 30 40 COVg in %
soil
15 25 35 COVn in % 15 25 35 COVn in %
2 4 6 COVK0 in % 2 4 6 COVK0 in %
Figure 5.10: Influence of the COV of the cohesion c′ , of the friction angel ϕ’ and of the
footing pressure on the probability of the failure by using the MC-criterion
in parametric study 2.
than for highly variable soils. One can conclude that the influence of the soil variability
is higher than the variability of the load. The results in figure 5.10 are evaluated for the
MC criterion.
Parametric study 2: The local and global sensitivities are shown in figure 5.12 within
parametric study 2. Both measures of sensitivity show the comparative importance of
the investigated random input variables. In the left part of figure 5.12 the results of
the local sensitivity analysis are shown. The negative local sensitivity indicates that the
width of the footing B, the unit weight of soil γsoil and the soil strength c′ and ϕ′ are resis-
tance parameters, whereas the footing pressure q is a load parameter. From the measures
[email protected]
of local and global sensitivity it can be concluded that the soil strength parameters have
107
Chapter 5 Selected case studies in uncertainty quantification using random properties
(a) (b)
0 0
local sensitivity of j’
local sensitivity of c’
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
MC -0.8
-0.8 MN
LADE
-1.0 -1
500 1500 2500 3500 500 1500 2500 3500
footing pressure footing pressure
Figure 5.11: Local sensitivities of the cohesion c′ and the friction angle ϕ′ with respect to
the deterministic footing pressure in parametric study 1.
(a) (b)
M=5
local sensitivity ai
0.8 MC
0.0
0.6 MN
-0.2
LADE
-0.4 0.4
-0.6 0.2
-0.8 0.0
B g’ j’ c’ q B g’ j’ c’ q
Figure 5.12: Local (a) and global (b) sensitivity factors of the full probabilistic analysis in
parametric study 2 for COVϕ′ = 20%, COVc′ = 40% and COVq = 20%.
the biggest importance on the probability of failure. This implies that additional infor-
mation on cohesion and friction angle have more impact than for other parameters e.g.
the unit weight of soil γsoil .
5.3.4 Conclusion
The bearing capacity problem is investigated by means of 2D FEM calculations and prob-
abilistic methods within this case study. The results clearly indicate that the choice of the
constitutive failure criterion within a linear elastic, perfectly plastic constitutive model
has a significant impact on the deterministic bearing capacity as well as on the failure
probability of strip footings. Therefore, this source of model uncertainty has to be con-
sidered within an uncertainty quantification. From a deterministic as well as from a
probabilistic point of view, the Mohr-Coulomb criterion is the most conservative choice
[email protected]
in comparison to Matsuoka-Nakai and the Lade-Duncan. This is investigated by means
108
5.3 Bearing capacity of vertically loaded strip footings
of fragility curves, which allow to visualize the influence of the different constitutive
failure criteria over a wide range of vertical footing pressures.
The load and resistance parameters of the studied problem are clearly identified by a
local sensitivity analysis and confirmed by the results of the global sensitivity analysis.
For sake of completeness, it has to be mentioned that the effects for a cross-correlation
ρc′ ,ϕ′ of the cohesion and the angle of friction are not investigated within the presented
parametric studies. This has been already been investigated in various publications e.g.
by Baecher & Christian [23] and Phoon [286] amongst others. Fenton [127] reports that
the probability of failure decreases with an increase of the negative correlation coefficient
ρc′ ,ϕ′ and vice versa.
It can be concluded that the presented findings are extending the state of the art from
a mechanical and a probabilistic point of view. The uncertainty of modelling the soil
behaviour should be investigated in further studies to derive tools to quantify this model
uncertainty in applied engineering.
109
Chapter 5 Selected case studies in uncertainty quantification using random properties
qs
H g’, c’, j’
D qt
[email protected]
Figure 5.13: Geometry of the tunnel.
110
5.4 Tunnel face stability
(a) (b)
qt0
support pressure qt
control point A A
qt
s’h = K0 s’v
qcollapse
Figure 5.14: Flow area at collapse (a) and typical pressure displacement curve (b).
ϕ′ the effective friction angle, γ ′ the unit weight of the soil and D the diameter of the
circular tunnel (figure 5.13).
1
Nc′ ,M C = Nγ ′ ,M C = 1
9 tan ϕ′
− 0.05 (5.12)
tan ϕ′
When the ratio of the overburden and the diameter of the tunnel D are bigger than H/D
= 1.5, the overburden and the load q at the surface have no influence on the failure
pressure qcollapse , nor on the stiffness, the dilatation angle or the Poisson’s ratio of the
soil, as reported by Vermeer et al. [392]. Additional 3D FEM studies on the influence of
different constitutive failure criteria are conducted by the author.
As a symmetrical tunnel is considered, the collapse-load calculations are based on only
half a circular tunnel, which is cut lengthwise along the tunnel axis. Figure 5.14 (a) shows
a typical finite element mesh as used for the calculations. The ground is represented by
10-noded tetrahedral volume elements. The boundary conditions of the finite element
calculations are as follows: the ground surface is free to displace, the side surfaces have
roller boundaries and the base is fixed.
It is assumed that the initial stresses follow a geostatic stress distribution according
to the rule σh′ = K0 · σv′ , where σh′ is the horizontal effective stress and σv′ is the vertical
effective stress; K0 is the coefficient of lateral earth pressure. Ruse [311] found that the
K0 -value influences the magnitude of the displacements, but not the pressure at failure.
The first stage of the calculations is to remove the volume elements inside the tunnel
and to activate the shell elements of the lining. This does not disturb the equilibrium
as equivalent pressures are applied on the inside of the entire tunnel. To get full equiv-
alence between the initial supporting pressure and the initial geostatic stress field, the
pressure distribution is not constant but increases with depth. This is obviously signif-
icant for very shallow tunnels, but a nearly constant pressure occurs for deep tunnels.
[email protected]
The minimum amount of pressure needed to support the tunnel is then determined by
111
Chapter 5 Selected case studies in uncertainty quantification using random properties
1
Nc,M N = Nγ,M N = 1
67 tan ϕ′
− 0.01 (5.13)
2 tan ϕ′
1
Nc,LADE = Nγ,LADE = 1
79 tan ϕ′
− 0.012 (5.14)
3 tan ϕ′
The system can be described with the limit state equation 5.15. This limit state equa-
tion compares the actual face pressure qt with the failure pressure qcollapse , as proposed
(a) (b)
soil weight stability number Ng
0.0 0.0
15 20 25 30 35 40 15 20 25 30 35 40
effective friction angle j’ effective friction angle j’
Figure 5.15: Variability of the stability numbers Nc (a) and Nγ (b) for different consti-
tutive failure criteria (Mohr-Coulomb, Matsuoka-Nakai, Lade-Duncan) in
[email protected]
comparison to experimental results from literature [196].
112
5.4 Tunnel face stability
Table 5.3: Properties of parametric study 1 on the tunnel heading using the Mohr-
Coulomb failure criterion.
113
Chapter 5 Selected case studies in uncertainty quantification using random properties
1.0
0.9
deterministic qcollapse
0.8 COVc’ = 10 % , COVj’ = 5 %
proebability of failure pf
0.7
COVc’ = 50 % , COVj’ = 25 %
COVc’ = 100 % , COVj’ = 50 %
0.6
0.5
0.4
0.3
0.2
0.1
0 qcollapse 50 100
face pressure qt [kN/m²]
Figure 5.16: Influence of the COV of the cohesion c′ and of the friction angle ϕ′ on face
pressure versus probability of failure using the Mohr-Coulomb criterion.
against the probability of failure pf . It can be seen in figure 5.16 that pf decreases with
an increasing deterministic face pressure qt and vice versa in the case of the reliability
index. It can also be concluded that, the higher the variability of ϕ′ and c′ , the higher is
the probability of failure pf . The deterministic value qcollapse shown in figure 5.16 is based
on the mean value of the cohesion and the friction angle.
The effects of positive and negative correlation between the cohesion c′ and the fric-
tion angle ϕ′ are shown in figure 5.17 (b). The ranges of ρc′ ,ϕ′ are chosen according to the
results of the literature study summarized in section 5.4.2. A positive correlation ρc′ ,ϕ′ in-
creases slightly the probability of failure. The influence of a negative correlation is more
obvious. In this parametric study it is found to be conservative to neglect a negative
correlation ρc′ ,ϕ′ , because the probability of failure is higher for ρc′ ,ϕ′ = 0.
114
5.4 Tunnel face stability
friction angel j’
25° (a)
20° qt = 40 kN/m²
15° qt = 70 kN/m²
10° qt = 100 kN/m²
0 1 2 3 4 5 qt = 130 kN/m²
cohesion c’ [kN/m²]
probability of failure pf
0 (b)
10
-50
10
-100
10
-60 -40 -20 0 20 40 60
correlation coefficient rc’,j’
Figure 5.17: Limit state surfaces for different levels of the face pressure qt in (a) and for
probability of failure versus correlation coefficient between cohesion c′ and
friction angle ϕ′ in (b) for the MC- criterion.
cess represented by the face pressure qt is evaluated in additional calculations. The re-
sults of this study are shown in figure 5.18 (a). Herein, the soil variability is increased
and compared to different levels of variability of the face pressure for a given face pres-
sure qt = 50 kN/m2 . Again, it can clearly be seen that a high level of variability of the
variables causes a high probability of failure. Moreover it can be seen that the system
behaviour of highly random properties (COVc′ > 60%, COVϕ′ > 30%) are very similar.
The influence of different constitutive failure criteria is shown in figure 5.18 (b). Herein,
the probabilities of failure differs significantly for MC, MN and LADE. Again, the MC-
criterion offers conservative results in comparison to the more realisitic MN criterion,
whereas the results of the LADE criterion are even lower.
Table 5.4: Properties of parametric study 2 on the tunnel heading using the Mohr-
Coulomb failure criterion.
115
Chapter 5 Selected case studies in uncertainty quantification using random properties
(a) 0
(b)
10
probability of failure pf
0 MC
probability of failure pf
10
-2
10
-4
COVqt = 20 % 10 MN
10
-4 COVqt = 40 %
COVqt = 60 % 10
-8
COVqt = 80 %
COVqt = 100 % -12 LADE
-5 10
10
20 30 40 50 60 70
coefficient of variation COVc’ [%] 10
-16
20 40 60 80 COVc’ [%]
10 15 20 25 30 35 10 20 30 40 50
coefficient of variation COVj’ [%] coefficient of variation COVj’ [%]
Figure 5.18: (a) Variation of the COVs for the cohesion c′ , the friction angle ϕ′ and
the tunnel face pressure qt together with the corresponding probabilities
of failure pf .
(b) Variation of the coefficients of variation for the cohesion c′ and the
friction angle ϕ′ for a given coefficient of variation COVqt = 10%
the tunnel face pressure qt , together with the corresponding probabilities
of failure pf .
(a) (b)
global sensitivity diPC
0.2
0.8
0.0
local sensitivity ai
-0.2 0.6
-0.4 MC βMC = 5.80
0.4
-0.6 MN βMN = 10.06
-0.8 LADE βLADE = 21.90 0.2
-1.0 0
c’ j’ D g’ qt c’ j’ D g’ qt
Figure 5.19: Reliability indices βi and local sensitivity αi (a) as well as the global sensitiv-
ity values δiP C values (b) for parametric study 2 on the tunnel face stability
using different constitutive failure criteria MC, MN, LADE.
116
5.4 Tunnel face stability
5.4.5 Conclusions
Within two parametric studies, the influence of soil variability, geometry and construc-
tion process on the probability of failure is studied. The limit state equation is derived
from 3D FEM studies using linear elastic, perfectly plastic constitutive models, which
are using the Mohr-Coulomb, the Matsuoka-Nakai and the Lade-Duncan failure criteria.
The assumption of uncorrelated shear strength parameters is found to be conservative
(i.e. it gives a greater probability of failure) in comparison to that of negatively correlated
parameters. The results of the parametric studies suggest that the strength parameters
of the soil and the tunnel face pressure have the major influence on the probability of
failure in comparison to the soil unit weight and tunnel diameter.
As a consequence of this, the influence of the different constitutive failure criteria is
investigated. One can clearly see that the well known Mohr-Coulomb criterion offers
very conservative results in relation to the more realistic Matsuoka-Nakai criterion and
to optimistic Lade-Duncan criterion. As pointed out in the appendix F, there are dif-
ferences in the definition of the failure surface, but the number of variables needed for
these criteria are the same. Therefore one can say that the complexity of these models
are similar.
It can be concluded from these results that the choice of the soil model and the consti-
tutive constitutive failure criteria has significant influence on the reliability of the tunnel
face. Additional investigation should focus on the model error and its considerations via
e.g. partial safety factors in applied engineering.
117
Chapter 5 Selected case studies in uncertainty quantification using random properties
Bearing capacity of vertically loaded strip footings: Within this case study on the
ultimate limit state of a footing, the soil behaviour is modelled via linear elastic, per-
fectly plastic models, which follow different constitutive failure criteria; namely Mohr-
Coulomb, Masuoka-Nakai and Lade-Duncan. These approaches are used within para-
metric 2D FEM studies, which are used to derive a semi-analytical limit state equations.
Fragility curves are used as deterministic tools to quantify deterministically the stochas-
tic system. Moreover, a local and global sensitivity analysis is conducted to investigate
the contributions of soil strength parameters and other random input variables to the
probability of failure.
Tunnel face stability: The ultimate limit state of the tunnel heading is investigated
by using the Mohr-Coulomb, the Matsuoka-Nakai and the Lade-Duncan failure criteria,
which are used within a parametric, 3D FEM study. A semi semi-analytical limit state
equation is derived from these results, which is used in the subsequent reliability analy-
sis. Similar to the results of the case study on the bearing capacity of strip footings, one
can also identify the Mohr-Coulomb criterion as conservative approach for these two
problems in comparison to the Masuoka-Nakai and the Lade-Duncan criteria.
118
Chapter 6
Selected case studies in uncertainty
quantification including spatial variability
It is stated in the E UROCODE 7 [119] that “characteristic values of soil and rock properties shall
take account of the variabilities of the property values”. Controversially, although statistical
methods are suggested as a possible way forward, there exists little guidance as to how
this should be achieved. This suggests a need to take soil variability into account within
a sound mathematical framework for geotechnical design.
The application of random fields, which represent the spatial variability of soil proper-
ties, is shown in different case studies in geotechnical engineering. The state-of-science
in uncertainty quantification is applied in the the estimation of tunnelling induced settle-
ments, slope stability and serviceability of footings. This chapter furthers the idea which
are presented in the three case studies in the last chapter.
Moreover, these applications also extend the common approaches with respect to mul-
tiple scales of soil variability within the framework of uncertainty quantification. Fi-
nally, the uncertainty of geology represented by macro-scale variability is investigated
by means of modern geostatistical simulation methods. This offers an insight into the
effects of geological uncertainty.
119
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
The ultimate rotation αultimate = 1/500 due to differential settlements is taken from the
DIN 4019 [102] to avoid cracks in masonry. This RFEM approach calculates the green-
field settlements and does not consider soil structure interaction. Moreover, this proce-
dure does not take into account the different convex and concave parts of the surface
settlements as described by Netzel [263].
120
6.1 Estimation of tunnelling induced settlements
Table 6.1: Material properties for the parametric study on tunnelling settlement in an
single-layered subsoil.
121
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
122
6.1 Estimation of tunnelling induced settlements
10 D
density function
probability
sZ sZ
X Y
pdamage mZ
Z = aultimate-a
Evaluation of the probability of damage Evaluation of the system response
100 % 600
denstiy function
Probability
function
400
50 %
200
0% 0
0 0.002 0.004 0 0.002 0.004
(b) Limit state function g Limit state function g (c)
Figure 6.1: (a) Evaluation scheme of the probability of damage pdamage due to
differential settlements,
(b) cumulative distribution function and
(c) histogram with fitted probability density function of the normally
distributed limit state function g, based on an underlying random field
(B/D = 2, H/D = 1, µ = 60 MN/m2 , COV = 50 %, θh = θv = 2D).
[email protected]
123
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
L/D
0 1 2 3
0
(a)
B / D = 1.0 qv / D = qv / D = 2
log(pdamage)
B / D = 2.0 H/D=1
-2 B / D = 3.0 COVE = 50 %
B / D = 4.0
-4
L/D
0 1 2 3
0
(b)
log(pdamage)
-2 COVE = 10 % qv / D = qv / D = 2
COVE = 50 % H/D=1
COVE = 75 % B/D=2
-4
L/D
0 1 2 3
(c ) 0
log(pdamage)
q / D = 0.5 = qv / D = qv / D
-2
q/D=1 H/D=1
q/D=2 B/D=2
-4 q/D=5 COVE = 50 %
L coordinate / diameter D
0 1 2 3
(d) 0
log(pdamage)
qh / qv = 1 qh / D= 2
-10 qh / qv = 5 H/D=1
qh / qv = 10 B/D=2
-20 qh / qv = 50 COVE = 50 %
124
6.1 Estimation of tunnelling induced settlements
maximum probability of
0
10
-15
10
0 5 10 50 100
ratio of anisotropy qh / qv
Figure 6.3: Effects of the ratio of anisotropy θh /θv on the maximum pdamage .
125
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
L = (P1+P2)/2
D D
B
B3 B1 B2 P1 P2
4D
10 D
Figure 6.4: FEM mesh and location of the boreholes (B1,B2 B3).
Table 6.2: Material properties for the parametric study on tunnelling settlement in a lay-
ered soil.
126
6.1 Estimation of tunnelling induced settlements
scaling factor b
reliability index b
2 boreholes (B1+B2) H /D = 1 qh / D = 1
q /D=2 1.5
10
3
3 boreholes (B1+B2+B3) boundary 1 borehole
qv / D = 1
2
qh / D = 1 1.0
10
10
1
0.5
0
10
0 0.5 1 1.5 2 2.5 0 1.0 2.0 3.0 4.0 5.0
L/D qboundary / D
(c)
1.0 B/D=2
scaling factor b
0.8 H/D = 1
qboundary / D= 2 D
0.6 1 borehole
0.4
0.2
0.0
1 2 10 20 50
anisotropy qh/qv
process is conditioned to one borehole to simulate the boundary between the isotropic,
homogeneous random fields of the upper and lower layers. Note that these random
fields are unconditional and not conditioned to the borehole data. The effects of in-
creasing values of θboundary are quantified via β̄. This ratio β̄ quantifies the effects of
different values of θboundary , which is normalized by the reliability index β with respect
to θboundary /D = 2 using one borehole for conditioning. It is clearly indicated in figure
6.5(b) that the boundary between both layers has a significant influence on the probabil-
ity of damage, which is shown via the reliability index β. It can be concluded that a long
correlation length of the stochastic boundary implies a larger reliability against damage
of the building.
Another issue of interest is the effects of the anisotropy of the spatial correlations in
the upper and lower layers. For this reason the spatial anisotropy θh /θv of the upper and
lower layers are simultaneously stepwise increased. The boundary correlation length
of θboundary /D = 2 is kept constant. The results in figure 6.5 (c) show that, the larger
the horizontal correlation lengths are, the bigger is the pdamage and the smaller is the
[email protected]
reliability index β. Comparing the results of the figure 6.5 (b) and (c), one can see that
127
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
the macro-scale variability of the boundary between the soil layers has more effect than
the spatial anisotropy θh /θv within the layers.
