You are on page 1of 227

196807-os-Engin_DEF 1.

indd 1 23-5-13 17:43


PROPOSITIONS
accompanying the thesis
Modelling Installation Effects - A Numerical Approach
by Harun Kürşat Engin, Delft, 10 June 2013

1. Load settlement behaviour of displacement piles cannot be modelled properly


without considering the installation effects.
2. The Press-Replace technique is a simple and stable method that can be used
to model pile penetration problems in standard finite element software.
3. The strength of interface elements ought to be directly correlated to the
strength of continuum elements.
4. The stresses around a wished-in-place pile that are imposed to approximate
the installation effects, reduce during the nil step in which the unbalance is
removed. This reduction can be minimised by using a stiff material when
imposing the stresses and restoring equilibrium.
5. The accuracy at which BVPs such as pile penetration (which involve com-
plicated stress paths) can be simulated, is not guaranteed by tuning the con-
stitutive model only for well-defined stress/strain paths.
6. The relative density, Id is not a reliable input parameter; however, if it could
be determined accurately, it is well suited for estimating the properties of
cohesionless soils.
7. The pleasure of doing research stands on the motivation to overcome frustra-
tion and to obtain the final result.
8. When a novelty is introduced, emphasizing the capabilities should not supress
the discussion on the limitations.
9. Engineers and managers are alike; the former use methods, the latter use
people that, in both cases, have not been developed by themselves.
10. The current scientific publishing system shows how academia is exploited twice:
by letting articles be reviewed for free and successively asking money to access
the published articles. Since the review process is done without any cost, there
should be no concern for the quality of open-access articles.
11. An overdose of agitation makes people indifferent.

These propositions are regarded as opposable and defendable, and have been approved
as such by the supervisors Prof.ir.A.F. van Tol and Dr.ir.R.B.J. Brinkgreve.

196807-st-Engin.indd 1 22-5-13 16:17


STELLINGEN
behorende bij het proefschrift
Modelling Installation Effects - A Numerical Approach
van Harun Kürşat Engin, Delft, 10 Juni 2013

1. Last-verplaatsingsgedrag van grondverdringende palen kan niet goed worden


gemodelleerd zonder de installatie effecten te beschouwen.
2. De Press-Replace techniek is een eenvoudige en stabiele methode die kan worden
gebruikt om het inbrengen van palen in standaard eindige-elementen software
te modelleren.
3. De sterkte van interface elementen dient direct gecorreleerd te worden aan de
sterkte van continuüm elementen.
4. De spanningen rondom een wished-in-place paal die worden aangebracht om de
installatie effecten te benaderen, nemen af tijdens de nil step waarin de onbalans
wordt verwijderd. Deze afname kan worden geminimaliseerd door gebruik te
maken van stijf materiaal tijdens het aanbrengen van de spanningen en het
herstellen van evenwicht.
5. De nauwkeurigheid waarmee randvoorwaardeproblemen zoals het inbrengen
van palen (hetgeen gecompliceerde spanningspaden met zich mee brengt) kan
worden gesimuleerd, is niet gegarandeerd door het constitutieve model alleen
af te stemmen op goed-gedefinieerde spanning/rek paden.
6. De relatieve dichtheid, Id , is geen betrouwbare invoer parameter; echter, als
deze nauwkeurig zou kunnen worden bepaald, is deze zeer geschikt om de ei-
genschappen van cohesieloze gronden te schatten.
7. Het plezier aan het doen van onderzoek staat met de motivatie om frustratie
te overwinnen en het eindresultaat te behalen.
8. Wanneer een noviteit wordt geı̈ntroduceerd dient het benadrukken van de mo-
gelijkheden de discussie over de beperkingen niet te onderdrukken.
9. Ingenieurs en managers zijn vergelijkbaar; de eersten gebruiken methoden, de
laatsten gebruiken mensen die, in beide gevallen, niet door hen zelf zijn ont-
wikkeld (gevormd).
10. Het huidige wetenschappelijk publicatie system laat zien hoe de academische
wereld dubbel wordt uitgebuit: door onbetaald artikelen te laten reviewen en
door vervolgens te laten betalen om toegang te verkrijgen tot de gepubliceerde
artikelen. Omdat het reviewen toch onbetaald is, zijn zorgen over de kwaliteit
van artikelen in vrij toegankelijke journals, ongegrond.
11. Een overdosis aan onrust maakt mensen onverschillig.

Deze stellingen worden opponeerbaar en verdedigbaar geacht en zijn als zodanig goed-
gekeurd door de promotoren Prof.ir.A.F. van Tol en Dr.ir.R.B.J. Brinkgreve.

196807-st-Engin.indd 2 22-5-13 16:17


Modelling Pile Installation Effects
A Numerical Approach
Modelling Pile Installation Effects
A Numerical Approach

PROEFSCHRIFT
ter verkrijging van de graad van doctor
aan de Technische Universiteit Delft,
op gezag van de Rector Magnificus prof.ir. K.C.A.M. Luyben,
voorzitter van het College voor Promoties,
in het openbaar te verdedigen op maandag 10 juni 2013 om 12.30 uur

door Harun Kürşat ENGİN

Master of Science in Civil Engineering


Middle East Technical University, Ankara, Turkey
geboren te Ankara (Turkije).
Dit proefschrift is goedgekeurd door de promotoren:
Prof. ir. A.F. van Tol

Copromotor:
Dr. ir. R.B.J. Brinkgreve

Samenstelling promotiecommissie:
Rector Magnificus, voorzitter
Prof. ir. A.F. van Tol, Technische Universiteit Delft, promotor
Dr. ir. R.B.J. Brinkgreve, Technische Universiteit Delft, copromotor
Prof. Dr. M.A. Hicks, Technische Universiteit Delft
Prof. Dr. ir. L.J. Sluys, Technische Universiteit Delft
Prof. Dr. ing. habil. I. Herle, Technische Universität Dresden
Prof. Dr. ing. H.P. Jostad, Norges Teknisk-Naturvitenskapelige Universitet
Dr. L. Zdravković, Imperial College London

This research is supported by the Dutch Technology Foundation STW, which is part
of the Netherlands Organisation for Scientific Research (NWO) and partly funded
by the Ministry of Economic Affairs (project number 10189).

The research leading to these results has also received funding from the 7th Frame-
work Programme (FP7/2007-2013 under grant agreement PIAG-GA-2009-230638).
The findings reflect only the authors views and the EC is not liable for any use that
may be made of the information contained therein.

ISBN 978-94-6186-140-5

© 2013 by H.K. Engin. All rights reserved. No part of the material protected by
this copyright notice may be reproduced or utilized in any form or by any means,
electronic or mechanical, including photocopying, recording or by any information
storage and retrieval system, without written consent of the publisher.

Cover: Erjona Engin & H. Kürşat Engin (Concept); A. Burak Veyisoğlu (Design)
Printed by : Ipskamp Drukkers / E-mail: info@ipskampdrukkers.nl
To my grandfather İbrahim Hakkı Bilici (RIP)
&
all others who have supported me
vi
Summary

One of the most traditional methods for supporting structures resting on soft soils
is the use of piles. They generally work by transferring the loads to deeper soil
layers, which can provide sufficient bearing capacity when mobilised. This type of
foundations has been commonly used throughout the world and also in the western
part of the Netherlands due to the typical soil profile which consists of a thick (10 -
20m) soft soil layer underlain with a stiff bearing stratum composed mostly of quartz
sand. In this perspective, this study sheds a light on the behaviour of piles installed
in silica sand.
Piles can be installed in different ways. The most common way of installing the
prefabricated piles is by using an impact hammer. With this driving method, the
pile is penetrated by each drop or blow of the ram until the desired penetration level
is achieved. Another alternative is the vibratory driving technique where the pile is
forced to penetrate by a heavy vibratory head on top. Recent developments have al-
lowed the piles to be jacked thus resulting in reduced nuisance and in higher capacity
and stiffness when compared to the hammering and vibratory driving techniques.
Yet, the method is generally limited to open ended piles or piles with relatively small
diameters for closed ended piles.
The bearing capacity of a pile depends on the soil properties and the stress state
it is surrounded with. This is because the behaviour of granular material is governed
by the packing of the grains and the contact stresses in between. The mean stress and
the density can be described as the soil state, and the soil behaviour is determined
on the basis of this state and the loading conditions. In the case of a displacement
pile, the installation process causes a considerable amount of soil displacement and
high levels of (reaction) stresses. These effects of pile installation are transmitted to
soil through the interaction between sand grains and the pile, resulting in an altered
soil state and properties.
It is therefore obvious that one should have different installation effects for dif-
ferent installation techniques. For example, the confining stresses resulting from
the dynamic installation technique (hammering or vibratory driving) are lower than
those of the jacking technique. Depending on the density, dilation would be expected
near the pile shaft during jacking. On the other hand, dynamic installation methods
cause compaction near the pile. The installation effects may have two important
consequences in the form of the influence on the performance of the pile (e.g. load
displacement behaviour, capacity) in its service life and the influence (e.g. noise,
vibrations) on the neighbourhood /environment.

vii
viii Summary

A more realistic behaviour and therefore an improved design would be achieved


by considering the installation effects in the analyses than performing the analysis
considering a geostatic stress state around the pile(s) modelled. In current practice,
the installation effects are taken into account by some empirical design methods
in order to estimate the bearing capacity of foundation piles. Several field and
model tests performed to investigate the influence of pile installation on the bearing
capacity, have led to an evolution of the empirical models to estimate the bearing
capacity of displacement piles. Recent attempts to investigate the change in the soil
state were also limited either to the measurement domain (generally close to the
pile) or resolution as well as the variables (e.g. displacement, strain, stress, density)
that can be quantified. However, the behaviour of piles during installation, the
interaction with the surrounding soil and the resulting alteration of soil properties
during installation are still not well known. This information is essential, not only
to make better predictions of the pile bearing capacity and its behaviour in the soil
under different loading conditions, but also to be able to predict the (side) effects of
pile installation on the neighbourhood.
The increase in the computational power allowed improved numerical techniques
like the finite element method (FEM); however the installation effects of displace-
ment piles have not been incorporated in the numerical models for engineering pur-
poses. Because the installation effects have been discarded there will be considerable
difference between the FE predictions and the real behaviour (performance) of the
pile. Therefore it is important to model the installation effects of displacement piles
numerically so that the load displacement behaviour of the pile during its service life
as well as the secondary effects on neighbouring structures can be predicted more
appropriately.
For an accurate analysis of the pile installation and its consequent effects, all
important aspects (e.g. large deformation effects, pore pressure effects, inertia effects
and use of a suitable constitutive model) should be included in the numerical model.
Nevertheless, simplifications are unavoidable. For example the inertia effects can be
discarded and drained conditions (no excess pore pressure effects) can be assumed
to hold for the pile jacking case, which is seen as quasi-static penetration. Therefore
in the resulting FE model, the level of complexity will be reduced.
This study is the numerical part of a larger project in which the installation
effects of driven piles are investigated, both experimentally and numerically. The
objective of this numerical study is to investigate and model the installation effects
of pile jacking in sand in a numerical framework. Since any jacking operation can be
considered as a quasi- static loading operation, the dynamic effects are not included
in the analyses. For the same reason, drained conditions are assumed. As a common
tool used in engineering practice, FE modelling is selected as the numerical method.
In order to model pile jacking using the FEM in a standard FE code, some addi-
tional aspects have to be considered to account for the large deformation effects (e.g.
change of geometry due to pile penetration). Bearing in mind the aforementioned
points and the ease of applicability in a standard FE package, the Press-Replace
(PR) technique was introduced. The technique benefits the use of a simple and
robust small deformation formulation on one side and models the large deforma-
ix

tion effects (e.g. the penetration process) by using updated geometry on the other
side. The capabilities and limitations of the method were investigated by means
of a parametric study. Based on this study, an optimised modelling technique was
proposed.
Another important aspect of FE analyses is that of modelling the soil behaviour.
First of all, the stress and density dependency is an important feature of the soils
to be modelled. This feature is quite well established by the Critical State theory in
which the state (stress and density) determines the soil response. There are several
Critical State models mentioned in the literature. Hypoplasticity was employed as
the constitutive model in the simulation considering its capabilities (e.g. the void
ratio and stress dependent stiffness).
It has been previously observed that due to the high stress levels encountered
at the pile base, particle or grain crushing may also occur. As a result of crushing,
not only the behaviour but more importantly the material properties change. The
constitutive model should therefore account for this modification. In order to model
the crushing effects, the hypoplastic model was modified. In the modification the
descriptive input parameters were allowed to alter during the simulations. However,
this gave rise to problems especially in regions with high gradients of stresses. As
a result, a converged solution with the modified version could not be obtained. In
order to keep the model simple and still applicable in a standard FE code, the grain
crushing effect was ignored in the FE simulations.
As the aim of the study was to model installation effects of piles, the PR tech-
nique was employed to investigate the installation effects for different pile geometries
and different soil densities. The results have shown that the installation field is qual-
itatively similar for all of these variations. This makes the generalisation of all these
installation fields possible.
In view of the results obtained from FE simulations, the possibilities of incor-
porating the installation fields (e.g. Cartesian stress and void ratio distributions)
around a jacked pile were investigated. The installation fields were mapped to dif-
ferent meshes to investigate the possibility of imposing the installation fields. The
idea was to account for the installation effects without simulating the penetration
process and, as a result, model the service (e.g. load -displacement) behaviour of the
displacement pile properly. The results have shown that the axial load displacement
response of the pile at the end of the PR analysis was successfully captured by the
imposed installation field around a wished-in-place pile modelled with a different
discretisation than the PR model.
For ease of applicability in engineering practice, the possibility of describing the
installation effects in a mathematical (functional) form was investigated. This allows
a versatile method for imposing the installation field around a wished-in-place pile
so that it captures the jacked pile behaviour. Firstly a multiple regression algorithm
was employed to fit functions for the Cartesian stresses and the void ratio for a
reference case. The exponential form of the functions closely fit the installation
surfaces. Secondly, these functions were extended to include the pile geometry and
relative density effects.
x Summary

Despite the limitations and simplifications, it was shown that the idea of de-
scribing the installation field in terms of functional forms works reasonably well and
can easily be applied in a standard FE analysis. In order to validate the functions
proposed, a centrifuge pile was modelled. As a result of incorporating the install-
ation effects around the wished-in-place pile, the load displacement response could
be modelled reasonably well.
In practice, if the installation effects are considered in the FE analyses, better
insight into foundation design can be obtained. As a result of considering increased
capacity and stiffness of the displacement pile analysed, the foundation costs would
be reduced. Furthermore, if there are any buildings in the vicinity of a pile, possible
disturbance or damage to that structure could be foreseen in the FE analysis.
Samenvatting (in Dutch)

Een van de meest traditionele methoden voor het ondersteunen van constructies op
slappe grond is het gebruik van palen. Die werken in het algemeen door de belasting
naar dieper gelegen grondlagen over te brengen, waardoor voldoende draagkracht
kan worden gemobiliseerd. Dit soort funderingen is veelvuldig toegepast over de
hele wereld en zo ook in het westen van Nederland door het typische grondprofiel
met een dikke (10 - 20 m) slappe grondlaag met daaronder een stijve draagkrachtige
laag bestaande uit voornamelijk kwarts zand. Vanuit dat perspectief bezien is deze
studie bedoeld om inzicht te geven over het gedrag van palen in silicium zanden.
Palen kunnen op verschillende manieren worden geı̈nstalleerd. De meest gebrui-
kelijke manier om prefab palen te installeren is door middel van een heiblok. Met
deze inbrengmethode wordt de paal bij elke slag een stukje ingebracht totdat het
vereiste inheiniveau is bereikt. Een alternatief is de trilmethode waarbij de paal de
grond wordt ingebracht door een zwaar trilblok bovenop de paal. Recente ontwikke-
lingen hebben geleid tot het in de grond drukken van de paal waarbij overlast worden
beperkt en draagkracht en stijfheid hoger (kunnen) zijn vergeleken met de hei- en
trilmethode. Toch wordt de methode nog beperkt tot open buispalen of massieve
palen met een relatief kleine diameter.
De draagkracht van een paal hangt af van de grondeigenschappen en de span-
ningstoestand rondom de paal. Dat komt omdat het gedrag van granulair materiaal
wordt bepaald door de pakking van de korrels en de contactspanningen ertussen.
De gemiddelde spanning en de dichtheid kunnen worden aangegeven als de toestand
van de grond, en het grondgedrag wordt bepaald op basis van deze toestand en de
belastingsituatie. Bij een grondverdringende paal veroorzaakt het installatieproces
een aanzienlijke grondverplaatsing en hoge (reactie-)spanningen. Deze paal installa-
tie effecten worden overgebracht naar de grond door middel van de interactie tussen
de zandkorrels en de paal, waardoor de toestand van de grond en eigenschappen
wijzigen.
Het is duidelijk dat verschillende installatietechnieken tot verschillende instal-
latie effecten leiden. Bijvoorbeeld, de spanningen rondom de paal die voortkomen
uit dynamische installatiemethoden (heien of trillen) zijn kleiner dan bij het in de
grond drukken. Afhankelijk van de dichtheid wordt dilatantie verwacht langs de
paalschacht tijdens het indrukken. Daar staat tegenover dat dynamische installa-
tiemethoden compactie langs de paal veroorzaken. De installatie effecten kunnen
twee belangrijke consequenties hebben in de zin van invloed op het gedrag van de
paal (t.w. last-zakkingsgedrag, draagkracht) in de gebruiksfase en de invloed op de

xi
xii Samenvatting (in Dutch)

omgeving (o.a. lawaai, trillingen).


Een meer realistisch gedrag en derhalve een verbeterd ontwerp kan worden be-
reikt door de installatie effecten in de analyse mee in beschouwing te nemen ten
opzichte van het modelleren van een zuiver geostatische spanningstoestand rondom
de paal (palen). In de huidige praktijk worden installatie effecten meegenomen in
empirische ontwerpmethoden ter bepaling van de draagkracht van funderingspalen.
Diverse veld- en model testen, uitgevoerd om de invloed van de installatie op de
draagkracht van de paal te onderzoeken, hebben geleid tot veranderingen in em-
pirische modellen ter bepaling van de draagkracht van grondverdringende palen.
Recente pogingen om de verandering van de toestand van de grond te onderzoeken
waren beperkt in het meetgebied (in het algemeen rondom de paal) of de meet-
puntdichtheid, als ook het aantal gemeten variabelen (verplaatsing, rek, spanning,
dichtheid). Echter, het gedrag van palen tijdens installatie, de interactie met de om-
liggende grond en de resulterende verandering van de grondeigenschappen tijdens
installatie zijn nog steeds niet goed bekend. Deze informatie is essentieel, niet alleen
om betere voorspellingen te maken van de paal draagkracht en het gedrag in de
grond onder verschillende belastingcondities, maar ook om de (neven-)effecten van
de paal installatie op de omgeving te kunnen voorspellen.
De toename in rekenkracht van computers heeft het gebruik van geavanceerde
numerieke methoden zoals de eindige-elementenmethode (EEM) mogelijk gemaakt;
echter de installatie effecten van grondverdringende palen zijn nog niet meegenomen
in numerieke modellen voor ontwerpdoeleinden. Vanwege het negeren van installa-
tie effecten is er een aanzienlijk verschil tussen de EEM predicties en het werkelijke
gedrag van de paal. Daarom is het belangrijk om de installatie effecten van grond-
verdringende palen numeriek te modelleren zodat het last-verplaatsingsgedrag van
de paal tijdens de gebruiksfase zowel als de secundaire effecten op omliggende con-
structies beter kunnen worden voorspeld.
Voor een nauwkeurige analyse van de paalinstallatie en daaruit voortkomende ef-
fecten is het van belang om alle belangrijke aspecten (zoals grote deformatie effecten,
waterspanningen, massa traagheid en het gebruik van een geschikt constitutief mo-
del) mee te nemen in het numerieke model. Desalniettemin zijn vereenvoudigingen
onvermijdelijk. Daarom wordt uitgegaan van in de grond gedrukte palen. Traag-
heidseffecten kunnen dan bijvoorbeeld worden weggelaten en gedraineerde condities
(geen wateroverspanningseffecten) kunnen worden verondersteld te gelden, hetgeen
wordt gezien als quasi-statische inbrenging. Daardoor zal de complexiteit in het
resulterende EEM model worden beperkt.
Deze studie is het numerieke deel van een groter project waarin installatie effec-
ten van geheide palen worden onderzocht, zowel experimenteel als numeriek. Het
doel van deze numerieke studie is om de installatie effecten van in de grond ge-
drukte palen in zand in een numerieke context te onderzoeken en te modelleren.
Omdat elke indrukking als een quasi-statische belasting kan worden beschouwd wor-
den dynamische effecten niet meegenomen in de berekeningen. Om dezelfde reden
worden gedraineerde condities aangenomen. Als een veelgebruikt gereedschap in de
ingenieurspraktijk wordt de EEM modellering als numerieke methode gehanteerd.
xiii

Om het indrukken van palen met de EEM in een standaard programma te model-
leren moeten enkele aanvullende aspecten worden beschouwd om de grote deformatie
effecten in rekening te brengen (t.w. de verandering van de geometrie als gevolg van
het inbrengen van de paal). Vanuit deze gedachte en de eenvoudige toepasbaarheid in
een standaard EEM pakket is de Press-Replace (PR) techniek geintroduceerd. Deze
techniek laat enerzijds het gebruik van een eenvoudige en robuuste kleine-deformatie
formulering toe en modelleert anderzijds de grote vervormingseffecten (het inbreng-
proces) door gebruik te maken van geometrie- aanpassing. De mogelijkheden en
beperkingen van de methode zijn onderzocht door middel van een parameterstudie.
Op basis van deze studie is een geoptimaliseerde modelleringstechniek voorgesteld.
Een ander belangrijk aspect van de EEM analyses is de modellering van het
grondgedrag. Ten eerste is de spannings- en dichtheidsafhankelijkheid een belang-
rijke eigenschap van de te modelleren gronden. Deze eigenschap is redelijk verdis-
conteerd in de Critical State theorie waarbij de toestand (spanning en dichtheid)
de grondrespons bepaalt. Er worden diverse Critical State modellen genoemd in de
literatuur. Hypoplasticiteit is hier gebruik als constitutief model in de simulaties
vanwege zijn eigenschappen (t.w. de spanning- en poriëngetal-afhankelijkheid van
de stijfheid).
Het is eerder waargenomen dat ten gevolge van hoge spanningsnivieaus aan de
paalvoet deeltjes of korrels kunnen verbrijzelen. Als gevolg van verbrijzeling ver-
andert niet allen het gedrag maar veranderen ook de materiaaleigenschappen. Het
constitutieve model moet daarom rekening houden met deze verandering. Om de
verbrijzelingseffecten te modelleren is het hypoplastische model aangepast. In het
aangepaste model mogen de modelparameters tijdens de simulatie veranderen. Ech-
ter dit gaf aanleiding tot problemen, in het bijzonder in de gebieden met hoge span-
ningsgradiënten. Als gevolg daarvan kon een geconvergeerde oplossing met het aan-
gepaste model niet worden verkregen. Om het model eenvoudig en toepasbaar te
houden in een standaard EEM programma is het verbrijzelen van korrels in de EEM
simulaties verder buiten beschouwing gelaten.
Met oog op het doel van deze studie om installatie effecten van palen te model-
leren is de PR techniek toegepast om de installatie effecten bij verschillende paal
geometrieën en verschillende dichtheden te onderzoeken. De resultaten hebben la-
ten zien dat de verdeling van installatie effecten kwalitatief gelijk is voor al deze
variaties. Dit maakt het mogelijk om deze effecten te generaliseren.
In het licht van de met de EEM simulaties verkregen resultaten zijn de mo-
gelijkheden om de installatie effecten (t.w. Cartesische spanningen en poriëngetal
verdeling) rondom een in de grond gedrukte paal mee te nemen, onderzocht. De
verdeling van installatie effecten is op verschillende meshes geprojecteerd om de mo-
gelijkheden voor het aanbrengen van deze effecten te onderzoeken. Het idee is om
de installatie effecten in rekening te brengen zonder het inbrengproces te simuleren,
met als resultaat om het (last-verplaatsings-)gedrag van een grondverdringende paal
in de gebruiksfase juist weer te geven. De resultaten laten zien dat de axiale last-
verplaatsingsrespons aan het eind van de PR analyse succesvol wordt weergegeven
met de aangebrachte installatie effecten rondom een zogenaamde wished-in-place
paal gemodelleerd met een andere discretisatie dan het PR model.
xiv Samenvatting (in Dutch)

Voor een eenvoudige toepassing in de ingenieurspraktijk is de mogelijkheid on-


derzocht om de installatie effecten in een wiskundige vorm (functie) weer te geven.
Dit zorgt voor een algemeen toepasbare methode voor het aanbrengen van instal-
latie effecten rondom een wished-in-place paal die het gedrag van een in de grond
gedrukte paal juist weergeeft. In eerste instantie is een meervoudig regressie algo-
ritme toegepast ter benadering van de functies voor de Cartesische spanningen en
het poringetal voor een referentie situatie. De exponentiële vorm van de functies
sluit nauw aan bij de verdelingen. Vervolgens zijn deze functies uitgebreid met de
paal geometrie en de relatieve dichtheidseffecten.
Ondanks de beperkingen en vereenvoudigingen is aangetoond dat het idee om
installatie effecten te beschrijven in functie-vorm redelijk goed werkt en eenvoudig
kan worden toegepast in standaard EEM berekeningen. Om de voorgestelde functies
te valideren is een centrifugetest gemodelleerd. Als gevolg van het meenemen van de
installatie effecten rondom een wished-in-place paal kon het last-verplaatsingsgedrag
redelijk worden gemodelleerd.
Wanneer in de praktijk de installatie effecten in EEM berekeningen worden mee-
genomen kan een beter inzicht in het funderingsontwerp worden verkregen. Als
gevolg van de toegenomen draagkracht en stijfheid van de geanalyseerde grond-
verdringende paal kunnen de funderingskosten worden gereduceerd. Verder, als er
gebouwen in de omgeving van de paal aanwezig zijn, kan mogelijke verstoring van
of schade aan de constructie worden voorzien middels de EEM analyse.
Contents

Summary vii

Samenvatting (in Dutch) xi

List of Symbols xxix

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Scope and Limitations . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2 Background on Pile Installation 7


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Estimation of Bearing Capacity . . . . . . . . . . . . . . . . . . . . . 7
2.2.1 Empirical Correlations . . . . . . . . . . . . . . . . . . . . . . 8
2.2.2 Cavity Expansion and Strain Path Methods . . . . . . . . . . 9
2.2.3 Numerical Analysis . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Investigation of Installation Effects . . . . . . . . . . . . . . . . . . . 16
2.3.1 Stress Change . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3.2 Density Change . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3.3 Material Change . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3.4 Effects on Pile Behaviour . . . . . . . . . . . . . . . . . . . . 26
2.4 Including Installation Effects in the Analyses . . . . . . . . . . . . . 31
2.4.1 Empirical Methods . . . . . . . . . . . . . . . . . . . . . . . . 31
2.4.2 Numerical Methods . . . . . . . . . . . . . . . . . . . . . . . 33
2.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

3 Constitutive Modelling 37
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2 Hypoplasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.3 Hypoplastic Grain Crushing Model . . . . . . . . . . . . . . . . . . . 42
3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

xv
xvi Contents

4 Press-Replace Technique 51
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.2 Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.3 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.3.1 Cone Penetration in Undrained Clay . . . . . . . . . . . . . . 58
4.3.2 Pile Penetration in Undrained Clay . . . . . . . . . . . . . . . 63
4.4 Parametric study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.4.1 Element Type and Mesh Density . . . . . . . . . . . . . . . . 66
4.4.2 Step Size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.4.3 Interface Properties . . . . . . . . . . . . . . . . . . . . . . . 70
4.5 Discussion and Conclusions . . . . . . . . . . . . . . . . . . . . . . . 76

5 Investigation of Installation Effects 79


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.2 Numerical Modelling - Application of the PR Technique . . . . . . . 80
5.3 Results of the FE Simulations . . . . . . . . . . . . . . . . . . . . . . 83
5.4 Discussion of Results and Conclusions . . . . . . . . . . . . . . . . . 92

6 Generalisation of Installation Effects 97


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
6.2 Approximation of the Installation Effects by Interpolation . . . . . . 98
6.2.1 Method - Triangular Mesh Interpolation . . . . . . . . . . . . 98
6.2.2 Results and Discussions . . . . . . . . . . . . . . . . . . . . . 105
6.2.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
6.3 Approximation of the Installation Effects by Bivariate Nonlinear Re-
gression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6.3.1 Theoretical Background . . . . . . . . . . . . . . . . . . . . . 109
6.3.2 Describing Installation Effects by Surface Fitting . . . . . . . 110
6.3.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
6.3.4 Discussion of Results and Conclusions . . . . . . . . . . . . . 125
6.4 Generalisation of the Installation Effects . . . . . . . . . . . . . . . . 125
6.4.1 Determination of the Coefficients of Generalised Functions . . 126
6.4.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

7 Application 133
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
7.2 Verification Cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
7.3 Validation Case - Centrifuge Pile Jacking in Baskarp Sand . . . . . . 136
7.4 Discussion of Results and Conclusions . . . . . . . . . . . . . . . . . 143

8 Conclusions 145
8.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
8.2 Outlook for Further Research . . . . . . . . . . . . . . . . . . . . . . 150

Bibliography 152
Contents xvii

A Calculation of Unbalance 167

B Implementation of Functions 171

C Calculation of Pile Bearing Capacity Using Empirical Methods 173

Acknowledgements 185

Curriculum Vitae 187

Publications 189
xviii Contents
List of Figures

2.1 a. Definition of normalising parameter pcs ; Normalisation of the b.