6.1.5 Conclusions
This case study investigates the effects of spatial soil variability on a complex soil-structure
interaction problem. The surface settlements are used for a simplified analysis of differ-
ential settlements of a rigid building, which are introduced by shallow tunnelling. These
surface settlements are calculated on the basis of the linear elastic, perfectly plastic Mohr-
Coulomb model and they are used to estimate the probability of damage of a building.
Within this, the interaction between subsoil and the weightless and rigid building is in-
vestigated.
Although these assumptions simplify this complex soil-structure interaction problem,
the effects of the spatial variability of the soil is quantified via the probability of dam-
age by varying geometrical properties (location of the building L, width of the building
B) and stochastic soil properties (coefficient of variation COVE , horizontal correlation
length θh , vertical correlation length θv and the ration θh /θv ). Within this, the single lay-
ered subsoil is assumed to have spatially correlated properties.
In addition to this, the effects of spatial variability of the stiffness at different scales
are investigated. This is done in a simplified way: The medium-scale spatial variability
of the stiffness is represented by random fields. The large scale spatial soil variability is
considered by introducing two different layers. Herein, a stochastic process is separating
the two layers, which are represented by random fields. At first, the stochastic process
is conditioned to one borehole. It is found that additional boreholes for conditioning
the stochastic process lower the probability of damage. Moreover, it is shown that a
large correlation length of the stochastic process also lowers the probability of damage.
Similar effects are observed for a large ratio of anisotropy θh /θv .
The presented case studies indicate the effects of spatial soil variability by using sim-
plified approaches for estimating the probability of damage. Additional investigations
on the soil-structure interaction should be carried out using advanced constitutive soil
models. These studies would contribute quantitatively to the presented results.
Above all, it has to be pointed out that spatial soil variability is three dimensional
and it cannot be fully described by two dimensional investigations. Therefore, only 3D
RFEM studies can fully evaluate the impacts of the spatial variability of soil properties
at different scales.
However, the uncertainty quantification, used in the case studies in chapter 5, is not
fully performed. Although random fields are employed to represent the spatial soil vari-
ability and to evaluate the response of the mechanical model, the sensitivity analyses are
not performed due to the absence of adequate methods presented in literature. This asks
for the development of new approaches for sensitivity analyses in order to quantify the
contribution of spatially correlated variables to the system response.
128
6.2 Slope stability
The slope stability problem is a very general geotechnical problem. Not only in urban,
but also in rural areas slope stability problems can occur because slope stability is con-
cerned with the stability of natural slopes, excavations, embankments, dams, road cuts,
mining pits or landfills, [86].
The main objectives of slope stability analysis are finding endangered areas, investiga-
tion of potential failure mechanisms, determination of the slope sensitivity to different
triggering mechanisms, designing of optimal slopes with regard to safety, reliability and
economics, designing possible remedial measures, e.g. barriers and stabilization, [355].
Due to the importance of the slope stability problem, lots of scientists have been deal-
ing with this issue from a deterministic as well from a probabilistic perspective. The
probabilistic analysis of slope stability has been in the focus of science since more than
40 years e.g. [114, 371, 383, 407, 414] and is still a topic of ongoing research [157, 158, 254,
266, 355, 391]. These investigations are focusing on the effects of uncertain soil parame-
ters including spatial variability as well as time dependent seepage forces.
This case study progresses this line of research. The influence of soil variability and
spatial variability on a single-layered soil slope is investigated in two dimensional slope
stability calculations. These calculations are compared to each other using different ap-
proaches including different correlation functions describing spatial variable soil proper-
ties and different random field generators. In addition to this, the effects of the anisotropy
of spatial correlation on the slope reliability are investigated. Also a two-layered slope is
investigated incorporating different scales of spatial variability. The sensitivity analysis
of this problem helps to understand the contributions of the different sources of uncer-
tainty and the different scales of variability to the failure probability of the systems.
0.3
1.5 H
0.4
δmax
(a) (b)
129
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
Finite Element Method for slope stability analysis: The finite element method (FEM)
is a powerful technique for slope stability analysis. Herein, the strength parameters
of the soil are reduced until the collapse of the system, which is called the strength-
reduction method. The strength reduction method is illustrated by the strength reduc-
tion factor versus maximum settlement plot in figure 6.6 (a) for the FEM mesh shown in
figure 6.6 (b). Although the FEM has been commonly used in the deformation analysis
of embankments and other geotechnical problems, it is still not widely used for the sta-
bility analysis of slopes as compared with the conventional limit equilibrium methods
[86, 147]. This can be deduced to the simplicity of the latter approach and to the available
computer programs usually providing a quick and accurate estimation of the FOS of a
slope.
In contrast, the FEM involves more complex theory and it usually requires more time
for developing model parameters, performing the computer analyses and interpreting
the results as described in detail in Smith & Griffiths [342].
Despite this, the FEM as in several advantages over the conventional limit equilibrium
methods, as stated by Griffiths & Lane [147]:
• No assumption is required in advance with respect to the shape and location of the
slip surface. Therefore, the failure finds its way through weak zones of the soil.
130
6.2 Slope stability
Smith & Griffiths [342] state that the FOS computed by the FEM is in good agreement
with that calculated by limit equilibrium methods, which is proven by various studies.
Influence of spatial variability on slope reliability: Therefore, the Random Finite Ele-
ment Method (RFEM) approach [127] is used to quantify the effects of spatial variability.
Within this RFEM study, random fields represent the spatial variability of a purely cohe-
sive soil, which is represented by an idealised linear-elastic, perfectly-plastic soil model.
Soil properties
soil unit weight γ’ = 21 kN/m3
friction angle ϕ = 0◦
cohesion c’ lognormal distribution
µ = 50 kN/m2
COVc’ = 10-100 % exp. corr. function
Θ = θ/H = 1-10
angel of dilatancy ψ = 0◦
Poisson’s ratio ν = 0.3
[email protected]
Geometry height H = 10 m
131
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
probability of failure pf
2.5 pf 10
reliability index b
2.0
-1
10
1.5
1.0 -2
10
0.5
b
-3
0.0 10
10 20 30 40 50 60 70 80 90
COVc’ in %
Figure 6.7: Effect of varying COVc on the reliability index β and on the probability of fail-
ure pf using lognormally distributed random variables for µc = 50 kN/m2 .
On the basis of the Monte-Carlo approach in chapter 4.3.2, each random field realisation
is represents a possible spatial distribution of higher and lower values, which influence
the system response.
Following the concepts of Fenton & Griffiths [127], the COV and the correlation length
of an isotropic random field of a cohesive slope are investigated using two different
random field generators, namely Sequential Gaussian Simulation Method (SGSIM) and
Sequential Indicator Simulation Method (SISIM). SGSIM and SISIM are basically similar
in the way of random field generation, but differ in the representation of spatial corre-
lation, as explained in appendix E. In the case of SISIM random fields, also the extreme
high and low values have a spatial correlation in contrast to SGSIM random fields. This
can be considered in the SISIM approach via different indicator correlation lengths. A
more detailed explanation of the differences and similarities of these approaches can be
found in Appendix E. The results of the case studies on the evaluation of the spatial vari-
ability of soil properties in chapter 3 clearly indicated that the investigated measurement
data cannot be fully described by a mean value, a standard deviation and only one single
spatial correlation function. Therefore, the SISIM approach is employed to investigate
the effects of this.
The libraries of GSLIB [97] are used for the generation of random fields, which are used
as input for the modified Monte-Carlo approach inside the RFEM framework, described
in section 6.1.3. After checking the convergence of the mean and standard deviation of
the REFM results, a normal distribution function is fitted to the simulation results via
the best-fit criterion in order to estimate the pf . Via this simplification it is possible to
investigate also small failure probabilities within reasonable computation times.
In the figure 6.8 (b) the effect of the normalized isotropic correlation length Θ = θ/H
on reliability index β for COVc = 20% can be seen. The reliability index of small normal-
ized correlation lengths is significantly higher than for large correlation lengths. Similar
results are reported Fenton & Griffiths [127].
[email protected]
Moreover, it is interesting to compare the outcomes of the different random field gen-
132
6.2 Slope stability
(a) (b)
SGSIM
reliability index b
reliability index b
30 30
SISIM
20 20
10 10
0 0
0 20 40 60 80 100 0 2 4 6 8 10
COVc’ in % isotropic correlation length Θ = q/H
Figure 6.8: (a) Effects of varying COVc on the probability of failure using different
random field generators with a Θ = θ/H = 2 and random variables
for µc = 50 kN/m2 ;
(b) effects of varying Θ = θ/H on the probability of failure for different
random field generators in comparison to random variables with
a COVc’ = 20%.
erators SGSIM and SISIM. As pointed out before, the SGSIM and LAS approach are used
in well known publications e.g. [127, 156, 158] due to their simplicity in describing a ran-
dom field by the mean value, a standard deviation and one correlation function. Com-
paring the results in figure 6.8, one can deduce that the influence of the spatial correlation
structure simulated via SGSIM and SISIM is offering comparable results. Although the
differences between the two approaches are relatively small, the SISIM approach offers
slightly lower reliability indices for the same spatial correlation lengths; the SGSIM ran-
dom fields only differ in terms of the indicator correlation length, as described in detail
in appendix E.4.
Additional conclusions can be drawn from figure 6.8 (a). Herein, the COVc has been
increased while keeping the isotropic correlation length constant. This offers a clear in-
sight into the effect of increasing COVc , which is similiar to the presented uncertainty
quantification of a single-layered soil slope using random variables instead of random
fields, which represent soil variability. The bigger the COVc of the soil the more unreli-
able is the slope.
By comparing figures 6.7 and 6.8, one can deduce that the consideration of spatial vari-
ability results in a higher slope reliability. In the case of an infinite correlation length, all
elements of a random field are fully correlated; therefore, the results of the RFEM and
reliability methods are assumed to converge. If there is a very small spatial correlation
of the cohesive soil, every single element of the random field is theoretically indepen-
dent. As a consequence, there will be no variation of the system response between each
realisation of a random field. In this case the slope stability problem is just influenced by
[email protected]
the mean value and the reliability index is infinity.
133
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
Cumulative probability of FOS
1
lognormal distribution
m = 50 kN/m² , COV = 20 %
0.8 Q = q / H = 2.0
0.6
exponential correlation function
-7
0.4 pf = 4.51 10
Gaussian correlation function
-9
0.2 pf = 1.85 10
Figure 6.9: Influence of the spherical, exponential and Gaussian correlation function on
the probability of failure of a slope.
Sensitivity analysis of the reliability of a single layered soil slope: Within this pre-
sented investigation, the influence of the stochastic properties is shown. One can clearly
see the effect of an increasing or decreasing spatial correlation and COVc’ . The transfer
of these findings to a practical situation is rather difficult, because at one site there will
not be a wide range of correlation lengths or COVs as shown in figure 6.8. As shown
in the results of previous chapter 3, one will get a probability density function (pdf) of
the correlation length as a result of the Bayesian Model Averaging approach of measure-
ment data, results presented in literature, correlation lengths derived from experiments
in comparable soils and expert knowledge. Linking this with the results of the para-
metric studies shown in figure 6.8, it is possible to calculate the most probable system
behaviour due to the probability density function of the correlation length. This idea
offers the calculation of one single value of pf of the single-layered soil slope. Also the
surrogate model of polynomial-chaos-expansion (PCE) polynomial can be employed to
approximate the system response, which offers as a by product the basis for the global
[email protected]
sensitivity analysis.
134
6.2 Slope stability
For this reason, the PCE approach is used within the concept of the Random Finite
Element Method (RFEM). This PCE mimics the system behaviour and represents the
mechanical system in a mean way as shown in appendix G. As outlined in section 4.4.2,
the global sensitivity indices can be calculated analytically by the presented formulae,
which quantify the sensitivity of the input parameters on pf .
Within this, the sensitivity of the uncertain variables for a 2D reliability analysis of the
slope is shown in figure 6.16. The mean value, the coefficient of variation and the correla-
tion lengths Θhor and Θver are considered as stochastic variables as indicated in table 6.4.
The system behaviour is approximated by a PCE at the different PCE evaluation points.
300 random field realisations are used for the computation of a mean system behaviour
at each PCE evaluation point. The results are approximated by a normal distribution
function, which allows one to extrapolate also small failure probabilities. Although this
approach is speeding up the evaluation of the system response significantly, there is still
a big computational effort for this study.
The fitting of the PCE to the system response is carried out with the approaches de-
scribed in section 4.3.3 and shown in figure 6.10. It can be derived from figures 6.10 (a)
and (b) that the expansion order M = 3 is enough to represent the system. By looking
at the determination coefficient Q2 , one can see that the fitting of the PCE is accurate
to almost 100 % for the expansion order M = 3. Besides this, it can be clearly derived
from these results that the determination coefficient, the adjusted determination coeffi-
cient and the Q2 determination coefficient quantify the fitting of the PCE-approximation
clearer in comparison to the empirical error or PRESS. Therefore, the author recommends
to used the determination coefficient and the Q2 determination coefficient in this context.
One can clearly deduce form the global sensitivity measures in figure 6.15 (c) that the
mean and COV of the cohesive soil are - as expected - the most influencing variables;
the vertical correlation length is far less important than the horizontal one. From this,
one can conclude that the vertical correlation length has a lower influence than the more
dominating horizontal correlation length. As a consequence, more effort should be put
into the investigation of the horizontal correlation length.
135
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
(a) (b)
PRESS
empirical error
determination coefficient
-4
x 10
PRESS 1.000
empirical error
0.01
1 0.999 det. coeff.
adj. det. coeff.
Q2 coeff.
0 0.000 0.998
1 2 3 1 2 3
expansion order M expansion order M
(c)
98 %
100
global sensitivity
index di in [%]
50
0.21 % ~0 % 1.75 %
0
mc COVc Qver Qhor
Figure 6.10: Different measures for the accuracy of the PCE to the system response of
a 2D slope stability analysis (a,b) and (c) global sensitivity measures of the
input variables.
different three failure modes, as illustrated by the typical deformed meshes and contours
of horizontal (out-of-face) displacement shown in figure 6.11. These modes depend on
the value of θhor relative to slope geometry, as defined by the slope height (HS ) and length
(L), and are summarised as follows [158, 266]:
• Mode 1: For θhor <HS there is little opportunity for failure to develop through semi-
continuous weaker zones. Hence, failure goes through weak and strong zones
alike, there is considerable averaging of property values over the failure surface,
and the slope fails along its entire length. This case is analogous to a conventional
2D analysis based on the mean value, [266].
• Mode 2: For HS < θhor < L/2, there is a tendency for failure to propagate through
semi-continuous weaker zones, leading to discrete 3D failures and a decrease in
reliability as the slope length increases. Hicks & Spencer [158] showed how proba-
bilistic theory could be used to predict the reliability of longer slopes based on the
detailed 3D stochastic analysis of shorter slopes.
• Mode 3: For θ
hor > L/2, the failure mechanism reverts to along the slope length. Al-
[email protected]
though it is similar in appearance to Mode 1, it is a fundamentally different mecha-
136
6.2 Slope stability
Mode 1
100 %
80 %
probability of failure pf
Mode 2
60 %
Mode 1
40 % Mode 2
Mode 3
Mode 3
20 %
0%
0.8 1.0 1.2 1.4 1.6 1.8
Factor of safety (FOS)
Figure 6.11: Different failure modes derived in Hicks & Spencer [158].
nism. In this case, failure propagates along weaker layers and there is a wide range
of possible solutions that depend on the locations of these layers. The solution for
this mode is analogous to a 2D stochastic analysis.
137
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
(a) (b)
3H Ls 3H 3H Ls 3H
Hs
HB 1.5 H HB
Figure 6.12: Geometry of the layered slope including one borehole using random vari-
ables (a) and random fields (b).
138
6.2 Slope stability
Table 6.5: Lognormal distributed input parameters of the C ASE STUDIES I, II and III.
139
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
(a) (b)
4 random variables 0.6
0.4
5
4 0.3
3 0.2
2
0.1
1
0 0.0
I.A I.B I.C c’ j’ c’ j’ HB LS HS
CASE STUDIES upper layer lower layer geometry
CASE STUDY I.A
CASE STUDY I.B
CASE STUDY I.C
Figure 6.13: C ASE STUDY I: Reliability evaluation of a two layered slope stability problem
(a) and the corresponding global sensitivity measures (b).
investigations like in reservoir engineering, which are usually not present in geotechni-
cal engineering.
Therefore, a simplified approach is set up. The combination of macro- and meso-scale
variability of soil properties is captured via a two step procedure: at first, the borderline
between the upper and lower layer is simulated via a one-dimensional random field,
which is conditioned to the soil investigation as shown in figure 6.12 (b). The SGSIM-
random fields of the strength properties of the upper and lower layer are generated and
mapped on the FEM mesh via the local averaging approach. The stochastic properties
are summarized table 6.5.
C ASE STUDY II is conducted to simulate the soil layer boundary as a simple horizontal
line and alternatively as a one dimensional random process, which is conditioned to a
soil investigation at the head of the slope as indicated in figure 6.12 (b). The properties
of both layers are shown in table 6.5. This illustrative example is showing the influence
of spatial variability at two scales; Figure 6.14 shows the scaling factor β̄ with respect to
the correlation length Θboundary = θboundary /H and the variance σboundary
2
of the stochastic
process, which is representing the boundary between the lower and upper layer. The
normalized reliability index β̄ is scaled by the reliability index β of C ASE STUDY I.A.
One can clearly see that the reliability is decreasing with a bigger variance σboundary
2
and a
longer correlation length Θboundary . The influence of the stochastic process between both
layers is vanishing in the case of an infinite correlation length Θboundary . The probability
of failure is significantly higher in the presence of spatial variability of both layers. It can
[email protected]
be concluded that the consideration of spatial variability is an important step towards
140
6.2 Slope stability
200
2
σ boundary= 0.2
reliability index b in %
scaling factor of the
190 2
σ boundary= 1.0
180
170
160
150
0 20 40 60 80 100
correlation length Θboundary
Figure 6.14: C ASE STUDY II: Normalized reliability index of different RFEM approaches
for a two layered slope with respect to a varying Θboundary = θboundary /H.
6.2.4 Conclusions
This case study investigates reliability of single- and a two-layered soil slopes. Herein, a
linear elastic perfectly plastic constitutive model on basis of the Mohr-Coulomb criterion
is used in a 2D FEM model.
At first, the effects of spatial soil variability are quantified for single layered soil slopes,
[email protected]
which are compared to slopes with random and not spatially correlated soil properties.
141
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
determination coefficient
1.00
0.98
determination coefficient
0.96 adjusted determination coefficient
Q2 coefficient
0.94
1 2 3 4 5
expansion order M of the PCE
Figure 6.15: C ASE STUDY III: Different determination coefficients for the PCE accuracy.
glolbal sensitivity of the parameters
0.3
0.2
0.1
0 PC PC PC PC PC PC PC
d3 d1 d13 d2 d23 d12 d123
Figure 6.16: C ASE STUDY III: Global sensitivity measures δiP C for the isotropic correlation
length of the upper layer (1) and the lower layer (2) as well as for the layer
boundary (3) for the PCE expansion order M = 5.