base and c. shaft bearing capacity factors for Leighthon Buzzard sand
(after [90]) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Comparison of MPM and ALE methods: 1. Beginning of calculation
step in which the position, velocity, current strain and stress, state
variables and local element coordinates are known at time t 2. Map-
ping of particle velocities to nodes 3. (Updated) Lagrangian phase
in which the element is deformed and 4. the position, velocity and
strain, stress and state variables are updated at the material points.
5.a. The Lagrangian mesh is reset. 5.b. The mesh is updated arbit-
rarily (after [27]) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Large deformation (UL) analyses using frictional contact, where smooth
geometries for pile tip are used a. [63] b. [112] c. [77] . . . . . . . . . 15
2.4 Average horizontal and vertical stress distribution based on photoelastic
measurements ([56]) . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5 Radial stress distribution after 7 m of continuous pile jacking in a.
dense and b. loose Karlsruhe sand; c. radial stress distribution at 4m
depth (after [112]) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.6 Radial stress distribution of pile hammering in a. dense and b. loose
Karlsruhe sand; c. radial stress distribution at 3 m depth (after [112]) 17
2.7 Radial stress distribution of vibratory pile driving in a. dense and b.
loose Karlsruhe sand; c. radial stress distribution at 3 m depth (after
[112]) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.8 Change in porosity during pile penetration for different initial condi-
tions at different measurement locations (after [56]) . . . . . . . . . . 19
2.9 a. Test setup for the centrifuge pile tests; b. The sketch of the dens-
ity measurement principle, where the potential drop was measured
between the conductors; the continuous lines are the electric field and
the arrows show the current flow. . . . . . . . . . . . . . . . . . . . . 20
2.10 Displacement and volumetric observations in 1 g models using a. ra-
diograph measurement ([128]), b. thermal conductivity ([38]) and c.
imaging techniques (in crushable carbonate sand, [180]) . . . . . . . 20
2.11 Void ratio distribution after pile penetration simulations based on a.
FE with remeshing [78] and b. CEL [124] in medium dense Karlsruhe
and Mai-Liao sands, respectively . . . . . . . . . . . . . . . . . . . . 21

xix
xx List of Figures

2.12 Void ratio distributions of medium-dense (Id = 0.45) Karlsruhe sand


at several distances from the pile shaft for a. jacking, b. hammering
and c. vibratory driving installation methods ([112]) . . . . . . . . . 22
2.13 Grain crushing for flat pile base a. [164] b. [103] and for c. conical
base [103] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.14 Results of DEM simulation a. without and b. with particle breakage
[104] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.15 DEM results incorporating particle crushing for different pile types
[105] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.16 Idealisation of induced tensile stress in a particle and the crushing
mechanism [105] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.17 A general critical state line for sands [132] . . . . . . . . . . . . . . . 25
2.18 a. End bearing mobilisation, and b. shaft friction transfer curves
of displacement and non-displacement piles obtained at different seg-
ments of on the pile for a medium dense FF sand (after [45]) . . . . 26
2.19 Load transfer (t-z) curves for a. simplified elastic-perfectly plastic soil
and b. for strain hardening and softening soil (after [125]); dashed
lines represent the confinement and the material change effects. Mo-
bilisation of shear strength at different h/R ratios for c. simplified
elastic-perfectly plastic soil and d. for strain hardening and soften-
ing soil. Field observations on the mobilisation of e. shear strength
mobilisation and f. shaft friction [98] . . . . . . . . . . . . . . . . . . 28
2.20 a. Typical shaft friction distributions [56]; b. Residual loads during
and after installation [181] . . . . . . . . . . . . . . . . . . . . . . . . 29
2.21 Results from static loading tests on 12.8 m pile ([61] after [14]) . . . 30
2.22 Extended modification factor Ff∗ for a. Chiibishi sand, b. Dogs Bay
and c. Toyoura sand ([97]) . . . . . . . . . . . . . . . . . . . . . . . . 32

3.1 a. Comparison of the limit surfaces Matsuoka-Nakai and Mohr Cou-


lomb for different triaxial compression friction angles, φtc ; b. Geo-
metrical representation of the invariants tanψ and cos3ϑ in principal
stress space (after [158]) . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.2 a. Decrease of the maximum void ratio ei , the critical void ratio ec
and the minimum void ratio ed with an increase of the mean effective
pressure p b. Interaction between e, p and the density factor fd [22] 40
3.3 Different intergranular strains δ related with different deformation
histories. Only the recent part of the previous strain path (bold arrow)
has an influence on δ. Current stress, void ratio and strain at current
state * may be same for all three cases (after [120] ). . . . . . . . . . 41
3.4 Evolution of intergranular stiffness M a. for χ = 6, b. for χ = 2; ρ
varies from 0 to 1 [25]. . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.5 Comparison of original (dashed lines) and modified (solid lines) ref-
erence ratios: a. change of e∗i0 , e∗c0 and e∗d0 by σv ; b. change of e∗i , e∗c
and e∗d evolutions (after [129]). . . . . . . . . . . . . . . . . . . . . . 44
List of Figures xxi

3.6 Reference void ratio evolutions for primary loading - unloading - re-
loading cycles a. [129]; b. this study . . . . . . . . . . . . . . . . . . 47
3.7 Comparison of Oedometer compression curves for reference hypo-
plastic model [158] and grain crushing models, [129] and improved
crushing model (this study) . . . . . . . . . . . . . . . . . . . . . . . 48

4.1 Details on the Press-Replace modelling technique and progress of pen-


etration of the pile . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.2 Oedometer curves for different stress rate definitions and the PR
method (after [153]) . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.3 FE model for the volumetric analysis of a cone penetrating in an
undrained clay: a. global view of the mesh configuration at the be-
ginning of analysis; b. details on the mesh: green wedge indicating
the embedded cone head, inclined slices prepared for the cone to be
penetrated. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.4 Surface heave near the cone at the end of 0.40 m penetration (max.uy =
6.67 · 10−3 m) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.5 Relation between Ir and Nc obtained from PR and other techniques
(α = 1 full frictional/ rough, α = 0 frictionless/smooth contact). . . 60
4.6 Current plastic zones during steady state for soil rigidity indices: a)
Ir = 150 and b) Ir = 300, full frictional contact; c) Variation of
plastic zone with soil rigidity index based on simulation results of
RITSS (after Lu et al. [106]) . . . . . . . . . . . . . . . . . . . . . . 61
4.7 Plastic zones around the penetrating cone simulated using explicit
adaptive meshing with contact (after Walker & Yu [160] at different
penetration depths (Ir = 100 and α = 1.0, full frictional contact). . . 62
4.8 Plastic zones around the penetrating cone simulated using PR tech-
nique at different penetration depths (Ir = 100 and α = 1.0, full
frictional contact). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.9 Penetration of a frictionless embedded conical pile. Normalised hori-
zontal (left column) and vertical (right column) stresses (σxx /su and
σyy /su , respectively) for a. zipper technique by prescribed displace-
ments (Plaxis 2D) b. zipper technique by contact algorithm (Abaqus)
c. PR technique (Plaxis 2D) . . . . . . . . . . . . . . . . . . . . . . . 65
4.10 Comparison of distribution (smoothness) of effective mean stresses
(p ) of cases with high order (15-noded) elements (coarse mesh) and
lower order (6-noded) elements (medium mesh) at 1m penetration level. 67
4.11 FE meshes of slice thickness of a. ts = 1 cm, b. ts = 2.5 cm, c. ts
= 5 cm, and d. ts = 10 cm used in the comparison of step size and
mesh density (Dark grey is the pile material). . . . . . . . . . . . . . 68
4.12 Penetration resistance curves of cases with varied step sizes for two
different meshes uy = 2.5cm (blue lines) and uy = 5cm (red lines),
namely. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
xxii List of Figures

4.13 Penetration resistance curves of cases with step sizes of a. uy = 5 cm,


and b. uy = 10 cm for different slice thicknesses (ts = 2.5 cm, ts = 5
cm and ts = 10 cm) . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.14 Penetration resistance curves of cases with varied interface stiffness . 71
4.15 Normalised CPU times of all cases. . . . . . . . . . . . . . . . . . . . 72
4.16 a. Mobilised friction angle, φmob (first row) b. Void ratio, e (second
row) distribution at penetration level, z= 1.00 m for Cases 2, 5, 7 & 10. 73
4.17 Assumed stress-strain curves for the interfaces a. at the extensions b.
at the pile tip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.18 Comparison of penetration curves for cases with peak (φpeak ) and
critical (residual) interface strength parameter (φcritical ) . . . . . . . 74
4.19 Unloading issue evident from the load displacement curves (D = 0.40
- Id = 0.80) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.20 Stress distributions at previous phase (uy = 3.12 m) a. τ k−1 ; b σyy
k−1 k−1 
or σyy for horizontal interfaces) and c. σxx or σn for vertical
interfaces to be activated in the current phase uy = 3.16 m . . . . . 75
4.21 Comparison of available interface strength and the shear stress at the
beginning of step a. for horizontal interfaces at the pile tip, b. for
vertical interfaces at the pile shaft . . . . . . . . . . . . . . . . . . . 76
4.22 Improvement of load displacement (FT - uy ) behaviour by using very
strong interface extensions . . . . . . . . . . . . . . . . . . . . . . . . 77

5.1 a. Sketch of the problem modelled b. General view of the FE model 82


5.2 FE meshes prepared for PR analyses for cases with a. D = 0.30 m,
b. D = 0.40 m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.3 a. Illustration of the method for the calculation of the equivalent base
force at the pile tip; b. Top view of one radian slice and the area of
ith slice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.4 Total pile shaft, Fs , and base, Fb reaction curves obtained for piles
having diameters and penetrating in sands with different soil densities,
Id . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.5 Load test curves for different pile diameters and total penetration
lengths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.6 Comparison of results of the analysis using the PR technique with
the design load test a. base resistance, b. shaft resistance curves
suggested by the Dutch code (NEN 9997-2010) . . . . . . . . . . . . 87
5.7 Sketch of the problem analysed and the description of 3D surface plots 88
5.8 Idealised stiffness profiles of typical a. overconsolidated clay, b. sand,
c. soft clay (after [111]) . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.9 Comparison of a. the effect of density; b. penetration length, and
c. pile diameter on the distribution of normalised horizontal stresses,
Rrr . In the leftmost column the 3D surface plots from a reversed angle
of view, in the middle column the vertical cross-section at 0.01D from
the shaft, and in the rightmost column the horizontal cross-sections
at 0.5D below the tip are presented. . . . . . . . . . . . . . . . . . . 90
List of Figures xxiii

5.10 Comparison of a. the effect of density ; b. penetration length, and c.


pile diameter on the distribution of normalised vertical stresses, Rzz .
In the leftmost column the 3D surface plots from a reversed angle of
view, in the middle column the vertical cross-section at 0.01D from
the shaft, and in the rightmost column the horizontal cross-sections
at 0.5D below the tip are presented. . . . . . . . . . . . . . . . . . . 91
5.11 Comparison of a. the effect of density; b. penetration length, and
c. pile diameter on the distribution of normalised tangential stresses,
Rθθ . In the leftmost column the 3D surface plots from a reversed angle
of view, in the middle column the vertical cross-section at 0.01D from
the shaft, and in the rightmost column the horizontal cross-sections
at 0.5D below the tip are presented. . . . . . . . . . . . . . . . . . . 92
5.12 Comparison of a. the effect of density; b. penetration length, and
c. pile diameter on the distribution of normalised shear stresses, Rrz .
In the leftmost column the 3D surface plots from a reversed angle of
view, in the middle column the vertical cross-section at 0.01D from
the shaft, and in the rightmost column the horizontal cross-sections
at 0.5D below the tip are presented. . . . . . . . . . . . . . . . . . . 93
5.13 Comparison of a. the effect of density; b. penetration length, and
c. pile diameter on the distribution of normalised void ratio, Re . In
the leftmost column the 3D surface plots from a reversed angle of
view, in the middle column the vertical cross-section at 0.01D from
the shaft, and in the rightmost column the horizontal cross-sections
at 0.5D below the tip are presented. . . . . . . . . . . . . . . . . . . 94
5.14 Determination of the equivalent critical state mean stress, pcs in hy-
poplasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
5.15 Comparison of normalised a. base and b. shaft resistances during pile
penetration together with trend lines obtained for Leighton Buzzard
sand*[90]; and c. average void ratio evolution under pile base with ec
reference curve (dashed line) of the hypoplastic model using Baskarp
sand parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

6.1 Convex Hull of a PR discretisation (only sand region considered). . . 99


6.2 Voronoi diagram and Delaunay tessellation (after [13]) . . . . . . . . 100
6.3 Nearest Neighbour algorithm for a. 1D uniform, b. 2D uniform data,
defined by stepwise discrete functions. . . . . . . . . . . . . . . . . . 101
6.4 Nearest Neighbour algorithm for a scattered surface data represented
by stepwise discrete functions based on the regions defined by Voronoi
diagram; a. 2D representation, b. 3D representation . . . . . . . . . 101
6.5 a. Voronoi diagram based on original data points 1,2,3,4 & 5; b.
Second order Voronoi cells around the sample point X where the ratio
of the enclosed areas Aabf, Aaehf, Aehd, Adcgh & Abcgf to Aabcde
are used to determine the percent contribution from neighbouring
cells of original data as weights [144]. . . . . . . . . . . . . . . . . . . 102
xxiv List of Figures

6.6 The extended line of points and original mesh stress points (D = 30,
L = 10D). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
6.7 FE discretisation of a. PR, b) PR coarse (PR-C), and two extreme
discretisation cases: c. finest mesh (F3) and d. coarsest mesh (C3)
for case D = 0.30 m L = 10D . . . . . . . . . . . . . . . . . . . . . . 104
6.8 Comparison of normalised load displacement behaviour of different
discretisations to which the results of the equalisation level of the PR
analysis were imposed by interpolation. Lower curves are the load
test results for K0 state, i.e. installation effects were not imposed. . 106
6.9 Comparison of a. imposed, and b. equalised interpolation distribution
of σzz . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
6.10 Deviations from imposed (interpolated) state. . . . . . . . . . . . . . 108
6.11 Geometrical properties of the grid prepared to precondition the dis-
tribution of surface data for bivariate nonlinear regression calculations.113
6.12 Comparison of window size effect of the median filter on the σrz dis-
tribution of D = 0.30 m; L = 2.5D; Id = 0.40 . . . . . . . . . . . . . 115
6.13 Comparison of filtered data with PR data of σrz distribution of D =
0.30 m; L = 2.5D; Id = 0.40 a. σrr , b. σzz , c. σθθ , d. σrz ,and e. e . 116
6.14 Fitting results of normalised vertical stresses ψ̂rr of cases: a. D =
0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . 117
6.15 Fitting results of normalised vertical stresses ψ̂zz of cases: a. D =
0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . 118
6.16 Fitting results of normalised hoop stresses ψ̂θθ of cases: a. D = 0.30m;
L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . . . . . 118
6.17 Fitting results of normalised shear stresses ψ̂rz of cases: a. D =
0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . 119
6.18 Fitting results of normalised void ratios ψ̂e of cases: a. D = 0.30m;
L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . . . . . 119

6.19 Comparison of the surface fit of σrr with PR data of cases: a. D =
0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . 120

6.20 Comparison of the surface fit of σzz with PR data of cases: a. D =
0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . 120

6.21 Comparison of the surface fit of σθθ of cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . . . . . . . . . . . 121

6.22 Comparison of the surface fit of σrz with PR data of cases: a. D =
0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . 121
6.23 Comparison of the surface fit of e with PR data of cases: a. D =
0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . 122
6.24 Comparison of numerical load test curves of the imposed fields using
the fitting results with PR result and K0 state of cases: a. D = 0.30m;
L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . . . . . 122
6.25 Change of approximated field imposed (red), after equilibrium (soil,
pink) and after pile and interfaces were activated (pile+NiL, green) a.
radial and b. vertical directions of case D = 0.30m; L = 10D; Id = 0.60123
List of Figures xxv

6.26 Distribution of radial unbalanced stresses for cases: a. D = 0.30m;


L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . . . . . 124
6.27 Distribution of vertical unbalanced stresses for cases: a. D = 0.30m;
L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . . . . . 124
6.28 Approximation result for the coefficient a11 and a31 (of ψ̂rr ) in L/D,
Id space. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.29 Approximation result for the coefficient b31 and r31 (of ψ̂rr ) in L/D,
Id space. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.30 Comparison of radial stress, σrr distribution of generalised form with
PR results of the cases: a. D = 0.30m; L = 10D; Id = 0.60, b.
D = 0.40m; L = 10D; Id = 0.80 . . . . . . . . . . . . . . . . . . . . . 128

6.31 Comparison of vertical stress, σzz distribution of generalised form
with PR results of the cases: a. D = 0.30m; L = 10D; Id = 0.60, b.
D = 0.40m; L = 10D; Id = 0.80 . . . . . . . . . . . . . . . . . . . . . 129

6.32 Comparison of tangential stress, σθθ distribution of generalised form
with PR results of the cases: a. D = 0.30m; L = 10D; Id = 0.60, b.
D = 0.40m; L = 10D; Id = 0.80 . . . . . . . . . . . . . . . . . . . . . 129
6.33 Comparison of shear stress, σrz distribution of generalised form with
PR results of the cases: a. D = 0.30m; L = 10D; Id = 0.60, b.
D = 0.40m; L = 10D; Id = 0.80 . . . . . . . . . . . . . . . . . . . . . 130
6.34 Comparison of void ratio, e distribution of generalised form with PR
results of the cases: a. D = 0.30m; L = 10D; Id = 0.60, b. D =
0.40m; L = 10D; Id = 0.80 . . . . . . . . . . . . . . . . . . . . . . . 130
6.35 Comparison of radial unbalanced stresses for cases: a. D = 0.30m;
L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . . . . . 131
6.36 Distribution of vertical unbalanced stresses for cases: a. D = 0.30m;
L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80 . . . . . . . 131
6.37 Comparison of the numerical load test curves based on the equalised
approximate installation effects with the state obtained using PR and
the reference K0 state of cases: a. D = 0.30m; L = 10D; Id = 0.60,
b. D = 0.40m; L = 10D; Id = 0.80 . . . . . . . . . . . . . . . . . . . 132

7.1 FE mesh used for the verification cases a. D = 0.30 m and b. D =


0.40 m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
7.2 Comparison of the numerical load test curves of the PR results (in-
terpolated on the reference mesh), the equalised approximate install-
ation effects and the reference K0 results of the verification cases a.
D = 0.30 m; Id = 70; and b. D = 0.40 m; Id = 50, both having a pile
length, L = 6D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
7.3 Comparison of the numerical load test curves of the PR results (in-
terpolated on the reference mesh), the equalised approximate install-
ation effects and the reference K0 results of the verification cases a.
D = 0.30 m; Id = 70; and b. D = 0.40m; Id = 50, both having a pile
length, L = 8D. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
xxvi List of Figures

7.4 a. Physical model test setup [56] (all units are in mm; acceleration
level, Ng = 35), and b. FE model of the validation case . . . . . . . 137
7.5 Comparison of the numerical load test curves with the centrifuge test
result [56] based on the equalised approximate installation effects with
the state obtained using formulas and the reference K0 state of the
dense case, e0 = 0.637; Id = 0.77. . . . . . . . . . . . . . . . . . . . . 138
7.6 Comparison of the numerical load test curves with the centrifuge test
result [56] based on the equalised approximate installation effects with
the state obtained using formulas and the reference K0 state of the
medium dense case, e0 = 0.709; Id = 0.58. . . . . . . . . . . . . . . . 138
7.7 Comparison of the numerical load test curves with the centrifuge test
result [56] based on the equalised approximate installation effects with
the state obtained using formulas and the reference K0 state of the
loose case, e0 = 0.783; Id = 0.38. . . . . . . . . . . . . . . . . . . . . 139
7.8 Comparison of total pile capacities . . . . . . . . . . . . . . . . . . . 141
7.9 Comparison of pile a. base and b. shaft capacities . . . . . . . . . . 141

8.1 Comparison of the numerical load test curves using formulas employed
with using original stiffness (hs ) and high stiffness (100 · hs ) for the
medium dense case, e0 = 0.709;Id = 0.58. . . . . . . . . . . . . . . . 151
List of Tables

3.1 Hypoplastic soil model parameters for Baskarp sand [8] . . . . . . . 47


3.2 Estimated grain crushing parameters [129] for the modified grain
crushing model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

4.1 Parameters and results of the cases analysed for volume preservation
check (c = 50 kPa) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.2 Parameters and results of the cases analysed for volume checking control 63
4.3 Hypoplastic soil model parameters for Baskarp sand (after [8]) . . . 66
4.4 Summary of analysis for the investigation of step size . . . . . . . . . 68
4.5 Summary of analysis for the investigation of mesh density . . . . . . 68
4.6 Summary of cases analysed in parametric study on the interface stiffness 71

5.1 Hypoplastic soil model parameters for Baskarp sand (after [8]) . . . 82
5.2 Summary of variations analysed for the investigation of pile diameter
(D), penetration length (L) and density (Id ) effects on the installation
field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

6.1 Geometrical properties of the regions defined for the interpolation grid 113
6.2 Summary of regression analysis of a)D = 0.30m; L = 10D: Id = 0.60
and b)D = 0.40m; L = 10D: Id = 0.80 . . . . . . . . . . . . . . . . . 117
6.3 Summary of regression analysis on the generalisation of the normal-
ised radial stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

7.1 Summary of improvement levels of the verification cases at uy = 0.10D136


7.2 Summary of improvement levels of the validation cases at uy = 0.10D 137
7.3 Summary of pile total capacity prediction accuracies of proposed func-
tions and empirical methods . . . . . . . . . . . . . . . . . . . . . . . 142
7.4 Summary of pile base and shaft capacity prediction accuracies of pro-
posed functions and empirical methods . . . . . . . . . . . . . . . . . 142

xxvii
xxviii List of Tables
List of Symbols

Latin

c cohesion
Cu coefficient of uniformity
D pile diameter
Deq radius of the area of influence for density change effects
d50 mean grain size
d60 , d10 sieve opening size through which 60% and 10% of the particles pass
e void ratio
e d , ec , ei minimum, critical, and maximum void ratio at current stress
ed0 , ec0 , ei0 reference minimum, critical, and maximum void ratio at low stress
ref
Eoed , Eoed oedometric stiffness at current and reference stress levels
f load vector
Ff ∗ modification factor considering compressibility of the sand
Ft , Fs , F b pile head, shaft, and base reaction
g gravitational acceleration
G shear modulus
Gi , Gs shear modulus of interface, and soil
h distance from pile tip
hs granulate hardness
Id density index
Ir rigidity index
K global stiffness matrix
k S , kN interface shear and normal stiffness
K0 at rest earth pressure coefficient
K0N C at rest earth pressure coefficient for normally consolidated soil
L pile (penetration) length
m exponent of stress dependent stiffness
mR , mT hypoplastic model stiffness modification factors
n number of piles within the influence area or
exponent of granulate hardness
Ng centrifuge acceleration ratio
Nq bearing capacity factor
p effective mean stress

xxix
xxx List of Symbols

pa atmospheric pressure (100 kPa)


pref reference mean stress
p0 effective mean stress at geostatic stress state
P1 , P2 , P3 contact forces on a particle P1 > P2 > P3
q b , qc , qs base, cone and shaft resistance
qb , max pile base resistance at maximum settlement level
Qu ultimate load capacity
r radial distance or normalised centre to centre distance b/w
adjacent piles
R pile radius or particle roundness
Rs normalised shaft resistance
s settlement
S particle sphericity
SD normalised displacement
smax maximum settlement
SP T − N the standard penetration resistance number
te , ts element, and slice thickness
u displacement
ur displacement normal to pile shaft
uy vertical prescribed displacement applied at pile head
v specific volume
vc r, vf offset of crushing and end of crushing lines

Greek

α exponent controlling volumetric behaviour in hypoplastic


law or interface strength reduction factor
α p , αq , βp , βq modification factors for reference void ratios
β normalised shaft resistance or
exponent controlling volumetric behaviour in hypoplastic
law
δ frictional coefficient of pile-soil interface
δ intergranular strain tensor
Δe change in void ratio
Δσr change in normal effective stress on pile shaft during
loading due to change in soil volume
ε̇ strain rate
H Henky strain
η improvement factor
φ internal friction angle of the soil
φc critical friction angle
φtc triaxial compression friction angle
ψ state parameter
ψ̂rr , ψ̂zz , ψ̂θθ , ψ̂rz , ψ̂e model functions of Cartesian stress and void ratio fields
λ0 , λcr , λf slope of the critical, crushing and end of crushing state
xxxi

lines
λ load scaling factor
σ Cauchy stress
σ̊ Jaumann stress rate
σm mean normal stress
 
σrf , σrc radial effective stress acting on pile shaft at failure and
after installation
σt tensile stress in a particle

σv0 geostatic effective vertical stress
σxx horizontal stress
σyy vertical stress
σ 1 , σ3 major and minor principal stress
τmax , τres peak and residual shear strength
τrz shear stress at pile shaft, at axisymmetric conditions
τs shear strength of the pile shaft
xxxii List of Symbols
Chapter 1

Introduction

1.1 Background
One of the most traditional methods to support structures resting on soft soils is
the use of piles. They generally work by transferring the loads to deeper soil layers,
which can provide enough bearing capacity when mobilised. This type of foundations
has been commonly used around the world. In the western part of The Netherlands
this type of foundation is often used due to the typical soil profile which consists of a
thick (10-20 m) soft soil layer underlain by a stiff bearing stratum composed mostly
of quartz sand.
Piles need to be strong and durable. In the past, wooden piles were used some-
times decay in time and hence lose their carrying capacity. Quite common tilted
Dutch houses are clear indication of the significance of this issue or overloading due
to the neglected downdrag. In current practice, piles made up of pre-cast reinforced
concrete or steel are used.
Piles can be installed in different ways. The most common way to install the
prefabricated piles is the use of an impact hammer. In this driving method, the pile
is displaced by each drop or blow of the ram until the desired penetration level is
achieved. Another alternative is the vibratory driving technique where the pile is
forced to penetrate by a heavy vibratory head on top. The vibrations degradate
the strength of (or liquefy if saturated) the surrounding soil and due to the heavy
weight of the vibrator the pile penetrates. Recent developments have allowed the
piles to be jacked resulting in reduced nuisance and higher capacity and stiffness as
compared to the hammering and vibratory driving techniques ([55], [76]).
The bearing capacity of a pile depends on the soil properties and the stress state.
The behaviour of granular material is governed by the packing of the grains and the
contact stresses between. The stress and density dependent behaviour determines
the soil response. The effects of the installation operation are transmitted to the
soil through the interaction of sand grains and the pile, resulting in an altered soil
state and properties.
The installation effects depend on the applied installation technique. The change
of stress and density as well as change of the physical properties (e.g. grain size dis-

1
2 1. Introduction

tribution) of the soil in the vicinity of the pile installed can be quite different for
different installation methods. The resulting confining stresses of a dynamic install-
ation technique (hammering or vibratory driving) are lower than those of the jacking
technique. Depending on the density, dilation would be expected near the pile shaft
during jacking (e.g. [38], [56]). On the other hand, dynamic installation methods
can cause compaction or loosening near the pile depending on the method and state
of the soil. For example impact driving, and for the loose and medium dense sands
vibratory driving, causes compaction ([83]), while for the dense sand vibratory driv-
ing causes loosening in the vicinity of the pile. The numerical simulations of [112]
indicated compaction around the pile for all dynamic installation methods and all
soil densities.
Two important consequences of the installation effects that should be considered
in the design are:

ˆ The influence on the performance of the pile (e.g. load–displacement beha-


viour, capacity) in its service life ([102], [56]). For example, if the density of
the soil as well as the confining stress around the pile is high, a higher bearing
capacity will be mobilised with less displacement.
ˆ The influence on the neighbourhood. For example, dynamic installation tech-
niques have often nuisance (e.g. noise, vibrations in the neighbourhood) issues
([171], [118]) in populated areas. Furthermore, the technique can cause dam-
age on the neighbouring structures [172]. The induced vibrations, depending
on the soil type, might amplify the vibrations and cause compaction of found-
ation strata. That can cause structural damage to the neighbouring struc-
tures, mostly due to induced settlements of the foundation soil. In the static
penetration case (e.g. jacking) the neighbouring underground structures (e.g.
prefabricated pile, sheet pile wall) may be influenced by the resulting displaced
soil and the increased stress level ([76]).

A more realistic behaviour and therefore an improved design would be achieved


by considering the installation effects in the analyses than performing the analysis
considering a geostatic stress state around the pile(s) modelled. In the current
practice, the installation effects are taken into account by some empirical design
methods in order to estimate the bearing capacity of foundation piles [123]. Several
field and model tests performed to investigate the influence of pile installation on the
bearing capacity, have led to only an evolution in empirical models to estimate the
bearing capacity of displacement piles [87], [125]. Coop et al. [46] attempted to take
the soil state around the pile into account by measuring both the contact stresses on
the pile as well as displacements of the sand around the pile in model tests, but did
not directly measure density changes in the zone influenced by the pile installation.
Recent attempts to investigate the change in the soil state were also limited to
either the measurement domain (generally close to the pile) or resolution as well
as the variables (e.g. displacement, strain, stress, density) that can be quantified
[164], [56], [103]. However, the behaviour of piles during installation as well as
the interaction with the surrounding soil and the change of soil properties during
1.1. Background 3

installation are still not well known. This information is essential, not only to make
better predictions of the pile bearing capacity and its behaviour in the soil under
different loading conditions, but also to predict the (side) effects of pile installation
on the neighbourhood.
The increase in computational power allowed improved numerical techniques like
the finite element method (FEM); however installation effects of displacement piles
have not been incorporated in the numerical models for engineering purposes. As a
result of discarding the installation effects there will be a large difference between
the FE predictions and the real behaviour (performance) of the pile [34]. Therefore
it is important to model the installation effects of displacement piles numerically so
that the load displacement behaviour of the pile during its service life as well as the
secondary effects on neighbouring structures would be predicted more accurately.
For a proper analysis of the pile installation and its consequent effects, all import-
ant aspects should be included in the numerical model. These can be summarised
as:

ˆ Large deformation effects, which require geometry update and use of objective
stress-strain measures (e.g. Truesdell, Green-Naghdi, Jaumann). When mod-
elling the penetration there are local large deformations that cause stress con-
centrations and distortion of the mesh. Unless properly modelled the analysis
will result in quite a different state than that would be expected. For example
geometry should be updated in order to account for penetration and obtain
more realistic displacement, strain and therefore stress field. Furthermore, the
stress (and conjugate strain) measures should be objective (e.g. Jaumann rate
of stress) so that the rigid body rotations would not cause any stress or strain
increase [24]. Additional volumetric and shear deformation effects in the ob-
jective stress measure such as Hill and Truesdell rate of stress, respectively,
would be more appropriate. However, to obtain a realistic response, these ob-
jective stress rates should be checked for certain stress–strain paths (like 1-D
compression, [153]).
ˆ Pore pressure effects, which have to be considered when the loading is fast
enough to cause undrained behaviour of the soil. If the soil is fully saturated
the pore pressure generation and dissipation should also be modelled.
ˆ Inertia effects, which depend on the rate of loading, should also be considered
for a proper modelling. The intrinsic properties of the material, the coupled
system (e.g. saturated dense sand) as well as the rate of loading determine
the response of the system. The displacements, velocities and accelerations
mobilize corresponding components of the system. As a conclusion, inertia
effects should be considered in the analysis of dynamic installation techniques.
ˆ Use of an appropriate constitutive model is quite important to mimic com-
mon characteristics of the soil considered. As most common bearing strata is
quartz sand in the Netherlands, the model used should have a state dependent
behaviour (e.g. Critical State Models) in which the soil stiffness and strength
is dictated by current state. Furthermore, at high stress levels the sand grains
4 1. Introduction

can be crushed. As a result, the material properties will change. Therefore,


the effect of grain crushing should be also considered in the constitutive model.

Nevertheless, simplifications are inevitable. For pile jacking, which is considered


as a quasi-static penetration, the inertia effects can be discarded and drained con-
ditions can be assumed (no excess pore pressure effects). Therefore in the resulting
FE model, the level of complexity will be reduced.

1.2 Scope and Limitations


This study is the numerical part of a larger project in which the installation effects
of displacement piles are investigated both experimentally and numerically. The ob-
jective of this numerical study is to investigate the installation effects of pile jacking
in sand in a numerical framework. Since the jacking operation can be considered as
quasi-static loading, the dynamic effects are not considered in the analyses. For the
same reason, drained conditions are assumed such that the pore pressures are only
taken into account as a reduction of total stress (giving the effective stresses). As a
common tool used in engineering practice, FE modelling is selected as the numerical
method.
In order to model pile jacking using the FEM in a standard FE code, some addi-
tional aspects have to be considered to account for the large deformation effects. In
this perspective possibilities to simulate installation effects as well as the installation
process were investigated and a simplified technique (the Press-Replace technique)
is proposed. After a sensitivity study, the method is employed to investigate the
installation effects around a jacked pile for different sand densities.
In the view of the results obtained from FE simulations, the possibilities of
incorporating the installation fields were investigated. Furthermore, describing the
installation effects in a functional form, which allows flexibility for imposing the
installation field around a so-called wished-in-place pile to model it as a jacked pile,
was studied.
A novel technique is introduced to describe the installation effects of a jacked pile
without simulating the penetration process. This way a jacked pile can be modelled
wished-in-place and corresponding installation effects (i.e. stress, density and stiff-
ness change) can be imposed. By the proposed method, there will be an enormous
gain in term of computational effort and time in the analyses of displacement piles.

1.3 Outline
The thesis consists of 8 chapters. First current practice on investigation and design
methods of the pile installation and consequent effects are explained (Chapter 2).
The important features of the constitutive model used in the analyses and the grain
crushing modifications are presented (Chapter 3). The Press-Replace technique,
which is developed for standard FE algorithm, is explained (Chapter 4). The method
is employed to investigate geometric and material effects on the installation field
1.3. Outline 5

Chapter 5. The possibilities of the representation of the installation effects and


their incorporation in a standard FE code is presented in Chapter 6. The method
for incorporation of the installation effects is validated by a centrifuge model test
in Chapter 7. Finally, in Chapter 8 the findings of this study are summarised and
some recommendations for future research are given.
6 1. Introduction
Chapter 2

Background on Pile Installation

2.1 Introduction
This chapter presents a summary of current practice on the prediction of pile bearing
capacity, the available methods for the investigation of the installation effects and
the existing methods for the incorporation of these effects in the analyses. As stated
in the previous chapter, considering installation effects in the analyses is important
to obtain a realistic behaviour, which results in a better design. However, most of the
current analysis techniques include little or none of the important components of the
installation effects (e.g. stress increase, density change). Section 2.2.1 presents the
empirical correlations used in practice to include the installation effects to determine
the bearing capacity of a pile. Analytical and semi-analytical methods like cavity
expansion and strain path methods (SPM) have been used to determine the bearing
capacity of piles (Section 2.2.2). Section 2.2.3 summarizes the first finite element
models for the investigation of the bearing capacity of a pre-embedded pile. Section
2.3 summarizes the effects of pile installation in the soil (Section 2.3.1 - 2.3.3) and
on the pile behaviour (Section 2.3.4). Lastly Section 2.4 highlights the essential
components that were missing in the methods used for calculation of pile bearing
capacity and stiffness, by the field, laboratory and numerical observations on the
effects of the installation process on the pile behaviour. In Section 2.4.2, available
empirical and numerical methods to include the installation effects in the analyses
are examined.

2.2 Estimation of Bearing Capacity


Pile bearing capacity is defined as the sum of shaft and tip resistances when fully
mobilised. The best way to assess the capacity and more importantly, the load-
displacement behaviour (stiffness) is to carry out a (static) pile load test. The test
provides valuable information on pile behaviour under expected loading conditions.
It is known that the shaft is mobilised with a small displacement (few millimetres),
much before the tip is mobilised (displacement greater than 0.10D, where D is the

7
8 2. Background on Pile Installation

pile diameter). However, many pile types do not show a distinct failure level as the
capacity increases by further mobilisation of the pile (i.e. increasing load on pile
head). In order to estimate the failure load from a static load test or the bearing
capacity of a pile, some empirical methods can be employed (e.g. [32], [51], [50]). In
general, the ultimate limit state capacity of a pile, Qu is assumed to be the capacity
at 0.10D displacement of the pile during the load test. The design capacity can be
determined either by a displacement criterion (e.g. 0.01D) or by a load criterion
(e.g. 1/1.5 − 1/2.5 · Qu ) [67]. It is not always feasible to carry out a pile load test.
Other field tests, can also provide valuable information.
In the following sections, several methods and correlations are summarised, some
of which include the installation process or the resulting effects. In Section 2.2.1 pile
bearing capacity correlations based on field tests (direct methods) are presented. In
Section 2.2.2, analytical and semi-analytical cavity expansion and strain path meth-
ods (SPM) are presented. These analyses use soil parameters that are determined in
standard laboratory tests and the bearing capacity is indirectly calculated. There-
fore these methods are called indirect methods. Lastly, in Section 2.2.3, numerical
methods (e.g. FEM, DEM), which are used to analyse the pile penetration process
and estimate the pile bearing capacity and behaviour, are presented.

2.2.1 Empirical Correlations


A common and efficient way to determine the bearing capacity of a pile is using
empirical correlations based on field tests (e.g. SPT, CPT). As the strength of the
soil is directly measured and correlated to the pile base and shaft capacity, these
methods are called ’Direct Methods’.
A first selection of pile capacity estimation methods can be grouped under Stand-
ard Penetration Test (SPT) correlations. Besides the SPT N-value [126], some of
these methods include the pile geometry [116], [117], the soil type as well as the pile
type [11].
The second group, the most commonly used pile capacity estimation technique,
comprises the cone penetration test (CPT) correlations. These methods, such as the
Dutch Method ([154], [54], [3]), the French Method ([36], [1], [65]) and the Belgium
method ([51], [2]) differ mostly in the determination of a representative qc value
for the pile base (e.g. which qc values to consider, what kind of averaging should
be used, etc.). In these methods different factors can be applied for the resistances
obtained from field tests or static pile load tests [51], [2]. The representative qc value
for the pile base is then multiplied by several factors to account for the degree of
mobilisation, the direction of loading (e.g. Fugro-05 method [92] for tensile loading),
the pile type (e.g. [11], [137]), the pile size (e.g. ICP-05 method [88]), the soil density
and the soil type (e.g. the Dutch Method) or a combination of these aspects (e.g.
LCPC method, UWA-05 method [100]).
Due to similar failure pattern observed at the CPT cone, the cone resistance,
qc , and the pile base bearing capacity are strongly correlated [101]. From the same
perspective, it would be expected that the shaft friction of the CPT, qf , and the
pile shaft resistance, qs , are also correlated. However, it was shown that qc gives a
2.2. Estimation of Bearing Capacity 9

Figure 2.1: a. Definition of normalising parameter pcs ; Normalisation of the b. base and c. shaft
bearing capacity factors for Leighthon Buzzard sand (after [90])

better prediction of qs [36], [139], [100]. Furthermore, the use of qc for the prediction
of shaft capacity was suggested. That is a good indication of the inaccuracy of the
shaft friction measurements.
Based on their study on model piles tested in a centrifuge, Klotz & Coop [90]
presented that the initial state of the soil around the pile, which is defined as the
combination of density and stress condition relative to the critical state line, was a
controlling factor in determining the pile capacity. They indicated that the design
based on only relative density would be misleading. Furthermore, they proposed to
define the state as a ratio of stresses rather than the common description based on
void ratio or specific volume (Fig.2.1.a). Fig.2.1.b&c shows the trend of Nq and β for
Leighton Buzzard sand tested at different stress and density levels. For example, the
more the in-situ void ratio deviates away from the Critical State Line (CSL), (i.e. the
denser the soil), the higher the pile bearing capacity is. The base resistance factor,
Nq , and shaft resistance factor, β, can be estimated using the proposed correlations
given in Figure 2.1.a&b. Then, using the Nq and β factors the pile end bearing and
unit shaft capacities can be calculated using:

qb = Nq · σvo (2.1)

qs = β · σvo (2.2)
The methods mentioned above include the installation effects either directly (as
in the case of SPT or CPT) or indirectly by empirical factors accounting for the
installation type (e.g. driven or jacked). The predictive performances depend on
pile type and loading type ([100], validations: [176]).

2.2.2 Cavity Expansion and Strain Path Methods


A more rigorous way to determine the pile bearing capacity is using an analytical
or semi-analytical method such as Cavity Expansion and Strain Path Methods.
Depending on the mechanics of the physical process, cavity expansion solutions
10 2. Background on Pile Installation

assume either spherical or cylindrical expansion from a finite radius. As the cavity
expands to its final level, which is the surface of the pile installed, it mobilises the
surrounding soil and increases the stress level. In this way the installation stresses
are assumed to be simulated. In the following paragraphs, the introduction and
development of the cavity expansion models are given. Later, an improved semi-
analytical method, the Strain Path Method (SPM), and its extensions are presented.
The first solid mechanics applications of the cavity expansion theory can be found
for metal indentation problems with cylindrical punches with conical heads [28], [82].
The method was later extended to predict pile or cone end bearing capacity [70],
[156], [17], [157], [107], [37]. Yu & Houlsby [185] introduced solutions using large-
strain theory for the plastic region and small strain theory for the elastic region.
A more realistic constitutive model for the soil, like elastic-perfectly plastic critical
state model based on the state parameter concept [23] is employed to solve cylindrical
and spherical cavity expansion problems [44]. A similar critical state model is used
by Shuttle & Jefferies [141] to solve the cavity expansion problem to interpret CPT
results. More advanced spherical and cylindrical cavity expansion solutions using
large-strain analysis by employing a hypoplastic constitutive model, which is a type
of critical state model for sand, are presented by Cudmani & Osinov [47]. The model
is used for the interpretation of CPT and pressuremeter tests.
The cavity expansion theory provides useful analytical solutions for determin-
ing the tip resistance of a penetrating cone especially in undrained cohesive soil;
however, as van den Berg [152] stated, the penetration process is not identical to
expansion of a spherical or cylindrical cavity. In the cavity expansion solutions pen-
etration of a frictionless cone is assumed. This assumption results in the lack of
an important component, i.e. the shearing effect of the penetrating cone. In order
to overcome this limitation, other methods are combined with the cavity expansion
theory. Sagaseta & Houlsby [133] assumed steady penetration of an infinite rigid
cone in an elastic-plastic material with von Mises yield criterion and corresponding
solutions are given for different cone roughness (friction). The cavity expansion the-
ory was also combined with bearing capacity theory: Baligh [17] divided the cone
penetration process into two parts: the work required to give a virtual vertical dis-
placement of the cone tip (bearing capacity theory) and the work required to expand
the soil around the cone and shaft in horizontal direction (cavity expansion theory).
However, this type of combined behaviour results in an overestimated total capa-
city. The cavity expansion theory was also combined with FEM [4] for the analysis
of penetration in clay. The penetration process was simulated in two phases using
the modified Cam Clay model. First the expansion of the cone from a small radius
to its actual radius was simulated. This part of the simulation is quite similar to
cylindrical cavity expansion. Then, the continuous penetration was modelled by
applying incremental vertical displacements at the cone and soil interface.
Another important contribution is the Strain Path Method (SPM) by Baligh
[18], in which a steady state flow around the cone was assumed. The strain paths
around the cone were calculated using the strain rates obtained by differentiating the
estimated velocity fields. A constitutive relation was introduced and the stress field
around the cone was obtained by enforcing equilibrium condition. Acar & Tumay
2.2. Estimation of Bearing Capacity 11

[5] give an improved strain field during penetration in clay. Teh [148] combined
the SPM with a Finite Element model by introducing the strain field predicted
from SPM as an initial state and calculating the corresponding stresses using the
FE model. Yu & Whittle [186] used SPM for the simple pile model solution of
Baligh [19] to estimate the size of the plastic zone and then used the spherical
cavity expansion theory to determine the stress field. [71] obtained strain fields for
different fluid viscosities using FEA for different penetrometer roughness (boundary
conditions) as an extension of SPM. They showed that the shape and extent of
the strain field depend on the viscosity of the material and the roughness of the
penetrometer. Although the mechanism around a penetrating probe can be assumed
as flow, the behaviour of soils is quite different than fluid flow. The latter is generally
modelled by the Navier-Stokes equations. On the other hand, soils exhibit dilative-
contractive as well as history and stress dependent behaviour. Therefore, the strain
paths assumed in the SPM solutions are probably different for soils. More recently,
Einav & Randolph [58] combined upper bound solutions and SPM (UBSPM) to
evaluate penetration resistances of a new generation of penetrometers, which have a
greater projected area than the cone shaft. The incompressibility condition assumed
for the SPM to model the flow field was optimized for an ideal perfectly plastic
material (Tresca or Von Mises). They predicted the bearing capacity better than
the one obtained by a standard SPM solution.
Yu et al. [183] developed a novel finite element formulation for the analysis of
steady state cone penetration in undrained clay using von Mises and modified Cam
Clay models. In this approach the cone is fixed and the stress-strain state of soil
elements flowing around the cone is integrated along the flow path (similar to SPM).
However there are two main approximations in this method, the tip nodal force is
zero and the stress history is integrated over columns of soil elements where the
radial movement of the soil is ignored.
All these methods summarised in this section provide a good approximation for
the determination of bearing capacity of a cone or displacement pile. The effects of
installation are considered inherently. Therefore the bearing capacities calculated
by these methods do not require an extra correction for the installation effects.
One of the important limitations of these methods, except for the ones combined
with FE analyses, is the inability to obtain load-displacement behaviour of the pile.
Therefore, even though the bearing capacity can be obtained satisfactorily, it is not
possible to check the stiffness or load-displacement behaviour of the installed pile by
these methods. The next section focuses on the methods employing the FEM.

2.2.3 Numerical Analysis


The bearing capacity of a pile has been analysed by FE methods first by small-strain
formulations, e.g. [53], [72] and [143] for cohesive materials. In these simulations
a pre-embedded cone in an in-situ stress state is forced to reach its collapse load.
Although these models provide a good basis for the penetration problem, they cannot
capture the correct initial conditions that is a result of large deformations.
12 2. Background on Pile Installation

One of the attempts to capture large deformation behaviour is proposed by Hibbit


et al. [81]. They developed an incremental FE formulation for the large displace-
ment, finite strain problem. Later [177] and [178] introduced a generalised Updated
Lagrangian (UL) procedure, in which 2nd Piola-Kirchoff stress and Green’s strain
rates were employed. This method was further elaborated and extended to dynamic
problems by Bathe et al. [20]. By this approach the mesh is modified after each
calculation step according to the calculated displacements, which captures the geo-
metric nonlinearity in a better way. However it is limited in the sense that the
method is not suitable for very large (and especially local) deformations. Excess-
ive mesh distortion results in highly inaccurate solutions or even divergence of the
FE solution. The procedure is improved by employing interface elements as done
by Cividini & Gioda [42]. In this way it was also possible to investigate different
roughnesses of the cone.
The Discrete Element Method (DEM) enables a more realistic modelling of the
soil medium by simulating the mechanism of the interaction of soil particles. How-
ever, it requires a proper description of the contact between particles. Furthermore,
the huge computational demand of DEM limits the element (particle) size for a
typical boundary value problem (BVP), such as modelling a pile penetration [104].
The Finite Element Method (FEM), on the other hand, requires less computational
capacity. Yet, the standard FEM faces the following challenges in modelling the
installation effects of displacement piles:

ˆ The stress-strain behaviour has to be defined by a constitutive model that


captures important aspects such as density and stress dependency. The con-
stitutive model parameters are mostly obtained from standard laboratory tests.
Field observations indicate that the stress-strain paths around a penetrating
pile can be quite different than in standard laboratory tests [99]. This is a ma-
jor drawback in the validation of the constitutive model. Moreover, particle
crushing observed at high stress levels (e.g. under the pile tip, [164]) intro-
duces additional complexity for the constitutive model (e.g. [132]) due to the
changes in the material characteristics.

ˆ The interaction between the structural elements (having high stiffness and
strength) and the soil (having relatively low stiffness and strength) should be
properly described in order to achieve a more representative mobilisation of
stresses in the soil.

ˆ Large deformations cause stress concentrations and distortion of the mesh,


which results in an inaccurate or diverging solution.

ˆ The flow of soil around the pile tip cannot be modelled.

Unlike DEM, FEM requires a constitutive model to incorporate the material be-
haviour. It is important to use an adequate constitutive model such as e.g [170];
[161] that represent the important characteristics of the soil behaviour (e.g. stress
dependency, hardening-softening, dilatancy- contractancy). The complex mechan-
ism observed during installation (e.g. high compaction, shearing, crushing, stress
2.2. Estimation of Bearing Capacity 13

rotations, etc.) is not yet completely understood. Due to the limitations on the
measurement techniques, the strain and stress paths in the soil are still not well
known [56]. In the light of this fact, it is appropriate to employ a constitutive model
based on the Critical State Theory. The state dependent mobilisation of strength,
volumetric behaviour and stiffness can be achieved by the elegant and robust for-
mulation of Hypoplasticity after Gudehus [73] and von Wolffersdorff [158]. One of
the most practical aspects of this constitutive model is the stress dependent stiffness
modelled in relation to the current state (e.g. based on current void ratio, relative
density and mean stress level). However, the model parameters have to be determ-
ined for an expected stress range during the analyses. Considering the high stress
range encountered during a pile penetration simulation, the accuracy of the model
decreases due to a lower order fit for the model parameters. Moreover, the model
does not account for crushing. One might try to fit a set of parameters to include
the crushing stress levels as well. However, it is not possible to capture the whole
stress range as the soil properties are also changing (due to crushing).
Modelling the soil-pile interaction is essential for a proper mobilisation of stresses
in the soil resulting in realistic simulation of pile penetration. The interface elements
used to model the interaction (e.g. [153]) by allowing relative displacement between
the very stiff pile elements and the relatively soft soil elements. The mobilisation
of the shear stresses and the sliding when the capacity is exceeded at the pile shaft
can be modelled using an elastic-plastic behaviour. The use of interface elements
is generally limited to conventional FEM. For most of the advanced methods (e.g.
ALE, CEL) a more elaborate contact algorithm is mostly employed. The contact
detection and interaction algorithms demand a high computational effort. In order
to reduce calculation time, the contact elements are generated only on the possible
contact surfaces. The interaction is defined based on a master-slave concept (e.g.
[140]).
The FE simulations falling under ’zipper type techniques’ also use contact models
[48], [112]) except the ones using direct boundary conditions on the pile-soil interface
(e.g. [59]). These methods resemble cavity expansion methods as the mesh is pushed
aside from a finite cavity. The methods differ from the cavity expansion solutions
firstly, because the mesh is gradually opened up from the pile axis like a zipper.
Secondly, the expansion is accompanied by shear and different levels of compaction
at the pile base and shaft, which are all dictated by the boundary conditions: either
directly by prescribed displacements or by contact models. For the contact models,
an aiding frictionless prehole of 1 mm is present right under the smoothed pile base
before the penetration commences. As the rigid pile penetrates, it pushes the con-
tinuum elements aside like a zipper while shearing the surface. The contact models
are quite elegant in modelling the interaction however; they are computationally
expensive and require relatively strenuous implementation in the FE code. In most
of the applications (e.g. [140]) relatively small frictional coefficients for the pile soil
contact were used to reduce local distortion of the mesh. Remeshing or other ad-
vanced techniques (e.g. ALE) might be needed when higher frictional coefficients
are used and the shearing cause mesh distortion.
14 2. Background on Pile Installation

Different approaches can be found in the literature to avoid the mesh distortion.
Intermittent remeshing (remeshing after a couple of simulation steps) can be ap-
plied to avoid mesh distortion [146]. However, each remeshing step introduces extra
computational effort and accuracy will be lost due to the remapping of the state
variables.
Another powerful FE method to overcome the mesh distortion is the Arbitrary
Lagrangian Eulerian (ALE) formulation (e.g. [162]). The method allows a Lag-
rangian deformation phase followed by a convective Eulerian phase in which material
(Gaussian) points are allowed to move freely through elements. As in the case of
remeshing, the remapping of variables in the convective phase introduces additional
errors to the calculations.
The Material Point Method (MPM) (e.g. [168], [26]), which is a variation of
the particle-in-cell (PIC) method (e.g. [145]) is quite similar to the ALE technique.
The material points are allowed to move throughout the pre-constructed mesh. In
contrast to ALE, MPM can contain a (background) mesh having full, partly-full
(active) and empty elements (inactive), which can be deactivated or activated by
the (freely) moving particles. In each step the mesh is allowed to deform as in the
case of ALE and Lagrangian calculations and in the convective phase the mesh is
reset and the nodal variables (e.g. velocity, displacement) are updated based on
the particles surrounding the node. This is also a sort of smoothing algorithm that
introduces an additional error.
The ALE and MPM are compared in Figure 2.2. It can be seen in the figure
that only the last Lagrangian phase differs for each method. In MPM, generally,
a fixed background mesh is used for the calculation of each Lagrangian step. In
fact, some versions of MPM also include remeshing [27] rather than using a fixed
background mesh. One of the reasons is to increase the accuracy of the method
by allowing synchronous deformation of the mesh and material points as in the
case of ALE and therefore reduce the convective steps. Nevertheless the gain in
the accuracy is hampered by extra computational effort and reduced accuracy due
to remeshing steps. One of the advantages of MPM is the possibility of adding
new material (points). This feature allows modelling of problems involving flow as
in the case of filling silo by a granular material. Furthermore, particle contact is
described intrinsically in MPM, which also allows for separation and gap closure
within the continuum that is modelled using particles. In short, MPM combines the
Lagrangian and Eulerian schemes effectively. On the other hand, a large calculation
time because of the stability condition of the time integration scheme, is a drawback
of the model [169].
Another effective way of avoiding mesh distortion is to couple Lagrangian and
Eulerian mesh by a contact description (e.g. CEL method see [124] for applications).
In this method the material is allowed to flow in an Eulerian mesh until a contact
is detected. The interaction is calculated using a contact algorithm (see [30] for
further reference). As for all FE analyses the CEL and MPM suffers from mesh
dependency (i.e. different response for different mesh sizes). Using a finer mesh
increases accuracy.
2.2. Estimation of Bearing Capacity 15

Figure 2.2: Comparison of MPM and ALE methods: 1. Beginning of calculation step in which
the position, velocity, current strain and stress, state variables and local element coordinates are
known at time t 2. Mapping of particle velocities to nodes 3. (Updated) Lagrangian phase in
which the element is deformed and 4. the position, velocity and strain, stress and state variables
are updated at the material points. 5.a. The Lagrangian mesh is reset. 5.b. The mesh is updated
arbitrarily (after [27])

Figure 2.3: Large deformation (UL) analyses using frictional contact, where smooth geometries
for pile tip are used a. [63] b. [112] c. [77]

Even though these advanced methods are quite successful in preventing the mesh
distortion problem, the implementation into standard FE code can be quite labor-
ious. Furthermore, smoother geometries (see Figure 2.3) are required for the nu-
merical stability (e.g. reducing fluctuations). Yet, the geometry does have influence
on the mobilisation of stress and strains. The singularity point introduced with a
sharp corner is the limitation of the Lagrangian methods and can be tackled by
using interface elements or more elegantly using Eulerian methods, which allow flow
around any obstacle with any shape.
In summary, the advanced methods (e.g. ALE, MPM and CEL) are quite power-
ful in simulating the pile penetration. One of the problems with the advanced meth-
ods is their relatively high computational demand. Secondly, the implementation of
these techniques into available standard FE codes is relatively complex. As a result,
these methods are still not commonly used in the geotechnical practice.
16 2. Background on Pile Installation

2.3 Investigation of Installation Effects


Previous section presented the current methods to determine the bearing capacity
of a pile. Henceforth, the missing components in these methods will be addressed
based on the observations in the stress (Section 2.3.1), density (Section 2.3.2) and
material (Section 2.3.3) changes.

2.3.1 Stress Change


In order to assess the installation effects properly, the stress evolution in the soil
during pile installation should be monitored. In reality, the measurement techniques
limit the assessment of stress evolution to only points where the measurement devices
(e.g. transducers, load cells) are placed. One exception is the use of photoelastic
tests (e.g. [6]), which enable determination of the stress distribution at a reasonable
level. One important limitation is the ambient stress level, which is restricted to
constant stress ([56]) as in the case of a calibration test. Therefore, the shaft friction
mobilised at the model pile shaft does not represent the real case. Besides, Dijkstra
[56] showed that the interpreted vertical stresses (σyy ) for four different tests were
more consistent than the horizontal stresses. This can be explained by the coaxiality
of the major principal stresses with the vertical stresses. Nevertheless, the average
of the four tests (Fig 2.4) gives a quantitative indication of the stress distributions.
In these tests, a steady state was reached and at this level 70% of the pile head force
was transferred to the pile base.

Figure 2.4: Average horizontal and vertical stress distribution based on photoelastic measure-
ments ([56])

Similarly, FE simulations indicate an increase of stresses (both horizontal and


vertical) due to pile jacking (e.g. [140]). For example in the continuous pile jacking
simulations for Karlsruhe sand [112] radial stresses were increased near the pile shaft.
Fig.2.5 shows the final radial stresses at a depth of 4.0 m near the pile shaft after a
continuous jacking of 7.0 m considering a pile penetration length to diameter ratio
of L/D ≈ 23 . Fig.2.5.c indicates that the initial radial stresses are increased by
approximately 2.5 times for the dense case (Id = 0.80) and 2.0 times for the very
loose case (Id = 0.10).
2.3. Investigation of Installation Effects 17

Figure 2.5: Radial stress distribution after 7 m of continuous pile jacking in a. dense and b. loose
Karlsruhe sand; c. radial stress distribution at 4m depth (after [112])

Figure 2.6: Radial stress distribution of pile hammering in a. dense and b. loose Karlsruhe sand;
c. radial stress distribution at 3 m depth (after [112])

Figure 2.7: Radial stress distribution of vibratory pile driving in a. dense and b. loose Karlsruhe
sand; c. radial stress distribution at 3 m depth (after [112])
18 2. Background on Pile Installation

In contrast to the radial stress increase observed for the jacking cases, the vi-
bratory driving and hammering simulations indicated a stress reduction near the
pile shaft. Mahutka et al [112] and Hanke & Grabe [77] explain the stress reduction
by the compaction of the soil in the vicinity of the pile shaft. Quantitatively, for the
vibratory driving case, the radial stresses reduced to approximately half of the K0
level for both dense and loose cases (Figure 2.7).
As an intermediate case, Figure 2.6 presents the results of the hammering simula-
tions. The results show a reduction in radial stresses for both densities to approxim-
ately 37% of the K0 level. The region of the reduced stresses start from the pile shaft
and extend up to 1.5D distance for the dense case and 2.3D for the loose case. The
stress reduction region is followed a zone of increased radial stresses extending up to
approximately 20D (Figure 2.6.c). In this region, the radial stresses were increased
by 100 % and 30 % of the K0 level, for the dense and loose cases, respectively. A
redistribution of the radial stresses can be expected in the equalisation period, in
which the driving equipment is removed from pile head. As a result of the stress
release at the pile head the higher stresses in the soil will shift towards the pile shaft.
All the jacking, hammering and vibratory driving simulations point out a dra-
matic change of the geostatic state and the relevance of the installation methods.
One important aspect these simulations lack is to consider the final equalisation (or
relaxation) state where the pile head is free, i.e. no force is applied, and the sys-
tem goes into a new equilibrium. If the equalisation phase is added at the end of
these simulations, it would be recognised that the stresses around the pile reduce.
In fact, this is the state where the pile installation effects should be quantified and
considered in the design.
In an ALE framework, using hypoplasticity, Dijkstra [56] modelled a centri-
fuge pile penetration test. The force displacement comparison of the numerical and
physical model tests showed that the numerical simulations differ quite from the
experimental observations. As pointed out by [56], the discrepancy is mainly due
to the limitation of the constitutive model parameters, which are determined and
calibrated for small stress and strain level. This is a standard procedure for numer-
ical modelling of BVPs. However, the simulations involving high stress and strain
levels and gradients, as in the case of pile penetration, require a constitutive model
covering a range of small to large strains and stresses. Determination of the model
parameters for this large range may not be possible or accurate enough for most
of the constitutive models. Therefore, the constitutive model should capture the
large strain and stress range (e.g. [132], in which high stress effects were captured
incorporating the grain crushing effects).