142
6.2 Slope stability
Within the investigation of the effects of different correlation function on the reliability
of single layered soil slopes, the exponential correlation function is identified as the most
conservative one. Moreover, it is shown that different complex correlation structures of
spatial variability, which are simulated by SGSIM and SISM algorithms, do not have a
major effect on the calculated probability of failure.
The influences of mean value, coefficient of variation, correlation function and random
field generators are investigated within the RFEM framework. Besides this, the effects
of the mean value, the coefficient of variation and the anisotropy of the horizontal and
vertical correlation length are quantified within a novel sensitivity analysis within the
RFEM framework. It is shown that the mean valued, the coefficient of variation and
the horizontal correlation length have the biggest influence on the probability of failure,
whereas the contribution of the vertical correlation length is very small.
The effects of spatial variability at different scales is in the focus of additional paramet-
ric studies on two-layered soil slopes. Again the importance of spatial soil variability is
highlighted by comparing the effects of random properties and spatially correlated ran-
dom properties. Within this study the effects of an uncertain geometry are investigated.
It is shown that the slope geometry and the soil layering have a bigger influence in com-
parison to the stochastic soil properties. The effects of spatial soil variability at different
scales are quantified with respect to random soil properties. The spatial variability at the
large scale and the spatial variability of the upper layer are contributing the most to the
probability of failure of the investigated two layered soil slope.
The presented studies are fundamental investigations on the effects of soil variabil-
ity. These studies can be enriched by additional investigations using 3D slope stability
analyses, which would contribute to the quantification of the effects of spatial variability
for slope analyses. In addition to this, the investigation of the uncertain slope geometry
would offer additional insight into 3D slope stability analyses.
The presented findings can be further extended by additional 3D slope stability anal-
yses, which would represent more exactly the spatial variability of the subsoil.
143
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
144
6.3 Risk based characterisation of an urban site
Table 6.6: Mean values and standard deviations of the lognormally distributed measure-
ment and literature data.
Information from more than 2,000 boreholes penetrating these deposits was used by
Marconi [238] to model their lithological and textural associations in a conceptual geo-
logical model. This model has been used as the basis for the geostatistical simulations.
For this case study, only geotechnical boreholes with geotechnical information penetrat-
ing these deposits were selected to derive geomechanical properties of the subsoil. This
set includes 283 boreholes and 719 samples.
Five main soil types are derived from this large dataset and summarized in table 6.6,
which have been used in the latter to simulate the soil behaviour. S OIL TYPE I consists
of gravel, medium and coarse sand, S OIL TYPE II of silty fine sand, and sandy silt, S OIL
TYPE III of silty clay, clayey silt and organic clay and S OIL TYPE IV of bedrock. Table
6.7 provides the number of measurement data for the strength and stiffness properties
of the different soil types. The author combined the measurement data with literature
[346] data using the Bayesian approach. Via this, the description of the variability of the
subsoil is enriched by the combination of measurements and expert judgement.
Simulation of the geological uncertainty. The uncertainty of the geology is seen as un-
certainty of different soil types, which includes the uncertainty of the domain boundaries
of the spatially distributed soil types. This is done via the Pluri-Gaussian Simulation ap-
proach, which is simulating spatially distributed categorical variables. The principle of
the Pluri-Gaussian Simulation is to simulate one or several continuous Gaussian fields
and to truncate them in order to produce a categorical variable. Armstrong et al. [14]
[email protected]
describe the concept of the Pluri-Gaussian simulation approach in depth and illustrate
145
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
soil types c′ ϕ′ E
number of measurement samples
I gravel, medium and coarse sand 8 8 4
II silty fine sand, sandy silt 40 40 3
III silty clay, clayey silt, organic clay 86 86 92
IV bedrock 2 2 1
the basic idea, as shown in figure 6.17: Two Gaussian random fields (figure 6.17 a,b)
are used to describe the main characteristics of the soil, e.g. anisotropies. The different
soil types are generated by truncating several Gaussian random fields (figure 6.17 a,b).
This provides flexibility to handle complex transitions and anisotropy among the soil
types. The truncation rule is called the lithotype rule (figure 6.17 c) because it synthe-
sizes transitions between the soil types. One can easily deduce from this that the choice
of a lithotype rule is a major step of the methodology. In this case study, the lithotype
rule was derived from borehole logs providing a good basis, on which soil types can
and cannot be in contact. These proportions are not constant over the domain, but vary
vertically and laterally because of the existence of trends in the geological processes as
shown in Appendix H. The mathematical theory of the Pluri-Gaussian [215, 241] meth-
ods is described in detail in by Marconi [238] amongst others [14, 118]. This includes the
conditioning of the unconditional Pluri-Gaussian random field to borehole data via the
Gibbs sampler.
In Appendix H, the details of the Pluri-Gaussian Simulations are listed, which have
been provided by Marconi [238]: The plan view of the investigated area, the spatial
statistics of the soil types, the lithotype rules and illustrative illustrations of one realisa-
tion of the Pluri-Gaussian Simulations are shown in Appendix H.
Mechanical description of the problem: Figure 6.19 shows the 2D FEM mesh of a
rigid building (B/W = 30/15 m), which is used for the evaluation of the load - displace-
[email protected]
ment curves. Within the stepwise construction of the rigid building (B=30 m) differential
146
6.3 Risk based characterisation of an urban site
(a) Y1 Y2 (b)
Y2
(c) (d)
(2)
(1)
(3)
Y1
antropic backfill
Soil type I
Soil type II
Soil type III
Soil type IV
Figure 6.18: Part of the realisation of the Pluri-Gaussian random field (200/200/75m) in-
cluding all soil types.
147
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
4.5 B B 4.5 B
q
P1 P2
3B
Figure 6.19: 2D FEM model used in this case study.
settlements between points P1 and P2 are observed, which are related to the spatial dis-
tribution of the 5 different soil types, as shown in table 6.6. Herein, the contact between
the rigid building and the subsoil is assumed to be rigid with full bounding.
The soil behaviour is simulated by a small strain, double hardening model, which was
developed by Benz [39]. This advanced, non-linear soil model is able to take the small
strain behaviour as well as the non-linear stress-strain relationship of soil into account.
The input variables of this soil model are given in the Appendix H in table H.1.
Stochastic soil properties: Although there is a large number of borehole data, it is quite
difficult to estimate the strength and stiffness parameters of the different soil types: Most
of these boreholes aimed for a geological description and only a relative small number
of boreholes focused on geotechnical characterisation as indicated in table 6.7. There-
fore, the measurement data are combined with expert knowledge from the Geotechnical
Handbook [346] via the Bayesian approach as proposed by Ching et al. [80]. The results
of this combination are given in table 6.6. One can clearly see that the combined values
of the cohesion, friction angle and E-modulus show a smaller standard deviation. These
lognormal distributed values are used for the random number generation of the strength
and stiffness parameters of each soil type. These random variables, which are based on
the mean values and standard deviations in table 6.6, are used in combination with the
300 Pluri-Gaussian random field realisations for the description of the geological hetero-
geneity. Each Pluri-Gaussian random field realisation has a length of 620 m, a width of
436 m and an overall-depth of 75 m with a horizontal discretisation of ∆X = ∆Y = 5 m
and a vertical discretisation of ∆Z = 0.05 m. These Pluri-Gaussian random fields were
generated by Marconi [238].
Due to the scarcity of data describing the mechanical behaviour of anthropic backfill
[email protected]
material, this material is not considered in the subsequent investigations.
148
6.3 Risk based characterisation of an urban site
Random Finite Element approach & fragility curves: As mentioned above, the 3D
Pluri-Gaussian random fields are used for the description of the geological uncertainty.
A modified RFEM approach is used to quantify the effects of geological uncertainty.
Within the preprocessing phase of this case study, a simple mapping procedure has been
used to map the 3D random field onto the 2D FEM mesh. The Pluri-Gaussian random
field is divided into blocks (5B × 10B × 75 m). The transformation from 3D to 2D is
carried out by averaging the soil properties in the third direction. Then the finer 2D
random field is averaged over the coarser FEM mesh.
Within the processing phase, the differential settlements of the rigid building due to a
load up to q = 2, 000 kN/m2 are evaluated by using the RFEM framework.
The boundaries of the mechanical model are chosen so as to minimize the influence on
the settlement prediction as indicated in figure 6.19. This follows the recommendations
for numerical modelling in geotechnical engineering in [291, 292].
As the first step within the postprocessing phase, the fragility curves are evaluated in the
middle of the building, which is shifted over the entire field. The fragility curves describe
the probability of exceeding the ultimate differential settlements due to a load, which is
increased to q = 2, 000 kN/m2 . As shown in literature e.g. [286, 289], this fragility curve
can be approximated by a cumulative lognormal distribution function, which offers the
possibility of a continuous definition of fragility due to differential settlements. For this
purpose the probabilities of exceeding the ultimate differential settlement (αultimate =
1/300, 1/500, 1/600 and 1/1, 000) for the different load levels are evaluated. Due to the
fact that especially for small load levels the probability of exceeding αultimate is very
small, the system response was approximated via a normal distribution function. Via
this, it is possible to estimate the probability of exceeding αultimate . One has to keep in
mind that the absolute numbers of these probabilities are just approximations and offer
a good basis to compare two different locations of the building.
149
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
are indicated by black points. The orange squares indicate the location of the boreholes
in the figures 6.21 and 6.22. The ultimate loads qSLS , which are calculated for four dif-
ferential settlement limits αultimate = 1/300, 1/500, 1/600, 1/1, 000. These ultimate loads
qSLS are plotted as grey-shaded regions in the background of figures 6.21 and 6.22. The
grey-shaded regions between the points are interpolated using the Ordinary Kriging al-
gorithm. By looking at the figures 6.21 and 6.22, one can clearly identify areas, which are
less and more endangered to differential settlements under a specific load.
6.3.4 Conclusion
The presented case study presents a methodology for risk based site characterisation.
This approach uses geostatistical simulation techniques and FEM modelling to derive
the ultimate load over a site. Via the presented methodology, the geological uncertainty
is combined with stochastic, geotechnical soil properties, which are based on measure-
ment data and on expert judgement. The uncertainty of the complex geological condi-
tions is quantified within a mathematical framework. By using RFEM in combination
with fragility curves, these results are translated into into maps, which offer another in-
sight into the effects of soil heterogeneity and the resulting risk, which is a novelty at the
interface between engineering geology, geostatistics and geotechnical engineering. A
by-product of this methodology of risk based site characterisation is the quantification
of the effects of macro-scale soil variability.
By using the Pluri-Gaussian Simulation algorithm one has to be aware that modelling
geological uncertainty by using different Gaussian random fields is a simplification. As
pointed out in chapter 2, a Gaussian random field is a simple description of spatial vari-
[email protected]
ability using a mean value, a standard deviation and one single covariance function.
150
6.3 Risk based characterisation of an urban site
(a)
300
relative y-coordinate in meter
250
200
midpoints of
150 the building
boreholes
100
50
0 qSLS =
4,500 kN/m²
0 50 100 150 200 250
4,000 kN/m²
relative x-coordinate in meter
(b) 3,500 kN/m²
2,500 kN/m²
250
2,000 kN/m²
200
1,500 kN/m²
150 1,000 kN/m²
50
Figure 6.21: Map of allowable loads pSLS on a footing with αulitmate = 1/300 (a) and
αulitmate = 1/500 (b).
[email protected]
151
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
(a)
300
relative y-coordinate in meter
250
200
midpoints of
150 the building
boreholes
100
50
0 qSLS =
4,500 kN/m²
0 50 100 150 200 250
4,000 kN/m²
relative x-coordinate in meter
(b) 3,500 kN/m²
2,500 kN/m²
250
2,000 kN/m²
200
1,500 kN/m²
150 1,000 kN/m²
50
Figure 6.22: Map of allowable loads pSLS on a footing with αulitmate = 1/600 (a) and
αulitmate = 1/1, 000 (b).
[email protected]
152
6.3 Risk based characterisation of an urban site
Moreover, one has to keep in mind that the lithotype is a subjective interpretation of ge-
ological conditions, although the lithotype rule is based on the analysis of a large set of
boreholes.
However, the geological uncertainty can also be simulated by other simulation ap-
proaches in this context, which are able to simulate categorical variables described in
literature [305].
Besides this, the Pluri-Gaussian simulation algorithm needs a large number of bore-
holes and laboratory tests to perform well. The stochastic properties of the Gaussian
random fields and the definition of the lithotype rule needs a quite large database of
boreholes. Otherwise, this simulation algorithm cannot be recommended for similar
investigations in the case of a small dataset.
The presented case study on the risk based characterisation of an urban site extends
the state of research presented in different contributions. Amongst others [24, 172, 312],
Fenton [127] focused on the effects of meso-scale soil variability and neglected the macro-
scale variability. On the basis of these results, Chen et al. [73] combine the micro- and
the meso-scale variability via a multi-scale random field approach, which is applied to
the ultimate limit state and to the serviceability limit state of strip footings.
However, the studied contributions [24, 127, 172, 312] do not take into account the
available prior information like literature knowledge or soil investigations, a concep-
tual geological model or subjective engineering judgement. Therefore, the presented
case study offers a promising scheme of evaluating the effects of geological, macro-scale
uncertainty in the presence of a large borehole database. Moreover, the presented ap-
proach offers a realistic interpretation of spatially variable site conditions via the adopted
fragility curve approach, which allows a realistic estimation of the risk of differential set-
tlements. The results of this are transformed back to a limiting load, which is obeying
the probability for SLS given in the E UROCODE 7. Via this, it is possible to identify
endangered areas of a building site and this procedure quantifies the consequences of
geological uncertainty.
153
Chapter 6 Selected case studies in uncertainty quantification including spatial variability
Risk based characterisation of an urban site: This case study investigates the effects of
geological uncertainty. The different soil types are simulated as categorical variables by
means of the Pluri-Gaussian simulation approach. This approach considers soil investi-
gations and geostatistical subsoil characteristics as well as engineering judgement. These
quantifications of macro-scale soil variability are used for the calculation of the footing
serviceability. The footing serviceability is investigated by differential settlements and
means of fragility curves. These fragility curves are the basis for risk maps, which vi-
sualize the effects of geological uncertainty and identifies endangered area. This study
clearly quantifies the influence of macro scale soil variability and indicates clearly the
need to consider this.
154
Chapter 7
Summary and conclusions
155
Chapter 7 Summary and conclusions
lation) may be studied via parametric studies, and this will bring confidence to the
predicted behaviour in a given design situation. The way to conduct an uncertainty
quantification in geotechnical engineering is shown in case studies onto tunnel lining,
tunnel face stability, settlements introduced by tunnelling, slope stability, bearing capac-
ity of vertically loaded strip footings and serviceability of footings. In these case studies
semi-analytically defined and numerically simulated limit states are investigated with
respect to their behaviour due to random variables and random fields. The contribution
of the input variables is analysed by means of local and global sensitivity analyses. The
presented case studies will show the potential of the framework of uncertainty quantifi-
cation.
It is also very important to be aware of the limitations that lie in a probabilistic analy-
sis. The presented approaches in chapter 2 allow to consider pre-knowledge from expert
judgement, literature and/or available field tests, which can enrich a stochastic site char-
acterisation. However, there is a need to conduct a bigger soil investigation campaign
compared to the state-of-the-art because more input data are needed to derive statistical
distributions and parameters describing spatial variability.
Besides this, another difficulty in using probabilistic concepts in applied engineer-
ing is the difficult statistical background, which is needed to understand the results of
probabilistic analyses. Within chapter 3, a large number of field tests is analysed and
contributes to the database on stochastic soil properties, which is assembled by the re-
sults of a large literature review in appendix A. The basics of safety and uncertainty
are summarized in chapter 4 and shall help to develop an understanding of uncertainty
quantification and to interpret results of probabilistic analyses presented in chapters 5
and 6. These case studies show the application of uncertainty quantification and shall
guide the reader to a comprehensive understanding of the presented approaches. These
different case studies in tunnelling and foundation engineering show the effects of soil
variability and spatial variability at different scales. Moreover, sensitivity analyses are
carried out to investigate the contribution of each random parameter to the probability
of failure.
156
7.3 What is the best way to evaluate the effects of soil variability?
isation. Different sampling concepts are offered in literature [257] e.g. simple random
pattern, stratified random pattern or cluster sampling, and are used in groundwater
ecology and geostatistics to guide the choice of additional soil samples. These concepts
mainly focus on the properties of different homogeneous soil layers, which can be evalu-
ated according to the statistical methods described in guidelines e.g. [195, 390]. It can be
deduced from the case studies in chapter 6 that the spatial variability of soil properties
has significant influence on the behaviour of geotechnical structures. Therefore, it is nec-
essary to quantify theses effects in a rational manner. Not only literature data but also
field investigations have to be conducted to investigate the the spatial correlation of soil
properties. Consequently, the sampling schemes for detecting spatial variability at dif-
ferent scales have to be used from geostatistics [79, 248] or soil science [314] to evaluate
the these properties in an economic and cost effective way.
157
Chapter 7 Summary and conclusions
ability do not play a major role. As shown in chapter 6, micro-scale variations (e.g. at
the grain-size level) have minor effects on the overall failure probability of a structure,
whereas large geological spatial variations (e.g. soil layers) can have significant effects
as shown in the case studies. One can deduce from the presented case studies that the
effects are significant for the combination of meso- and macro-scale spatial soil variabil-
ity. The presented engineering approach is taking these two scales of spatial variability
into account: The soil layer boundary is represented by a conditional, stochastic process,
which is separating layers with spatial variable properties. By using the Polynomial-
Chaos-Expansion (PCE) in connection with the Random Finite Element approach for the
representation of the system response, it is possible to consider not only a single corre-
lation length, but a lognormal distribution of correlation lengths. Moreover, the global
sensitivity measures of the uncertain variables can be derived from the PCE analytically.
This offers a possible way to quantify the effects of multi-scale spatial variability of soil
properties.
Science: Additional studies should be conducted to gain more knowledge and experi-
ence in the of stochastic quantification of soil properties. The presented literature study
and the results of the case studies analysing a large set of CPT tests offer a good basis
for this. This would form a good starting point for the formulation of a guideline for the
evaluation of spatial variability and heterogeneity comparable to recommendations like
[11, 17, 67, 95]. The need for cost effective sampling schemes is closely linked with this.
Apart from the presented Methods of Moments and Maximum Likelihood approaches
in chapter 2, the author recommends inverse methods like in geostatistics or earthquake
engineering to investigate the effects of soil variability. These methods are based on
Bayesian approaches [164] or sequential approaches[171, 304] amongst others.
Moreover, the quality of the random field representation within the framework of un-
certainty quantification and reliability based design is an important task. The forecast
quality in meteorology [309] might be a staring point for further developments in this
context .
Apart from this, the partial safety factors in the EC7, which are mainly based on expe-
rience, should be enriched and extended with the results from additional investigations.
This would lead to a more precise separation between soil uncertainties and human re-
lated errors.
Another interesting part in the context of economic design is the optimization of a
geotechnical structure including uncertainties. Reliability based design optimization
includes different failure modes of complex structures incorporating the variability of
[email protected]
loads and resistance forces in a proper way.