2.3.2 Density Change


It is well established that the behaviour of sand under shear is a function of density
and the confining stress [110]. Been & Jefferies [23] introduced the state parameter,
ψ, which expresses the current density relative to the critical state at a mean effective
stress. As one of the most important components of the state is the density, it
has a significant influence on the pile behaviour. The density, together with the
2.3. Investigation of Installation Effects 19

stress level (therefore the soil state, ψ) determines the stiffness of the soil. Hence,
it is important to know the density change (if possible together with the stress
evolution) in order to determine the stiffness and capacity change of the pile during
installation. Unfortunately, only a few researchers (e.g. [56]) could investigate both
density and stress evolution at the same time. However, the resolution of the density
measurements was relatively low.

Figure 2.8: Change in porosity during pile penetration for different initial conditions at different
measurement locations (after [56])

There is a limited amount of experimental work on the investigation of density


change. A comprehensive overview of these contributions can be found in [56]. In
model pile tests performed at low (1g) stress levels, both for initially loose and dense
conditions, loosening was observed near the shaft (e.g. [38], Figure 2.10.b). This
behaviour can be explained by the high level of shearing close to the pile shaft. The
dilative behaviour near the pile was also observed at scaled (Ng) stress levels. For
example, in the geo-centrifuge tests, [56] observed dilative behaviour of the soil near
the penetrating pile regardless of the initial density (loose or dense). As expected,
the dilation during the pile installation was the largest for the dense samples.
Figure 2.8 shows the change in porosities of samples with different initial porosit-
ies measures at 3 different levels on the model pile. The density change was determ-
ined based on the potential (electrical resistivity) change recorded at the conductors.
The locations of the instrument levels and a sketch of the extent of density change
measurement are given in Figure 2.9. The measurements provide a volumetric av-
erage in a zone approximately 1D around the measurement point. Therefore it
is possible that the amount of dilation in the centrifuge tests is higher than the
interpretations based on the electrical resistivity measurements.
Dijkstra [56] explains the unexpected dilation of the loose samples by the un-
certainty of the initial conditions near the pile. The dilation might also be due to
the relatively large shear bands caused by the particle size in the centrifuge test,
although properly scaled (D/D50 = 115).
Porosity measurements of all samples with different densities showed that the
dilation reached a limit state and then remained constant. At the beginning of
penetration, different mobilisation levels at measurement locations of the pile were
observed. During loading no additional density change was detected. For the case
when the pile was unloaded a density increase was recorded. Additional displacement
controlled load cycles caused the largest densification. The densification around the
20 2. Background on Pile Installation

Figure 2.9: a. Test setup for the centrifuge pile tests; b. The sketch of the density measurement
principle, where the potential drop was measured between the conductors; the continuous lines are
the electric field and the arrows show the current flow.

shaft due to load cycles resulted in reduction in the confining stresses and therefore
a reduction of shaft friction.

Figure 2.10: Displacement and volumetric observations in 1 g models using a. radiograph meas-
urement ([128]), b. thermal conductivity ([38]) and c. imaging techniques (in crushable carbonate
sand, [180])

In contrast to the aforementioned observations, compression was observed near


the pile shaft (e.g. [91] and [165]) in the calibration chamber tests. In these types of
physical models, the stress level is only scaled correctly in the zone near the pile tip.
Therefore, interpretations on the behaviour of the shaft should be done cautiously.
This compaction resulted in a far field loosening. By the Digital Image Correlation
(DIC) technique the aforementioned mechanism could be visualised [165]. Similarly,
Dijkstra [56] observed compression around the pile shaft in a photoelastic setup at
increased constant stress level (similar to the calibration chamber).
2.3. Investigation of Installation Effects 21

For the pile base contradictory observations were reported. For initially dense
sand, a dilative response was observed at low stress levels, even for the calibration
test with relatively low length to diameter ratio, i.e. L/D = 6 [49]. This might
indicate that no particle crushing had occurred. Conversely, compaction at the pile
base was observed as a cone that was composed of crushed material ([164]). [180]
also observed a compaction zone under the pile base during penetration of a pile in
crushable carbonate sands (Figure 2.10.c). However, the shape of the compaction
zone was spherical rather than a ’nose of cone’ observed by White [164]. The com-
paction due to particle crushing, however, should be evaluated as a material change
(see Section 2.3.4) rather than a density or volumetric change. Radiograph meas-
urements of Robinsky & Morrison [128] also showed a main compaction zone under
the pile tip (Figure 2.10.a). It is important to note that the L/D ratio determines
the mobilisation and therefore contractive or dilative behaviour is observed at the
pile tip. The stress levels and gradients in the aforementioned studies were different
than a normal geostatic K0 stress distribution. Therefore the mechanisms observed
in these studies are probably different than the one that would be observed in the
field.

Figure 2.11: Void ratio distribution after pile penetration simulations based on a. FE with
remeshing [78] and b. CEL [124] in medium dense Karlsruhe and Mai-Liao sands, respectively

Similar to Dijkstra’s measurements [56], [78] indicated a zone of dilation, near


the pile (for medium dense soil Id = 0.45) based on their pile jacking simulations. A
relatively confined zone of compaction was observed 1D away from the pile shaft. As
expected, the level of compaction is higher close to the tip, but not directly below
the tip (Figure 2.11.a). In these simulations a relatively low frictional coefficient
of the contact was used (δ = 1/3φ). In reality, the friction level between the pile
and soil is much higher (δ = 0.75φ). Using a lower frictional coefficient results in a
reduced transmission of stresses and strains to the soil. Therefore the level of shear
strength mobilisation in the soil is expected to be different.
[124] carried out CEL simulations in Mai-Liao sand, which has relatively low
stiffness compared to Karlsruhe sand. The results indicated a zone of dilation near
the pile. However the dilation zone observed in these simulations was not as clear
22 2. Background on Pile Installation

as the FE results of [78]. Furthermore, a zone of compaction at the upper part of


the pile was observed for CEL simulations (Figure 2.11.b).
It is difficult to assess the degree of approximation of the mobilisation levels
obtained by Lagrangian and Eulerian simulations. In Lagrangian methods the mo-
bilisation rate as well as the area of influence is higher. On the other hand, relatively
low mobilisation rates observed in Eulerian methods limit the influential area and
more confined failure mechanisms are evident.
Logically, for a very loose Karlsruhe sand (Id = 0.10), the pile jacking model of
Henke & Grabe [78] indicated only compaction near the pile shaft. However, it is
quite difficult to find such loose sand deposits or even to prepare these in laboratory
conditions.
Dijkstra [56] also modelled the centrifuge pile penetration by incorporating large
deformation effects using both fixed (full Eulerian) and moving pile (ALE) mod-
els. Hypoplasticity was used to model the Baskarp sand used in the centrifuge pile
test. Both results of the simulation of the effects of 13D pile jacking also indicate
considerable amount of dilation at the pile shaft.

Figure 2.12: Void ratio distributions of medium-dense (Id = 0.45) Karlsruhe sand at several
distances from the pile shaft for a. jacking, b. hammering and c. vibratory driving installation
methods ([112])

Mahutka et al. [112] clearly indicated the effects of different installation methods
using an adaptive remeshing technique (Figure 2.12). For both of the dynamic
installation types a zone of compaction was observed (highest for vibro-driving). As
indicated before, for the jacking case, dilation was observed close to the pile shaft
(r = 0.5D), whereas at distances from r = 1D a compaction of approximately 10%
for the hammering case was observed.
As a conclusion, several experimental and numerical methods suggest different
patterns and levels of density change around a displacement pile. The observed
density changes should be cautiously evaluated considering the limitations and as-
sumptions in these simulations. During penetration of a pile shearing, compaction
and dilation mechanisms are dominant at the shaft and base of the pile. Therefore,
depending on the initial state of the soil, different levels of compaction and dilation
as well as the size of the influence area can be expected.
2.3. Investigation of Installation Effects 23

2.3.3 Material Change


The stress and density are not the only variables that change due to pile installation.
Very high stresses (in the order of tens of MPa’s) can be observed under the pile
base. Besides, it was shown that high levels of confinement can result in material
crushing (e.g. [109], [151]). Therefore, even for the stronger silica sands some ma-
terial crushing might be expected. For example, [164] in his plane strain calibration
test setup, observed a ’nose of cone’ accumulating underneath the pile tip and con-
taining highly crushed sand (Fig.2.13.a). The crushed material was identified to flow
around the pile shaft, crushed further and infiltrated into the soil. The diffusion of
the crushed particles causes a volume and consequently, stress reduction near the
pile shaft. The partial crushing, penetration and redistribution of the grains, causes
a reduction of stresses at the pile shaft, which is called friction fatigue as observed
in the field tests and mentioned in the previous paragraph. [103], in a geo-centrifuge
test, also observed a crushed zone under the pile base (Fig.2.13.b & c). However,
for a conical base the crushing was much lower. On the contrary, [56] reported no
grain crushing for the geo-centrifuge pile tests.

Figure 2.13: Grain crushing for flat pile base a. [164] b. [103] and for c. conical base [103]

In the DEM simulations with crushable particles (idealised) similar behaviour


was observed [104]. In these simulations the pile was continuously jacked at a velo-
city of 10−6 m/step, in order to avoid any crushing due to pile impact. The Fig.2.14
clearly presents the redistribution and reduction of stresses due to particle breakage.
The unrealistic high stresses observed in FE simulations of pile penetration can be
reduced to realistic levels by incorporating the grain crushing effects. The particle
crushing and migration and diffusion of particles resulted in reduced stresses. There-
fore it is important to consider the crushing effects in the analyses. Although the
model has many simplifications (e.g. particle size, crushing description, etc.), the
results of [164] and [165] could be validated qualitatively. However, there was no
clear zone of crushing forming ’nose of cone’ Nevertheless, the DEM method gives
the possibility to compare the stress distribution using the same material with and
without crushing.
In their follow-up study, [105] showed that the pile base geometry has an influence
on the crushing level (see Fig.2.15). A pile with a flat base induced higher stresses
and therefore more crushing than the one with a conical tip.
24 2. Background on Pile Installation

Figure 2.14: Results of DEM simulation a. without and b. with particle breakage [104]

Figure 2.15: DEM results incorporating particle crushing for different pile types [105]

The crushing mechanism in reality is quite different than in the idealised model.
It can be clearly seen that the particles used in the models are larger than real soil
particles. Besides, an idealised crushing mechanism was employed in these studies:
once a predefined tensile capacity of a particle is reached, it is divided into segments
fitting into the initial geometry of the particle (Fig.2.16.c). As a result of this
idealised mechanism, some material is lost. Furthermore, the induced tensile stress
within a particle was also idealised as shown in Fig. 2.16.a & b. These idealisations
result in a deviation from the real soil behaviour.
The particle crushing is an important phenomenon, which has to be considered
when analysing displacement piles and their effects. Ideally the change of material
properties should be assessed in the field, or in properly scaled physical models,
during and after the installation phases. A relatively easy way is to employ particle
2.3. Investigation of Installation Effects 25

Figure 2.16: Idealisation of induced tensile stress in a particle and the crushing mechanism [105]

models, which provide a good basis for understanding the mechanics of pile-soil
interaction and its consequences including particle crushing effects.

Figure 2.17: A general critical state line for sands [132]

Alternatively, the crushing can be modelled in the constitutive model. For ex-
ample, [132] introduced a bounding surface model for sands. The model captures
the general critical state behaviour for a wide range of stresses including particle
crushing. The model includes the CSL (the initial line segment with slope λ0 ), the
particle crushing (second line segment with slope λcr ) and the specific volume at
the onset of particle crushing νcr , and finally, the end of crushing (last line segment
with slope λf ) with an onset of for the very high stress levels. The continuity of the
line segments was ensured by a nonlinear polynomial fit. The model was validated
by element tests of different stress paths.
Using a constitutive model in an FE simulation results in a more realistic stress
and strain distribution. However, the crushing mechanism cannot be investigated
unless extra state variables are defined to represent the level of crushing. For example
the combination of a specific volume and the mean effective stress can be correlated
to the change in particle size distribution that can be represented by the mean
particle size, d50 , and the coefficient of uniformity Cu .
26 2. Background on Pile Installation

2.3.4 Effects on Pile Behaviour


The installation effects observed in the soil are due to the alteration of the stress
and density state. These changes have an influence on the pile bearing capacity and
stiffness depending on the method of installation. For example, it was previously
shown that displacement piles have behave stiffer behaviour than bored piles (e.g.
[45]). Based on a series of field tests Deeks et al. [55] indicated similarly, stiffer
behaviour for jacked piles than bored piles. It is concluded that a higher base
stiffness compared to bored piles is due to preloading during jacking and resulting
residual stresses. As a consequence, the required mobilisation level for displacement
piles (0.10D) is much less than what is required for non- displacement piles (1D; [52];
[62]; [66]). As the critical base capacity is generally determined based on a relative
settlement s/D = 0.05 − 0.10, for the same amount of displacement, a displacement
pile will have a higher capacity.

Figure 2.18: a. End bearing mobilisation, and b. shaft friction transfer curves of displacement
and non-displacement piles obtained at different segments of on the pile for a medium dense FF
sand (after [45])

Figure 2.18 shows the effect of pile installation on the normalised pile base,
qbase /qbase,max , and shaft friction, τ . Displacement curves are plotted for a medium
dense FF sand, i.e. Id = 0.40 - 0.42, for acceleration levels of Ng = 30, 80 & 130,
which scale the model pile to different diameters and penetration lengths). The
FF Sand is a very fine silica powder with pit rocks origin [45]. The qbase /qbase,max
curves show a clear stiffness increase for the displacement piles. Up to a very high
mobilisation level, i.e. s/smax (smax = D), the normalised base capacities for the
displacement piles are larger than the non-displacement piles. The convergence of the
mobilisation curves of both type of piles to a same level of qbase /qbase,max is an indic-
ation of a steady state for the pile base. Obviously, the unit shaft capacities reached
to their steady states at different levels, which depend on the confinement level.
Therefore lower levels of shaft capacities were obtained for the non-displacement
piles compared to those obtained for the displacement piles.
It is quite common to measure the horizontal contact stresses on a number of
discrete positions on the shaft in field and model tests. An easier way to deduce shaft
friction is to measure axial stresses in the pile at different levels and at the base.
The observations on the radial stresses after pile installation are summarised in [56].
2.3. Investigation of Installation Effects 27

For the jacked piles the horizontal stresses were increased due to the intrusion of the
pile volume. However, in almost all of the cases, except for the long term effects, a
reduction in radial stresses was observed with further penetration of the pile (e.g.
[167]).
The shaft friction degradation during pile driving named as ’Friction Fatigue’ [75],
which is later named as the ’h/R effect’ [40], was investigated using instrumented
field tests (see [39] for details). [166] describe the mechanism of friction fatigue by
considering a soil element, which is first sheared to its peak strength, τmax . By
further penetration of the pile, this element is sheared to its residual strength, τres .
The reduction from peak to residual strength is another reason of the reduction of
shaft friction besides friction fatigue (Fig.2.19.d). Furthermore, during penetration,
the material exhibits dilation. As a result of this loosening, the stiffness and strength
of the soil reduce. This behaviour can be explained using the t-z curves showing
hardening and softening behaviour. In Fig.2.19 the consequence of assuming elastic
perfectly-plastic behaviour or strain hardening and softening behaviour near the
shaft can be seen. In the former, the strength remains constant once mobilised.
In the latter case, the strength reduces by further shearing. The resulting shear
strength mobilisations and their evolutions for both cases can be seen in Fig.2.19.c &
d. The shear strength mobilisations change under different confinement levels. The
material change due to grain crushing certainly alter the shear strength mobilisation.
The dashed and dotted curves in Fig.2.19.a represent the confinement and crushing
effects. All these mechanisms shed a light on the ’h/R effect’. [98] also observed
the ’h/R effect’ in the field tests. In these tests 102 mm diameter cone ended piles
were installed with a series of 250 mm jacking strokes. As a result of increasing h/R
value, in other words further penetration of the pile, a reduction of shear strength
and shaft friction was observed. The difference in the mobilised shear strengths is
due to the spatial variation of soil type and therefore stiffness and strength. [99]
suggest that the principal stress rotations reduce the radial stresses. Compression
tests showed small reductions in the radial stresses. On the other hand, tension tests
showed significant reductions in the radial stresses.
The friction fatigue behaviour could also be represented by the numerical models.
In their FE model with contact, [140] observed a noticeable increase in the radial
stresses just above the cone as well as a low stress zone (lower than K0 level) at a
distance below the tip.
In contrast to the possible mechanism during penetration and the friction fatigue
presented in Fig.2.19, [167], showed in their centrifuge tests,that friction fatigue
occurs only with loading and unloading cycles. The rotation of principal stresses as
suggested by [99] is one of the major causes for the reduction in the cyclic load tests.
However, not only the number of cycles but also the amplitude of the strains have
influence on the stress evolution around the pile. For example, in the high amplitude
cyclic tests with a displacement, u > 0.033 · Dpile , [56] observed a densification near
the pile shaft. This volume reduction near the shaft relieved the radially stressed
zone (locked in or residual stresses). As a result of the reduction in the radial
stresses, the shaft capacity was reduced. On the contrary, for low amplitude cycles
(u < 0.033 · Dpile ), an increase in shaft capacity was observed both for the dense
28 2. Background on Pile Installation

Figure 2.19: Load transfer (t-z) curves for a. simplified elastic-perfectly plastic soil and b. for
strain hardening and softening soil (after [125]); dashed lines represent the confinement and the
material change effects. Mobilisation of shear strength at different h/R ratios for c. simplified
elastic-perfectly plastic soil and d. for strain hardening and softening soil. Field observations on
the mobilisation of e. shear strength mobilisation and f. shaft friction [98]

and the loose samples. This is an indication of dilation during the loading cycles,
which increased the radial stresses and consequently the shaft capacity.
As mentioned in the previous paragraph, the installation method has influence
on the reduction of shaft friction. From this perspective, [167] suggested that the
empirical correlations relating the shaft capacity to h/D can be improved by incor-
porating the number of load cycles. They showed that the normalised horizontal
stresses acting on the pile shaft did not change during monotonic installation. How-
ever, this observation should not be misinterpreted. The normalisation was based on
the CPT cone resistance, which already included the installation effects. Therefore,
the reduction mentioned for the cyclic loading is the relative shaft friction reduction
when compared to continuous jacking, i.e. the CPT in this case. The compressive
loading followed by an unloading step was called as one way cycle, and the com-
pressive loading followed by a tensile (reversed) loading was called as two way cycle.
In their two way cyclic tests, the horizontal confining stresses were reduced to a
very low level after approximately 80 cycles. Therefore, the empirical correlations
presented to account for the friction fatigue (e.g. [98]; [39]) should be improved by
incorporating the cyclic effects.
The cyclic loading has influence on the base resistance as well. [56] observed a
reduction of 55% of initial capacity at the base and 60% of the initial total capacity
for the dense sample due to loading cycles. Moreover, the contribution of the base
capacity to the total capacity was also reduced from ≈ 60% to 50%. On the contrary,
the capacity was increased for the loose sand, i.e., an increase from ≈ 30% to 55%
of the total capacity for the base and a slight increase for the shaft.
2.3. Investigation of Installation Effects 29

Figure 2.20: a. Typical shaft friction distributions [56]; b. Residual loads during and after
installation [181]

There are basically two types of shaft friction distributions: exponential and
parabolic (Fig.2.20.a). [56] explains the exponential distribution by the exponential
increase of stiffness of the soil with depth. However, the stress dependency of the
soil stiffness can be represented by relating the oedometric stiffness, to the confining
stress that can be determined from the minor principal stress, σ3 , the normally
consolidated geostatic stress ratio, K0N C , the cohesion, c and the internal friction
ref
angle, φ based on the reference oedometer stiffness, Eoed calculated at a reference
vertical stress, pref as:
⎛ ⎞m
σ
c · cos φ − K N3C sin φ
ref ⎝ 0 ⎠
Eoed = Eoed (2.3)
c · cos φ − pref sin φ

The maximum value of the power, m, is 1 for a normally consolidated clay,


which relates the stiffness to the stress level linearly. For normally consolidated
sand, the power is generally taken as 0.5, which relates the stiffness to the square
root of the stress level. From this perspective the stiffness increase does not explain
the exponential shaft friction distribution. The stress increase near the pile tip
due to installation, however, can be a good explanation for the exponential stress
distribution on the pile shaft. [90] observed a stress reduction around the shaft
in carbonate sands as a result of grain crushing. Based on this observation, [56]
suggested that the parabolic shaft friction distribution could be an indication of
the grain crushing. Otherwise, an exponential increase would be obtained if the
measurements of the lateral stresses on the pile shaft are corrected for the residual
stresses as suggested by [7] and [61].
[7] and [61] pointed out the importance of residual stresses that are locked-in
after pile driving (Fig.2.21.b). They claim that the ’critical depth’ concept, which
suggests that the unit shaft resistance becomes constant with further embedment,
does not reflect the physical situation. Furthermore, the observations in the field
tests [155], were a result of measurement error in which the residual stresses are not
30 2. Background on Pile Installation

Figure 2.21: Results from static loading tests on 12.8 m pile ([61] after [14])

taken into consideration. It is claimed that if the measurements are corrected for the
residual stresses, a realistic shaft friction distribution can be obtained. However, the
residual stresses arising from the installation effects (or simply locked-in stresses)
are generally not measured. The correction is based on an assumption of true shaft
friction distribution and involves an iterative scheme for the prediction of true load
distributions. [181] applied simple t-z analyses to estimate the residual stresses by
unloading from the stress level at the end of penetration. In these analyses the
residual base stress was found to be of the base stress at the end of the jacking
phase.
Besides the change in stress due to installation, long term changes in stress were
reported by several researchers. The gradual increase in soil stresses near the pile
after installation resulting in increased pile capacity is called pile setup. [41] and [15]
reviewed the setup phenomenon extensively. Some of the load tests were performed
several times on the same pile at different time intervals and based on these tests
an increase of pile capacity was reported (Fig.2.21). However, as stated by [61],
this behaviour cannot be explained only by the setup phenomenon. It is known
that the loading-reloading cycles without any pause would yield similar behaviour.
Nevertheless, the horizontal pressures measured at different time intervals indicate
a clear increase by time. Besides, a previous study of [147] indicates a total capacity
increase of in 20 days for a concrete pile in medium-dense fine sand. Moreover, the
effect is more for piles in denser, finer sands and silts than that in looser, coarser
sands and gravels. For tubular piles in dense marine sand an increase of in shaft
capacity in an interval of 6 months to 5 years after the installation was found [41].
The setup phenomenon can be explained by soil ageing, the breakdown with
time of hoop stresses and kinematically restrained dilation mechanisms [31]. The
effect was observed for strong (silica) sands and not for weak (calcareous) sands.
[31] indicated that the rate of soil disturbance has also influence on the ageing or
capacity increase. Faster loading (e.g. pile hammering) results in quicker and higher
dilatant creep than the slow loading (e.g. pile jacking). As the setup relates highly
to restrained behaviour, marginal pile setup was observed in the top few meters of
the soil [31]. [179] studied empirical correlations (e.g. logarithmic time dependent
capacity increase) to incorporate setup into design of driven piles. In fact there is
a large scatter in the observations. Additionally, it was shown that the increased
2.4. Including Installation Effects in the Analyses 31

capacity could not be obtained for a prefailed pile during tensile load tests [40].
[39] presented the effect of increase of shaft friction (51%) due to installation
of a second pile. In contrast, the bearing capacity of the previously installed pile
was reduced by 43%. This reduction is an indication of a stress relief at the pile
base probably due to upward movement of the pile [56]. Therefore, it can be con-
cluded that when considering the installation effects a direct superposition might
overestimate the group behaviour.
[78], in their FE simulations, investigated the effect of pile installation on neigh-
bouring ’wished-in-place’ piles. As the installation effects were missing for the previ-
ously activated piles, the aforementioned behaviour could not be reproduced. How-
ever, in the same study, it was concluded that the pile jacking had more influence
on the neighbouring piles than the vibratory driving technique.

2.4 Including Installation Effects in the Analyses


In order to benefit from the installation process in the analyses and design, the in-
stallation effects should be incorporated in the analyses. However, the stress and
density state after the installation process is generally not included in the FE ana-
lyses when the pile is ’wished in place’. The calculated capacities are generally
increased by empirical factors to account for the installation effects. However, some
techniques that impose the strain field calculated by SPM into FE models to ac-
count for the installation process can be found. Section 2.4.1 and Section 2.4.2
present current empirical and numerical methods that account for installation ef-
fects, respectively.

2.4.1 Empirical Methods


Most of the CPT methods use empirical factors to account for the installation type
and therefore the installation effects. However, these empirical factors are used only
to reduce or increase the capacities determined using these methods.
Jardine and Chow [87] improved the interface model of Wernick [163] to estimate
shaft capacity by considering contact (interface) properties Uesugi and Kishida [150]
and radial effective stress change Δσr (also observed experimentally, e.g. [29]) based
on cavity expansion theory.

τrz = σrf tan δ  = (σrc

+ Δσr ) tan δ  (2.4)

where,

σrf : radial effective stress acting on pile shaft at failure

σrc : radial effective stress acting on pile shaft after installation process
(before loading)
32 2. Background on Pile Installation

Δσr : change in normal effective stress during loading, due to soil volume
change

tan δ : friction angle at soil-rigid inclusion interface

And based on cavity expansion theory:

4G
Δσr = · ur (2.5)
D
where,

G : operational shear modulus, function of the pressure and strain level


ur : displacement normal to pile shaft

The installation effects are partly considered by incorporating equalisation ra-



dial stress, σrc . As mentioned in Section 2.3.1, at the equalisation (or relaxation)
state the pile head is free, i.e. no force is applied, and the system goes into a new

equilibrium. The pre-loading radial effective stress, σrc can be determined as [40]:

 
0.13  −0.38
 σv0 h
σrc = 0.029qc (2.6)
Pa R

The proposed relation predicted the total capacities mostly within ∼


= ±20% of
those measured. The method is therefore quite successful in predicting the pile
capacities.

Figure 2.22: Extended modification factor Ff∗ for a. Chiibishi sand, b. Dogs Bay and c. Toyoura
sand ([97])

Kuwajima et al. [97] extended the idea of modification of the bearing capa-
city factor based on compressibility by considering the mean normal stress and the
normalized displacement (settlement) of the pile S/D as:

qp = SD · Ff∗ · Nq · σm

(2.7)
2.4. Including Installation Effects in the Analyses 33

They proposed a modification factor Ff , which is a function of both compressib-


ility Cp and S/D (Fig.2.22). The Ff∗ decreased with increasing soil compressibility
and increased with normalized penetration ratio (Fig.2.22.d-f), S/D as:
 
∗ 3
Ff = F f (2.8)
2K0 + 1
Although the method accounts for density change, i.e. mostly due to grain
crushing, only the pile tip resistance is modified. Furthermore, the model was not
validated for different sands. Therefore, the predictive capability of the model cannot
be assessed.
In Dutch practice, the density change due to pile installation is only considered
for the tension piles and especially to account for the group effects based on CUR
2001-4 (ontwerpregels voor trekpalen). This method is proposed to account for
densification due to installation of several piles (i.e. group effect). The stress increase
is then, indirectly captured by using a modified cone resistance, qc value based on
the empirical correlations (see [108]). In this method, the compaction of the soil due
to pile installation in a presumed area of influence (6Deq ) is estimated by:

n
(r − 6) (1 + e0 )
Δe = − (2.9)
1
5.5 50

where,

Δe0 : decrease in void ratio (increase in density) within the area of influence
6Deq
n : number of piles within the influence area
r : normalised (R/Deq ) center to center distance between the neighbouring
piles (if more than one)

The pile volume introduced is calculated as a volume reduction of the pores


within the area of influence. However, as mentioned previously, porosity change
around a displacement pile can be quite different than as assumed here, i.e. the
dilation observed near the pile shaft.

2.4.2 Numerical Methods


There is quite limited amount of work on the incorporation of installation effects
numerically. [7] suggested using the residual stresses as an integral part of the
finite element analysis to obtain more realistic load-displacement behaviour. They
also considered the degradation of shaft resistance with cyclic loading, which cause
compaction. However, no relation was considered for density and stiffness change of
the soil.
As mentioned in 2.2.2, a combination of SPM and FEM solutions can be found
in the literature. For example, Teh [148] combined the SPM with a Finite Element
34 2. Background on Pile Installation

model by using the strain field obtained from SPM. Then the strain field was imposed
as an initial state and the corresponding stresses were calculated using the FE model.
Similarly, Gill & Lehane [71] obtained strain fields for different fluid viscosities using
FEA for different penetrometer roughness as an extension of SPM. They showed
that the shape and extent of the strain field depend on the viscosity of the material
and the roughness of the penetrometer. The mechanism around a penetrating probe
can be assumed as flow, however, the behaviour of soils is quite different than fluid
flow. The latter is generally modelled by the Navier -Stokes equations. On the
other hand, soils exhibit dilative-contractive as well as history and stress dependent
behaviour. Therefore, the strain paths assumed in the SPM solutions are probably
different for soils.
Another way to tackle the penetration problem is trying to simulate installation
effects rather than the installation process itself. These types of simulations are
based on standard FEAs, which are basically employment of a prescribed bound-
ary causing radial or volumetric expansion. Using a standard FE package, [34]
investigated the possibility of simulating installation effects by means of displace-
ment boundary conditions and volumetric strains imposed at the pile surface and
volume, respectively. Successful load displacement behaviour could be obtained but
no clear correlation for the level of applied displacements or volumetric strains could
be found.
Similarly, [135] considered the installation effects using empirical correlations
based on field data. These correlations have been used in order to account for
the change in radial stress, shaft friction and base resistance. This strategy has
been proved adequate in reproducing in a simple way the changes in the state of
stress produced by the installation procedure. Despite the good agreement between
experimental data and numerical predictions, some aspects (e.g. densification of the
sand below the pile tip ad increased soil stiffness) still require a refined analysis able
to reproduce the real installation process.

2.5 Conclusions
In this chapter, current methods to investigate and incorporate pile installation ef-
fects were presented. Although the effects are quite important for the behaviour of
the pile, it is still not possible to quantify the effects as a continuous field experi-
mentally. Therefore the numerical models could not be completely validated.
Not many techniques incorporate installation effects. However, using the altered
state (installation effects) in pile loading analyses will give more realistic results.
Based on these results, better grasp on the pile behaviour in service life and there-
fore pile design can be obtained. There are quite advanced models available to
simulate the installation process, like CEL, MPM and ALE; however, they are not
yet widely used in engineering practice. Therefore a simpler technique to investigate
the installation effects roughly and incorporate these effects in the analyses will be
an important step for the engineering practice. For this reason, a simplified FE
technique (the Press - Replace technique) is introduced and explained in Chapter 4.
2.5. Conclusions 35

As a step further, the important components of the installation effects, such as:

ˆ Modified stress state

ˆ Modified density state (or both can be quantified by the state parameter; [23])

ˆ New stiffness of the soil

ˆ Change of properties due to crushing

can be described as a new state around the pile to be modelled. This way
the displacement piles can be modelled easily without discarding the important
installation effects.
Few empirical methods consider the installation effects for a particular loading
case; either compressive, tensile or lateral loading. However, the installation effects
are based on the installation and not the loading type. If the installation effects are
properly modelled, the resulting state should work in any loading case. Hence, it
is important to have the entire installation field rather than only the reflection of
these effects on the pile.
In the next chapter (Chapter 3), a summary of the constitutive model employed
in the numerical analyses is given.
36 2. Background on Pile Installation
Chapter 3

Constitutive Modelling

3.1 Introduction
In the previous chapter, a summary of current practice on the prediction of pile
bearing capacity, the available methods for the investigation of the installation ef-
fects and the existing methods for the incorporation of these effects in the analyses
was presented. This chapter presents a summary of the constitutive model employed
in the numerical analyses, which will be discussed in Chapters 4 & 5. Section 2.2.1
outlined the important aspects of finite element modelling (FEM) paying particular
attention to the installation of displacement piles. As mentioned in the same sec-
tion one crucial component of the FEM is the constitutive modelling, by which the
stress-strain behaviour is defined. The constitutive model should capture the typical
behaviour of soils (e.g. stress and density dependency), which can be expressed in
the framework of the Critical State Theory.
This study focuses on the effects of the installation of displacement piles in sand.
In order to represent the essential characteristics of sand (e.g. stress-dependency,
hardening-softening, dilatancy-contractancy) and to improve numerical stability, es-
pecially in large deformation problems (e.g. [77]), the modified version of von Wolf-
fersdorff’s hypoplastic model [114] was employed in the numerical analyses. As
described in the following sections, hypoplasticity provides flexibility and modelling
capabilities in terms of capturing not only hardening but also softening behaviour,
besides void ratio and stress dependent stiffness. The following section (3.2) gives
a brief background on hypoplasticity. Furthermore, the hypoplastic model of von
Wolffersdorff’s [158] and the small strain extension of Niemunis and Herle [120] are
explained.
It is important to consider the soil behaviour at high stress levels since during
pile penetration high levels of stresses have been observed. Due to high levels of
confining stress, the dilatancy is suppressed and therefore the stresses are increased.
As a result of further increase in stresses grain crushing takes place [109]. In general
the constitutive models are tuned for relatively low stress levels compared to the
stress levels encountered at the pile tip during penetration. In Section 2.3.3, the
experimental evidence on the material change due to pile installation was presented.

37
38 3. Constitutive Modelling

Based on these observations, in Section 2.3.1 the importance of grain crushing and
therefore the necessity of modelling the corresponding behaviour was mentioned. De-
pending on the material origin (e.g. silica or carbonate sand), it might be necessary
to account for possible grain crushing by further modifications on the model. From
this perspective Section 3.3 presents a modified version of hypoplasticity to account
for the grain crushing [129]. Some improvements on Rohe’s model are proposed.
The numerical stabilities of the crushing models are discussed briefly.

3.2 Hypoplasticity
Starting from the hypoelasticity model of Truesdell and Noll [149], Kolymbas [93]
introduced an incrementally nonlinear constitutive relation for soils of the form:

σ̊ = h (σ, ε̇) (3.1)

where σ̊ is the Jaumann stress rate, σ is the Cauchy stress, h is an isotropic tensorial
function linear in strain rate (stretching) ε̇.
Kolymbas [94] presented the first form of the hypoplastic model, which consists
of terms linear in strain rate similar to hypoelasticity as well as additional terms
nonlinear in strain rate. The model has an explicit interdependence of strain rate
and stress rate directions, which results in a path dependent and dissipative beha-
viour. Due to its elegant formulation and a simple calibration procedure [95], several
researchers (e.g. [174]; [173]; [21]; [74]; [158]; [175]) worked on the model extensively.
For example, the incorporation of the Critical State concept in the model en-
hanced the history dependence of instantaneous stress. Furthermore, the mater-
ial parameters could be defined such that a particular type of soil is characterised
by a unique set of parameters, whereby the void ratio fully defines the density-
dependence. Besides, the important features of the soil behaviour such as density
dependent strength and stiffness, critical state behaviour and nonlinearity from the
very beginning of loading, stiffness dependency on loading direction, could be rep-
resented. The critical state was embedded in hypoplasticity by multipliers defined
based on current void ratio, e (e.g. [173]; [21]; [73]). The extended form of the
hypoplastic constitutive model therefore had the following form:

σ̊ = h (σ, ε̇, e) (3.2)

As mentioned earlier, in this study von Wolffersdorff [158]’s hypoplastic model in-
corporating the Matsuoka - Nakai failure criterion was employed in the finite element
calculations. The hypoplastic frame independent stress tensor of von Wolffersdorff’s
model is of the form:

σ̊ = fs (L : ε̇ + fd N ||ε̇||) (3.3)

where, L and N are fourth and second order constitutive tensors, both functions
of stress, respectively. The first part is linear in ε̇ and the
second part is nonlinear
in ε̇ (due to the Euclidean norm of strain rate ||ε̇|| = tr2 (ε̇)). Both are posit-
ively homogeneous of first degree in ε̇. The function fs (trσ, e) is the product of
3.2. Hypoplasticity 39

the pycnotropy (density dependence) function fe (e) and the barotropy (stress level
dependence) function fb (trσ). The function fd (e) describes the pycnotropy only.

Figure 3.1: a. Comparison of the limit surfaces Matsuoka-Nakai and Mohr Coulomb for different
triaxial compression friction angles, φtc ; b. Geometrical representation of the invariants tanψ and
cos3ϑ in principal stress space (after [158])

The linear and nonlinear parts of the model can be deducted from the complete
form:
1
2
σ̊ = fb fb 2
F ε̇ + a2 tr(σ̂ ε̇)σ̂ + fd aF [σ̂ + σ̂ ∗ ] ||ε̇|| (3.4)
tr(σ̂ )
as:

L = F 2 ε̇ + a2 tr(σ̂ ε̇)σ̂ (3.5)

N = aF [σ̂ + σ̂ ∗ ||ε̇||] (3.6)


where, a and F incorporate the Matsuoka-Nakai failure criterion as:

3(3 − sinφc )
a= √ (3.7)
2 2sinφc
and

1 2 − tan2 ψ 1
F = tan2 ψ + √ − √ tanψ (3.8)
8 2 + 2tanψcos3ϑ 2 2

with invariants (Figure 3.1 .b)



tanψ = 3 ||σ̂ ∗ || (3.9)
√ tr(σ̂ ∗3 )
cos 3ϑ = − 6 3/2
(3.10)
[σ̂ ∗2 ]
40 3. Constitutive Modelling

where the normalised Cauchy stress (σ̂) and the normalised deviatoric stress (σ̂ ∗ )
ratios are defined as:
σ
σ̂ = (3.11)
trσ
1
σ̂ ∗ = σ̂ − I (3.12)
3
and I is the unit tensor.
Bauer [21] and Gudehus [73] define the characteristic void ratios shown in Figure
3.2.a as:
 n 
ei ec ed −trσ
= = = exp − (3.13)
ei0 ec0 ed0 hs

Figure 3.2: a. Decrease of the maximum void ratio ei , the critical void ratio ec and the minimum
void ratio ed with an increase of the mean effective pressure p b. Interaction between e, p and the
density factor fd [22]

The pycnotropy functions fe (e) and fd (e) are


 e β
c
fe (e) = (3.14)
e
 α
e − ed
fd (e) = (3.15)
ec − ed
and the barotropy function fb (trσ) is
  β  1−n  α −1
hs ei0 ei0 −trσ √ ei0 − ed0
fb (trσ) = 3+a −a 3
2
(3.16)
n ec0 ec0 hs ec0 − ed0

The multiplication of fb and fe can be also be written in the form of one function
(e.g. [113]) as:

    1−n  √  −e α −1
e β −trσ
fs (trσ, e) = hs
n
1+ei
ei
i
e hs 3 + a2 − a 3 eec0
i0 d0
−ed0 (3.17)
3.2. Hypoplasticity 41

The the pycnotropy function fd controls the maximum stress ratio and its vari-
ation is shown in Figure 3.2.b. It can be seen that the pycnotropy functions fe
and fd become unity once the current void ratio reaches to the critical level, ec and
remain constant. The fd controls the maximum stress ratio at peak σ̂p and the
dilatancy. The pycnotropy factor fe , a kind of measure of current relative density,
controls the influence on the incremental stiffness. The barotropy factor fb ensures
the constitutive equation given in Eqn. 3.4 follows the isotropic compression law
(Eqn. 3.13) at isotropic loading.
The constitutive equation requires the following input parameters: the granular
hardness hs ; the critical friction angle φc ; the maximum, critical and minimum void
ratios at σ = 0: ei0 , ec0 and ed0 , respectively; the constants n, α and β appearing
as power terms in the equations.

Figure 3.3: Different intergranular strains δ related with different deformation histories. Only
the recent part of the previous strain path (bold arrow) has an influence on δ. Current stress, void
ratio and strain at current state * may be same for all three cases (after [120] ).

In order to improve the performance of the hypoplastic model in the range of


small load cycles a new state variable of the tensor form called intergranular strain,
δ was introduced [120]. At very small strain levels (< 0.01%), the model uses
a hypoelastic relation. At larger strain levels (>> 0.01%), the model uses the
normal hypoplastic equation (Eqn. 3.4 ). Niemunis and Herle [120] introduced an
interpolant to determine the hypoplastic relation between the two extreme strain
levels (Eqn. 3.19). In this function, the effect of the recent history of deformation
was also considered. The general form of the improved hypoplastic model is given
as:

σ̊ = M : ε̇ (3.18)

where M is the tangential stiffness tensor and calculated by the summation of the
hypoplastic tensors L(σ, e) and N (σ, e), which are modified according to the nor-
malised intergranular strain ρ(= δ/ ||δmax ||) and δ : ε̇. Using stiffness modification
factors mR and mT and a weighting factor ρχ , Niemunis and Herle [120] proposed
the following interpolation to determine M :
 →
−→− →
− →

ρχ (1 − mT )L : δ δ + ρχ N δ for δ : ε̇ > 0
M = [ρ mT + (1 − ρ )mR ]L + x +
χ χ

− →
− →
− (3.19)
ρχ (mR − mT )L : δ δ for δ : ε̇ ≤ 0
42 3. Constitutive Modelling

Figure 3.4: Evolution of intergranular stiffness M a. for χ = 6, b. for χ = 2; ρ varies from 0 to


1 [25].

Figure 3.4 shows the evolution of intergranular stiffness for different weighting
factor ρχ and strain directions. In this figure the strain directions are presented in
terms of the angle between the current and previous strain direction in degrees. For
example, the cases sketched in Figure 3.3a-b have the angles 180, 90 (or 270) and
0 degrees, respectively. The value of ρ depends on the strain level. For example,
if ρ = 0, then the maximum stiffness multiplier, mR and the linear constitutive
tensor, L will be active, yielding a fully hypoelastic response. If ρ = 1, then the
small strain stiffness diminishes and the constitutive tensors would evolve back to
the forms presented in Eqns. 3.5 and 3.6, yielding full the hypoplastic response.
The ||δmax || is the small strain stiffness range. mR and mT are the stiffness
multipliers for reversed (Figure 3.3.a) and tangential (Figure 3.3.b) loading.
Detailed information on the determination of the hypoplastic model parameters
can be found in literature (e.g. [79]; [130]; [8]; [114]). The parameters represent
the material properties independent of the density state. This is based on the as-
sumption that the size and shape of the particles do not change. At high stress
and strain levels, this assumption does not hold as the particles may be rounded or
even crushed. Therefore, for a realistic representation, the model should be mod-
ified so that the change in the geometry of particles is considered. In the next
section, a modified hypoplastic model [129], which account for grain crushing effects
is discussed. Furthermore some improvements on the model are proposed.

3.3 Hypoplastic Grain Crushing Model


Rohe [129] summarised the effects of microscopic properties of single grains (e.g.
mineralogy, particle shape, size and hardness, internal friction between particles)
as well as grain assemblies (e.g. geometrical distribution of particle sizes, number
of contacts between particles) and macroscopic properties of the continuous grain
assembly (e.g. particle size distribution, relative density, degree of saturation, ageing,
creep) on grain crushing. Furthermore, external influences on the grain crushing such
as the macroscopic normal and shear stresses was discussed. As a result of grain
3.3. Hypoplastic Grain Crushing Model 43

crushing properties such as density, grain size distribution, and internal friction angle
change.
In order to link the grain properties to macroscopic mechanical behaviour after
crushing, Rohe [129] traced the following properties of sand:

ˆ d50 , the mean particle size, where d50 is the sieve opening size through which
50% of the sand particles pass, respectively.

ˆ Cu , the coefficient of uniformity, Cu = d60 /d10 , where d60 and d10 are the
sieve opening size through which 60% and 10% of the sand particles pass,
respectively.

ˆ R, the particle roundness, which is the ratio of the average radius of the surface
features to the radius of maximum sphere inscribing the particle.

ˆ S, the particle sphericity, which is defined as the ratio of the diameter of the
largest inscribed sphere to the diameter of the smallest circumscribed sphere.

Rohe [129] proposed to modify the material properties of the hypoplastic model
to account for the effects of grain crushing. First of all the initial reference void
ratios, ei0 , ec0 and ed0 were modified as a function of uniformity coefficient, Cu .
Based on the data from Youd [182] and Nakata [119] following correlations were
proposed:

ˆ for R = 0.19 (angular)

emax = 1.3392 − 0.2437 · ln(Cu ) (3.20)


emin = 0.7429 − 0.1613 · ln(Cu ) (3.21)

ˆ for R = 0.37 (sub-rounded)

emax = 0.85 − 0.1730 · ln(Cu ) (3.22)


emin = 0.4807 − 0.1072 · ln(Cu ) (3.23)

ˆ for silica sand [119]

emax = 0.9501 − 0.0702 · ln(Cu ) (3.24)


emin = 0.6768 − 0.1698 · ln(Cu ) (3.25)
Furthermore, Rohe [129] performed a regression analysis on the results presented
by Nakata [119]. Based on the regression results, the following relations between the
change of uniformity coefficient, Cu and vertical stress were proposed:

Cu (σv ) = 0.0003 · σv2 + 0.0256 · σv + 2.1669 (3.26)

0.0001 · σv2 + 0.0065 · σv + 1.5742


Cu (σv , d50 ) = (3.27)
d50
44 3. Constitutive Modelling

Note that the effective vertical stress, σv is in MPa’s and the coefficients of the
σv have units in m4 /M N 2 .
Rohe [129] combined the Eqns. 3.24, 3.25 with 3.26, assuming:

eco = emax (3.28)

edo = emin (3.29)


eio = 1.15 · emax (3.30)
to obtain modified reference void ratios:

e∗co = 0.9501 − 0.0702 · ln(0.0003 · σv2 + σv + 2.1669) (3.31)

e∗do = 0.6768 − 0.1698 · ln(0.0003 · σv2 + σv + 2.1669) (3.32)


e∗io = 1.0926 − 0.0807 · ln(0.0003 · σv2 + σv + 2.1669) (3.33)

Figure 3.5: Comparison of original (dashed lines) and modified (solid lines) reference ratios: a.
change of e∗i0 , e∗c0 and e∗d0 by σv ; b. change of e∗i , e∗c and e∗d evolutions (after [129]).

Figure 3.5 shows the effect of the modifications on the reference void ratios.
Figure 3.5a compares the reference void ratios of the reference hypoplasticity, ei0 ,
ec0 and ed0 and the hypoplastic grain crushing model, e∗i0 , e∗c0 and e∗d0 . It can be
seen that in the crushing model the e∗i0 , e∗c0 and e∗d0 are reduced by increasing σv .
As a result of this modification, the reference void ratio curves, ei , ec and ed also
change (Figure 3.5b).
The formulation based on the regression analyses has shown the dependency of
the reference void ratios on the uniformity coefficient, Cu , which is changing with
vertical stress level. In order to have an objective assessment of stress level and
change in Cu , Rohe [129] introduced the following generalised forms by defining the
uniformity coefficient, Cu , the particle roundness, R and the particle sphericity, S
as functions of mean effective stress, p and shear stress, q:

R∗ = R0 + λRp ln(−p ) + λRq ln(−q), 0 < R < 1 (3.34)

S ∗ = S0 + λSp ln(−p ) + λSq ln(−q), 0 < S < 1 (3.35)


3.3. Hypoplastic Grain Crushing Model 45

where R0 and S0 are initial roundness and sphericity, respectively. The fitting coef-
ficients (λ) correlates the effect of p and q. The Cu is generalised in the form of:

Cu∗ = αp (−p )2 + αq2 + βp (−p ) + βq (−q) + Cu0 (3.36)

where, the α and β coefficients relate linear and quadratic contributions of p and q,
and have units in m4 /M N 2 .
Similarly, defining the general form of maximum and minimum void ratios based
on the maximum and minimum void ratios, emax,0 and emin,0 as:

e∗max = emax,0 − γmax · ln(Cu ) (3.37)

e∗min = emin,0 − γmin · ln(Cu ) (3.38)


where, γmax and γmin are the factors controlling the rate of emax,0 and emin,0 change.
Substituting Eqn. 3.36 into Eqn. 3.37 and Eqn.3.38, and using the relations
given in Eqns. 3.28-3.30 following form of initial reference void ratios can be found:

e∗c0 = emax,0 − Δemax


ref (3.39)

e∗d0 = emin,0 − Δemin


ref (3.40)
e∗i0 = 1.15 · {emax,0 − Δeref } (3.41)
where,
  2 

ref = γmax · ln αp (−p ) + αq (−q) + βp (−p ) + βq (−q) + Cu0
Δemax 2
(3.42)
  2 

ref = γmin · ln αp (−p ) + αq (−q) + βp (−p ) + βq (−q) + Cu0
Δemin 2
(3.43)
Rohe [129] suggests finding the fitting parameters αp , αq , βp , βq , γmax and γmin
by experimental tests. However, the method of determining these parameters was
not shown.
Considering the state at the beginning of loading, where p = q = 0, one can
obtain:

e∗c0 = emax,0 − γmax · ln(Cu0 ) (3.44)

e∗d0 = emin,0 − γmin · ln(Cu0 ) (3.45)


e∗i0 = 1.15 · {emax,0 − γmax · ln(Cu0 )} (3.46)
The Eqns.3.44 - 3.46 show that the reference void ratios are modified even at
stress-free state, which is not consistent. In order to circumvent this issue, these
initial reductions can be added to the Eqns.3.44 - 3.46 to obtain the improved for-
mulations given in Equations 3.47-3.49. For a more general description the coefficient
1.15 used to find e∗i0 is defined as ei0 /ec0 .

e∗c0 = emax,0 − Δec,max


ref (3.47)

e∗d0 = emin,0 − Δec,min


ref (3.48)
46 3. Constitutive Modelling

ei0  
e∗i0 = · emax,0 − Δec,max
ref (3.49)
ec0
where,

αp (−p )2 + αq (−q)2 + βp (−p ) + βq (−q) + Cu0
Δec,max
ref = γmax · ln (3.50)
Cu0

αp (−p )2 + αq (−q)2 + βp (−p ) + βq (−q) + Cu0
Δec,min
ref = γ min · ln (3.51)
Cu0
Another point to consider in these formulations is the behaviour during unload-
ing. The Eqns.3.39-3.41 proposed by Rohe [129] follow the current mean effective
stress, by which the modifications are reversed during unloading. However, grain
crushing is irreversible. Therefore, one would expect the modified parameters re-
main unchanged during unloading. In order to improve the model proposed by
Rohe, the history of p and q should be traced to register the maximum levels the
sand has been subjected to. Therefore instead of using the mean
 effective stress, p 
and the shear stress

at current step, one can use pmax = max p1 , p2 , · · · , pi−1 , pi
 

and qmax = max q1 , q2 · · · , qi−1

, qi in Eqns. 3.47-3.49. Employing these max-
imum stresses to modify the reference void ratios would prevent the reversal of grain
crushing modifications and the following improved formulations can be obtained:
e∗c0 = emax,0 − Δec,max,h
ref (3.52)

e∗d0 = emin,0 − Δec,min,h


ref (3.53)
ei0  
e∗i0 = · emax,0 − Δec,max,h
ref (3.54)
ec0
where,

c,max,h αp (−pmax )2 + αq (−qmax )2 + βp (−pmax ) + βq (−qmax ) + Cu0
Δeref = γmax ·ln
Cu0
(3.55)

αp (−pmax )2 + αq (−qmax )2 + βp (−pmax ) + βq (−qmax ) + Cu0
Δec,min,h
ref = γmin ·ln
Cu0
(3.56)
Figure 3.6 presents the evolution of reference void ratios of the reference hypo-
plastic model [158], the grain crushing model of Rohe [129] and the improved grain
crushing model proposed in this study. It can be seen that the improved model
preserves the crushing modifications on the reference void ratios at unloading. Fur-
thermore, there is no modification at zero stress level. It can be argued how close
the improved model predicts the crushing behaviour. The improved model provides
more realistic response than Rohe’s crushing model for unloading.
Figure 3.7 presents the comparison of high stress oedometer simulations of the
standard hypoplastic model, original [129] and improved (this study) grain crushing
models. The model parameters used in both simulations are given in Tables 3.1 &
3.2.
3.4. Conclusions 47

Figure 3.6: Reference void ratio evolutions for primary loading - unloading - reloading cycles a.
[129]; b. this study

Table 3.1: Hypoplastic soil model parameters for Baskarp sand [8]

ϕc hs n ed0 ec0 ei0 α β


◦ MPa - - - - - -
30 4000 0.42 0.55 0.93 1.08 0.12 0.96

Table 3.2: Estimated grain crushing parameters [129] for the modified grain crushing model

αp αq βp βq Cu0 γmin γmax


- - - - - - -
0.0003 0.0000 0.0256 0.0000 2.17 0.1700 0.070

3.4 Conclusions
In this Chapter first, the important aspects of the constitutive model employed in the
numerical analyses have been introduced. The capability of the model in representing
the important features of the soil behaviour such as density dependent strength
and stiffness, critical state behaviour and nonlinearity from the very beginning of
loading, stiffness dependency on loading direction has been discussed. Furthermore,
the small strain extension of the hypoplastic model has been summarized. The
reference hypoplastic model provides a flexible formulation, by which a large range
of density and stress states can be represented by one set of parameters. However,
at very high stress levels (> 5 − 10 MPa’s), grain crushing becomes evident. As a
result of crushing, it is likely to have a change in the properties of the soil and hence
the hypoplasticity model parameters.
Rohe’s [129] crushing model incorporating high stress behaviour has been sum-
marized. The model modifies the reference void ratios based on current stress level,
coefficient of uniformity, Cu and coefficients correlating the stress level to the change
of reference void ratios. Although the crushing model enables the improvement of the
48 3. Constitutive Modelling

Figure 3.7: Comparison of Oedometer compression curves for reference hypoplastic model [158]
and grain crushing models, [129] and improved crushing model (this study)

high stress behaviour, the formulation has two non-physical characteristics. First,
the model modifies the reference void ratios immediately, at zero stress level due to
γ · ln(Cu0 ) term. In this chapter, an improved formulation has been proposed so
that the reference void ratios are equal to the original reference void ratios.
The second non-physical property is the recovery of the modifications upon un-
loading. In other words, the modifications are functions of current stress level and
are independent of the stress history. In order to solve this issue, a history dependent
modification has been proposed (i.e. keeping the history of maximum stress level).
The reference hypoplastic model, Rohe’s original crushing model and the improved
model proposed in this study were compared. The crushing models have shown
slightly higher compaction levels at high stresses (> 20 MPa). At unloading stages
the proposed model has shown a more realistic response than the crushing model of
Rohe [129].
The idea of considering the crushing effects was quite promising and could be
simulated at stress point level. The performance of the model on the other hand,
has shown convergence issues during finite element simulations of the boundary
value problems. Moreover, the crushing models incorporate only the change in
the reference void ratios. Since the crushing results in a new material one would
expect a change in all hypoplastic parameters. The concept of modifying all input
parameters was discussed by Rohe [129]. The prediction of crushing behaviour might
be improved by incorporating the changes in all hypoplastic parameters; however
3.4. Conclusions 49

the assessment of the changes in these parameters would then become unpractical.
It has been shown that the crushing can be represented by a modified hypoplas-
ticity model. However, the model had convergence issues during FE simulations.
Therefore it needs a thorough investigation and further improvement in order to get
a stable and robust formulation, which is beyond the scope of this study.
As mentioned in the introduction, the modified version of von Wolffersdorff’s
hypoplastic model [114] is employed in the FE simulations in the following two
chapters. The small strain extension is not used since at large deformations such
as pile penetration; the small strain stiffness has less influence on the overall be-
haviour. In Sections 5.3 and 6.2.1, the state variables that are used to represent
the installation field are presented. Using the small strain extension would require
the assessment of the change in intergranular strain components due to installation
process. As a result, six additional state variables, δ11 , δ22 , δ33 , δ12 , δ13 and δ23
would be required for the representation of the installation effects. In order to limit
the number of state variables representing the installation field and for the sake of
practicalities, the reference model without the small strain extension will be used
throughout this study.
50 3. Constitutive Modelling
Chapter 4

Press-Replace Techniquei

4.1 Introduction
The previous chapters presented both empirical and numerical techniques, which
account for installation effects. As previously mentioned, it is important to consider
the installation effects in the analyses. It is also important to employ an objective
way to model the installation effects without using a back - calculation scheme for
each pile penetration case that is modelled. The installation effects are generally not
considered when displacement piles are analysed using commercially available or in-
house FE codes in the current engineering practice. Most of the available FE codes
used to analyse geotechnical problems employ the small displacement formulation.
However, when modelling geotechnical processes involving large displacements and
rotations, such as pile penetration, the large deformation effect should somehow be
taken into account. As an alternative to a true large deformation analysis, small
deformation formulations can be combined with a geometry or mesh update step to
account for the large deformation effects. Depending on the scheme, the technique
can be quite cumbersome, especially for practical use.
A simplified approach based on small displacement formulation, on the other
hand, can promote the modelling of large deformation effects. As a result, the
installation effects would be modelled at a reasonable accuracy. Most importantly,
once these effects are included in further FE calculations, such as modelling load
displacement behaviour of a displacement pile, and, as a result, more accurate load
- displacement behaviour can be obtained.
In this chapter, a simplified FE technique to model penetration of a pile is in-
troduced - the Press-Replace (PR) technique (Section 4.2). The PR technique can
be employed in a standard FE code. By this technique large deformation effects can
be accounted for to a certain extent by the geometry update. In Section 4.2, the
technique is validated by a cone penetration model. This section shows that reas-
onable results were obtained when compared to other simple methods (e.g. bearing
i A large part of this chapter is accepted for publishing in the International Journal for Nu-

merical and Analytical Methods in Geomechanics

51
52 4. Press-Replace Technique

capacity solutions, cavity expansion methods, SPM) and more advanced methods
(e.g. ALE, MPM).
The importance of the constitutive model in conjunction with its possibilities
and limitations was explained in Chapter 3. Considering the behaviour of sand and
the possibilities of the available constitutive models, hypoplasticity is selected and
employed in the analyses. The PR technique is further investigated by a parametric
study in Section 4.3. Based on the results of this study, optimum parameters are
suggested. In Chapter 5 the PR technique is employed to investigate the installation
effects for different pile geometries, soil densities and soil types.