158
7.4 Recommendations for research
Applied engineering: Guidelines and worked examples would be the best arguments
to convince applied engineers to used reliability based design approaches. This has al-
ready been started by different groups like GeoSNET [360] or JCSS [399] amongst others
[262, 280]. It has to be pointed out that user friendly software tools would support this,
which would allow one to estimate the reliability of complex systems in a fast and effi-
cient way by using robust and efficient algorithms.
159
[email protected]
Bibliography
[1] M. Aboufirassi and M.A. Marino. Cokriging of aquifer transmissivities from field
measurements of transmissivity and specific capacity. Mathematical Geology, 16(1):
19–35, 1984.
[2] M. Abramovich and I.A. Stegun. Handbook of Mathematical Functions, Graphs and
Mathematical Tables. National Bureau of Standards Applied Mathematics, Wash-
ington, 1964.
[3] M. Abramovich and I.A. Stegun. Handbook of Mathematical Functions with Formulas,
Graphs and Mathematical Tables. Dover, 1974.
[5] H. Ahrens, E. Lindner, and K.H. Lux. Zur Dimensionierung von Tunnelausbauten
nach den Empfehlungen zur Berechnung von Tunneln im Lockergestein (1980).
Die Bautechnik, 8:260–311, 1982.
[6] H. Akaike. Information theory and an extension of the maximum likelihood prin-
ciple. In Proceedings of Second International Symposium on Information Theory, vol-
ume 1, pages 267–281. Springer Verlag, 1973.
[7] H. Akaike. A Bayesian analysis of the minium AIC procedure. Ann. Inst. Statist.
Math., 30:9–14, 1978.
[8] R. Al-Khoury, K.J. Bakker, P.G. Bonnier, H.J. Burd, G. Soltys, and P.A. Vermeer.
PLAXIS 2D Version 9. R.B.J. Brinkgreve and W. Broere and Waterman, 2008.
[9] D. Alber and W. Reitmeier. Beschreibung der räumlichen Streuungen von Bodenken-
nwerten mit Hilfe der Zeitreihenanalyse, volume 7. Technische Universität München,
1986.
[10] E.E. Alonso and R.J. Krizek. Stochastic formulation of soil properties. In Proc. 2nd
Int. Conf. on Applications of Statistics & Probability in Soil and Structural Engineering,
Aachen, volume 2, pages 9–32, 1975.
[11] American Society of Civil Engineers. Standard Guideline for the Geostatistical Esti-
mation and Block-averaging of Homogeneous and Isotropic Saturated Hydraulic Conduc-
tivity. American Society of Civil Engineers, Environmental and Water Resources
[email protected]
Institute, 2010.
161
Bibliography
[13] A.H.S. Ang and W.H. Tang. Probability concepts in engineering planning and design,
vol. 2: Decision, risk, and reliability. JOHN WILEY & SONS, INC., 1984.
[15] A. Asaoka and D.A. Grivas. Spatial variability of the undrained strength of clays.
Journal of the geotechnical engineering division, 108(5):743–756, 1982.
[17] ASTM D 5549-94. Standard Guide for the Contents of Geostatistical Site Investiga-
tion Report, 2001.
[19] R. Azzouz, C. Bacconnet, and J.-C. Faugeras. Analyse geostatistique dune cam-
pagne de reconnaissance au penetrometre statique. In Proc. 5th Int. Conf. on Appli-
cations of Statistics and Probability in Soil and Structural Engineering, pages 821– 828,
Vancouver, 1987.
[20] G.L. Babu and S.M. Dasaka. The effect of spatial correlation of cone tip resistance
on the bearing capacity of shallow foundations. Geotechnical and Geological Engi-
neering, 26(1):37–46, 2008.
[21] G.B. Baecher. Site Exploration: A probabilistic approach. PhD thesis, Massachusetts
Instiute of Technology, 1972.
[22] G.B. Baecher. Analyzing exploration strategies. In C.H. Dowding, editor, Site Char-
acterization and Exploration, New York, 1978. American Society of Civil Engineers.
[23] G.B. Baecher and J.T. Christian. Reliability and statistics in geotechnical engineering.
John Wiley & Sons Inc., 2003.
[24] G.B. Baecher and T.S. Ingra. Stochastic FEM in settlement predictions. Journal of
the Geotechnical Engineering Division, 107(4):449–463, 1981.
[25] G.B. Baecher, M. Chan, T.S. Ingra, T. Lee, and L.R. Nucci. Geotechnical reliability
of offshore gravity platforms. Technical Report Report MITSG 80-20, Sea Grant
[email protected]
College Program, Mass. Inst, of Tech, 1980.
162
Bibliography
[26] G.B. Baecher, J.D. Ditmars, D.E. Edgar, and C.H. Dowding. Radioactive waste iso-
lation in salt: A method for evaluating the effectiveness of site characterization
measures. Technical report, Report ANL/EESTM-342, Argonne National Labora-
tory, Ill, 1988.
[27] S. Bakhtiari. Stochstiac Finite Element Slope Stability Analysis. PhD thesis, Manch-
ester Centre for Civil and Construction Engineering, University of Manchester,
England, 2011.
[28] A.A. Bakr, L.W. Gelhar, A.L. Gutjahr, and J.R. MacMillan. Stochastic analysis of
spatial variability in subsurface flows. 1. comparison of one-and three-dimensional
flows. Water Resources Research, 14(2):263–271, 1978.
[30] A. Bárdossy and Z.W. Kundzewicz. Geostatistical methods for detection of outliers
in groundwater quality spatial fields. Journal of Hydrology, 115(1-4):343–359, 1990.
[32] R. J. Barnes. Bounding the required sample size for geological site characterization.
Mathematical Geology, 20(5):477–490, 1988.
[34] J.S. Bendat and A.G. Piersol. Random data analysis and measurement procedures, vol-
ume 11. IOP Publishing, 2000.
[35] J.P. Benkendorfer, C.V. Deutsch, P.D. LaCroix, L.H. Landis, Y.A. Al-Askar, A.A.
Al-AbdulKarim, and J. Cole. Integrated Reservoir Modelling of a Major Arabian
Carbonate Reservoir. Middle east oil show, 298(96):1–26, 1995.
[38] T. Benz. Small-strain stiffness of soils and its numerical consequences. PhD thesis,
University of Stuttgart, Institute of Geotechnical Engineering, 2007.
[39] T. Benz. Small strain double hardening model. Technical report, NTNU, Trond-
[email protected]
heim, 2012.
163
Bibliography
[41] D.T. Bergado and L.R. Anderson. Stochastic analysis of pore pressure uncertainty
for the probabilistic assessment of the safety of earth slopes. Soils and Foundations,
25(2):87–105, 1985.
[42] J.O. Berger, V. De Oliveira, and B. Sansó. Objective Bayesian analysis of spatially
correlated data. Journal of the American Statistical Association, 96(456):1361–1374,
2001.
[43] M. Berveiller, B. Sudret, and M. Lemaire. Presentation of two methods for com-
puting the response coefficients in stochastic finite element analysis. In Proceedings
of 9th ASCE specialty Conference on Probabilistic Mechanics and Structural Reliability,
Albuquerque, USA, 2004.
[45] P. Biçer and P.T. Özener. Vertical variability of the undrained shear strength of
golden horn clay. Marine Georesources and Geotechnology, 27(4):309–321, 2009.
[46] A. A. Binsariti. Statistical analyses and stochastic modeling of the Cortaro aquifer in
Southern Arizona. PhD thesis, University of Arizona, 1980.
[48] R. Borja, editor. Multiscale and multiphysics processes in geomechanics. Springer, 2011.
[49] J.-M. Bourinet, C. Mattrand, and V. Dubourg. A review of recent features and
improvements added to FERUM software. In Proceedings of 10th International Con-
ference on Structural Safety and Reliability (ICOSSAR’09), Osaka, Japan, 2009.
[50] D.S. Bowles and H.H. Ko, editors. Probabilistic Characterization of Soil Properties:
Bridge Between Theory and Practice. ASCE, 1984.
[51] G.E.P. Box and G.M. Jenkins, editors. Time Series Analysis: Forecasting and Control.
Holden-Day, San Fransisco, 1970.
[52] H. Bozdogan. Model selection and akaike’s information criterium: The general
tehory and its analytical extensions. Psychometrika, 52:345–370, 1987.
164
Bibliography
[55] P.I. Brooker, J.P. Winchester, and A.C. Adams. A geostatistical study of soil data
from an irrigated vineyard near Waikerie, South Australia. Environment Interna-
tional, 21(5):699–704, 1995.
[56] D.J. Brus and J.J. De Gruijter. Random sampling or geostatistical modelling?
choosing between design-based and model-based sampling strategies for soil
(with discussion). Geoderma, 80(1-2):1–44, 1997.
[59] P.A. Burrough. Multiscale sources of spatial variation in soil. I - The application
of fractal concepts to nested levels of soil variation. Journal of Soil Science, 34(3):
577–597, 1983.
[60] P.A. Burrough. Multiscale sources of spatial variation in soil. III - The application
of fractal concepts to nested levels of soil variation. Journal of Soil Science, 34(3):
578–599, 1983.
[62] E. Byers and D.B. Stephens. Statistical and stochastic analyses of hydraulic con-
ductivity and particle-size in a fluvial sand. Soil Science Society of America Journal,
47(6):1072, 1983.
[63] D.G. Cacuci, M. Ionescu-Bujor, and I.M. Navon. Sensitivity and Uncertainty Analy-
sis: Applications to large-scale systems, volume 2. CRC, 2005.
[65] F. Cafaro and C. Cherubini. Large sample spacing in evaluation of vertical strength
variability of clayey soil. Journal of geotechnical and geoenvironmental engineering,
128:558, 2002.
[66] J.A. Calgaro. Safety philosophy of eurocodes. In Proceedings of the 3rd International
Symposium on Geotechnical Safety and Risk, 2011.
[67] E.O.F. Calle, G.A.M Kruse, M.T. van der Meer, W.R. Halter, and B.M. Effing.
Schematiseren Geotechnische Faalmechanismen bij Dijken - SBW Faalmechanis-
[email protected]
men TR Grondonderzoek, (in Dutch). Technical report, Deltares, 2009.
165
Bibliography
[68] R.G. Campanella, D.S. Wickremesinghe, and P.K. Robertson. Statistical treatment
of cone penetration test data. In Proceedings of the 5th International Conference on
Reliability and Risk Analysis in Civil Engineering, 1987.
[69] S.F. Carle and G.E. Fogg. Transition probability based indicator geostatistics. Math-
ematical Geology, 28(4):453–476, 1996.
[70] R. Caspeele. Proabilistic evaluation of conformity control and the use of the Bayesian up-
dating techniques in the framework of safety analyses of concerte structures. PhD thesis,
University Gent, 2009.
[71] CEN. Eurocode: Basis of structural design. European standard, EN 1990: 2002,
2002.
[73] Q. Chen, A. Seifried, J.E. Andrade, and J.W. Baker. Characterization of random
fields and their impact on the mechanics of geosystems at multiple scales. Inter-
national Journal for Numerical and Analytical Methods in Geomechanics, 36(2):140–165,
2010.
[74] Q. Cheng. Multifractality and spatial statistics. Computers & Geosciences, 25:949–
961, 1999.
[77] C. Cherubini, I. Giasi, and L. Rethati. The coeffcient of variation of some geotech-
nical parameters. In Proceedings of Probabilistic Methods in Geotechnical Gengineering,
1993.
[78] P. Chiasson and Y.J. Wang. Spatial variability of sensitive Champlain sea clay and
an application of stochastic slope stability analysis of a cut. In Proceedings of the 2nd
International Workshop on Characterisation and Engineering Properties of Natural Soils.
Singapore, November, 2006.
[80] J. Ching, J.R. Chen, J.Y. Yeh, and K.K. Phoon. Updating uncertainties in friction
angles of clean sands. Journal of Geotechnical and Geoenvironmental Engineering, 138
[email protected]
(2):217–229, 2011.
166
Bibliography
[81] Y.H. Chok. Modelling the effects of soil variability and vegetation on the stability of
natural slopes. PhD thesis, University of Adelaide, School of Civil, Environmental
and Mining Engineering, 2008.
[82] G. Christakos. Random field models in earth sciences. Academic Press, 1992.
[83] J. Chu, W. Hsču, H. Zhu, and A.G. Journel. The Amoco case study. Report 4,
Stanford Center for Reservoir Forecasting, 1991.
[84] W.G. Cochran. Sampling techniques. Wiley, London, 3rd edition, 1977.
[85] P.D. Coddington. Random number generators for parallel computers. The NHSE
Review, 2:1–26, 1997.
[90] D.J. DeGroot and G.B. Baecher. Estimating autocovariance of in-situ soil proper-
ties. Journal of Geotechnical Engineering, 119(1):147–166, 1993.
[91] J.P. Delhomme. Spatial variability and uncertainty in groundwater flow parame-
ters: a geostatistical approach. Water Resources Research, 15(2):269–280, 1979.
[92] F.A. Delory, A.M. Crawford, and M.E.M. Gibson. Measurements on a tunnel lining
in very dense till. Canadian Geotechnical Journal, 16(1):190–199, 1979.
[93] A. Der Kiureghian, H-Z. Lin, and S.-J. Hwang. Second-order reliability approxi-
mations. J. Engrg. Mech. Div., ASCE, 113(8):1208–1225, 1987.
[96] C.V. Deutsch. Geostatistical Reservoir Modeling. Oxford University Press, New York,
[email protected]
USA, 2002.
167
Bibliography
[97] C.V. Deutsch and A.G. Journel. GSLIB - Geostastical Software Library and Users’s
Guide. Oxford University Press, 1998.
[98] J.L. Devary and P.G. Doctor. Pore velocity estimation uncertainties. Water Resources
Research, 18(4):1157–1164, 1982.
[99] J. Diaz-Padilla and E.H. Vanmarcke. Settlement of Structures on Shallow Founda-
tions: A Probabilistic Analysis. Dept. of Civil Engineering, Massachusetts Institute
of Technology, 1974.
[100] C.R. Dietrich and M.R. Osborne. Estimation of covariance parameters in kriging
via restricted maximum likelihood. Mathematical Geology, 23(1):119–135, 1991.
[101] DIN 4017. Baugrund-Berechnung des Grundbruchwiderstands von Flachgrün-
dungen, 2006.
[102] DIN 4019. Baugrund-Setzungsberechnungen, 2001.
[103] DIN 4094-1. Subsoil - field investigations - part 1: Cone penetration tests, 06 2002.
[104] DIN 4094-5. Geotechnical field investigations - part 5: Borehole deformation tests.,
2001.
[105] O. Ditlevsen and H.O. Madsen. Structural reliability methods, volume 178. Wiley
Chichester, 1996.
[106] D.N. Veritas. Recommended practice report dnv-rp-c207: Statistical representation
of soil data. Technical report, D.N. Veritas, 1997.
[107] G.R. Dodagoudar and G. Venkatachalam. Reliability analysis of slopes using fuzzy
sets theory. Computers and Geotechnics, 27(2):101–115, 2000.
[108] H. Duddeck. Empfehlungen zur Berechnung von Tunneln im Lockergestein. Die
Bautechnik, 10:349–356, 1980.
[109] A.S. Dyminski, J. Howie, D. Shuttle, and A. Kormann. Soil variability study for
embankment design of port of navegantes, brazil. In Proceedings of GeoCongress
2006: Geotechnical Engineering in the Information Technology Age, pages 1–6. ASCE,
2006.
[110] M.D. Ecker and A.E. Gelfand. Bayesian variogram modeling for an isotropic spa-
tial process. Journal of Agricultural, Biological, and Environmental Statistics, 3:347–
369, 1997.
[111] B. Efron. Bootstrap methods: another look at the jackknife. The annals of Statistics,
7(1):1–26, 1979.
[112] J. Eggleston and S. Rojstaczer. Identification of large-scale hydraulic conductiv-
ity trends and the influence of trends on contaminant transport. Water Resources
[email protected]
Research, 34(9):2155–2168, 1998.
168
Bibliography
[114] H. El-Ramly, N.R. Morgenstern, and D.M. Cruden. Probabilistic slope stability
analysis for practice. Canadian Geotechnical Journal, 39(3):665–683, 2002.
[115] A.M.M. Elfeki. Stochastic characterization of geological heterogeneity and its impact
on groundwater contaminant transport. PhD thesis, TU Delft, Delft University of
Technolog, 1996.
[117] B.R. Ellingwood, O.C. Celik, and K. Kinali. Fragility assessment of building struc-
tural systems in mid-america. Earthquake Engineering & Structural Dynamics, 36
(13):1935–1952, 2007.
[122] M.H. Faber and M.G. Stewart. Risk assessment for civil engineering facilities: crit-
ical overview and discussion. Reliability engineering & system safety, 80(2):173–184,
2003.
[124] M.A. Fardis and D. Veneziano. Estimation of SPT-N and relative density. Journal
of the Geotechnical Eengineering Division, 107(10):1345–1359, 1981.
[125] G.A. Fenton. Random field modeling of CPT data. Journal of Geotechnical and Geoen-
vironmental Engineering, 125(6):486, 1999.
[126] G.A. Fenton. Estimation for stochastic soil models. Journal of Geotechnical and
Geoenvironmental Engineering, 125(6):470, 1999.
[127] G.A. Fenton and D.V. Griffiths. Risk assessment in geotechnical engineering. John
[email protected]
Wiley & Sons, 2008.
169
Bibliography
[128] S. Ferson. Constructing probability boxes and Dempster-Shafer structures. Sandia Na-
tional Laboratories, 2002.
[130] R.V. Field and M. Grigoriu. On the accuracy of the polynomial chaos approxima-
tion. Probabilistic Engineering Mechanics, 19(1):65–80, 2004.
[131] B. Fiessler, H.-J. Neumann, and R. Rackwitz. Quadratic limit states in structural
reliability. J. Engrg. Mech. Div., ASCE, 105(4):661–676, 1979.
[132] G.E. Fogg, F.J. Lucia, and R.K. Senger. Stochastic simulation of interwell-scale
heterogeneity for improved prediction of sweep efficiency in a carbonate reservoir.
In NIPER/DOE Second International Reservoir Characterization Technical Conference,
Dallas, TX, pages 355–381, 1989.
[135] L.W. Gelhar, C. Welty, and K.R. Rehfeldt. A critical review of data on field-scale
dispersion in aquifers. Water Resources Research, 28(7):1955–1974, 1992.
[136] A. Gelman, J.B. Carlin, H.S. Stern, and D.B. Rubin. Bayesian data analysis. CRC
press, 2004.
[137] M.G. Genton. Highly robust variogram estimation. Mathematical Geology, 30(2):
213–221, 1998.
[138] R.G. Ghanem and P.D. Spanos. Stochastic finite elements: a spectral approach.
Springer, 1991.
[140] D.I. Goggin. Patterns of Permeability in Eolian Deposits: Page Sandstone (Juras-
sic), Northeastern Arizona, SPE, U. of Texas ll. A. Chandler, U. of Texas. SPE reprint
series, 2(27):46, 1988.
[141] D.J. Goggin, M.A. Chandler, G. Kocurek, and L. Lake. Permeability transects of eo-
lian sands and their use in generating random permeability fields. SPE Formation
[email protected]
Evaluation, 7(1):7–16, 1992.