4.2 Technique
The ’Press-Replace’ (PR) technique is derived from [9] who modelled penetration of
a suction pile in clay. While the method of [9] involves a load controlled scheme, the
proposed technique has an easier and automated form of modelling pile penetration
by displacement control. In the PR technique the initial FE mesh is preserved. The
material properties of the penetrated volume are updated at the beginning of each
phase resulting in a change of the global stiffness matrix without the need to update
the mesh. Hence the calculations are relatively fast compared to other algorithms
with mesh updating schemes.
The PR technique involves a step-wise updated geometry, which consist of a
straining phase followed by a geometry update. The geometry update is to model
the penetrated part of the pile (or any kind of penetrating object), which can be
achieved by modifying the global stiffness matrix at the beginning of each phase
(step 0). Subsequently, a Dirichlet boundary condition is applied to mobilise the
soil under the tip and near the shaft. The displacement controlled FE algorithm is
a kind of staged construction, the details of which are given as follows:
In each phase an updated global stiffness matrix K ph and corresponding system
with appropriate boundary conditions can be formed as:

K ph Δu = Δf ph (4.1)

where the load increment is equal to the total unbalance of phase Δf ph at the
beginning of the phase, as a result of the updated geometry:
ph,0
Δf ph = fex
ph
− fin (4.2)
ph
Here, fex is the external load vector in phase ph. The internal reaction vector
ph,0
fin can be calculated as:
ˆ
ph,0
fin = B T σ ph,0 dV (4.3)

where B T is the matrix containing the derivatives of the interpolation (shape) func-
tions and σ ph,0 is the stress state at the beginning (step 0) of the phase (= stress
at the end of the previous phase). In order to obtain an accurate and converged
4.2. Technique 53

solution, the total unbalance Δf ph is solved in multiple steps. In each step (i), the
global system:

K ph Δuph,i = Δf ph,i (4.4)

is formed, where:
ˆ
Δf ph,i = fex
ph,i
− B T σ i−1 dV (4.5)

and
 
ph,i
fex ph
= fex − λi − 1 Δf ph (4.6)

λ is a scale factor ranging from 0 at the beginning of a phase to 1 at the end


of the phase. Considering an increment, Δλi of the scale factor in step i, the scale
factor is updated according to:

λi = λi−1 + Δλi (4.7)

Since the penetration is dictated by the prescribed boundary, its degrees of


freedoms (dof’s) can be used to reduce the number of equations of the global system.
Consider the global (phase) system:

⎡ ph ph ph ⎤ ⎧ ⎫ ⎧ ph ⎫
K11 .. K1j .. K1n ⎪
⎪ Δuph
1 ⎪⎪ ⎪
⎪ Δf1 ⎪ ⎪
⎢ : ⎪ ⎪ ⎪ ⎪
⎢ .. : .. : ⎥⎥⎪


⎨ :



⎬ ⎨


⎪ :




⎢ ph ⎥
⎢K .. ph
Kjj .. ph ⎥
Kjn ⎥ Δuj ph
= Δfj ph
(4.8)
⎢ j1
⎢ ⎥⎪⎪ ⎪ ⎪
⎪ ⎪ ⎪

⎣ : .. : .. : ⎦⎪ ⎪
⎪ : ⎪ ⎪



⎪ : ⎪ ⎪


⎩ ⎪
⎭ ⎪
⎩ ⎪
ph ph ph ph ph ⎭
Kn1 .. Knj .. Knn Δun Δfn

If the prescribed boundary at j th dof has a value of aph


j , the j
th
equation can
ph ph
be simplified to Δuj = aj [122]. This operation can simplify the equation system
given in Eqn. 4.8 to :

⎡ ph ph ⎤
⎧ ph ⎫ ⎧ ph ph ⎫
K11 .. 0 .. K1n ⎪
⎪ Δu1 ⎪ ⎪ ⎪
⎪ Δf1 − K1j ⎪

⎢ ⎥⎪⎪










⎢ : .. : .. : ⎥⎪ : ⎪ ⎪ : ⎪
⎢ ⎥ ⎨ ph ⎬ ⎨ ⎬
⎢ 0 .. 1 .. 0 ⎥ = a (4.9)
⎢ ⎥ ⎪Δuj ⎪ ⎪ j

⎢ ⎥⎪⎪ : ⎪ ⎪ ⎪
⎪ ⎪

⎣ : .. : .. : ⎦⎪ ⎪ ⎪
⎪ ⎪
⎪ : ⎪

⎪ ⎪ ⎪ ⎪
ph ph ⎩ ph ⎭ ⎩ ph ph ⎭
Δfn − Knj
Kn1 .. 0 .. Knn Δun

Substituting zeros into j th equation and reduces the solution without destroying
the symmetry of the stiffness matrix. However, the reaction force Δfjph cannot be
solved. This disadvantage can be avoided by keeping the original stiffness matrix
K ph , but skipping the equations of the prescribed dof’s during decomposition and
54 4. Press-Replace Technique

Figure 4.1: Details on the Press-Replace modelling technique and progress of penetration of the
pile

back substitution process. Once all the displacements are calculated, the reaction
force Δfjph can be calculated using a partitioned form as described by Potts and
Zdravković [122].
In each incremental step i the force equilibrium is checked by an iterative scheme.
The iterative scheme is used to bring each step in ’equilibrium’. Equilibrium, in this
respect, is defined for the part of the unbalance that is supposed to be solved in
this step (λi Δf ph ). Hence, for the phase as a whole, a part of the total unbalance
(1 − λi )Δf ph still exists.
If the tolerance is satisfied, the displacement increment of the current step can
be represented by Δuph,i . The displacement of the current step uph,i can therefore
be updated by:

uph,i = uph,i−1 + Δuph,i (4.10)

The calculation procedure described by (Eqn. 4.1) to (Eqn. 4.10) is applied for
each phase until the desired penetration depth is reached.
In Figure 4.1, a general view of the model and four subsequent calculation phases
are presented. The penetration of the pile (indicated by blue material) can be also
seen in the same figure. The calculation phases continue until the desired level of
penetration of the pile is achieved. The straining is realized by a Dirichlet boundary
condition (vertical prescribed displacement) on the surface so that proper shearing
and bearing resistances are mobilized. During the geometry update, the properties
of the region that is just displaced are switched to pile properties. This is necessary
since the calculations are based on small deformation theory such that the reference
geometry does not change. The geometry update is also necessary in order to have
the pile installed in the mesh together with the installation effects at the end of the
analysis. In this way, the analysis does not suffer from mesh distortion effects. More
importantly, the installation effects, which are necessary to model service behaviour
of the jacked pile, are reasonable modelled.
In order to facilitate the PR technique, an axisymmetric FE mesh with small
slices (i.e. tslice ≈ D/8, where D is the pile diameter) is introduced in the region
where the pile will be jacked (Figure 4.1.a&b). Interface elements are also defined
between the pile tip and the underlying soil to model proper interaction with the
continuum (Figure 4.1.b). In order to avoid stress oscillations at the corner of the
4.2. Technique 55

pile tip, interfaces should be extended into the soil volume [153]. In this study, the
effect of the interface extension length on the stress distribution and CPU times
int
is investigated considering four different extension lengths (lext = 0.5, 1, 1.5 and
2·tslice ). No distinct influence on the overall behaviour was observed. Therefore, for
practical purposes, vertical as well as horizontal interface extension lengths equal to
int
the slice thickness (lext = tslice ) are used in the analyses.
The smoothness of stress and other state variable distributions increases using a
smaller step size and finer mesh [59]. Moreover, if the step size is too large, some
of the soil elements, which are not displaced to the desired extend, will be switched
to elastic pile material during the material update phase. This will result in lower
mobilisation of the soil and therefore lower capacity than expected during a real pile
penetration case. However, there is also a practical limit for the smallest step size
due to computational requirements. If the step size is too small, the resulting mesh
will be too fine and therefore the total number of elements will be too large. As a
result, the system will not be practically solvable. The optimum slice thickness and
step size is determined by a parametric study, which is presented in Section 4.4.
In large deformation formulations like Updated Lagrange (UL), objective stress
measures, such as Jaumann, Hill or Truesdell rate of stress, are used. The aim is to
determine the stress in an objective manner in a deforming geometry where large
rotations, volumetric changes and shearing are expected.
The Jaumann stress rate can be formulated as:
J
σ̊ij = σ̇ij − σik Ωkj − σjk Ωki (4.11)
J
where σ̇ij is the Cauchy stress rate and Ωkj is the spin tensor.
The Hill rate of stress is given as:
H J
σ̊ij = σ̊ij + ε̇kk σij (4.12)

where the effect of the volumetric strain rate ε̇kk is also included.
The Truesdell rate of stress, where all deformation effects are considered is given
as:
T
σ̊ij J
= σ̊ij + ε̇kk σij − σik ε̇kj − σjk ε̇ki (4.13)

The velocity field u̇i is interpolated from FE nodal displacements with shape
functions Nij as:

u̇i = Nij ȧj (4.14)

where ȧj are the nodal velocities. The velocity gradient u̇i,k , the strain rate tensor
ε̇ik and the spin tensor Ωik can be described as [153]:

u̇i,k = Nij,k ȧj = Lijk ȧj (4.15)


56 4. Press-Replace Technique

Components of the velocity gradient are used to calculate the strain rate tensor
and the spin tensor:
1
ε̇ik = (Lijk + Lkji ) ȧj = Bijk ȧj (4.16)
2
1
Ωik = (Lkji + Lijk ) ȧj = Cijk ȧj (4.17)
2
In the PR technique, the standard Cauchy stress is used. Unlike Jaumann and
Hill rates of stress, the rigid body rotations, the volume change and shearing are
not taken into account. In order to compare the PR technique with large strain
solutions using these objective stress rate measures an oedometer test is simulated.
For an oedometer test [153] showed that the Truesdell stress strain relationship
can be described by:
 
σ = E eε H − 1 (4.18)

where, E is Young’s Modulus; Poisson’s ratio is chosen equal to zero and εH is the
logarithmic (Henky) strain that can be defined as:
 
l
εH = ln (4.19)
l0
Using the Hill stress rate, the stress-strain relation is given as:
 
H
σ = E 1 − e−ε (4.20)

Since the strain is defined as logarithmic and there is no rotation of stresses in


the oedometer test, the Jaumann stress rate reduces to the Cauchy stress rate and
the stress-strain relation becomes:

σ = EεH (4.21)

The linear stress-strain relation is used the PR technique is simply:

σ = Eε (4.22)

The strain ε is the summation of linear (engineering) strain of each step resulting
in a discrete form like:
 li − li−1
ε= (4.23)
li
where the current length can be defined as:

li = l0 − iΔl (4.24)

If the step size is infinitesimally small the total strain can be defined as:
 li − li−1
ε|Δl→0 = lim (4.25)
Δl→0
i
li
4.2. Technique 57

Figure 4.2: Oedometer curves for different stress rate definitions and the PR method (after [153])

Substituting Eqn.4.24 into Eqn.4.25 will result in the logarithmic strain Eqn.4.26
as:
 ˆ l  
Δl dl l
lim = = ln = εH (4.26)
Δl→0
i
l0 − iΔl l0 l l0

It is proven in Eqn.4.23 to Eqn.4.26 that the strain obtained by the PR technique


converges to logarithmic strain when the step size is infinitesimally small (at least
under the condition of a one-dimensional compression test).
Recalling the engineering (linear) strain ε, (Eqn.4.27), gives the same stress-
strain description as Truesdell when the logarithmic strain description Eqn.4.19 is
substituted into Eqn.4.18.
l Δl
ε= −1= (4.27)
l0 l0
In order to compare the PR with the objective stress rate measures with instant
logarithmic strain descriptions (Eqn.4.26), an oedometer test is simulated by PR
technique in a FE analysis for three different step sizes 0.01L, 0.05L and 0.10L,
where L is the height of the specimen. In addition to numerical simulations, an
analytical form is also used to check convergence of the method. This enabled
analysing quite small step sizes, which otherwise would not be possible to simulate
(due to meshing issues) in a FE calculation.
Figure 4.2 presents the results of different rates of stress as well as the PR
technique (both FE and analytical) simulations. It can be seen that the PR method
58 4. Press-Replace Technique

converges to the Jaumann stress rate result. This shows that the step-wise geometry
update is similar to a UL formulation using a common stress measure (Jaumann)
when there is no (or less significant) stress rotation. As mentioned earlier, volumetric
effects as considered in the Hill stress rate are not taken into account in the PR
technique.

4.3 Validation
In order to check the capabilities and limitations of the PR technique, some veri-
fication and validation cases from literature have been selected and the results are
discussed separately in the following subsections.

4.3.1 Cone Penetration in Undrained Clay


One important issue is to verify of the PR technique’s capability to properly model
the introduction of the pile in the soil. In a large deformation analysis (e.g. UL,
ALE, MPM) the volume intrusion of the penetrating pile is dictated by the pen-
etrating object. However, in the PR technique, the volume intrusion is indirectly
captured due to straining under the tip of the pile that is pushed into the soil and
the corresponding Poisson’s effect.
In order to check the significance of this issue, an undrained analysis was per-
formed for a cone penetrating in clay and the volume conservation was investigated
(Figure 4.3). The cone had a diameter, Dcone = 36 mm. The slice thickness equal
to a step size of ts = uy =8 mm (≈ D/5) was selected. The slices were defined
compatible with the cone tip, which had an apex angle of 60◦ . For the sake of
simplicity, calculations started when the cone head was already embedded in the
soil. For the modelling of the clay an elastic - perfectly plastic material with Tresca
yield surface was employed. The effect of soil stiffness was investigated based on the
cases analysed by [152]. Undrained behaviour was modelled by using two different
Poisson’s ratios, ν = 0.490 (as in [152]) and ν = 0.499, which is closer to the limiting
value ν = 0.500 for zero volumetric strain of undrained materials. The parameters
used for different cases analysed and the corresponding results are summarized in
Table 4.1.
The volume displaced due to a penetration of 0.40 m is calculated by integrating
the surface heave uy (r) (Figure 4.4) considering the axisymmetric geometry. Since
2
the penetrated volume is known (i.e. Vcone = πrcone Lpen = π(0.018)2 (0.400) =
−4
4.07 · 10 m ), the volume loss, Vloss can be calculated as:
3

ˆ R
Vloss = uy (r)dr − Vcone (4.28)
Dcone /2

where R is the FE model width.


In Table 4.1, it can be seen in that there is a minor volume loss arising from
the PR technique when a Poisson’s ratio of ν = 0.499 is used. The (inevitable)
compressibility introduced by using Poisson’s ratio, ν = 0.499 instead of ν = 0.500,
4.3. Validation 59

Figure 4.3: FE model for the volumetric analysis of a cone penetrating in an undrained clay: a.
global view of the mesh configuration at the beginning of analysis; b. details on the mesh: green
wedge indicating the embedded cone head, inclined slices prepared for the cone to be penetrated.

Figure 4.4: Surface heave near the cone at the end of 0.40 m penetration (max.uy = 6.67 · 10−3 m)

Table 4.1: Parameters and results of the cases analysed for volume preservation check (c = 50
kPa)

E ν G Ir = G/c Vheave Vloss Vloss = (VCP T − Vheave /Vcpt )


kPa - kPa - m3 m3 %
3000 0.490 1007 50 5.63E − 04 4.39E − 05 10.8
3000 0.499 1001 50 4.03E − 04 3.76E − 05 0.9
6000 0.490 2013 101 3.78E − 04 2.94E − 05 7.2
6000 0.499 2001 100 4.05E − 04 2.33E − 05 0.6
12000 0.490 4027 201 3.88E − 04 1.96E − 05 4.8
12000 0.499 4003 200 4.05E − 04 1.87E − 05 0.5
30000 0.490 10067 503 3.96E − 04 1.15E − 05 2.8
30000 0.499 10007 500 4.06E − 04 1.61E − 05 0.4
60 4. Press-Replace Technique

is the major cause of the volume loss calculated. Additionally, there is a minor
loss due to the small compressibility of the interfaces under the tip. The observed
volume loss can be entirely attributed to the sum of these effects. Therefore it can
be concluded that the penetrating volume of a pile is properly modelled using the
PR technique.

Figure 4.5: Relation between Ir and Nc obtained from PR and other techniques (α = 1 full
frictional/ rough, α = 0 frictionless/smooth contact).

In order to validate the PR technique a cone penetration in undrained clay is


analysed. The aim is to compare the cone tip resistances (qc ) for different soil
stiffness and shear strength values. Actually, the verification case analysed here,
has already been studied extensively to derive a relation between the rigidity index
of the soil, Ir , (Eqn.4.29) and the normalised cone resistance, Nc , (Eqn. 4.30) by
several researchers (e.g. [156]; [148]; [152]). The aforementioned quantities can be
formulated as:
G
Ir = (4.29)
cu
qc
Nc = (4.30)
cu
cint
α= (4.31)
cu
where, α is the interface strength reduction factor and cint is the shear strength of
the interface.
The surface roughness has also been analysed and compared with previous studies
(e.g. [106]; [159]). As expected, increasing the strength reduction factor, α, also
increases the normalised cone resistance, Nc (Figure 4.5).
4.3. Validation 61

Figure 4.6: Current plastic zones during steady state for soil rigidity indices: a) Ir = 150 and
b) Ir = 300, full frictional contact; c) Variation of plastic zone with soil rigidity index based on
simulation results of RITSS (after Lu et al. [106])

For this validation part, the cases listed in Table 4.1 were analysed. However, to
be compatible with [152]’s simulations, only the Poisson’s ratio, ν = 0.490 was used
in all cases (except for a few published MPM cases that employed a Poisson’s ratio
of ν = 0.400) to model nearly undrained behaviour.
In the PR technique, the cone tip resistance values were calculated by integrating
the vertical stresses acting on the cone tip. In Figure 4.5, Ir vs. Nc values are plotted
in comparison with results reported by several researchers based on analytical/semi-
analytical as well as ALE and MPM solutions.
It can be seen that the PR technique gives an average behaviour within other
results. The rate of increase in Nc values with increasing Ir is slightly higher than
the general trend. This is mostly due to the pronounced effect of the stress concen-
trations at the cone tip and shoulder. Using a smoothed discretisation of the cone
(or conical pile), has been proven to avoid stress oscillations [140].
The plastic zones obtained at different depths of cone penetration for a rigidity
index of Ir = 100 for rough contact (α = 1.0) by adaptive remeshing and contact
algorithm and PR technique are given in Figures 4.7 and 4.8. Figure 4.6.a and 4.6.b
shows the plastic points for rigidity indices of Ir = 150 and Ir = 300 for rough cone,
smooth shaft. Lu et al. [106] summarized the size of plastic zones depending on the
soil rigidity index in Figure 4.6.c. It can be seen that the PR simulations reach to an
approximate steady state around a penetration depth of z/D = 9, which confirms
Walker & Yu [160]. However, the plastic zones obtained by PR (Figure 4.8 about
3.5 R) are less than Walker & Yu [160] simulations results (Figure 4.7 about 5 R) as
well as the Lu et al. [106] (RITSS) results (Figure 4.6.c about 5 R). Smaller plastic
zone obtained for the PR simulation can be explained by the removal of plastic zone
after each material switch (i.e. plastic soil to elastic pile). The amount of removal,
hence the reduction in plastic zone can be minimised by using smaller steps in the
PR technique.
62 4. Press-Replace Technique

Figure 4.7: Plastic zones around the penetrating cone simulated using explicit adaptive meshing
with contact (after Walker & Yu [160] at different penetration depths (Ir = 100 and α = 1.0, full
frictional contact).

Figure 4.8: Plastic zones around the penetrating cone simulated using PR technique at different
penetration depths (Ir = 100 and α = 1.0, full frictional contact).
4.3. Validation 63

4.3.2 Pile Penetration in Undrained Clay


In this second validation case, the results of the study on two types of zipper tech-
niques were compared with the proposed PR technique. The zipper type techniques
with contact elements open up a pre-bored small cavity by further penetration. The
method was employed to model a pressuremeter test [47] and pile penetration ([112];
[77]; [140]). In this part, 0.30 m penetration of a pile of diameter 0.10 m, which was
pre-embedded at 2 m depth in an undrained soft clay was investigated. An elastic
- perfectly plastic model with Tresca yield criterion was used as constitutive rela-
tion. The undrained shear strength, cu was 4.5 kPa. For Poisson’s ratio, ν = 0.49
was used to model nearly incompressible behaviour of the soil. An elastic modulus,
E = 2025 kPa giving a rigidity index of Ir = G/cu ≈ 150 was used.
The first type of modelling technique investigated by [59] is a kind of zipper
technique employed in a standard finite element (FE) model to model the penetra-
tion of a suction anchor with a tapered steel skirt tip. Penetration was modelled
by employing prescribed horizontal displacements that incrementally update the
mesh geometry to model the penetration of the pile. The second type of modelling
technique used a contact algorithm. The corners were smoothed to prevent stress
concentrations. A frictionless contact surface, which behaves like a zipper, was used.
The PR technique is employed with a step size equal to the slice thickness, tslice , of
0.02 m.
Table 4.2: Parameters and results of the cases analysed for volume checking control

Average vertical stress Bearing capacity factor


Solution according to
at pile tip σ̄ tip kPa Nc = σ̄ tip /cu
Cavity expansion [70], [184]
27 6.0
Cavity expansion [134] 30.5 6.8
SPM+FEM [148] 47.16 10.5
Empirical [142] 40.5 9.0
Zipper with prescribed disp.
35 7.8
[59]
Zipper with Contact
40 8.9
[59]
This study (PR technique) 40 8.9

The normalized horizontal and vertical stresses are compared with the two zipper
techniques reported by [59] in Figure 4.9. The results of [59] shown here are the
ones in which large deformation theory (updated Lagrange) is employed. As can
be seen from this figure, the vertical and horizontal stress distributions are quite
comparable with the ones obtained from the PR technique. The stress distributions
obtained from the PR analysis are even better than the large deformation solution of
the first zipper technique in which prescribed displacements are employed to model
the penetration. This is because the PR method mobilizes the soil stresses more
appropriately and does not suffer from the mesh distortion problem. However, a
rough contact model was used in the PR technique due to a numerical difficulty
in the analysis (concentration of stresses at sharp corners, i.c. pile tip and pile
shoulder). The contact model, on the other hand, does not suffer from this numerical
issue since it uses a smoothed geometry and smooth contact.
64 4. Press-Replace Technique

The cone resistance can be validated by some analytical solutions based on the
classical bearing capacity solution [142], cylindrical cavity expansion theory ([70];
[184]) and as well as the Strain Path Method (SPM) solutions [148]. The bearing
capacity values and the corresponding bearing capacity factors obtained from several
analytical techniques, the two zipper techniques studied by [59] and the PR technique
are summarised in Table 4.2. The two cavity expansion solutions give relatively low
tip resistances. Even though the two zipper techniques resemble cavity expansion
behaviour, as the far field elements [35] were not employed, higher bearing values are
obtained. The tip resistances obtained in this study lie within this range (≈ 27 − 47
kPa).
The strain path method (SPM) give higher tip resistance. The SPM solution is
based on the integration of the strain field around a pile with a flat tip. In reality,
after the pile tip resistance is fully mobilised, a cone is formed and from this level
on the flat and conical pile behaves similarly. However numerical solutions vary in
different geometries of pile tip to overcome stress concentrations. Besides, depending
on the method used different mobilisation patterns around the model pile can be
observed.
Based on the validation cases, the PR technique seems to give reasonable resist-
ances compared to other analytical and numerical solutions. Since the constitutive
model used in the simulations (elastic-perfectly plastic model with Tresca yield sur-
face) does not have complex features like hardening-softening behaviour, the influ-
ence from the constitutive model is minimised and therefore, the performance of the
PR technique could be verified more objectively.
As the main objective of this research is to model pile penetration in sand and
its consequent effects, proper sand models should be employed. Therefore, the per-
formance of the PR technique should also be verified in conjunction with advanced
sand models. For a proper modelling of sand the main ingredients of the Critical
State Theory such as the state dependent mobilisation of strength, volumetric be-
haviour and stiffness should be considered. From this perspective Hypoplasticity
after [74] and [158] is an appropriate choice as it has the required Critical State
description. Moreover an elegant and robust formulation allows modelling of a par-
ticular sand type with one set of parameters. The void ratio controls the volumetric
behaviour and the stiffness is also based on the stress level. It provides the advant-
age of having one parameter set to describe all densities, when compared to most
other strain-hardening models. Moreover, the hypoplastic model also accounts for
softening behaviour. All the features mentioned above are essential to model the
sand behaviour during and after the pile installation process. In the next part, the
PR technique is investigated thoroughly using the hypoplastic soil model.

4.4 Parametric study


In the previous part the ’Press-Replace’ technique was explained and the validation
cases were presented. In order to clarify the possibilities and limitations of the
PR technique, this section presents a parametric study. In all of the variations, an
axisymmetric FE model was employed with a 7.175 m embedded pile with a diameter
4.4. Parametric study 65

Figure 4.9: Penetration of a frictionless embedded conical pile. Normalised horizontal (left
column) and vertical (right column) stresses (σxx /su and σyy /su , respectively) for a. zipper tech-
nique by prescribed displacements (Plaxis 2D) b. zipper technique by contact algorithm (Abaqus)
c. PR technique (Plaxis 2D)

D = 0.524 m and 1 m of penetration was analysed.


Another important objective of this parametric study was to find an optimized
way of selecting appropriate geometrical settings like, step size (or geometry update
frequency), element type, mesh density and interface stiffness properties, to increase
66 4. Press-Replace Technique

Table 4.3: Hypoplastic soil model parameters for Baskarp sand (after [8])

ϕc hs n ed0 ec0 ei0 α β



MPa - - - - - -
30 4000 0.42 0.548 0.929 1.080 0.12 0.96

numerical performance without losing accuracy. In the following sections these para-
meters and their influence on the model behaviour are explained in detail. For the
parametric variations, a fixed set of hypoplastic parameters of Baskarp sand [8],
given in Table 4.3, was used. Chapter 3 presented more details on Hypoplasticity.

4.4.1 Element Type and Mesh Density


In this part, the effect of element type as well as mesh density is investigated by two
different element types, quadratic (6-noded triangular) and fourth-order (15-noded
triangular) elements having three different mesh densities, named coarse, medium
and fine. The mesh density definitions are based on the relative size of the slice (ts )
and the size of the element (te ) in each slice defined such that:

Coarse: te = ts

Medium: te < ts ≤ 3 · te
Fine: 3 · te < ts ≤ 6 · te

It was possible to complete the analysis for all coarse mesh cases in a reasonable
calculation time (≈ 6 hrs/m penetration) compared to large deformation simulations
using advanced soil models. The convergence rates of the medium and especially
fine mesh cases were quite slow.
It can be seen from Figure 4.10 that the case with higher order elements has
smoother stress distributions even though the medium lower-order mesh has a higher
total number of nodes (and integration points) than the coarse higher-order element
mesh, in the area of influence (i.e. pile tip).
Hughes & Belytschko [84] state that for a smooth problem, the rate of conver-
gence depends on the order of completeness (order of interpolation). They also
indicated that for non-smooth materials, the problem accuracy does not improve.
Considering relative smoothness of the hypoplastic constitutive relation, it can be
expected that the convergence rate would improve by increasing the order of inter-
polation, or in other words, by using higher order elements. A recent study by [131]
has shown an increased accuracy with the same number of unknowns when compared
with a lower-order FE model for the solution of the eigenvalue problem for arbitrary
inhomogeneous dielectric axisymmetric resonators. They also noted that for a given
accuracy, higher-order elements could significantly reduce the memory requirement
and calculation time when compared to lower-order ones. In the view of the findings
presented herein and the suggestions from literature, it can be concluded that using
higher order elements for PR calculations would be a proper choice in order to have
4.4. Parametric study 67

Figure 4.10: Comparison of distribution (smoothness) of effective mean stresses (p ) of cases with
high order (15-noded) elements (coarse mesh) and lower order (6-noded) elements (medium mesh)
at 1m penetration level.

not only smoother distribution of the state variables but also improve numerical
performance (i.e. increased calculation speed).
Considering the convergence rates obtained for the coarse mesh and the smooth-
ness for the higher order elements, it is suggested to use a coarse mesh with high
order (15-noded) elements when the PR technique is employed.