170
Bibliography
[142] J. Goldsworthy. Reducing the Risk of Foundation Failures by Improving the Effective-
ness of Geotechnical Investigations. PhD thesis, School of Civili and environmental
Engineering, University of Adelaide, 2006.
[143] J.J. Gómez-Hernández and X.H. Wen. Multigaussian models: The danger of par-
simony. Statistical Methods & Applications, 4(2):167–181, 1995.
[145] P. Goovaerts. Geostatistics for natural resources evaluation. Oxford University Press,
USA, 1997.
[146] P.A. Graniero and V.B. Robinson. A real–time adaptive sampling method for field
mapping in patchy, heterogeneous environments. Transactions in GIS, 7(1):31–53,
2003.
[147] D.V. Griffiths and P.A. Lane. Slope stability analysis by finite elements. Geotech-
nique, 49(3):387–403, 1999.
[149] R. Hack, B. Orlic, S. Ozmutlu, S. Zhu, and N. Rengers. Three and more dimen-
sional modelling in geo-engineering. Bulletin of Engineering Geology and the Envi-
ronment, 65(2):143–153, 2006.
[150] I.S. Halim and W.H. Tang. Bayesian method for characterization of geological
anomaly. In Proceedings of Uncertainty Modeling and Analysis, pages 585–590. IEEE,
1990.
[152] A.M. Hasofer and N.C. Lind. Exact and invariant second-moment code format.
Journal of the Engineering Mechanics Division, 100(1):111–121, 1974.
[153] T. Haukaas and A. Der Kiureghian. Strategies for finding the design point in non-
linear finite element reliability analysis. Probabilistic Engineering Mechanics, 21(2):
133 – 147, 2006. ISSN 0266-8920.
[154] Y.A. Hegazy, P.W. Mayne, and S. Rouhani. Geostatistical assessement of spatial
variability in piezocone tests. In Proceedings of Uncertainty in the geological environ-
ment: from theory to practice, New York, 1996. ASCE.
171
Bibliography
[157] M.A. Hicks and K. Samy. Influence of anisotropic spatial variability on slope re-
liability. In Proceedings of the 8th International Symposium on Numerical Models in
Geomechanics, Rome, Italy, pages 535–9, 2002.
[158] M.A. Hicks and W.A. Spencer. Influence of heterogeneity on the reliability and
failure of a long 3D slope. Computers and Geotechnics, 37(7):948–955, 2010.
[160] K. Hoeg and W.H. Tang. Probabilistic considerations in the foundation engineer-
ing for offshore structures. In Proceedings of 2nd International Conference on Struc-
tural Safety and Reliability, ICOSSAR, Munich, pages 267–269, 1977.
[161] R.J. Hoeksema and P.K. Kitanidis. Analysis of the spatial structure of properties of
selected aquifers. Water Resources Research, 21(4):563–572, 1985.
[162] J.A. Hoeting, D. Madigan, A.E. Raftery, and C. T. Volinsky. Bayesian Model Aver-
aging: a tutorial. Statistical Science, 14(14):382–417, 1999.
[163] W. Hof. Zum Begriff Sicherheit. Beton- und Stahlbetonbau, 86(12):286–289, 1991.
[167] Y. Honjo and K. Kuroda. A new look at fluctuating geotechnical data for reliability
design. Soils and Foundations, 31(1):110–120, 1991.
[168] Y. Honjo, T. Hara, M. Suzuki, M. Shirato, T.C. Le Kieu, and Y. Kikuchi. Code
calibration in reliability based design level i verification format for geotechnical
structures. In Proceedings of 2nd International Symposium on Geotechnical Safety and
[email protected]
Risk, pages 435–452, 2009.
172
Bibliography
[170] L. Houy, D. Breysse, and A. Dents. Influence of soil heterogeneity on load redis-
tribution and settlement of a hyperstatic three-support frame. Geotechnique, 55(2):
163–170, 2005.
[171] L.Y. Hu, G. Blanc, and B. Noetinger. Gradual deformation and iterative calibration
of sequential stochastic simulations. Mathematical Geology, 33(4):475–489, 2001.
[173] M. Huber, P.A. Vermeer, and A. Bardossy. Untersuchungen zur Variabilität des
Untergrundes. In H. Schad and C. Vogt, editors, Proceedings of Bauen in Boden und
Fels, 2009.
[174] M. Huber, M.A. Hicks, P.A. Vermeer, and C. Moormann. Probabilistic calcula-
tion of differential settlement due to tunnelling. In L. Gucma, P. van Gelder, and
D. Proske, editors, Proceedings of 8th International Probabilistic Symposium, 2010.
[176] M. Huber, M.R.S. Sadaghiani, C. Moormann, and K.J. Witt. Beitrag zur räumlichen
Beschreibung der Heterogenität des Untergrundes. In J. Tiedemann, editor, Pro-
ceedings of 18. Tagung für Ingenieurgeologie, 2011.
[177] M. Huber, P.A. Vermeer, M.A. Hicks, and C. Moormann. Reliability based design
in tunnelling. In Ch. Brenguer, editor, Proceedings of the European Safety and Relia-
bility Conference: Advances in Safety, Reliability and Risk Management, Troyes, France,
pages 308–317, 2011.
[178] M. Huber, P.A. Vermeer, and C. Moormann. Evaluation and implications of soil
variability in tunneling. In M.H. Faber, editor, Proceedings of Applications of Statistics
and Probability in Civil Engineering, Proceedings of the 11th International Conference
on Applications of Statistics and Probability in Civil Engineering (ICASP11), Zürich,
Switzerland, pages 2785–2791, 2011.
[180] J. Iqbal, J.A. Thomasson, J.N. Jenkins, P.R. Owens, and F.D. Whisler. Spatial vari-
ability analysis of soil physical properties of alluvial soils. Soil Science Society of
[email protected]
America Journal, 69:1338–1350, 2005.
173
Bibliography
[181] E.H. Isaaks. Indicator simulation: application to the simulation of a high grade
uranium mineralization. In Geostatistics for Natural resources characterization; P II.
D. Reidel, Dordrecht, pages 1057–1069, 1984.
[182] E.H. Isaaks and R.M. Srivastava. Applied geostatistics, volume 2. Oxford University
Press New York, 1989.
[183] ISO 2394. General principles on the reliability of structures, 1998.
[184] M. B. Jacksa. The influence of spatial variability on the geotechnical design properties of
a stiff, overconsolidated clay. PhD thesis, The University of Adelaide, 1995.
[185] M.B. Jaksa, P.I. Brooker, and W.S. Kaggwa. Modelling the spatial variability of
the undrained shear strength of clay soils using geostatistics. In Proceedings of 5th
International Geostatistics Congress, pages 1284–1295, 1997.
[186] M.B. Jaksa, W.S. Kaggwa, and P.I. Brooker. Experimental evaluation of the scale of
fluctuation of a stiff clay. In Proceedings of 8th International Confernece on the Appli-
cation of Statistics and Probability, volume 1, pages 415–422. Sydney, AA Balkema,
Rotterdam, 1999.
[187] M.B. Jaksa, W.S. Kaggwa, G.A. Fenton, and H.G. Poulos. A framework for quanti-
fying the reliability of geotechnical investigations. In Proceedings of 9th International
Conference on the Application of Statistics and Probability in Civil Engineering, pages
1285–1291, 2003.
[188] A.G. Journel. Resampling from stochastic simulations. Environmental and Ecological
Statistics, 1(1):63–91, 1994.
[189] A.G. Journel and F. Alabert. Non-gaussian data expansion in the earth sciences.
Terra Nova, 1(2):123–134, 1989.
[190] A.G. Journel and C.J. Huijbregts. Mining geostatics. Academic Press, London, 1978.
[191] H. Karadeniz and T. Vrouwenvelder. SAFERELNET : Task 5.1: Overview reliability
methods. Technical report, TU Delft, The Netherlands, October 2003.
[192] J.M. Keaveny, F. Nadim, and S. Lacasse. Autocorrelation functions for offshore
geotechnical data. In Proceedings of the 5th International Conference on Structural
Safety and Reliability (ICOSSAR), pages 263–270, 1994.
[193] R. Kerry and M. A. Oliver. Determining the effect of asymmetric data on the vari-
ogram II - Outliers. Computers & Geosciences, 33(10):1233–1260, 2007.
[194] R. Kerry and M.A. Oliver. Determining the effect of asymmetric data on the vari-
ogram I - Underlying asymmetry. Computers & Geosciences, 33(10):1212–1232, 2007.
[195] L. Kirkup and R.B. Frenkel. An Introduction to Uncertainty in Measurement: Using the
GUM (Guide to the Expression of Uncertainty in Measurement). Cambridge University
[email protected]
Press, 2006.
174
Bibliography
[196] A. Kirsch. On the face stability of shallow tunnels in sand. PhD thesis, Universtität
Innsbruck, 2009.
[200] P.K. Kitanidis and R.W. Lane. Maximum likelihood parameter estimation of hy-
drologic spatial processes by the Gauss-Newton method. Journal of hydrology, 79
(1-2):53–71, 1985.
[203] M. Kleiber and T.D. Hien. The Stochastic Finite Element Method. John Wiley & Sons
Inc, 1992.
[206] A.S.D.A.C.M. Kormann and J.H.D. Shuttle. Variability Analysis of SPT and CPT
data for a Reliability-based Embankment Design of a Southern Brazilian Port Site.
In Proceedings of XIII Brazilian Conference on Soil Mechanics and Geotechnical Engi-
neering. Curitiba, Brazil, 2006.
[209] P.H.S.W. Kulatilake and J.G. Um. Spatial variation of cone tip resistance for the
clay site at texas university. Geotechnical and Geological Engineering, 21(2):149–165,
[email protected]
2003.
175
Bibliography
[210] H. Kupfersberger and C.V. Deutsch. Methodology for integrating analog geologic
data in 3-d variogram modeling. AAPG bulletin, 83(8):1262–1278, 1999.
[211] S. Lacasse and J.Y. de Lamballerie. Statistical treatment of CPT data. In Proceedings
of the International Symposium on Cone Penetration Testing, pages 4–5, 1995.
[212] S.M. Lacasse, C.C. Ladd, and A.K. Barsvary. Undrained behavior of embankments
on new liskeard varved clay. Canadian Geotechnical Journal, 14(3):367–388, 1977.
[213] P.V. Lade and J.M. Duncan. Elastoplastic stress-strain theory for cohesionless soil.
Journal of the Geotechnical Engineering Division, 101(10):1037–1053, 1975.
[214] R.M. Lark. Estimating variograms of soil properties by the method-of-moments
and maximum likelihood. European Journal of Soil Science, 51(4):717–728, 2000.
[215] G. Le Loc’h, H. Beucher, A. Galli, B. Doligez, HERESIM Group, et al. Improve-
ment in the truncated gaussian method: Combining several gaussian functions.
In ECMOR IV, Proceedings of the 4th European Conference on the Mathematics of Oil
Recovery, R oslash ros, Norway, 1994.
[216] E. Leca and L. Dormieux. Upper and lower bound solutions for the face stability
of shallow circular tunnels in frictional material. Geotechnique, 40(4):581–606, 1990.
[217] M. Lemaire, A. Chateauneuf, and J.C. Mitteau. Structural reliability. Wiley Online
Library, 2009.
[218] A. Lentz. Acceptability of civil engineering decisions involving human consequences.
PhD thesis, Technischen Universität München, München, 2007.
[219] C.C. Li and A. der Kiureghian. Optimal discretization of random fields. Journal of
Engineering Mechanics, 119:1136, 1993.
[220] J. Li, A. Bárdossy, L. Guenni, and M. Liu. A copula based observation network
design approach. Environmental Modelling & Software, 26(11):1349–1357, 2011.
[221] K.S. Li and I.K. Lee. The assessment of geotechnical safety. In Selected Topics in
Geotechnical Engineering-Lumb Volume, 1991.
[222] C. N. Liu and C.-H. Chen. Spatial correlation structures of CPT data in a liquefac-
tion site. Engineering Geology, 111(1-4):43 – 50, 2010.
[223] C.N. Liu and C.H. Chen. Estimating spatial correlation structures based on CPT
data. Georisk: Assessment and Management of Risk for Engineered Systems and Geohaz-
ards, 4(2):99–108, 2010.
[224] K. Loague and G.A. Gander. R-5 revisited: Spatial variability of infiltration on a
small rangeland catchment. Water Resources Research, 26(5):957–971, 1990.
[225] N. Luhmann. Die Moral des Risikos und das Risiko der Moral. In Proceedings of
[email protected]
Risiko und Gesellschaft, pages 327–338. Westdeutscher Verlag, Opladen, 1997.
176
Bibliography
[226] P. Lumb. Safety factors and the probability distribution of soil strength. Canadian
Geotechnical Journal, 7(3):225–242, 1970.
[227] P. Lumb. Application of statistics in soil mechanics. In Soil Mechanics: New Hori-
zons, pages 44–112, 1974.
[228] P. Lumb. Spatial variability of soil properties. In Proceedings of the 2nd International
Conference on Application of Statistics and Probability to Soil and Structural Engineering,
Aachen, volume 2, pages 397–421, 1975.
[229] P. Lumb and J.K. Holt. The undrained shear strength of a soft marine clay from
Hong Kong. Geotechnique, 18(1):25–36, 1968.
[230] H.J. Lund, H. Ates, E. Kasap, and R.W. Tillman. Comparison of Single and Multi-
Facies Variograms of Newcastle Sandstone: Measures for the Distribution of Bar-
riers to Flow. In Low Permeability Reservoirs Symposium, pages 507–522, 1995.
[231] R.J. Luxmoore, B.P. Spalding, and I.M. Munro. Areal variation and chemical modi-
fication of weathered shale infiltration characteristics. Soil Science Society of America
Journal, 45(4):687, 1981.
[232] A.V. Lyamin, R. Salgado, S.W. Sloan, and M. Prezzi. Two-and three-dimensional
bearing capacity of footings in sand. Géotechnique, 57(8):647–662, 2007.
[233] R.J. Mair and R.N. Taylor. Theme lecture: Bored tunneling in the urban envi-
ronment. In Proceedings of the 14th International Conference on Soil Mechanics and
Foundation Engineering, pages 2353–2385, 1997.
[234] H.B. Mann. Nonparametric tests against trend. Econometrica: Journal of the Econo-
metric Society, 2:245–259, 1945.
[236] B.P. Marchant and R.M. Lark. Estimating variogram uncertainty. Mathematical
Geology, 36(8):867–898, 2004.
[237] B.P. Marchant and R.M. Lark. Adaptive sampling and reconnaissance surveys for
geostatistical mapping of the soil. European Journal of Soil Science, 57(6):831–845,
2006.
[238] F. Marconi. Ricostruzione con metodi geostatistici di corpi alluvionali sepolti per la valu-
tazione delle pericolosita geologiche in aree urbane. PhD thesis, Universita degli studi
di Roma "Sapienza", Facolta di Ingegneria, Dipartimento Ingegneria Chimica Ma-
[email protected]
teriali Ambiente, 2013.
177
Bibliography
[239] G. Matheron. The theory of regionalized variables and its applications, volume 5. École
national supérieure des mines, 1971.
[244] H.G. Matthies and A. Keese. Galerkin methods for linear and nonlinear elliptic
stochastic partial differential equations. Computer Methods in Applied Mechanics
and Engineering, 194(12):1295–1331, 2005.
[247] A.B. McBratney and R. Webster. Choosing functions for semi-variograms of soil
properties and fitting them to sampling estimates. European Journal of Soil Science,
37(4):617–639, 1986.
[248] A.B. McBratney, R. Webster, and T.M. Burgess. The design of optimal sampling
schemes for local estimation and mapping of of regionalized variables: Theory
and method. Computers & Geosciences, 7(4):331–334, 1981.
[249] M.D. McKay. Evaluating prediction uncertainty. Technical report, Office of Nu-
clear Regulatory Research. Division of Systems Technology and Los Alamos Na-
tional Laboratory, 1995.
[250] S.A. McKenna. Probabilistic Approach to Site Characterization: MIU Site, Tono
Region, Japan. Technical report, Sandia National Labs., Albuquerque, NM (US);
Sandia National Labs., Livermore, CA (US), 2001.
[252] B. Möller, M. Beer, W. Graf, A. Hoffmann, and J.-U. Sickert. Modellierung von
[email protected]
Unschärfe im Ingenieurbau. Bauinformatik Journal, 3(11):697–708, 2000.
178
Bibliography
[253] S.C. Möller. Tunnel induced settlements and structural forces in linings. PhD thesis,
Institute of Geotechnical Engineering, University of Stuttgart, 2006.
[255] G. Mollon, D. Dias, and A.H. Soubra. Probabilistic analysis of circular tunnels in
homogeneous soil using response surface methodology. Journal of Geotechnical and
Geoenvironmental Engineering, 135:1314–1325, 2009.
[256] G. Mollon, D. Dias, and A.H. Soubra. Face stability analysis of circular tunnels
driven by a pressurized shield. Journal of Geotechnical and Geoenvironmental Engi-
neering, 136:215–229, 2010.
[257] Monitor Environmental Consultants Ltd. Secondary model procedure for the de-
velopment of appropriate soil sampling strategies for land contamination, 2012.
URL www.environment-agency.gov.uk.
[258] R.L. Muhanna, H. Zhang, and R.L. Mullen. Interval finite elements as a basis for
generalized models of uncertainty in engineering mechanics. Reliable Computing,
13(2):173–194, 2007.
[259] D.J. Mulla. Estimating spatial patterns in water content, matric suction, and hy-
draulic conductivity. Soil Sciences Society of America Journal, 52(6):1547–1553, 1988.
[260] D.S. Murty and G.L.S. Babu. Effect of spatial correlation structure and transforma-
tion model on the design of shallow strip footing. Geomechanics and Geoengineering,
5(2):91–97, 2010.
[261] J. Murzewski. Sicherheit der Baukonstruktionen. VEB Verlag für Bauwesen, Berlin,
1974.
[263] H.D. Netzel. Building response due to ground movements. PhD thesis, Delft University
of Technology, 2011.
[265] V.U. Nguyen and R.N. Chowdhury. Simulation for risk analysis with correlated
variables. Geotechnique, 35(1):47–58, 1985.
[266] J.D. Nuttall. Parallel Implementation and Application of the Random Finite Element
[email protected]
Method. PhD thesis, The University of Manchester, Manchester, UK, 2011.
179
Bibliography
[267] M. P. O’Reilly and B. M. New. Settlements above tunnels in the UK-their magni-
tude and prediction. In Proceedings of Tunnelling, 82, pages 173–181, 1982.
[268] J. Ortiz and C.V. Deutsch. Calculation of uncertainty in the variogram. Mathemat-
ical Geology, 34(2):169–183, 2002.
[269] J. Ouellet, D.E. Gill, and M. Soulié. Geostatistical approach to the study of induced
damage around underground rock excavations. Canadian Geotechnical Journal, 24
(3):384–391, 1987.
[273] R.L. Parsons and J.D. Frost. Evaluating site investigation quality using GIS and
geostatistics. Journal of Geotechnical and Geoenvironmental Engineering, 128(6):451–
461, 2002.
[274] R.B. Peck. Deep excavations and tunnelling in soft ground. In 7th Int. Conf. on Soil
Mech. and Found. Engrg., Mexico, pages 225–290. Sociedad Mexican de Mecanica de
Suelos, 1969.
[275] Pacific Earthquake Engineering Research Center (PEER). Strong motion database,
06 2012. URL http://peer.berkeley.edu.
[277] E.W. Perau. Ein systematischer Ansatz zur Berechnung des Grundbruchwiderstands von
Fundamenten. PhD thesis, Universität Gesamthochschule Essen, 1995.
[279] G.M. Peschl. Reliability Analysis in Geotechnics with the Random Set Element Method.
PhD thesis, TU Graz, Institute of Soil Mechanics, 2004.
[280] G.M. Peschl. Reliability analyses in geotechnics with the random set finite element
[email protected]
method. PhD thesis, TU Graz, Institute of Soil Mechanics, 2011.
180
Bibliography
[281] A.N. Pettitt and A.B. McBratney. Sampling designs for estimating spatial variance
components. Journal of the Royal Statistical Society. Series C (Applied Statistics), 42(1):
185–209, 1993.
[282] K.-K. Phoon and F.H. Kulhawy. Evaluation of geotechnical property variabaility.
Canadian Geotechnical Journal, 36:625–639, 1999.
[283] K.-K. Phoon and F.H. Kulhawy. Characterization of geotechnical variability. Cana-
dian Geotechnical Journal, 36:612–624, 1999.
[284] K.-K Phoon, S.-T. Quek, and P. An. Identification of statistically homogeneous oil
layers using modified Bartlett stastistics. Journal of Geotechnical and Geoenvironmen-
tal Engieneering, 120(7):649–659, 2003.
[285] K.K. Phoon. Reliability-based design of foundations for transmission line structures.
PhD thesis, Cornell University, 1995.
[287] J.S. Piersol. Random data analysis and measurement procedures. Measurement
Science and Technology, 11:1825, 2000.
[288] R. Popescu. Stochastic variability of soil properties: data analysis, digital simulation,
effects on system behavior. PhD thesis, Princeton University, 1995.
[290] R. Pöttler and H.F. Schweiger. Anwendung der Random-Set-Methode (RSM) für
Standsicherheitsuntersuchungen im Tunnelbau. Bautechnik, 83(11):754–759, 2006.
[291] D. Potts, K. Axelsson, L. Grande, H.F. Schweiger, and M. Long. Guidelines for the
use of advanced numerical analysis. Thomas Telford Services Limited, 2002.
[292] D.M. Potts and L. Zdravković. Finite element analysis in geotechnical engineering:
application, volume 2. Thomas Telford Services Limited, 2001.
[293] L. Prandtl. Über die Härte plastischer Körper. Göttingen Nachr. Math. Phys. Kl, 12:
74–85, 1920.
[294] W.H. Press, S.A. Teukolsky, W.T. Vettrling, and B.P. Flannery. Numerical Recipes in
FORTRAN - The Art of Scientific Computing, volume second edition. Cambridge
University Press, second edition edition, 1992.
[295] D. Proske. Catalogue of Risks - Natural, Technical, Social and Health Risks. Springer-
[email protected]
Verlag Berlin, 2008.
181
Bibliography
[296] D. Proske. Debris flow impact uncertainty modeling with Grey Numbers. In
L. Gucma, P. van Gelder, and D. Proske, editors, Proceedings of 8th International
Probabilistic Symposium, pages 285–294, 2010.
[298] Z.P. Qiu and I. Elishakoff. Antioptimization of structures with large uncertain-but-
non-random parameters via interval analysis. Computer methods in Applied Mechan-
ics and Engineering, 152(3):361–372, 1998.
[299] R. Rackwitz. Optimization and risk acceptability based on the life quality index.
Structural Safety, 24(2-4):297–331, 2002.
[303] K.R. Rehfeldt, J.M. Boggs, and L.W. Gelhar. Field study of dispersion in a heteroge-
neous aquifer 3. geostatistical analysis of hydraulic conductivity. Water Resources
Research, 28(12):3309–3324, 1992.
[304] N. Remy, A. Boucher, and J. Wu. Applied geostatistics with SGeMS: A user’s guide.
Cambridge University Press, 2008.
[305] N. Remy, Alexandre. Boucher, and Jianbing. Wu. Applied geostatistics with SGeMS.
Cambridge University Press, 2009.
[308] P.K. Robertson. Soil classification using the cone penetration test. Canadian Geotech-
nical Journal, 27(1):151–158, 1990.
[309] P.J. Roebber. Visualizing multiple measures of forecast quality. Weather and Fore-
[email protected]
casting, 24(2):601–608, 2009.
182
Bibliography
[311] N.M. Ruse. Räumliche Betrachtung der Standsicherheit der Ortsbrust beim Tunnelvor-
trieb. PhD thesis, Institute of Geotechnical Engineering, University of Stuttgart,
2004.
[312] C. Russelli. Probabilistic methods applied to the bearing capacity problem. PhD thesis,
Institute of Geotechnical Engineering, University of Stuttgart, 2008.
[314] D. Russo. Design of an optimal sampling network for estimating the variogram.
Soil Science Society of America Journal, 48(4):708–716, 1984.
[317] A. Saltelli. Sensitivity analysis in practice: a guide to assessing scientific models. John
Wiley & Sons Inc, 2004.
[318] A. Saltelli and I.M. Sobol. About the use of rank transformation in sensitivity
analysis of model output. Reliability Engineering & System Safety, 50(3):225–239,
1995.
[320] P. Samui and T.G. Sitharam. Site characterization model using artificial neural
network and kriging. International Journal of Geomechanics, 10:171–180, 2010.
183
Bibliography
[324] B. Schmidt. Settlements and ground movements associated with tunneling in soil. PhD
thesis, University of Illinois., 1969.
[327] G.I. Schueller, H.J. Pradlwarter, and P.S. Koutsourelakis. A comparative study of
reliability estimation procedures for high dimensions. In Proceedings of 16th ASCE
Engineering Mechanics Conference, pages 233 – –243, 2003.
[328] M.T. Schultz, B.P. Gouldby, J.D. Simm, and J.L. Wibowo. Beyond the factor of
safety: developing fragility curves to characterize system reliability. Technical re-
port, DTIC Document, 2010.
[329] B. Schuppener. Eurocode 7 geotechnical design. part 1: General rules and its latest
developments. Georisk, 4(1):32–42, 2010.
[331] T. Schweckendiek and E.O.F. Calle. A factor of safety for geotechnical charac-
terization. In Proceedings of Seventeenth Southeast Asian Geotechnical Conference
(17SEAGC)–Geo-Engineering for Natural Hazard Mitigation and Sustainable Develop-
ment, volume 2, pages 227–230, 2010.
[332] H.F. Schweiger and G.M. Peschl. Reliability analysis in geotechnics with the ran-
dom set finite element method. Computers and Geotechnics, 32(6):422–435, 2005.
[333] H.F. Schweiger, R. Thurner, and R. Pöttler. Reliability analysis in geotechnics with
deterministic finite elements. International Journal of Geomechanics, 1(4):389–413,
2001.
[334] H.F. Schweiger, G.M. Peschl, and R. Pöttler. Application of the random set finite
element method for analysing tunnel excavation. Georisk, 1(1):43–56, 2007.
[335] R.K. Senger, F.J. Lucia, C. Kerans, M. Ferris, and C. E. Fogg. Geostatistical/Geo-
logical Permeability Characterization of Carbonate Ramp Deposits in Sand Andres
Outcrop, Algerita Escarpment, New Mexico . In Permian Basin Oil and Gas Recovery
Conference, Midland, Texas, pages 287–301. Society of Petroleum Engineers, 1992.
[336] K. Sett, B. Jeremić, and M.L. Kavvas. Probabilistic elasto-plasticity: solution and
verification in 1D. Acta Geotechnica, 2(3):211–220, 2007.
[337] M.A. Shahin, M.B. Jaksa, and H.R. Maier. Artificial neural network applications in
[email protected]
geotechnical engineering. Australian Geomechanics, 36(1):49–62, 2001.
184
Bibliography
[338] J.E. Shigley and C.R. Mischke. Mechanical Engineering Design. McGraw Hill, New
York, 2001.
[339] T.W. Simpson, T.M. Mauery, J.J. Korte, and F. Mistree. Kriging models for global
approximation in simulation-based multidisciplinary design optimization. AIAA
journal, 39(12):2233–2241, 2001.
[340] J.B. Sisson and P.J. Wierenga. Spatial variability of steady-state infiltration rates as
a stochastic process. Soil Science Society of America Journal, 45(4):699, 1981.
[341] T.G. Sitharam, P. Samui, and P. Anbazhagan. Spatial variability of rock depth in
bangalore using geostatistical, neural network and support vector machine mod-
els. Geotechnical and Geological Engineering, 26(5):503–517, 2008.
[342] I.M. Smith and D.V. Griffiths. Programming the finite element method. John Wiley &
Sons, Inc., 1998.
[343] J.L. Smith and R.A. Freeze. A stochastic analysis of steady-state groundwater flow in a
bounded domain. 1978., 1978.
[347] I.M. Sobol. Global sensitivity indices for nonlinear mathematical models and their
monte carlo estimates. Mathematics and computers in simulation, 55(1-3):271–280,
2001.
[348] V.V. Sokolovski and J.K. Kushner. Statics of granular media. Journal of Applied
Mechanics, 33:239, 1966.
[349] P. Soos. Die Rolle des Baugrundes bei der Anwendung der neuen Sicherheitsthe-
orie im Grundbau. Geotechnik, 19:82–91, 1990.
[350] A.H. Soubra. Kinematical approach to the face stability analysis of shallow circular
tunnels. In Proceedings of 8th International Symposium on Plasticity, pages 443–445,
2000.
[351] A.H. Soubra, D.S. Youssef Abdel Massih, and K. Kalfa. Bearing capacity of foun-
dations resting on a spatially random soil. In Geocongress008, 2008.
185
Bibliography
[353] M. Soulié, P. Montes, and V. Silvestri. Modelling spatial variability of soil parame-
ters. Canadian Geotechnical Journal, 27(5):617–630, 1990.
[354] M.G. Speedie. Selection of design value from shear test results. In Proceedings of the
2nd Australia-New Zealand Conference on Soil Mechanics and Foundation Engineering,
pages 107–109, Wellington, 1965.
[355] W.A. Spencer. Parallel stochastic and finite element modelling of clay slope stability in
3D. PhD thesis, University of Manchester, UK, 2007.
[356] M.J. Spry, F.H. Kulhawy, and M.D. Grigoriu. Reliability-based foundation design
for transmission line structures: Geotechnical site characterization strategy. Tech-
nical Report EL-5507, Electric power research institute, Palo Alto, California, 1988.
[358] G. Stefanou. The stochastic finite element method: Past, present and future. Com-
puter Methods in Applied Mechanics and Engineering, 198:1031 – 1051, 2009. ISSN
0045-7825.
[359] A. Stein and J.W. van Groenigen. Constrained optimization of spatial sampling us-
ing continuous simulated annealing. Journal of Environmental Quality, 27(5):1078–
1086, 1998.
[361] D. Straub, A. Lentz, I. Papaioannou, and R. Rackwitz. Life quality index for as-
sessing risk acceptance in geotechnical engineering. In Proceedings of the 3rd Inter-
national Symposium on Geotechnical Safety and Risk, pages 37–46, 2011.
[363] E.A. Sudicky. A natural gradient experiment on solute transport in a sand aquifer:
Spatial variability of hydraulic conductivity and its role in the dispersion process.
Water Resources Research, 22(13):2069–2082, 1986.
[365] B. Sudret. Global sensitivity analysis using polynomial chaos expansions. Reliabil-
[email protected]
ity Engineering & System Safety, 93(7):964–979, 2008.
186
Bibliography
[366] B. Sudret and A. Der Kiureghian. Stochastic finite element methods and reliability:
A state-of-the-art report. Technical report, Department of Civil & Cnvironmental
Cngineering, University of California, Berkeley, 2000.
[367] B. Sudret and A. Der Kiureghian. Comparison of finite element reliability methods.
Probabilistic Engineering Mechanics, 17(4):337–348, 2002.
[370] W.H. Tang. Updating anomaly statistics-single anomaly case. Structural safety, 4
(2):151–163, 1986.
[371] J.M. Tantalla, J.H. Prevost, and G. Deodatis. Spatial variability of soil properties
in slope stability analysis: fragility curve generation. In Proceedings of ICOSSAR,
volume 1, pages 17–21, 2001.
[372] D.W. Taylor. Fundamentals of soil mechanics. Soil Science, 66(2):161, 1948.
[373] K. Terzaghi and R.B. Peck. Soil mechanics in engineering practice. Wiley, 1948.
[379] K. Unlu, D.R. Nielsen, and W. Biggar. Stochastic analysis of state unsaturated flow.
one dimensional monte carlo simulations and comparisons with spectral pertur-
bation analysis and field observations. Water Resources Research, 26(1):2207–2218,
[email protected]
1990.
187
Bibliography
[380] U.S. Geological Survey (USGS). Earthquake hazards program, 06 2012. URL
http://earthquake.usgs.gov/.
[381] M. Uzielli, G. Vannucchi, and K.K. Phoon. Random field characterisation of stress-
normalised cone penetration testing parameters. Geotechnique, 55(1):3–20, 2005.
[382] P.H.A.J.M. van Gelder. Stasticial methods fro the risk-based design of civil structures.
PhD thesis, Delft University of Technology, 2000.
[383] E.H. Vanmarcke. Reliability of earth slopes. Journal of Geotechnical Engineering, 103:
1247–1265, 1977.
[385] E.H. Vanmarcke. Random fields: analysis and synthesis. MIT Press, 1984.
[386] E.H. Vanmarcke and N.F. Fuleihan. Probabilistic prediction of levee settlements.
In Proceedings of the 2nd International Conference on Applications of Statistics and Prob-
ability in Soil and Structural Engineering, Aachen, volume 2, pages 175–190, 1975.
[387] E.H. Vanmarcke and M. Grigoriu. Stochastic finite element analysis of simple
beams. Journal of Engineering Mechanics, 109:1203, 1983.
[388] V.N. Vapnik. The nature of statistical learning theory. Springer-Verlag New York Inc.,
2000.
[389] Det Norske Veritas. Structural reliability analysis of marine structures. D.N.V., 1992.
[390] Det Norske Veritas. Statistical representation of soil data. Technical report, D.N.V.,
1997.
[392] P.A. Vermeer, N. Ruse, and T. Marcher. Tunnel heading stability in drained ground.
Felsbau, 20(6):8–18, 2002.
[393] A. Verruijt and J.R. Booker. Surface settlements due to deformation of a tunnel in
an elastic half plane. Geotechnique, 46(4):753–756, 1996.
[395] M.F. Versteeg. External safety policy in the netherlands: an approach to risk man-
[email protected]
agement. Journal of hazardous materials, 17(2):215–222, 1988.
188
Bibliography
[397] J.P. Visodic. Design Stress Factors, volume 55. ASME International, New York, 1948.
[399] T. Vrouwenvelder. The JCSS probabilistic model code. Structural Safety, 19(3):245–
251, 1997.
[400] T. Vrouwenvelder and E.O.F. Calle. Measuring spatial correlation of soil proper-
ties. HERON, 48(4):2003, 2003.
[402] R. Webster and M.A. Oliver. Sample adequately to estimate variograms of soil
properties. Journal of soil science, 43(1):177–192, 1992.
[403] R. Webster and M.A. Oliver. Geostatistics for environmental scientists. John Wiley &
Sons Inc., 2007.
[405] D.S. Wickremesinghe. Stasticial characterisation of soil profile using in situ tests. PhD
thesis, University of British Columbia, 1989.
[407] T.F. Wolff. Analysis and design of embankment dam slopes: a probabilistic approach. PhD
thesis, Purdue University, Lafayette, Ind., 1985.
[408] S.Y. Wong. Stochastic Characterization and Reliability of Saturated Soils. PhD thesis,
Ph. D. Thesis, Manchester Centre for Civil and Construction Engineering, Univer-
sity of Manchester, England, 2004.
[409] A.D. Woodbury and EA Sudicky. The geostatistical characteristics of the Borden
aquifer. Water Resources Research, 27(4):533–546, 1991.
[410] WS Atkins Consultants Ltd . Risk implications in site characterisation and analysis
for offshore engineering and design. Technical report, Health and Safety Execu-
[email protected]
tive, 2004.
189
Bibliography
[411] T.H. Wu. Uncertainty, safety, and decision in soil engineering. Journal of the Geotech-
nical Engineering Division, 100(3):329–348, 1974.
[412] T.H. Wu and A. El-Jandali. Use of time series in geotechnical data analysis. Geotech.
Test. J, 8(4):151–158, 1985.
[413] J. R. Ximenez de Embun and M. R. Romana. The formulation of soil model from
penetrometer probabilistic information - application to differential settlement pre-
diction. In Proc. 4th Int. Conf. on Applications of Statistics and Probability in Soil and
Structural Engineering, pages 1601–1613, Pitagora Editrice, 1983.
[414] M.S. Yucemen, W.H. Tang, and A.H.S. Ang. A probabilistic study of safety and
design of earth slopes. Technical report, Engineering Experiment Station, College
of Engineering, University of Illinois at Urbana-Champaign, 1973.
[415] H. Zhang, R.L. Mullen, and R.L. Muhanna. Interval Monte Carlo methods for
structural reliability. Structural Safety, 32(3):183–190, 2010.
[416] J. Zhang, L.M. Zhang, and W.H. Tang. Bayesian framework for characterizing
geotechnical model uncertainty. Journal of geotechnical and geoenvironmental engi-
neering, 135:932, 2009.
[417] Z.J. Zhang and M.T. Tumay. Statistical to fuzzy approach toward CPT soil classi-
fication. Journal of Geotechnical and Geoenvironmental Engieneering, 125(3):179–186,
1999.