4.4.2 Step Size


The Press-Replace technique is based on displacement (Press) and material switch
(Replace) of the predefined slices from hypoplastic soil material to linear elastic pile
material. Therefore it is important to investigate the effect of the step size (material
update or the penetration length achieved at each step). In the previous part it
was shown that using a relatively coarse mesh and higher order elements result in
a better convergence rate and smoother distribution of state variables. In this part,
models with four different slice thicknesses (ts ) of 1 cm, 2.5 cm, 5 cm and 10 cm for
the coarse mesh (Fig. 4.11) with higher order (15-noded triangular) elements with
compatible step sizes were considered (i.e. for ts = 2.5 cm step size of 2.5 cm, 5 cm
and 10 cm). The case with slice thickness, ts = 1 cm could not be analysed due to
extreme computational demand and convergence issues. Furthermore, meshing was
not possible for a reasonable pile length (L = 4m). This case is presented in Figure
4.11 as a reference lower bound for the method.
First, for each different mesh density with slice thicknesses of ts = 2.5 cm, 5
cm and 10 cm, the effect of the step size was investigated by employing prescribed
displacements (= penetration length) of uy = 1 cm and 5 cm for ts = 1 cm, uy =
2.5 cm, 5 cm and 10 cm for ts = 2.5 cm, and uy = 5 cm and 10 cm for ts = 5 cm.
The summary of these analyses is given in Table 4.4. Secondly, the effect of mesh
density was investigated using the same prescribed displacements in different mesh
68 4. Press-Replace Technique

Figure 4.11: FE meshes of slice thickness of a. ts = 1 cm, b. ts = 2.5 cm, c. ts = 5 cm, and d.
ts = 10 cm used in the comparison of step size and mesh density (Dark grey is the pile material).

configurations where different slice thicknesses, ts were generated (i.e. for uy = 5


cm cases with ts = 2.5 cm & 5 cm). The summary of these cases is given in Table
4.5.
Table 4.4: Summary of analysis for the investigation of step size

Mesh type / Step size /


Case
Slice thickness, ts Prescribed displacement, uy
cm D cm D
T10U10 1.0 ≈ 0.02 1.0 ≈ 0.02
T025U025 2.5 ≈ 0.05 2.5 ≈ 0.05
T025U050 2.5 ≈ 0.05 5.0 ≈ 0.10
T025U100 2.5 ≈ 0.05 10.0 ≈ 0.19
T050U050 5.0 ≈ 0.10 5.0 ≈ 0.10
T050U100 5.0 ≈ 0.10 10.0 ≈ 0.19

Table 4.5: Summary of analysis for the investigation of mesh density

Mesh type / Step size /


Case
Slice thickness, ts Prescribed displacement, uy
cm D cm D
T025U025 2.5 ≈ 0.05 5.0 ≈ 0.10
T025U050 5.0 ≈ 0.10 5.0 ≈ 0.10
T025U100 2.5 ≈ 0.05 10.0 ≈ 0.19
T050U050 5.0 ≈ 0.10 10.0 ≈ 0.19
T050U100 10.0 ≈ 0.19 10.0 ≈ 0.19

In Fig. 4.12 the effect of the step size uy for two different meshes with a slice
thickness of ts = 2.5 cm and ts = 5 cm, respectively is shown. An acceptable
(≈ 6.5%) difference in penetration resistance values at a depth of L = 1 m between
the cases with the lowest (T100U100) and the highest (T025U100) vertical force is
obtained.
It can be seen for both of the mesh types that there is an abrupt reduction in
reaction force close to the beginning of each calculation phase after the material
4.4. Parametric study 69

Figure 4.12: Penetration resistance curves of cases with varied step sizes for two different meshes
uy = 2.5cm (blue lines) and uy = 5cm (red lines), namely.

Figure 4.13: Penetration resistance curves of cases with step sizes of a. uy = 5 cm, and b. uy =
10 cm for different slice thicknesses (ts = 2.5 cm, ts = 5 cm and ts = 10 cm)

update was performed. The reduction amounts seem to increase with increasing
depth as well as with increasing step size, uy . This is mainly due to the relative
shearing at the pile tip interface. The interface shear strength assumed and the
mobilised shear strength of the soil are not equal. This issue is further investigated
and summarised in Section 4.4.3.
In Fig. 4.13 it can be seen that there is not much difference in total reaction to the
pile (max. 6.5%). Therefore selecting a larger element and step size would speed up
70 4. Press-Replace Technique

the calculations. This might be necessary especially for large penetration lengths.
However, using finer slices and employing small steps results in smoother stress
distributions around the tip of the pile and smoother load - displacement responses.
In this respect, the case T050U050 (ts = uy = 5 cm ≈ 0.10D) by which a reasonable
accuracy and an optimum numerical performance was obtained, is recommended.

4.4.3 Interface Properties


As previously mentioned, [153] introduced the concept of using interface elements to
overcome singularities at soil-structure interaction problems in FE. The idea is based
on introducing interfaces to improve the mesh first with potential slip planes and
secondly, with singular point treatment. The idea of using the interface extensions
to treat singularities and smoothed the stresses around the corner has proven itself
satisfactorily. On the other hand, the determination of the stiffness and strength
parameters for soil types showing hardening and softening behaviour can be com-
plicated. In this perspective, the intention of this section is to give more detailed
information on the determination of interface properties for the hypoplastic soil
model. Furthermore, a parametric study on the selection of appropriate interface
stiffness is given.

Interface Stiffness
The effect of interface stiffness is investigated by the finite element model with a slice
thickness of 5 cm (ts = uy = 5 cm) T050U050. Hypoplastic model parameters of
dense (ID = 0.77) Baskarp sand were used. In order to change the interface stiffness
values either the (virtual) interface thickness or the interface shear stiffness, Gi (and
therefore the interface oedometric stiffness Eoed,i ) has to be modified. The effect of
these modifications is purely numerical.
The interface shear kS and normal stiffness kN values can be determined as [153]:
Gi
ks = (4.32)
ti
Eoed,i
kN = (4.33)
ti
1 − νi
Eoed,i = 2Gi (4.34)
1 − 2νi
where the interface Poisson’s ratio is assumed to be νi = 0.45.
In Table 4.6 a summary of cases analysed is given. In these analyses, the (virtual)
interface thickness ti is a fixed value (ti = 0.157803 m) defined by the FE software
based on the average element size of the FE mesh (0.10·tave ). The oedometer and
shear stiffness of the soil was Eoed,s = 6.5 · 104 kN/m3 and Gs = 1.86 · 104 kN/m3 ,
respectively. The interface shear stiffness ks (and therefore interface normal stiffness
kN ≈ 11 · ks ) value was varied by using different values of Gi (and as a result Eoed,i ).
The Gi values were found using a stiffness multiplier, χ∗E = χE · ti /te , where te is the
average element size under the pile tip and χE = {16, 8, 5, 4, 3, 2.5, 1.5, 1, 0.5, 0.25},
4.4. Parametric study 71

for the cases 1 to 10, respectively. The equivalent soil stiffness Gs /ts and Eoed,s /ts
are defined in order to normalise the ks and kN values. A slice thickness ts = 0.05
m is used for this part.

Table 4.6: Summary of cases analysed in parametric study on the interface stiffness
  E 
Case χ∗E Gi Eoed,i ks kN ks / Gs
ts kN / oed,s
ts
- - kPa kPa kPa/m kPa/m - -
1 68.1 1.26 · 106 1.39 · 106 8.01 · 106 8.81 · 107 21.6 67.8
2 34.0 6.32 · 105 6.95 · 106 4.00 · 106 4.41 · 107 10.8 33.9
3 21.3 3.95 · 105 4.34 · 106 2.50 · 106 2.75 · 107 6.7 21.2
4 17.0 3.16 · 105 3.48 · 106 2.00 · 106 2.20 · 107 5.4 16.9
5 12.8 2.37 · 105 2.61 · 106 1.50 · 106 1.65 · 107 4.0 12.7
6 10.6 1.97 · 105 2.17 · 106 1.25 · 106 1.38 · 107 3.4 10.6
7 6.4 1.18 · 105 1.30 · 106 7.51 · 105 8.26 · 106 2.0 6.4
8 4.3 7.90 · 104 8.69 · 105 5.01 · 105 5.51 · 106 1.4 4.2
9 2.1 3.95 · 104 4.34 · 105 2.50 · 105 2.75 · 106 0.7 2.1
10 1.1 1.97 · 104 2.17 · 105 1.25 · 105 1.38 · 106 0.3 1.1

The variation of the stiffness values of the interfaces used in different parts (such
as tip and shaft of the pile and the extensions) independently is quite a rigorous task
and was not examined. Rather, the global response was investigated by varying the
interface stiffness value that is assigned to all interfaces at once.

Figure 4.14: Penetration resistance curves of cases with varied interface stiffness

The penetration resistance (or simply force - displacement, Fy - uy ) curves for


the cases with different interface stiffness are given in Fig. 4.14. For Case 1, in
which very high stiffness values of kN and ks were assigned, no convergence could be
obtained. This level of stiffness can, therefore, be considered as the upper limit. In
the other cases (Cases 2-10), it can be seen that there is not much difference in the
penetration resistance values. There is almost no difference up to 0.20 m penetration
depth. The maximum difference in the vertical resistance, Fy between Cases 2 and
10 at 1.0 m penetration level is 88 kN. At this penetration level the mean value of
the vertical resistance Fy is 2814kN. This indicates a (minor) difference of 3.14%.
72 4. Press-Replace Technique

Figure 4.15: Normalised CPU times of all cases.

Normalised CPU times of all cases are given in Fig.4.15. Case 1 is not included
in this plot as a converged solution could not be obtained. Case 2 can be selected as
the reference case, which has the most accurate solution with reasonable convergence
rate. It can be seen that the CPU times decrease by decreasing interface stiffness.
After Case 4, there are two distinct groups of analyses (Cases 5-6 and Cases 7-10)
having more or less the same CPU times. In this perspective and to save some space,
only the void ratio (e) and mobilised friction angle (φmob ) plots for Case 2, 5, 7 and
10 are compared in Fig. 4.16.a & b, respectively.
The influence of interface stiffness is quite evident for Case 10. The mobilisation
of soil friction is not sufficient and clearly affected by the low interface stiffness.
Considering the computational efficiency, one might prefer using Case 7 (or even
Case 9). However, for Case 7 the influence of the interface can be seen in Fig. 4.16
even though it is not as significant as the Case 10.
It can be deduced from Fig. 4.16.a that the interface stiffness has an influence on
the mobilisation of stresses in the soil. As the interface gets stiffer (i.e. Case 2), the
amount of deformation of the interface decreases and therefore the amount of stresses
transferred to the soil (elements) increases, which is more realistic. Furthermore, it is
found that the CPU time decrease with reduced interface stiffness. The selection of
interfaces stiffness is purely a numerical issue and the interfaces should not influence
the model behaviour. In view of the results obtained, Case 5 can be selected as
the one with optimum interface stiffness. In other words, using a stiffness ratio,
E
ks /[ Gtsi ] ≈ 4.0 and kN /[ oed,i
ts ] ≈ 12.7 is recommended.

Interface Strength
The interface extension length may have an influence if the interface strength is not
equal to the strength of the surrounding soil volume. This issue was already evident
(as reduction in force at the beginning of each phase) in the force-displacement
responses given in Fig. 4.12 & Fig. 4.13. This behaviour indicates the importance
of defining the interface strength equivalent to the soil strength. However, if the
interface behaviour is limited to elastic - perfectly plastic behaviour, as in most
FE codes and the FE code used in this study (Plaxis 9.02), appropriate strength
4.4. Parametric study 73

Figure 4.16: a. Mobilised friction angle, φmob (first row) b. Void ratio, e (second row) distribution
at penetration level, z= 1.00 m for Cases 2, 5, 7 & 10.

properties should be assigned according to the expected capacity of neighbouring


soil. For example, in order to avoid premature failure in the pile shaft (failure
in the interface before the soil reaches its capacity), the strength parameters of
the interfaces can be selected such that the peak strength of the surrounding soil
elements is adopted (see Fig. 4.17.a).

Figure 4.17: Assumed stress-strain curves for the interfaces a. at the extensions b. at the pile tip

On the other hand, at the tip of the pile the stresses will be mobilized to the
critical state due to high compressive (axial) strain. If peak strength parameters are
assigned to the interfaces, there would be unrealistic high shear stresses under the
tip. Therefore, residual strength parameters are assigned to the interfaces at the tip.
In this manner, a simplified yield curve is assigned to the interface at the tip for the
material in any state (see Fig. 4.17.b).
It can be noted that the use of residual strength parameters at the pile tip can
cause some numerical problems at the beginning of each new step in the PR tech-
nique. If the mobilised shear strength of the soil is higher than the one assigned to
74 4. Press-Replace Technique

Figure 4.18: Comparison of penetration curves for cases with peak (φpeak ) and critical (residual)
interface strength parameter (φcritical )

the interface, shear failure would occur in the interface until equilibrium is satisfied.
This failure will reduce the reaction force and result in the Fy - uz plots as given
in Fig. 4.12 & 4.13. The larger the step size, the more reduction in the reaction
force occurs. The reduction amount increases with depth where higher stresses are
mobilized. However, by further straining, the soil stresses will increase gradually
and catch up the previous stresses under the pile tip.
Alternatively, if peak strength parameters are used, the mobilisation would be
equivalent to the soil until the peak strength is reached. After this point the soil will
soften and the interface will be stronger. This will make the computations even more
difficult. The convergence rates of the calculations with peak strength parameters
at the pile tip interface were quite slow. Furthermore the analyses with residual
strength parameters at the pile tip interface have converged to the same stress level
at the end of each penetration level (Fig. 4.18). Therefore, the assumption for
the interface strength properties shown in Fig. 4.17 is proven to be valid for the
penetration model used. As a conclusion, it is recommended to use peak strength
for the pile shaft and residual strength for the pile tip.
Another issue with the interfaces, especially for dense cases and deep penetrations
(L > 5D), is the significant amount of reductions in the reaction force that can be
recognised in the load displacement curves (Fig. 4.19). These are due to sliding
of interface extensions when they are activated at the beginning of each phase.
The sliding is a result of the difference in the strength properties of the extended
interfaces and the soil. The soil is modelled using hypoplasticity and the friction
angle is density dependent. On the other hand, the interface model is elastic perfectly
plastic with MC yield criterion and the friction angle is constant. Therefore, due to
the density increase with penetration, the soil strength and stiffness values increase
but this increase cannot be followed by the interface extensions.
4.4. Parametric study 75

Figure 4.19: Unloading issue evident from the load displacement curves (D = 0.40 - Id = 0.80)

Two subsequent calculation phases and are considered; the interface shear strength,
can be calculated as:
τint = cint + σn tan φint (4.35)
where, cint and φint are the interface strength parameters and σn is the normal stress
in the interface at the beginning of the phase (stress from previous phase). The shear
stresses mobilised in the previous phase, τ k−1 (where k is the current phase number)
can be then compared with the available shear strength in the interfaces as given in
Fig. 4.20.

 k−1
Figure 4.20: Stress distributions at previous phase (uy = 3.12 m) a. τ k−1 ; b σyy or σyy for
k−1 
horizontal interfaces) and c. σxx or σn for vertical interfaces to be activated in the current phase
uy = 3.16 m

It can be seen in Fig. 4.21 that the shear strength is higher than the shear
stresses under the pile tip. This is due to high level of normal stresses σn , which
mobilise higher shear strength τint as shown in Eqn. 4.35.
76 4. Press-Replace Technique

In general, the shear strength is also, but slightly, higher than the shear stresses
at the beginning of the phase (step 0 of the current phase) at the interface extensions.
However, at some locations, the actual shear strength is lower than the shear stress
at the beginning of the phase.

Figure 4.21: Comparison of available interface strength and the shear stress at the beginning of
step a. for horizontal interfaces at the pile tip, b. for vertical interfaces at the pile shaft

One option is to assign high strength and stiffness values to the interface ex-
tensions (e.g. c = 10 MPa, E = 10 GPa) while keeping the rest as defined at the
beginning of this section. In this way the interaction between the pile and soil will
not be affected and the artificial interface extensions into the soil do not fail or de-
form in the soil. As a result, the force displacement behaviour can be improved (Fig.
4.22).

4.5 Discussion and Conclusions


A numerical analysis technique, the ’Press-Replace’ (PR) technique, to model pile
jacking has been introduced. The method is based on a geometry update of small
deformation phases. The method does not suffer from any mesh distortion effects.
Furthermore, it does not require any special treatment, such as defining contact or
advanced mesh update schemes. For a better description of the pile-soil interaction,
interface elements have to be used especially for the cases with flat tips that have
numerical singularity (corner). Another benefit of the method is the possibility of
using any constitutive relation for the soil. The method can be employed for any type
of jacking problems (e.g cone penetration, open ended piles, piles with any cross-
section). Another important benefit of the proposed technique is the possibility of
using sharp geometries for the penetrating object. Hence, there is no geometrical
limitation unlike most of the advanced numerical techniques, which require smoothed
corners for the penetrating object.
4.5. Discussion and Conclusions 77

Figure 4.22: Improvement of load displacement (FT - uy ) behaviour by using very strong interface
extensions

Some verifications as well as comparison with several analytical and numerical


techniques are also presented. The results show that the PR technique gives results
that are comparable with more advanced large strain methods. The PR technique
allows the use of non-smooth geometry (e.g. sharp corners at the pile or cone tip or
shoulder). However, a concentration of stresses at these locations cannot be avoided.
As a conclusion, the PR method can be used in a standard finite element program
to model large deformation problems, like pile jacking, benefiting from stability as
well as reduced calculation time without losing much accuracy.
The model has a limitation in the sense of not fully capturing the flow mechanism
of soil being pushed under the tip towards the shaft of the pile. Other numerical
techniques like Eulerian, ALE, CEL and MPM are more suitable to model the flow
mechanism. However, these methods are not yet generally available or common for
engineering applications. Furthermore, the limitation of soil flow may not be a major
issue if the shearing effects can be indirectly modelled. Section 4.3 has shown that
the Poisson’s effect during the straining phase introduces such shearing effects near
the pile tip and the shearing near the shaft has been sufficiently modelled by the
interfaces. The last but not the least, is that a proper constitutive model is essential
to account for density-dependent behaviour of the soil. When using such a model,
the limitation that soil flow is not modelled would not be a major issue.
78 4. Press-Replace Technique
Chapter 5

Investigation of Installation Effects

5.1 Introduction
In the previous chapter a practical FE technique to model pile penetration was in-
troduced. The method was proven to model the penetration process reasonably well
(Sections 4.2 & 4.3). As the final goal of this study is to investigate the possibility
of incorporating the installation effects around a wished-in-place pile without sim-
ulating the whole penetration process, possibilities to come up with a generalised
procedure are investigated.
The installation effects are the changes in mechanical (e.g. stress and density)
and physical (e.g. shape, smoothness of the grains) properties of the soil due to
penetration of a pile. In this study the focus is on the stress and density change
since the service behaviour of the pile installed strongly affected by these properties.
During penetration of the pile, the soil surrounding the pile is exposed to high level
of stresses and strains. The level of stresses and strains reduce by increasing distance
from the pile. The stress and density level is not effected outside the influence region.
The size of the influence region is generally around 8 - 10D. The size of the influence
region is also proportional to the initial density of the soil
The well-known bearing capacity solutions (e.g. [115]; [33]; [157]) show that the
bearing capacity of a foundation depends not only on the soil properties but also the
foundation geometry and depth. The pile penetration mechanism can be considered
as a combination of an infinite number of consequent bearing capacity problems.
Since the penetration of a pile can be achieved by reaching the bearing capacity of
the soil, one would therefore, expect that the factors influencing the bearing capacity
(e.g. pile diameter, penetration depth, soil density) should also affect the penetration
resistance of a pile and resulting installation effects in surrounding soil. From this
perspective, in this chapter, several variations on the pile diameter and penetration
depth are investigated (Section 5.2). The effects of each parameter are discussed
in Section 5.3. These results set a basis for Chapter 6 in which a compact way of
representing the installation field (e.g. geometrical functions) is investigated.

79
80 5. Investigation of Installation Effects

5.2 Numerical Modelling - Application of the PR


Technique
In the current practice, CPT based methods are proven to be a reliable way to
incorporate the installation effects in the design of displacement piles [102]. On the
other hand, as [138] points out, there is still no consensus on a preferred CPT based
pile design method (e.g. Dutch, LCPC, ICP, UWA methods) in the engineering
practice. A common ground in these methods is that the pile base resistance is
correlated to the cone resistance qc . However, there are different ways to determine
the qc value as well as the coefficient relating this value to the base resistance, qb . In
some methods (e.g. [36]) the shaft resistance is related to the sleeve friction, fs ; in
others to the cone resistance, qc ([99]; [39]; [85]). The CPT correlations to determine
soil properties (e.g. density, internal friction angle and stiffness) indicate the effect
of soil properties on the CPT cone and shaft resistances.
The strong correlation between CPT and pile resistance would therefore, imply
that the pile resistance and its behaviour during as well as after installation would
be affected by the soil properties. [136] support this proposition by stating that
the normalised base resistance (qb /qc ) is a function of relative density (or density
index, Id ), the confining stress and the coefficient of lateral earth pressure at rest
(K0 ). Furthermore, the design charts introduced by [64] indicate the importance of
the density effect on the end bearing capacity of piles. These plots incorporate the
density effect by increasing the bearing resistance for denser soils for a given initial
vertical stress level. This effect is pronounced at low stress levels (e.g. shallow
depths). Although the cases in this chapter were analysed for relatively shallow
penetration depths (L = 2.5D − 10D), the stress levels were not low due to the
overburden pressure of the clay overlying the sand layer. Nevertheless, the effect
of relative density were evident as the sand behaviour is mainly dictated by the
stress and relative density, in other words the ’state’ of the sand [23]. Therefore it is
important to investigate the effect of soil density on the installation field. The initial
stress state can be defined by the geostatic field conditions. Using the exponential
stress dependent the density [21] corresponding initial void ratio distribution can
be obtained. The initial void ratios, e0 , hence the initial relative densities, Id were
varied to investigate the effects on installation field. Three different densities are
considered in each diameter and penetration length analysed: loose (Id = 0.40),
medium (Id = 0.60) and dense (Id = 0.80), to investigate the effects on the soil
stress and density change. The variations in the relative density are investigated
together with the variations in the pile geometry (diameter and penetration length).
The pile shaft friction (τs ) reduces with further penetration, which is called as
’Friction fatigue’ ([80], [99], and [39]), until the residual capacity is reached. Hence,
the depth of penetration is important, especially at shallow depths where the failure
mechanisms are different than those observed for the deep failure mechanisms. In
this study relatively short piles 2.5D to 10D are examined to assess the effect of
depth of embedment.
In the Dutch code (NEN 9997-2010), the cone tip resistance, qc and the pile
base resistance, qb are assumed to be equal discarding the difference in the size of
5.2. Numerical Modelling - Application of the PR Technique 81

base areas. [138] compared six different CPT based end bearing capacity calculation
method and there it was pointed out that for jacked piles the full mobilisation of
the base pressure was generally around 0.10D. This implies that the effect of pile
diameter vanishes for jacked piles. In order to verify this statement, two different
diameter cases, which are D = 0.30 m and D = 0.40 m are investigated.
As the goal is to define the installation effects as a function of pile diameter,
penetration length and density of the soil and incorporate therefore the installation
effects in the FE analyses, this chapter focuses on the effects of these parameters.
The current practice follows some design rules to model pile bearing capacity by
bearing capacity factors Nq and β related to CPT results. As mentioned in Section
2.2.1, the effect of installations are taken into account indirectly for displacement
piles. In reality, the pile capacity is dictated by the state around the pile installed
and loading conditions as well as the interface properties. Therefore the focus is
given on the change in state around the pile due to installation process, which
also dictates the load displacement behaviour. For this, a limited number of pile
geometry and soil density variations were analysed using the PR method explained
in the previous chapter (Chapter 4). In these variations, the method is employed
to model continuous pile jacking in sand. The effects of the pile geometry and soil
density on the altered ’equalisation’ state due to installation are presented in the
following Section 5.3.
A simplified sketch of the problem considered and the corresponding axisymmet-
ric FE model are given in Fig. 5.1. This is a typical profile in The Netherlands and
places having similar geological sequences (e.g. Pleistocene glacial and interglacial
periods). The top 10 m very soft clay layer overlying sand layer is modelled as an
elastic layer of 1 m thickness having a unit weight of γ’=100kN/m3 . The purpose
of modelling the clay layer as such is to facilitate the generation of geostatic (K0 )
stresses in the sand layer. Therefore the clay layer is considered to generate proper
stress level on top of the sand layer. The interaction between the pile and the clay
layer is also ignored due to very low cohesion of the clay layer.
The bearing capacity of a pile is also a function of the rigidity index, which is the
ratio of stiffness to strength. The stiffness is a function of depth (which increases
with the mean effective stress level to the power generally around 0.5 for sand).
Therefore the rigidity index will reduce with depth. [96] discussed this reduction of
the stiffness and its effect on the pile end bearing capacity. The constitutive model
employed in our numerical simulations incorporates the stiffness change with stress.
As mentioned in Section 4.4, hypoplasticity is used in the analyses as the con-
stitutive model, due to state dependent stiffness and strength. As a result of proper
modelling of the stiffness and strength, different levels of stiffness and strength was
mobilised at different levels of penetration for different pile diameters and different
sand densities. Moreover, the effect of change in stiffness is considered intrinsically
in the analyses. For the sake of simplicity, only one soil type, the Baskarp sand,
is considered in the analyses. Table 5.1 presents once again the hypoplastic model
parameters given in Table 4.3 in Section 4.4 for the sake of completeness.
In order to investigate the effect of the soil density, the pile diameter and the
penetration lengths, a combination of the variations (24 different cases in total) sum-
82 5. Investigation of Installation Effects

Figure 5.1: a. Sketch of the problem modelled b. General view of the FE model

Table 5.1: Hypoplastic soil model parameters for Baskarp sand (after [8])

φc hs n ed0 ec0 ei0 α β



MPa - - - - - -
30 4000 0.42 0.548 0.929 1.080 0.12 0.96

marised in Table 5.1 were analysed using the PR technique. A maximum penetration
level of 10D is reached for each variation.
Table 5.2: Summary of variations analysed for the investigation of pile diameter (D), penetration
length (L) and density (Id ) effects on the installation field

D L Id
(m) in terms of D -
0.30
2.5 5 7.5 10 0.4 0.6 0.8
0.40

In the numerical analyses two different FE meshes were employed. These were
basically for two different diameter cases, D = 0.30 m and D = 0.40 m given in
Figure 5.2.a & .b, respectively. For the analyses using the PR technique, the mesh is
prepared for a 10D maximum length of penetration. Both axisymmetric FE meshes
have a size of 50D by 50D. Based on the parametric study results (Section 4.4.2), a
slice thickness of ts = 0.1D was used and each slice consists of one row of elements.
In other words, the slice thickness (ts = 0.1D) is equal to the element size in the
region, where the pile will be placed.
As mentioned in Section 2.3.1, in geotechnical engineering the most relevant state
for the design of the foundation pile is the final ’equalisation’ state ([98]; [39]), where
5.3. Results of the FE Simulations 83

Figure 5.2: FE meshes prepared for PR analyses for cases with a. D = 0.30 m, b. D = 0.40 m

the pile head is free, i.e. no force is applied, and the pile-soil system goes into a
new equilibrium. Therefore the focus is on the installation effects at the end of
’equalisation’ state. In order to obtain the equalisation state, an additional phase,
in which the pile is allowed to get into equilibrium with the installation stresses
without any load on pile head, is calculated after the desired level of penetration is
achieved (e.g. 2.5D, 5D, 7.5D or 10D).

5.3 Results of the FE Simulations


In order to have an objective assessment of each case normalisation is applied on the
results obtained. From this perspective, the displacements uy are simply normalised
with the pile diameter, D as:

ū = uy /D (5.1)

The pile shaft is fully mobilised much before the pile base capacity is reached
[68]. Further loading on the pile head mobilises the base pressure after the shaft
is fully mobilised [136]. Therefore, it is common to distinguish and investigate the
base (qb ) and shaft capacities (qs ) of a pile separately.
As described in Section 4.2, the prescribed displacement boundary condition is
maintained at pile head throughout the analyses using the PR technique. Recalling
the basic force displacement relation in the FE analyses:

Ku=f (5.2)
84 5. Investigation of Installation Effects

one could calculate the reaction force at the pile. The calculation of the base and
shaft resistances is not straightforward. For calculation of the average base resist-
ance, first vertical stresses below the pile tip were found by interpolation from the
nearest Gaussian points.

Figure 5.3: a. Illustration of the method for the calculation of the equivalent base force at the
pile tip; b. Top view of one radian slice and the area of ith slice

Then these normal stresses were integrated to find the equivalent base reaction
Fb . As illustrated in Figure 5.3, the integration was done based on a trapezoidal
rule considering area of each segment (see Figure 5.3.a) that is proportional to the
radius, ri such that due to axisymmetry:
ˆ ˆ 

 i i+1

Fb = σzz rdrdθ ≈ Fbeq = 2π 0.5 ri · σzz + ri+1 · σzz · Δri (5.3)
θ r

then the base resistance can be calculated by dividing by the base area as:
4 Fb
qb = Fs /Ab = (5.4)
πD2
The shaft resistance can be simply calculated by:
Ft − F b Ft − F b
qs = = (5.5)
As πD · L
Alternatively, the shaft resistance can be calculated by interpolating the shear
stresses at the pile shaft from nearest Gaussian points. Then these stresses can
simply be integrated by a trapezoidal rule to get the total shear force at the shaft.
Dividing the total shear force by the pile shaft area should give more or less the
same shaft resistance (Eqn. 5.7).
'  i  ' i 
2π 0.5 · σrz i+1
+ σrz · Δli π i+1
σrz + σrz · Δli
qs = = (5.6)
As πD · L
5.3. Results of the FE Simulations 85

Inspired by the normalisation method applied for the CPT cone resistance,qc

normalisation [127], qb can be normalised by the effective vertical stress σvo at current
i
penetration depth (L ) as:
qb qb qb
q̄b = i |L = Li
=  =  i (5.7)
σvo γs · Li + qsurch γs · uy + qsurch

where, qsurch is the surcharge load at L = 0.


The base resistances, qb were also normalised with the ultimate base resistances
qb,ult at 0.1D displacements during the load tests. In Figure 5.6, the normalised base
resistances, q̄b were compared with the normalised pile load test curves suggested by
the Dutch code (NEN 9997-2010).
qb qb
q̄b = = (5.8)
qb,ult qb |uy = 0.10D

Figure 5.4 shows the penetration resistance curves for different cases investigated.
The base resistance obtained for both diameter variations were quite similar. Hence
no effect of diameter size on the pile base resistance has been observed.