190
Appendix A
Database on stochastic soil properties
191
Table A.1: Database on the variability of rock
192
property soil type θver [m] θhor [m] applied spatial model reference
permeability outcrop of grainstone and dolomit 0.1 0.8 [201]
permeability sand and gravel aquifer 0.5 20 [179]
permeability sandstone aquifer 1 [28]
permeability subsurface of dolomitized packstone and sandstone 2.6 2.7 [201]
permeability eolian sandstone outcrop 3 8 [140]
permeability limestone aquifer 3,500 [91]
[email protected]
permeability limestone aquifer 6,300 [91]
permeability chalk 7,500 [91]
permeability sandstone aquifer 17,000 [91]
permeability sandstone aquifer 45,000 [161]
SPT soft rock, hard soils 4 spherical [202]
SPT soft rock, hard soils 5 spherical [202]
Appendix A Database on stochastic soil properties
A.2 Database on the variability of frictional soil
193
194
Table A.2: Database on the variability of frictional soils.
property soil type θver [m] θhor [m] applied spatial model reference
ϕ’ + c’ sand 5.1 242 [362]
buldk density medium gravel 0.2 exponential [375]
buldk density coarse sand, fine gravel 0.3 exponential [375]
buldk density medium gravel 0.4 exponential [375]
CPT sand 0 ∼2 exponential [10]
[email protected]
CPT alluvial deposits 0.1 ∼ 2.6 exponential [413]
CPT sand 0.1 ∼2 exponential [222]
CPT sand 0.3 ∼ 1.2 3.5 ∼ 15 Campanella [408]
CPT sand 0.3 LAS [405]
CPT sand 0.4 ∼ 2.4 LAS [27]
CPT sand 0.5 5.4 LAS [288]
CPT sand 1 12.1 LAS [288]
CPT clean sand 1 [208]
CPT soft glacial clay, North sea 1 [412]
Appendix A Database on stochastic soil properties
[email protected]
index property sand 0.2 ∼2 [148]
index property sandy clay 0.4 ∼ 1.1 1.2 ∼ 2.5 [349]
index property poorly sorted gravel 0.7 15 [307]
index property sand 0.9 exponential [109]
index property sand 1 ∼3 5 ∼ 15 [276]
permeability sand 3 25 exponential [303]
permeability fluvial sand 0.1 3 [62]
permeability braided river environment 0.3 10 [322]
permeability medium gravel 0.3 ∼ 0.9 2 ∼ 11.4 exponential [375]
permeability medium gravel 0.4 1.8 exponential [375]
permeability medium gravel 0.4 0.3 exponential [375]
permeability sand aquifer 0.5 21.5 [409]
permeability sandy gravel 0.6 1.2 [377]
permeability coarse sand, fine gravel 0.6 5 exponential [375]
permeability sandy soil 0.9 9 [335]
permeability fluvial sand 1.5 13 [302]
permeability sand 1.7 [344]
permeability sand 3.7 exponential [123]
permeability sand 3.7 exponential [124]
permeability strongly weathered rock 3.8 600 [132]
permeability sandstone 4.5 1.6 [141]
permeability heterogeneous alluvial aquifer 4.8 38.4 [303]
permeability sand 6 900 [94]
Continued on next page
195
A.2 Database on the variability of frictional soil
Table A.2 – continued from previous page
196
property soil type θver [m] θhor [m] applied spatial model reference
permeability sand 7.2 5.4 [357]
permeability sand 12 ∼ 16 40 [379]
permeability sand 12 1,200 [83]
permeability limestone 42 6,300 [35]
permeability sand » 40 [123]
permeability compacted clay 0.5 ∼2 [37]
permeability weathered shale subsoil 2 [231]
[email protected]
permeability Pleistocene Quadra sand 3.3 [344]
permeability sand 6 exponential [92]
permeability strony layer, medium gravel 9.5 exponential [375]
permeability sand and gravel aquifer 12.5 [112]
permeability mediterranean soil 20 [315]
permeability sand 40 ∼ 280 [159]
permeability and 67 exponential [123]
permeability sand 67 exponential [124]
permeability salt dome 1,500 [26]
Appendix A Database on stochastic soil properties
197
198
Table A.3: Database on the variability of cohesive soils.
property soil type θver [m] θhor [m] applied spatial model reference
CPT organic, soft clay 0.1 [206]
CPT organic clay 0.1 exp [109]
CPT clay 0.1 ∼ 0.7 exp + cos [68]
CPT clay, silty sand 0.1 ∼ 0.6 exp [381]
CPT deltatic soils 0.1 ∼ 0.4 [404]
[email protected]
CPT clay 0.2 ∼ 0.3 exp [209]
CPT organic, soft clay 0.2 [206]
CPT organic clay 0.2 exp [109]
CPT clay, silty sand 0.2 exp [381]
CPT clay 0.2 ∼ 0.7 exp [396]
CPT clay 0.2 ∼ 0.4 0.2 ∼ 0.7 [65]
CPT clay 0.2 ∼ 1.7 0.5 ∼ 3 [9]
CPT silty clay 0.2 exp [221]
CPT clay, silty sand 0.2 ∼ 0.5 exp [381]
Appendix A Database on stochastic soil properties
[email protected]
CPT varved clay 1 [90]
CPT clay 1.2 ∼ 5.6 exp [45]
CPT marine clay 1.5 exp [242]
CPT clay 2 5 exp [20]
CPT sea clay 2 40 [78]
CPT soft clay 2 40 [167]
CPT sensitive clay 2 [78]
CPT sensitive clay 2 [78]
CPT sensitive clay 2 [78]
CPT sensitive clay 2 [78]
CPT sandy clay 2.5 ∼ 4.5 20 ∼ 30 exp [207]
CPT clay 4 24 exp [16]
CPT varved clay 5 exp [383]
CPT alluvial clay 6.5 [167]
CPT clay 8.6 8.6 [411]
CPT varved clay 46 exp [212]
CPT marine clay 55 35 ∼ 60 exp [15]
CPT marine clay 6 exp [227]
CPT clay 0.2 [260]
CPT silty clay 0.2 ∼ 1.9 exp [223]
CPT sandy clay 25 [18]
CPT seabed deposits 30 ∼ 60 [160]
CPT seabed deposits 53 exp [368]
Continued on next page
199
A.3 Database on the variability of cohesive soil
Table A.3 – continued from previous page
200
property soil type θver [m] θhor [m] applied spatial model reference
CPT alluvial clay 200 [167]
CPT silty clay 212 ∼ 245 exp [264]
CPT clay 278 ∼ 286 exp [264]
CPT clay 0.6 ∼ 1.8 exp [186]
CPT silty clay 1 [211]
CPT sensitive clay 2 [78]
CPT clay 2 exp [310]
[email protected]
CPT clay 10 289 exp [185]
CPT deep glacial sands 20 ∼ 35 exp [400]
CPT clay 9.6 [211]
CPT North sea soil 13.9 [192]
CPT marine clay 400 exp [229]
CPT North sea clay 30 [368]
CPT North sea soil 37.5 [192]
CPT (friction ratio) clay 0 ∼ 0.6 exp [10]
CPT (friction ratio) clay to silty clay 0.1 exp [381]
Appendix A Database on stochastic soil properties
∼ 0.5
CPT (friction ratio) clay, silty sand 0.1 ∼ 0.5 exp [381]
CPT (friction ratio) clean sand, silty sand 0.2 ∼ 0.6 exp [381]
CPT (friction ratio) clay, silty sand 0.2 ∼ 0.3 exp [381]
CPT (friction ratio) claen sand, silty sand 0.3 ∼ 0.6 exp [381]
CPT (friction ratio) clay, silty sand 0.4 exp [381]
CPT (friction ratio) silty sand to sandy silt 0.4 exp [381]
CPT (friction ratio) claen sand,silty sand 0.5 ∼ 0.6 exp [381]
CPT (friction ratio) silty sand to sandy silt 0.6 exp [381]
CU (triaxial) offshore soil 1 [192]
CU (triaxial) offshore soil 3 [192]
density compacted clay 5 [25]
density compacted clay 4 ∼5 [25]
Continued on next page
Table A.3 – continued from previous page
property soil type θver [m] θhor [m] applied spatial model reference
density clay 10 ∼ 15 [9]
Index Properties marine clay 1.6 exp [227]
liquid limit soft silty loam 0 ∼ 3.6 18 ∼ 22 exp [10]
liquid limit clay »0 [227]
liquid limit clay »0 [228]
[email protected]
modulus of elasticity clay 2 ∼5 [186]
modulus of elasticity clay 6 298.5 spherical [185]
modulus of elasticity clay 400 [235]
permeability glacial Lacustirne sand 0.1 3 [363]
permeability glacial outwash sand 0.3 5 [155]
permeability clay 0.4 5 [343]
permeability fine sands 1.5 14 [323]
permeability clay, fine sands 5.4 3 [230]
permeability silty clay 0.1 [340]
permeability fluvial soil 7.6 [135]
permeability alluvial soil 15 [180]
permeability alluvial aquifer 150 [91]
permeability gravelly loamy sand 500 [313]
permeability alluvial clay 800 [46]
permeability alluvial aquifer 820 [98]
permeability alluvial aquifer 1,800 [91]
permeability clay 4,000 [1]
ϕ’ clay 800 [301]
preconsolidation pressure soft, silty clay 0.6 exp [99]
preconsolidation pressure soft, silty clay 180 exp [99]
sand fraction soft silty loam 0 ∼ 5.4 19.6 ∼ 23 exp + cos [10]
sand fraction clay, silty sand 0.3 exp [381]
sand fraction silty sand,clay 0.6 exp [381]
Continued on next page
201
A.3 Database on the variability of cohesive soil
Table A.3 – continued from previous page
202
property soil type θver [m] θhor [m] applied spatial model reference
sand fraction clay, silty sand 1 exp [381]
silt fraction soft silty loam 0 ∼ 5.4 19.5 ∼ 23 exp + cos [10]
silt fraction soft silty loam 0 ∼ 3.8 exp [10]
UCS Mexico City clay 0 ∼ 0.9 exp [10]
unit weight soft silty loam 0 ∼3 17 ∼ 22 exp + cos [10]
unit weight soft clay 1.2 [386]
void ratio soft silty loam 16.8 exp + cos [10]
[email protected]
0 ∼ 4.5 ∼ 22
Void ratio soft, silty clay 3 exp [99]
VST very soft clay 1 22 [41]
VST soft clay 1.2 [15]
VST sensitive clay 2 [78]
VST soft clay 2.4 exp [15]
VST sensitive clay 3 30 [353]
VST marine clay 3 [353]
VST organic soft clay 3.1 [352]
VST soft clay 6.2 exp [15]
Appendix A Database on stochastic soil properties
Figure A.2 shows the frequency of the used correlation functions and techniques used
in the different entries of the database entries. One can clearly deduce from figure A.2
that nearly 50% of the literature sources in the database only offered the experimental
variogram or engineering judgement, but not a theoretical covariance function. Many of
the studied publications fitted an exponential correlation function to the experimental
values.
29 %
48%
45 %
4% 47%
7% 4% 1%
exponential LAS
Gaussian, spherical, bilinear without
Figure A.2: Frequency of the correlation functions in the database for frictonal and cohe-
[email protected]
sive soils.
203
Appendix A Database on stochastic soil properties
4000 (a)
2000
0
0 10 20 30 40 50
vertical correlation length qver
0 0 0
0 5 10 20 30 50
qver qver qver
Figure A.3: Histogram of vertical correlation lengths in cohesive soils for all scales (a),
for small(b), medium (c) and large scales (d).
204
A.3 Database on the variability of cohesive soil
pdf
1500 (a)
1000
500
600 150
400 100
200 50
Figure A.4: Histogram of horizontal correlation lengths in cohesive soils for all scales (a),
medium (b) and large scales (c).
205
Appendix A Database on stochastic soil properties
pdf
1500
(a)
1000
500
0
0 5 10 15 20 25 30 35
vertical correlation length qver
pdf pdf
(b) (c)
400
50
200
0 0
0 5 25 30 35
qver qver
Figure A.5: Histogram of vertical correlation lengths in frictional soils for all scales (a),
medium (b) and large scales (c).
pdf
300
(a)
200
100
0 10 20 30 40 50 60 70 80 90
horizontal correlation length qhor
pdf pdf
80 (b) (c)
100
60
40
20
0
0 10 20 90
qhor qhor
Figure A.6: Histogram of horizontal correlation lengths in frictional soils for all scales
(a), medium (b) and large scales (c).
[email protected]
206
Appendix B
Measurement data of CPT - databases
207
Appendix B Measurement data of CPT - databases
5.00
10.00
Depth
(m)
15.00
20.00
Figure B.1: Plan view and typical CPT profile of NGES site A LAMEDA.
208
Friction Ratio Tip Resistance
Fs/Qt (%) Qt (MN/m²)
0.00 0.0 10.0 0.0 50.0
5.00
Depth
(m)
20.00
25.00
30.00
Figure B.2: Plan view and typical CPT profile of NGES site E VANSVILLE A REA.
209
Appendix B Measurement data of CPT - databases
Depth
(m)
10.00
20.00
Figure B.3: Plan view and typical CPT profile of NGES site L ANCESTER.
210
Friction Ratio Tip Resistance
Fs/Qt (%) Qt (MN/m²)
0.00 0.0 10.0 0.0 50.0
5.00
Depth
(m)
20.00
25.00
30.00
Figure B.4: Plan view and typical CPT profile of NGES site S AN B ERNADINO C OUNTY.
211
Appendix B Measurement data of CPT - databases
Depth
(m)
10.00
20.00
Figure B.5: Plan view and typical CPT profile of NGES site S AN L UIS O BISPO C OUNTY.
212
Friction Ratio Tip Resistance
Fs/Qt (%) Qt (MN/m²)
0.00 0.0 10.0 0.0 50.0
5.00
Depth
(m)
20.00
25.00
30.00
Figure B.6: Plan view and typical CPT profile of NGES site S ANTA C LARA C OUNTY.
213
Appendix B Measurement data of CPT - databases
5.00
Depth
(m)
20.00
25.00
30.00
Figure B.7: Plan view and typical CPT profile of NGES site S OLANO C OUNTY.
214
Tip Resistance Sleeve Friction
q (MPa) fs (kPa)
T
0 10 20 30 40 50 60 0 1000 2000
0
6
Depth (m)
10
12
Figure B.8: Plan view and typical CPT profile of PEER site A NSSALL.
215
Appendix B Measurement data of CPT - databases
qc (MPa) fs (MPa)
0 10 20 30 40 0.00 0.10 0.20
0 0
1 1
2 2
3 3
4 4
5 5
Depth (m)
6 6
7 7
8 8
9 9
10 10
11 11
12 12
Figure B.9: Plan view and typical CPT profile of PEER site B ERKELEY.
216
Appendix C
Basic definitions & statistical background
Coefficient of variation
σ
COV = (C.3)
µ
• Third moment of distribution & skewness
Z ∞
1 3 µ3 E (X − µ)3
γ1 = 3 (x − µ) f (x) dx = 3 = (C.4)
σ −∞ σ σ3
217
Appendix C Basic definitions & statistical background
• quantile values can be derived such as the 0.10 quantile or the 1st decile:
q0.10 = F −1 (0.10) = z − outcome value such that Prob(Z ≤ q0.10 ) = 0.10 (C.7)
Phoon [286] as well as Baecher & Christian [23] point out the normal and the lognor-
mal distribution function as widespread probability distribution functions in geotechni-
cal engineering, which are explained in detail afterwards.
218
C.2 Distribution functions
support x∈R
mean µ
median µ
mode µ
variance σ2
skewness 0
excess kurtosis 0
X = eµ+σ Z (C.11)
On a non-logarithmised scale, µ and σ can be called the location parameter and the scale
parameter respectively. The probability density function of a lognormal distribution is
as shown in the following equation.
1 (ln x − µ)2
fX (x; µ, σ) = √ exp x>0 (C.12)
xσ 2π 2 σ2
219
Appendix C Basic definitions & statistical background
support x ∈ µ + span(Σ) ⊆ Rk )
mean µ
median µ
mode µ
variance Σ
The main diagonal gives the variance and the off-diagonals are symmetrical covariances.
The covariance matrix is not singular and definite. In case of a singular covariance ma-
trix, the corresponding distribution has no density.
The probability density function f (z) can be written in the following form.
According to Sachs [316], no analytical expression exists for den cumulative density
function.
220
C.2 Distribution functions
In Gelder [382] a comparison of the different methods together with applications of these
approaches are presented.
221
[email protected]
Appendix D
Polynomial Chaos Expansion
He1 (x) = x
He2 (x) = x2 − 1 (D.5)
He3 (x) = x3 − 3 x
223
Appendix D Polynomial Chaos Expansion
For a simple use in mathematical formulas, the multivariate Hermite polynomials are
often renamed and sorted by using only one numerical index, for example:
Each polynomial ψi of the basis of the PCE of two variables ξ1 and ξ2 can be entirely
defined by two indexes i1 and i2 such that
With this notation, the following equations have been derived by Sudret & Kiureghian
[366].
E(ψi2 ) = i1 ! · i2 !
E(ψi · ψj · ψk ) = Di1,j1,k1 · Di2,j2,k2 (D.10)
E(ψi · ψj · ψk · ψl ) = Di1,j1,k1,l2 · Di2,j2,k2,l2
In these expressions, the D terms are obtained by:
(
i! j! (i + j + k)even
Ci,j,k = ((i+j−k)/2)! ((j+k−i)/2)! ((k+i−j)/2)!
k ∈ [|i − j| , i + j] (D.11)
0 otherwise
224
D.3 Two variables PCEs used for different values of the PCE order
225
[email protected]
Appendix E
Sequential simulation algorithms
The aim of sequential simulation, as it was originally constructed, is to reproduce de-
sirable multivariate properties through the sequential use of conditional distributions
[79, 97]. Consider a set {Z(uj ), j = 1, . . . N } of N random variables
l defined at N loca-
tions uj . The objective is to generate several joint realizations z (uj ), j = 1, . . . , N , l =
1, . . . , L conditional to the available data and to some structural model such as the vari-
ogram. It can be shown that the N - point multivariate distribution can be decomposed
into a set of N one-point conditional cumulative distribution function’s as
where F (uN ; zN | (n + N − 1)) is the conditional CDF of Z(uN ) given the set of n orig-
inal data values and the previous (N − 1) realizations z (l) (uj ) , j = 1, . . . , N − 1 [145].
This decomposition allows us to generate an image by sequentially visiting each node.
Sequential simulation, under a given multivariate distribution, amounts to read the de-
composition E.2 from left to right, i.e. the purpose of sequential simulation is to repro-
duce the properties of the given multivariate distribution. The simulation algorithm
proceeds as follows according to Goovaerts [145]:
• end loop
227
Appendix E Sequential simulation algorithms
Within the Sequential Gaussian Simulation algorithm, all conditional cumulative dis-
tribution F (·) are assumed Gaussian and their means and variances are given b a series
fo N simple kriging systems; in terms of the Sequential Indicator Simulation algorithm,
the conditional cumulative distribution are obtained directly by indicator kriging, as
shown below.
1. Ensure that the data are approximately normal; transform to a standard normal
distribution if necessary.
3. Specify the coordinates of the points at which you want to simulate. These will
usually be on a grid.
4. Determine the sequence in which the points, xj ; j = 1, 2, . . . will be visited for the
simulation. Choosing the points at random will maximize the diversity of different
realizations.
(c) Insert this value into the grid at xi , and add it to the data.
(d) Proceed to the next node and simulate the value at this point in the grid.
(e) Repeat steps (a) to (c) until all of the nodes have been simulated.
The SGSIM algorithm is very fast and straightforward because the modelling of the
ccdf at each location u requires the solution of only a sling kriging system at that loca-
[email protected]
tion. The implicit assumption is that the spatial variability of the attribute values can be
228
E.2 Kriging algorithm
229
Appendix E Sequential simulation algorithms
3. Kriging with a trend model considers that the unknown local mean m(u′ ) smoothly
varies with each local neighbourhood W (u), hence over the entire study area A.
The trend component is modelled as a linear combination of functions fk (u) of the
coordinates:
XK
′
m(u ) = ak (u′ ) fk (u′ ) (E.7)
k=0
• Define a random path visiting each node of the grid only once.
[email protected]
• At each node u′ :
230
E.4 Comparison of SGSIM and SISIM algorithms
1. Determine the K ccdf values F (uα ; zk |(n)) using an indicator kriging algo-
rithm e.g. simple indicator kriging. The conditioning information consists of
indicator transform of neighbouring original z-data and previously simulated
z-values.
2. Correct for any order relation deviations resulting from negative kriging weights
[145]. Then build a complete ccdf model (u′ ; z|(n)) , ∀z.
3. Draw a simulated valued z (l) (u′ ) form that conditional cumulative distribu-
tion function.
4. Add the simulated value to the conditioning data set.
5. Proceed to the next node along the random path, and repeat steps 1 to 4.
the SGSIM algorithm can be found in various publications [79, 96, 97, 143–145, 189, 304].
231
Appendix E Sequential simulation algorithms
-4
-2
0
2
4
qver
qhor
Figure E.1: Random field realisations of the SGSIM (a) and the SISIM (b) algorithm with
an isotropic, spatial correlation (θver = θhor ).
For the sake of completeness, figures E.4 and E.5 summarize different methods for the
simulation of random fields, which are taken from Chiles & Delfiner [79]
232
E.4 Comparison of SGSIM and SISIM algorithms
g(t) g(t)
Cumulative probability
0.9 0.6
0.7
0.4
0.5
SGSIM algorithm
0.3 0.2
Figure E.2: Cumulative probability distribution (a) and variograms (b) of the SGSIM and
SISIM random field in figure E.1.
cumulative distribution
-4 -2 0 2 4
modulus of elasticity [MN/m²] indicator correlation
length qind
SGSIM maximum indicator
SISIM correlation length qind,max
Figure E.3: Indicator correlation lengths for each threshold of the SGSIM and SISIM
algorithm.
233
Appendix E Sequential simulation algorithms
Figure E.4: Main characteristics of various random field simulation methods from Chiles
& Delfiner [79].
234
E.4 Comparison of SGSIM and SISIM algorithms
Figure E.5: Main characteristics of various random field simulation methods from Chiles
& Delfiner [79].
235
[email protected]
Appendix F
Failure criteria in constitutive modelling of soil
Benz [38] states, that the Mohr-Coulomb failure criterion for soils is one of the earliest
and most trusted failure criteria.
It is experimentally verified in triaxial compression and extension and is of striking
simplicity. However, the Mohr-Coulomb (MC) criterion is very conservative for inter-
mediate principal stress states between triaxial compression and extension. The Mohr-
Coulomb failure criterion in principal stress space is defined as:
Matsuoka & Nakai [243] (MN) proposed a failure criterion that is in better agreement
with the experimental data, which is shown in figure 5.7 with the MC and the Lade -
Duncan criteria [213] (LADE). Matsuoka & Nakai [243] propose the concept of a Spatial
Mobilized Plane (SMP), which defines the plane of maximum spatial, averaged particle
mobilization in principal stress space. The SMP is geometrically constructed by deriving
the mobilized (Mohr-Coulomb) friction angles for each principal stress pair separately
(figure F.1, left) and sketching the respective mobilized planes in principal stress space
(figure F.1, right). Matsuoka & Nakai derive their failure criterion by limiting the aver-
aged ratio of spatial normal stress to averaged spatial shear stress on this plane. Their
failure stress ratio can be expressed as a simple function of the first, second, and third
stress invariant, I1 , I2 , and I3 as shown in equation F.3 and F.3. With the SMP concept, the
Matsuoka-Nakai criterion automatically retains the well established material strength of
the Mohr-Coulomb criterion in triaxial compression and extension. The likewise well-
known failure criterion by Lade & Duncan appears compared to the Mohr-Coulomb
criterion and the Matsuoka-Nakai criterion rather optimistic in plane strain conditions
and triaxial extension. Benz [38] points out that using bifurcation analysis, progressive
failures would most likely "correct" for the Lade criterion’s overly optimistic, ultimate
material strength estimate.
Both failure criteria, Matsuoka-Nakai and Lade, are functions of the first, second, and
third stress invariants, I1 , I2 , and I3 respectively:
I1 I2 9 − sin2 ϕ
fMN = − c1 = 0 with c1 = (F.2)
I3 −1 + sin2 ϕ
I13 (−3 + sin ϕ)3
[email protected]
fLade = − c2 = 0 with c2 = (F.3)
I3 (−1 + sin ϕ) (−1 + sin ϕ)2
237
Appendix F Failure criteria in constitutive modelling of soil
(a) (b)
t sI
s1
SMP
Mobilized
fmob23
Planes fmob13
fmob13 45°+ 2
fmob12
fmob12 s3
45°+ 2
sIII
s3 s2 s1 s
fmob23
45°+ 2
s2
sII
where
I1 = σij
1
I2 = (σij σij − σii σjj ) (F.4)
2
1
I3 = (σii σjj σkk + 2 σij σjk σki − 3 σij σji σkk )
6
In principal stress, the stress invariants simplify to:
I 1 = σ1 + σ2 + σ3
I2 = −σ1 σ2 − σ2 σ3 − σ3 σ1 (F.5)
I 3 = σ1 σ2 σ3
The constants c1 and c2 in equation F.3 and F.3 are defined so that both failure criteria are
identical to the Mohr-Coulomb criterion in triaxial compression.
238
Appendix G
Application of non-intrusive SFEM to a
reliability problems
The non-intrusive Stochastic Finite Element Method (SFEM) is applied to a simple ex-
ample in order to visualize the steps described in section 4.3.3. The limit state eqution
5.11 is used for the comparison of the probability of failure between non-intrusive SFEM
and FORM, the statistical moments of the response and for the calculation of the Sobol
indices. Within this example the soil properties of the parametric study 1 in section 5.4
and table 5.3 and a COVϕ′ = 20% and COVc′ = 10% are used.
The deterministic results of the system response as a function of the random input
variables in the shown in figure G.1.
The probabilistic results of a parametric study are shown in the figure G.2. Generally
speaking, it can be deduced that the higher the face pressure, the higher the reliability of
the system.
FORM and non-intrusive SFEM represent this qualitatively in the same way. The in-
fluence of the cohesion c′ and of the friction angle ϕ′ remains constant while varying the
face pressure qt .
To enlighten the concept of non-intrusive SFEM, the collocation points for different
PCE orders are plotted in the Gaussian as well in the physical space (figure G.4). The
dependency of the statistical moments to the PCE order is shown in figure G.3. Together
with the results of figure G.5 and figure G.6, the reader can see the convergence of the
non-intrusive SFEM approximation. The mean value, the variance, the skewness and
also the kurtosis show only minor changes at the PCE order M = 5. This can also be
deduced from Figure G.5 : The PCE order M = 5 is acceptable for the probability of
failure pf . This is strengthened the graphs of the empirical error and the coefficient of
determination in figure G.3.
The difference of both approaches can be clearly seen in Figure 4. The lower the prob-
ability of failure, the bigger is the difference of FORM and CSRSM. This is due to the
approximation of the system response via a PCE within CSRSM, which becomes more
inaccurate with deceasing probability of failure. The dependency of the Sobol indices to
the PCE order M is rather low, which is enlightened in Figure G.7.
It can be concluded from the presented study that the non-intrusive SFEM cannot be
accurately applied to evaluate the small probability of failures, which was also reported
in literature e.g. Phoon [286]. This can be deduced to the concept approximation of the
stochastic system response using a high order polynomial, which does not represent the
small tails of the distribution of system response. But the concept of non-intrusive SFEM
[email protected]
is can be efficiently applied to evaluate the first and second statistical moments of the
239
Appendix G Application of non-intrusive SFEM to a reliability problems
system response as well as for the analytical evaluation of global sensitivities. These
benefits can be applied to any sort of uncertain input variable ranging from soil strength
properties, correlation lengths,... etc.
If this approach is applied to the system response assuming uncertain correlation
lengths, on as to keep in mind that the evaluation points are mean values and the fit-
ted PCE model will just approximate the real response surface; therefore, the sensitivity
factors derived from the PCE are not global factors but help to understand the impor-
tance between the input variables between each other.
system response
system response
qt [kN/m²]
40 30
35
30
25
25
20
15 20
10
5 15
0
45 10
fric 40 4.0 4.5 ²]
tion 35
ang 30 3.0 3.5
c’ [ kN/m
2.5 sion
el j
’ [° 25 1.5 2.0 cohe
]
Figure G.1: Input variables and system response of an PCE with order M = 5 and a face
[email protected]
pressure of qt = 40 kN/m2 .
240
0
10
probability of failure pf
-5
10
-10
10
non-intrusive SFEM (M= 5)
-15 FORM
10
10 20 30 40 50
face pressure qt in kN/m²
Figure G.2: Comparison of non-intrusive SFEM and FORM in terms of the probability of
failure for a COVϕ′ = 5 % and COVϕ′ = 10 % .
-3
x 10
0.01
empirical error
PRESS
2
0.005
1
0 0
1 2 3 4 5 1 2 3 4 5
expansion order M of the PCE expansion order M of the PCE
determination coefficient
1
0.9995
0.999 determination coefficient
adjusted determination coefficient
0.9985 Q2
0.998
1 2 3 4 5
expansion order M of the PCE
Figure G.3: Acccuracy of the PCE fitting for a face pressure qt = 40 kN/m2 .
[email protected]
241
Appendix G Application of non-intrusive SFEM to a reliability problems
order M = 2 45
4
40
ξ j’ 0 φ’ 35
30
-4
-4 -2 0 2 4 2 2.5 3 3.5
ξ c’ c
order M = 3 45
4
40
ξ j’ 0 φ’ 35
30
-4
-4 -2 0 2 4 2 2.5 3 3.5
ξ c’ c
c’
order M = 4 45
4
40
ξ j’ 0 φ’ 35
30
-4
-4 -2 0 2 4 2 2.5 3 3.5
ξ c’ c’
order M = 5 45
4
40
ξ j’ 0 φ’ 35
30
-4 25
-4 -2 0 2 4 1.5 2 2.5 3 3.5
ξ c’ c’
Figure G.4: Collocation points in Gaussian space and physical space for several PCE ex-
pansion orders M.
242
2 0.4
Mean value
1.5
Variance
0.3
1 0.2
0.5 0.1
0 0
1 2 3 4 5 6 1 2 3 4 5 6
PCE order M PCE order M
1 2
Skewness
0 0
1 2 3 4 5 6 1 2 3 4 5 6
PCE order M PCE order M
Figure G.5: Statistical moments of the approximated system response for a face pressure
qt = 40 kN/m2 .
proability of failure p f
-2
10
-4
10
-6
10
-8
10
0 2 4 6
PCE order M
243
Appendix G Application of non-intrusive SFEM to a reliability problems
-3
0.951 0.052 6 x 10
0.95 0.051 5
0.949 4
0.05
0.948
0.049 3
0.947
0.048 2
0.946
0.945 0.047 1
0.944 0.046 0
1 2 3 4 5 6 1 2 3 4 5 6 1 2 3 4 5 6
PCE expansion order M PCE expansion order M PCE expansion order M
Figure G.7: Sobol indices for several PCE expansion orders for the face pressure qt = 40
kN/m 2 .
244
Appendix H
Pluri-Gaussian Simulation Method
H.1 Introduction
Soil properties vary naturally through space as a result of the complex geological pro-
cesses through which soil evolve, [204]. Physical and chemical processes, including
structural deformation, deposition and diagenesis control the geometry and texture of
sedimentary deposits and create soil variability at different scales.
This multi-scale spatial variability of soil properties can be captured by geostatis-
tical simulation approaches. Koltermann & Gorelick [204] group the simulation ap-
proaches in three main categroies: structure imitating, process imitating and descriptive.
Structure-imitating methods rely on spatial statistics, probabilistic rules and determin-
istic constraints to depict geometric relations within aquifers and reservoirs. Process-
imitating methods solve governing equations to represent either the processes through
which soil form or the physics of subsurface fluid-flow and transport. Descriptive meth-
ods divide subsoil into zones by synthesizing measured soil properties and geologic
observations into a conceptual depositional model.
Structure-imitating methods are further subdivided into Gaussian and Non-Gaussian
methods. Gaussian-related methods produce images of a continuous variable with the
same mean, variance, covariance function and a Gaussian univariate distribution of val-
ues. The continuous distribution can be truncated or grouped into categories, [204].
Non-Gaussian algorithms include indicator-based methods, simulated annealing, Boolean
methods, and Markov chains. Boolean methods do not reproduce a covariance func-
tion but create a geologic image by random1y generating in space simple shapes com-
monly observed in sedimentary deposits. In contrast, indicator-based methods are based
on variogram models, simulated annealing solves an optimization formulation derived
from the problem of solidification of a solid from a hot liquid upon cooling, and Markov
chains are based on transition probabilities, [204]. Geologic information can be consid-
ered through training images that contain geologic features deemed important to the
investigated problem.
Geologic information can be considered through training images that contain geologic
features deemed important to subsurface fluid flow and transport. Kupfersberger &
Deutsch [210] report that training images can be geologic maps, cross sections, well logs
data, fence diagrams, outcrop maps, conditioning data grouped into zones or images
from quantitative depositional models. The form and parameters of a model of spatial
[email protected]
correlation must reflect the features observed in the training image, [79].
245
Appendix H Pluri-Gaussian Simulation Method
-0.6 0.5
246
H.2 Pluri-Gaussian simulation approach
identify two directions (N50 and N150) with a very high anisotropy.
Experimental variograms in the horizontal plane coordinate system of three indicator
variables considered, calculated with a pitch of 100 m, 50 m tolerance step angular toler-
ance of 45◦ . The variograms are calculated in both directions N60 and N150 (direction of
maximum anisotropy) and exhibit a variability of the spatial distribution of soil types,
especially medium-large, non-stationarity of the phenomenon, made more apparent by
the study of proportions in figure H.8 (b), is partly masked by the fact of having areas
with mixed distributions very different.
In figure H.8 (a) the vertical proportion curves of an area 500 m x 500 m and map of
the vertical proportion curves of the soil types 1, 2 and 3 H.8 (b).
These proportion curves in figure H.8 are analysed to derive the lithotypes, which are
used in the pluri-Gaussian simulations shown in figure H.9.
247
Appendix H Pluri-Gaussian Simulation Method
(a)
(b)
Figure H.2: (a) Simplified geological sketch map of the study area from Raspa et al.[301].
Legend:
(1) upper Pleistocene - Holocene alluvial deposits,
(2) middle-upper Pleistocene volcanic bedrock,
(3) Plio-Pleistocene sedimentary bedrock,
(4) boreholes with geotechnical samples,
(5) boreholes with samples endowed with the full set of
geotechnical information and
(6) track of the geological cross section in figure H.3 (a).
248
H.2 Pluri-Gaussian simulation approach
(a) Tenever NE
SW 30
River
20
10
0 m above
seal level
geotechnical
-10
borehole
-20
antropic backfill
-30
Soil type I
-40
Soil type II
-50
Soil type III
-60
Soil type IV 0 500 m
(b)
Figure H.3: (a) Geological cross section of the recent alluvial deposits filling the
Tevere valley from [301].
(b) 3D plot of the selected boreholes with location of the geotechnical
samples (black points) for location of the boreholes in figure
[email protected]
H.2 from [301].
249
Appendix H Pluri-Gaussian Simulation Method
boreholes
geological soil
investigation
Figure H.4: Plan view investigated area and area of the case study (400 m x 600 m) in
the city of Rome with locations of the boreholes and soil investigations from
[238].
[email protected]
250
H.2 Pluri-Gaussian simulation approach
gind(t)
251
Appendix H Pluri-Gaussian Simulation Method
gind,1(t)
t [m]
gind,2(t)
gind,1-2(t)
t [m] t [m]
gind,3(t)
gind,1-3(t)
gind,2-3(t)
t [m]
Figure H.6: Horizontal indicator variograms of the three different soil types from [238].
252
H.2 Pluri-Gaussian simulation approach
gind,1(t)
t [m]
gind,1-2(t)
gind,2(t)
t [m] t [m]
gind,3(t)
gind,2-3(t)
gind,1-3(t)
t [m]
Figure H.7: Vertical indicator variograms of the three different soil types from [238].
253
Appendix H Pluri-Gaussian Simulation Method
Soil type I
Soil type II
Soil type III
20
10
0
-10
-20
-30
-40
-50
0 50 100
4 646 000
4 645 000
4 644 000
4 643 000
4 642 000
4 641 000
4 640 000
288 000 289 000 290 000 291 000 292 000
Figure H.8: Vertical proportion curve of an area 500 m x 500 m (a) and map of the vertical
proportion curves of the soil types 1, 2 and 3, [238].
[email protected]
254
H.2 Pluri-Gaussian simulation approach
Y1 Y1 Y1
3 1 2 3 1 2
2 1 3
Y2 Y2 Y2
255
Appendix H Pluri-Gaussian Simulation Method
0 1 2
km
Figure H.10: Plan view of the Pluri-Gaussian mesh from Marconi [238].
256
H.2 Pluri-Gaussian simulation approach
(a)
(c)
(b)
Figure H.11: Part of the realisation of the pluri-Gaussian random field for soil type 1 (a),
soil type 2 (b).
257
Appendix H Pluri-Gaussian Simulation Method
(a)
(b )
(c)
Figure H.12: Part of the realisation of the pluriGaussian random field including soil type
3 (a), soil type 4 (b), antropogenetic top soil layer (c).
258
Table H.1: Input parameters of the soil model.
[email protected]
reference pressure pref [kN/m2 ] 100 100 100 100
failure ratio Rf [-] 0.9 0.9 0.9 0.9
Poisson’s ratio ν [-] 0.33 0.37 0.37 0.37
reference Young’s modulus E 50 [kN/m2 ] stochastic stochastic stochastic stochastic
reference Young’s modulus for unloading-reloading E ur [kN/m2 ] 105, 000 12, 000 8, 000 300, 000
ratio of shear moduli G0 / Gur [-] 2.0 2.0 2.0 2.0
threshold shear strain γ 0.7 [-] 10-4 10-4 10-4 10-4
259
H.2 Pluri-Gaussian simulation approach
[email protected]
Lebenslauf
30.09.1979 Geboren in Linz (Österreich)
261
[email protected]
Mitteilungen des Instituts für Geotechnik
der Universität Stuttgart
263
Mitteilungen des Instituts für Geotechnik
264
Mitteilungen des Instituts für Geotechnik
265
Mitteilungen des Instituts für Geotechnik
266
Mitteilungen des Instituts für Geotechnik
Mitteilungen des
Instituts für Geotechnik der Universität Stuttgart
Hrsg.: Prof. Dr.-Ing. Dr. h.c. U. Smoltczyk
Nr. 28 Kolb, H. (1988) Ermittlung der Sohlreibung von
Gründungskörpern unter
horizontalem kinematischen
Zwang
267
Mitteilungen des Instituts für Geotechnik
268
Mitteilungen des Instituts für Geotechnik
Mitteilungen des
Instituts für Geotechnik der Universität Stuttgart
Hrsg.: Prof. Dr.-Ing. P.A. Vermeer
Nr. 41 Vermeer, P.A. (1996) Deponiebau und Geotechnik
269
Mitteilungen des Instituts für Geotechnik
270
Mitteilungen des Instituts für Geotechnik
271
Mitteilungen des Instituts für Geotechnik
Mitteilungen des
Instituts für Geotechnik der Universität Stuttgart
Hrsg.: Prof. Dr.-Ing. habil. Ch. Moormann
Nr. 65 Moormann, Ch. (2011) 7. Stuttgarter Geotechnik-
Symposium
272