Figure 5.4: Total pile shaft, Fs , and base, Fb reaction curves obtained for piles having diameters
and penetrating in sands with different soil densities, Id

Figure 5.5 presents the pile load test curves performed after equalisation for
different cases. Obviously, the capacity increases with increasing diameter and pen-
etration lengths. Normalising the displacements with diameter, D results in similar
86 5. Investigation of Installation Effects

Figure 5.5: Load test curves for different pile diameters and total penetration lengths

F − uy response for the cases with different diameters. In the same way, the initial
stiffness increases with increasing pile diameter and length. The normalised plots
given in Figure 5.6 show the relative stiffness of each case in comparison with the
design curves suggested by NEN 9997-2010 [3]. For the sake of consistency, the ul-
timate base capacities were determined at 0.10D displacement, which is a common
assumption in practice. The normalised base resistance curves show a reasonable
agreement with the curves suggested in NEN 9997-2010 for driven piles. In general
they are in between the curves suggested for driven and auger piles, closer to the
former. At the beginning of loading, the shaft friction is negative due to the residual
stresses at the pile base that cause an upward movement of the pile at the equal-
isation phase. For compatibility with the NEN 9997-2010 curves, the shaft friction
curves were shifted on ss axis so that the unit shaft friction is zero at zero displace-
ment. The shaft mobilisation curves show a reasonable trend that is between the
driven and bored piles. At 15 mm displacement, the shaft friction of most of the
cases are completely mobilised.
The soil state (stress and density) around the pile is altered as a result of the
displacement of the soil dictated by the pile jacking operations. Concurrently, the
stiffness of the soil changes. Therefore, these three major installation effects should
be assessed in order to quantify the new state around the pile installed. Hypoplas-
ticity uses stresses and void ratio as state variables and the stiffness is formulated
based on these variables. Therefore assessing the stress and void ratio change would
be sufficient to describe the installation effects (i.c. stress, density and stiffness
change). From this point of view, in order to assess the installation effects for each
f f f
case analysed, the Cartesian stresses at the end of equalisation, σrr , σzz , σθθ and
f f
σrz and the void ratio, e distributions around the displacement pile are normalised
5.3. Results of the FE Simulations 87

Figure 5.6: Comparison of results of the analysis using the PR technique with the design load
test a. base resistance, b. shaft resistance curves suggested by the Dutch code (NEN 9997-2010)

with their corresponding geostatic (K0 ) states as:


f
σrr
Rrr = (5.9)
p0
f
σzz
Rzz = (5.10)
p0
f
σθθ
Rθθ = (5.11)
p0
f
σrz
Rrz = (5.12)
p0
ef
Re = (5.13)
eK0
For the K0 state,
 
 

σrr,0 σrr,0 3 · K0 · σzz,0 3 · K0
Rrr,0 =  =  =  = (5.14)
p0 1 
σrr,0 
+ σzz,0 
+ σθθ,0 σzz,0 · (1 + 2K0 ) 1 + 2 · K0
3

Similarly, for the other normalised Cartesian stresses the following can be ob-
tained:
3
Rzz,0 = (5.15)
1 + 2 · K0
3 · K0
Rθθ,0 = (5.16)
1 + 2 · K0
88 5. Investigation of Installation Effects


Since σrz,0 = 0,
Rrz,0 = 0 (5.17)
And since e = eK0 ,
Re,0 = 1 (5.18)
The normalised values can therefore be used as a multiplier to alter the geostatic
state to the installation state. Based on the new stress and void ratio, the hypoplastic
model alters the stiffness to a consistent level as a result of its efficient stiffness
formulation.
All length scales in the model were also normalised with the size of pile diameter,
D. This normalisation enables a more objective evaluation of the effects of pile
diameter and length on the Rrr , Rzz , Rθθ , Rrz and Re . In the following paragraphs
the effect of pile geometry on each of these variables is discussed.

Figure 5.7: Sketch of the problem analysed and the description of 3D surface plots

Figure 5.7 gives the description of surface plots, which will be presented and
discussed in the following paragraphs. Figures 5.8 - 5.12 present a comparison of all
the geometrical variations regarding the normalised Cartesian stresses and the void
ratio.
First of all, it can be seen in Figures 5.9.c - 5.13.c that the Rrr , Rzz , Rθθ , Rrz and
Re show a linear correlation to the diameter. Figures 5.9.a - 5.13.a show the effect
of density and pile length on the normalised stresses and void ratio. The increase
of Rrr , Rzz , Rθθ and Rrz (Figures 5.9.b - 5.13.b) confirms the idealised parabolic
stiffness increase with depth (Figure 5.8.b). It can be seen that the increase of the
peak and the decrease of spatial decay from the peak of Rrr , Rzz , Rθθ and Rrz
shows a similar parabolic trend.
5.3. Results of the FE Simulations 89

Figure 5.8: Idealised stiffness profiles of typical a. overconsolidated clay, b. sand, c. soft clay
(after [111])

In Figure 5.9, the distribution of normalised horizontal stresses, Rrr is given.


The results show that the equalisation state involves also a considerable amount of
stress change (e.g. a maximum value of Rrr ≈ 30 for case D = 0.30 m; L = 10D;
Id = 0.80). The Rrr value is proportional to soil density and penetration length. As
can be expected, the stress increase is most pronounced below the pile base. From
this maximum level under the pile base, it reduces exponentially to the geostatic
level approximately 8D both in radial and in vertical directions. These values are
comparable with the values suggested by other researchers (e.g. [77]).
Similarly, Figure 5.10 shows that the normalised vertical stresses, Rzz are in-
creased considerably to a maximum level of approximately 35 times of the geostatic
mean stress level, p0 (D = 0.30 m; L = 10D; Id = 0.80). Similar to the normalised
horizontal stresses, Rrr , the stress increase is most pronounced below the pile base
and reduces exponentially to the geostatic level approximately 5D both in radial
and in vertical directions.
The distribution of normalised tangential or hoop stresses, Rθθ is given in Figure
5.11. Similar to the horizontal and vertical stresses, the increase in tangential stresses
is most pronounced near the pile base and reduces exponentially by moving away
from the pile base.
Figures 5.12.a-c present the normalised shear stress, Rrz distributions for dif-
ferent densities, penetration lengths and pile diameters, respectively. Increase in
the peak values proportional to the relative density of the soil is evident. A slight
parabolic increase of the peak Rrz values with increasing penetration length can be
seen in Figure 5.12.b.
Similar to the normalised stress plots, Figures 5.13.a-c give a comparison of the
normalised void ratio, Re distributions for different densities, penetration lengths
and diameters. Re values larger than unity indicate dilation and values lower than
one show compaction relative to the K0 state. As previously observed for the nor-
malised stress plots, a linear correlation for the different diameters is once again
evident (Figure 5.13.c). Hence, by normalising the radial and vertical distances, al-
most the same surfaces have been obtained for diameters, D = 0.30 m and D = 0.40
m. The plots indicate a pronounced dilation near the pile shaft and base, for all
cases, which is in agreement with the experimental observations of [56]. The dila-
tion zone is limited to approximately 1.5D radial distance and 2.5D depth below the
90 5. Investigation of Installation Effects

Figure 5.9: Comparison of a. the effect of density; b. penetration length, and c. pile diameter
on the distribution of normalised horizontal stresses, Rrr . In the leftmost column the 3D surface
plots from a reversed angle of view, in the middle column the vertical cross-section at 0.01D from
the shaft, and in the rightmost column the horizontal cross-sections at 0.5D below the tip are
presented.

pile tip. The dilation zone is followed by a compaction zone similar to the results
obtained by [77]. A negligible amount of compaction can be seen near the pile shaft
at a radial distance 1.5D to 8D and at a region 2.5D to 7.5D depth below the pile
tip.
The base and shaft resistances were compared with the results of [90] in Figure
5.13. The equivalent critical state mean stress, pcs is the mean stress at the intercept
of the critical state line (See Figure 5.14).
The equivalent critical state mean stress, pcs is the effective mean stress corres-
ponding to the in-situ void ratio, e0 on the ec line. In hypoplasticity, the ec line is
defined as:
 n
3p

ec = ec0 · e hs
(5.19)

for any effective stress p . Substituting e0 for ec and pcs for p in Equation 5.19, one
can obtain the pcs as:
  1/n
hs e0
pcs = − ln (5.20)
3 ec0
5.3. Results of the FE Simulations 91

Figure 5.10: Comparison of a. the effect of density ; b. penetration length, and c. pile diameter
on the distribution of normalised vertical stresses, Rzz . In the leftmost column the 3D surface
plots from a reversed angle of view, in the middle column the vertical cross-section at 0.01D from
the shaft, and in the rightmost column the horizontal cross-sections at 0.5D below the tip are
presented.

And therefore the stress state parameter Rs can be calculated as:

ps
Rs = (5.21)
pcs

The bearing capacity factors for the base can be calculated using the average base
pressure, q̄b (Eqn. 5.7) and the normalised unit shaft friction, β s can be calculated
using the average shaft friction, qs (Eqn. 5.6) as follows:
q̄b
Nq =  (5.22)
σv0
qs
βs =  (5.23)
σv0
It can be seen that the results of the analyses using the PR technique show a similar
trend for all variations. As observed for spatial distribution of the installation effects
quantified by Rrr , Rzz , Rθθ , Rrz and Re , the variation on the diameter had no
effect. As expected the bearing capacity factors increase with density. The average
92 5. Investigation of Installation Effects

Figure 5.11: Comparison of a. the effect of density; b. penetration length, and c. pile diameter
on the distribution of normalised tangential stresses, Rθθ . In the leftmost column the 3D surface
plots from a reversed angle of view, in the middle column the vertical cross-section at 0.01D from
the shaft, and in the rightmost column the horizontal cross-sections at 0.5D below the tip are
presented.

mean stress and void ratio at the pile base were calculated using Equation 5.3 and

substituting effective mean stress and void ratio instead of σzz . Figure 5.13.c shows
the average volumetric behaviour during penetration for different cases. From this
figure, it can be seen that a considerable amount of dilation is followed by a relatively
small amount of compaction for all variations. Figure 5.13.b shows a similar trend
for normalised unit shaft friction, β s , i.e. decrease with further penetration. In the
Nq curves given in Figure 5.13.a, the compaction parts follow the trend observed
by [90]. For a deep penetration, it is foreseen that the compaction part will be
pronounced and a similar trend would be obtained.

5.4 Discussion of Results and Conclusions


In this chapter the PR technique has been employed to investigate the pile behaviour
during and the after jacking phase and the resulting installation effects for various
soil densities and pile geometries. This is an intermediate step to the main goal of
this dissertation, describing the installation effects of a displacement pile by simple
geometric functions. This chapter investigates the parameters that have signific-
5.4. Discussion of Results and Conclusions 93

Figure 5.12: Comparison of a. the effect of density; b. penetration length, and c. pile diameter on
the distribution of normalised shear stresses, Rrz . In the leftmost column the 3D surface plots from
a reversed angle of view, in the middle column the vertical cross-section at 0.01D from the shaft,
and in the rightmost column the horizontal cross-sections at 0.5D below the tip are presented.

ant effects on the installation field and quantifies their contributions by several FE
simulations using the PR technique.
The pile penetration can be thought as a continuous bearing capacity problem.
Therefore, the parameters affecting the bearing capacity of a foundation (e.g. soil
properties, foundation size and depth) are considered to affect the pile behaviour
during the jacking process and the installation effects resulting from the penetration
process. In this respect, a variety of diameter, penetration length and initial soil
density cases were analysed using the PR technique. The results enabled a quant-
itative assessment of the installation effects, which was assumed to be the change
in the normalised Cartesian stress components and the normalised void ratio. The
effect of the initial density of the soil, the penetration length and the pile diameter
on the Cartesian stress and the void ratio distributions around a jacked pile has been
investigated. Possible correlations with respect to diameter and length are presented
qualitatively. The behaviour of a jacked pile after installation has been compared
by well-established design charts. For almost all cases, a trend close to the design
curve suggested for driven piles were obtained for the base resistance. However,
the mobilisation of the shaft friction was underestimated. These responses can be
94 5. Investigation of Installation Effects

Figure 5.13: Comparison of a. the effect of density; b. penetration length, and c. pile diameter
on the distribution of normalised void ratio, Re . In the leftmost column the 3D surface plots from
a reversed angle of view, in the middle column the vertical cross-section at 0.01D from the shaft,
and in the rightmost column the horizontal cross-sections at 0.5D below the tip are presented.

Figure 5.14: Determination of the equivalent critical state mean stress, pcs in hypoplasticity

improved by using small strain stiffness (the intergranular strain for hypoplasticity)
for the loading phase by assuming the equalisation state is the new at rest state
around the jacked pile.
5.4. Discussion of Results and Conclusions 95

Figure 5.15: Comparison of normalised a. base and b. shaft resistances during pile penetration
together with trend lines obtained for Leighton Buzzard sand*[90]; and c. average void ratio
evolution under pile base with ec reference curve (dashed line) of the hypoplastic model using
Baskarp sand parameters.

In the light of the findings presented here a compact way of representing the
installation field (e.g. geometric functions) is investigated in Chapter 6.
96 5. Investigation of Installation Effects
Chapter 6

Generalisation of Installation
Effects

6.1 Introduction
In the previous chapter the PR technique was employed to investigate installation
effects for various pile jacking cases (e.g. pile geometries and soil densities). The
main purpose of this chapter is to investigate the possibility of incorporating these
installation effects around a “wished-in-place” pile to account for these installation
effects without simulating the whole penetration process. This will reduce compu-
tational demand, especially when there is more than one pile (e.g. pile groups and
piled raft).
The results of the reference FE calculation results presented in the previous
chapter are in discrete form (i.e. values at stress points). To be able to transfer
these results to different discretisations, it is necessary to represent these results as
a continuous field so that the values can be determined at any location (e.g. at the
stress points of the new discretisation). The continuous fields can be represented
by geometric functions. One way of determining these fields is using interpolation
functions. Section 6.2 describes the interpolation of the calculated installation ef-
fects (stress and density) to FE models having different discretisations to check the
applicability and possible mesh dependency of incorporating the installation effects.
The interpolation functions provide good approximations to the surface formed by
the guiding stress points. On the other hand, these functions require information of
all stress points (e.g. location and state).
A more practical way to represent the installation effects is using a model func-
tion. By a regression analysis the coefficients of the model function can be determ-
ined. Once having the fit, less data, i.e. the fitting coefficients, are required to
represent a continuous field of installation effects. From this perspective, Section
6.3 presents how the installation effects can be extended to simpler functional forms
than the interpolation functions. Finally Section 6.4 combines the functional forms
presented in Section 6.3 into a more generalised form. In this generalised form,

97
98 6. Generalisation of Installation Effects

the coefficients of the state variables such as normalised Cartesian stresses and void
ratios, representing the installation effects are expressed in terms of the pile dia-
meter, penetration length and soil density. Chapter 7 presents validation of these
generalised functions.

6.2 Approximation of the Installation Effects by


Interpolation
As mentioned in the introduction, the main objective of this study is to establish a
continuous description of the installation effects. As a result the calculated installa-
tion effects can be imposed to any FE discretisation and a displacement pile can be
modelled more realistically without simulating the installation process.
There are several ways to represent the installation effects by a continuous sur-
face. The easiest way is to use an interpolation technique (e.g. surface interpolation
with nearest neighbour algorithm). At the same time, the interpolation methods
have some disadvantages compared to the regression method, such as the require-
ment for more reference data and having a less smooth form than surface fitting.
Nevertheless, the possibilities and limitations of transferring installation effects to a
different FE mesh can be investigated using the interpolation methods.
The installation effects are defined by the stress and void ratio variables. As
discussed in Chapter 3 and Section 5.3 , the stiffness follows the stress and density
state due to the efficient formulation of the hypoplastic constitutive model. Other
changes, such as grain crushing, are more difficult to assess and model. Section
3.3 discussed a possible modification of the hypoplastic model to account for grain
crushing effects. As mentioned in Section 3.4 the modified model had convergence
issues when employed in boundary value problems. To find a stable and efficient
hypoplastic grain crushing model requires a thorough investigation and is beyond
the scope of this study.
Furthermore, due to the reasons discussed in Section 3.3 and to limit the number
of state variables assumed to represent the installation effects, the intergranular
strain extension is not employed in the FE analyses. Hence, only stress and density
changes are considered as installation effects. In the following sub-sections, first
the interpolation techniques are explained (Section 6.2.1). Then, the details of each
FE mesh are presented. Finally, the sensitivity of the mesh discretisation on the
transferred installation effects are discussed in details (Section 6.2.2).

6.2.1 Method - Triangular Mesh Interpolation


The results of the PR analyses (e.g. state variables) are stored at the Gaussian stress
points (with coordinate pairs of rk , zk , k = 1, 2, ··· , nGP ), which are scattered in a
non-rectangular pattern. These points form surface data and can be represented as
a continuous surface.
A surface can be represented in implicit (Eqn.6.1), explicit (Eqn.6.2) or paramet-
ric (Eqn.6.3) form. Since the explicit or functional form is useful for graph surfaces
6.2. Approximation of the Installation Effects by Interpolation 99

[86], the focus in this section is on an explicit representation of the (installation)


field.

f (x, y, z) = 0 (6.1)

z = f (x, y) (6.2)
(x, y, z) = (x(u, v), y(u, v), z(u, v)) (6.3)
Surfaces can be represented by several methods (e.g. Polygonal meshes, surface
patches) and interpolations can be made using these representations. An interpol-
ation can be performed within the convex hull of the reference data points (in this
case the stress points of the reference analysis). The convex hull can be visualised
as an elastic strip surrounding the cloud of data points (see Fig.6.1).

Figure 6.1: Convex Hull of a PR discretisation (only sand region considered).

As mentioned in the introduction, the only state variables considered for the
 
assessment of installation effects were the Cartesian stress components (σrr , σzz ,

σθθ , and σrz ) as well as the void ratio (e). The change in these state variables was
used to quantify the installation effects.
In this part the focus is only on the interpolation of these state variables such
that they can be represented by a continuous field. First the theory of multivariate
(bivariate) interpolation is explained very briefly. The method was used in the
sensitivity study for transferring the results of the PR analyses and imposing to
different meshes.
Interpolation is the estimation of the values of the unknown points, which lie
within the convex hull of the points with known values. The simplest method is the
’nearest neighbour interpolation’. In this method the unknown value is set equal
100 6. Generalisation of Installation Effects

to the value of the nearest point. Intrinsically, this algorithm does not provide a
smooth field. Nevertheless, it can be used to evaluate (extrapolate) the unknown
values of the points, which lie outside the convex hull.
In the interpolation methods used in this part, the scattered points with known
values are first decomposed into triangular segments based on Delaunay tessellation
(DT ). The DT is a special dual of Voronoi diagram, which describes the regions of
interest around the points with known values.

Figure 6.2: Voronoi diagram and Delaunay tessellation (after [13])

Aurenhammer [13] provide the following mathematical description for the Voro-
noi diagram and the Delaunay tessellation:
If S denotes a set of n ≥ 3 point sites (e.g. known points of the interpolation) of
p, q, s,··· in a plane, DT (S) can be obtained by connecting any two points p, q of S
with a line segment for which a circle exists passing through these points and does
not contain any other site of S inside or on the boundary. DT (S) is also the graph-
theoretical dual Voronoi diagram V (S). Fig.6.2 shows an example of the duality of
a Voronoi diagram and the Delaunay tessellation.
The Voronoi diagram of S is defined by V (S):
(
V (S) = V R(p, S) ∩ V R(q, S) (6.4)
p,q∈S,p=q

where V R(p, S) is the Voronoi region of p with respect to S and defined as:
)
V R(p, S) = D(p, q) (6.5)
q∈S,p=q

and D(p, q) is the halfplane:

D(p, q) = {x|d(p, x) < d(q, x)} (6.6)

Which is separated by the bisector of p and q:

B(p, q) = {x|d(p, x) = d(q, x)} (6.7)


6.2. Approximation of the Installation Effects by Interpolation 101

Figure 6.3: Nearest Neighbour algorithm for a. 1D uniform, b. 2D uniform data, defined by
stepwise discrete functions.

The regions defined by the Voronoi diagrams and the Delaunay tessellations set
basis for the interpolations. For example, the nearest neighbour algorithm uses the
regions defined by the V (S) to determine the nearest known value. In 3D space,
these regions form prismatic bars forming a discontinuous (discrete) surface. A
simple case for the nearest neighbour method is presented in Fig.6.3.a.

Figure 6.4: Nearest Neighbour algorithm for a scattered surface data represented by stepwise
discrete functions based on the regions defined by Voronoi diagram; a. 2D representation, b. 3D
representation

Any point to be interpolated will take the value of the known nearest point, the
regions of which are shown in blue lines in Fig.6.2. Similarly, for 2D uniform data,
any interpolated point will take the value of the nearest point (Fig.6.3.b).
Although the nearest neighbour algorithm provides a practical interpolation
method, the resulting surface is not continuous. In order to have a continuous dis-
tribution (at least C 0 continuous) of the interpolating surface, a natural neighbour
102 6. Generalisation of Installation Effects

algorithm can be used. The algorithm again makes use of the Voronoi tessellation
to represent sites (regions) of points with known values and the contributions from
neighbouring sites by weighting functions. The mathematical description of the
weighting functions and the calculation method is given in the following paragraphs
based on [144]’s description.

Figure 6.5: a. Voronoi diagram based on original data points 1,2,3,4 & 5; b. Second order Voronoi
cells around the sample point X where the ratio of the enclosed areas Aabf, Aaehf, Aehd, Adcgh &
Abcgf to Aabcde are used to determine the percent contribution from neighbouring cells of original
data as weights [144].

Let X(x) ∈ R2 be the sampling point and V (S) is the Voronoi diagram of the
known (reference) data points 1,2,3,4 and 5 (as shown in Fig.6.5). If the sampling
point is included in the Delaunay tessellation, the new Voronoi diagram V (S ∗ ) can
be obtained as shown in Fig.6.5.b. The resulting Voronoi cell of X and the original
Voronoi cells of the reference data can be used together to determine the contribution
from each data point such that the interpolated value of sampling point:

n
uh (x) = wl (x) · ul (6.8)
l=1

where, ul (l = 1,2,··· ,n) are the reference point values at the n (natural) neighbours,
wi (x) are the weights associated with each reference data point and can be determ-
ined as:
Area(V S(X) ∩ V S(l))
wl (x) = (6.9)
Area(V S(X))

By definition the following properties are self evident:



n
wl (x) = 1 (6.10)
l=1

0 ≤ wl (x) ≤ 1 (6.11)
6.2. Approximation of the Installation Effects by Interpolation 103

The natural neighbour coordinates also satisfy the Local Coordinate Property such
that the geometrical coordinates are interpolated exactly:

n
x= wl (x) · xl (6.12)
l=1

The ’natural neighbour interpolation’ technique is bounded by the convex hull of


reference data points, which is generally the case for all interpolation techniques. If
the Gaussian stress points of the new mesh to which the results would be interpol-
ated are outside the convex hull of the original mesh points, the ’natural neighbour
interpolation’ cannot determine the values at these points.
Data points of the reference analysis and the boundaries of the FE mesh are
given in Fig.6.6. Obviously, the area of the convex hull of the data points is smaller
than the FE mesh. A discretisation finer than the reference mesh might contain
points outside the convex hull of the reference data points. The natural neighbour
interpolation technique cannot estimate the value of any sample point, which lies
outside the convex hull of reference data points.

Figure 6.6: The extended line of points and original mesh stress points (D = 30, L = 10D).

Therefore the results of the original mesh were extended at the outer boundaries
of the mesh to ensure any mesh point regardless of the discretisation coarseness
should lie within the convex hull of the original analysis results. The extension, te
was one tenth of the pile diameter (0.1D) size. The extension of the boundary was
performed only by just one ring of nodes. The spacing of the nodes was adjusted such
that the spacing of the neighbouring original stress points is compatible. For example
regions close to the pile (Fig.6.6.b &c), the extrapolation points were generated at
points ri and zi based on the following criterion:
l1 : − te ≤ r ≤ D/2; te ≤ z ≤ −L|zi = te − (i − 1) · (L + te )/800; (6.13)
D
ri = zi · cot((L + te )/( + te ))
2
Similarly, in six more regions, the following criteria were defined for the extrapolation
points in regions I2 to I7 :
l2 : r = −te ; −2L ≤ z < −L|zi = −L − i · L/600 (6.14)
104 6. Generalisation of Installation Effects

l3 : r = −te ; −50D − te ≤ z < −2L|zi = −2L − i · L/100 (6.15)


l4 : z = −50D − te ; −te ≤ r < 50D + te |ri = −te + i · L/70 (6.16)
l5 : r = −50D + te ; −50D − te ≤ z < te |zi = −50D − te + i · L/100 (6.17)
l6 : z = te ; 8D ≤ r < 50D + te |ri = 50D + te − i · L/70 (6.18)
l7 : z = te ; D/2 ≤ r < 8D + te |ri = 8D − i · L/100 (6.19)
The values at the extended points were determined using the nearest neighbour
algorithm. The extended line of points, the convex hull and original mesh stress
points are shown in Fig.6.6. As a result of extending the reference discretisation, it
was ensured that any point from the new discretisation would be bounded by the
convex hull of the reference mesh points.

Figure 6.7: FE discretisation of a. PR, b) PR coarse (PR-C), and two extreme discretisation
cases: c. finest mesh (F3) and d. coarsest mesh (C3) for case D = 0.30 m L = 10D

The stress points inside the pile region were kept at K0 level and not altered by
the interpolation. After imposing the interpolated stresses and the void ratio fields
6.2. Approximation of the Installation Effects by Interpolation 105

on the rest of the stress points by interpolation, an equalisation step was performed.
In this NiL step, the pile takes over the surrounding stresses with marginal deform-
ation due to its high stiffness. The numerical pile load test was performed after this
equalisation step.
In order to check the effect of mesh discretisation on the interpolation of the in-
stallation effects different mesh discretisations were employed. Besides the reference
case (original PR discretisation, Fig.6.7.a) a coarser mesh is generated based on the
PR discretisation (Fig.6.7.b). These cases have interfaces in the pile region. The
reason having interfaces in the pile is due to the modelling technique (the PR tech-
nique) employed during penetration simulations. In a conventional FE modelling of
a pile using volume elements, the interfaces are placed at the outer boundaries of
the pile. For the sake of convenience, the rest of the mesh variations do not have
any interface within the pile region.
In the original mesh of the PR simulation (D = 0.30 m L = 10D) there are 6
elements beneath the pile tip and 100 elements next to the pile shaft. After the PR
simulation a numerical pile load test was performed and the result is presented in
Fig.6.8 as reference curve. The cases having 4 to 8 elements beneath the pile tip
and ≈70 to 100 elements next to the pile shaft were called as fine and indicated by
the capital letter ’F’ (see Fig.6.7.c). Similarly, the cases having 3 or less elements
beneath the pile tip and ≈50 elements next to the pile shaft were called as coarse
and indicated by the capital letter ’C’ (see Fig.6.7.d). In all discretisation variations
6-noded triangular elements were used for two cases of fine and coarse discretisations
6-noded triangular element formulations were used.

6.2.2 Results and Discussions


In order to investigate the level of installation effects inherited from the original
analysis, numerical pile load tests were performed. Fig.6.8 presents normalised load
displacement curves of all variations. Two sets of curves that represent the load-
displacement behaviour of a pile at K0 stress state (F̄Tmax < 1) and at equalisation
state after penetration (F̄Tmax ≥ 1) are presented. It can be seen that most of
the discretisations are lying in a narrow band slightly overestimating the original
PR result. This behaviour is also valid for the case in which the original mesh is
preserved. This is a consequence of the modification of tensile stresses (setting to a
slight compression level e.g.−5 kPa) due to the restriction of the constitutive model,
which does not allow any tensile stress at the beginning of the calculation step.
The variations without interface in the pile region also overestimated the original
PR curve as well as the interpolation cases (PR-int and PR-C) with interfaces inside
the pile region due to discretisation of the PR method. This is due to the fact that
the interfaces reduce the stiffness between two neighbouring nodes from infinity (no
relative displacement in between) to a finite value, which allows a relative displace-
ment in between. In the PR model 100 slices were generated with interfaces. These
interfaces, therefore, introduce an artificial softening, which was clearly observed in
the load displacement behaviour.
106 6. Generalisation of Installation Effects

Figure 6.8: Comparison of normalised load displacement behaviour of different discretisations to


which the results of the equalisation level of the PR analysis were imposed by interpolation. Lower
curves are the load test results for K0 state, i.e. installation effects were not imposed.

Figure 6.8 summarizes the numerical pile load test curves obtained for the ref-
erence case and all mesh variations. The load test curves of the K0 states are also
presented for the assessment of the effect of mesh density and the accuracy of in-
terpolations. The ’PR’ case is the reference load test curve that is obtained using
the special discretisation (i.e. interfaces in the pile region due to slices prepared
for the analysis) after PR simulations. The second reference, ’PR-int’, had a mesh
discretisation close to the mesh density of ’PR’ case but without interfaces in the
pile region. The stress and void ratio fields were interpolated from the ’PR’ results.
The letters ’C’ and ’F’ represent coarse and fine mesh variations, respectively. The
number following the letters indicate the level of mesh coarseness or fineness, e.g. 1
is the lowest and 4 is the highest level.
In general the load displacement curves are pretty much the same except for the
three most coarse discretisations (C2-6n, C3, and especially for C4). There is a clear
deviation from the reference curve and from the general trend of all other variations.
This is, mainly, due to discretisation effects (e.g. mesh locking). As for every FE
model, by a sensitivity analysis on the discretisation, a mesh dependent behaviour
can be avoided.
Second cause of the deviation from the load displacement behaviour of the coarse
mesh cases is the effect of discretisation coarseness on the distribution of the inter-
polated state. In other words, the number of the stress points were not enough to
represent the values and actual (PR) distribution of the state variables might be lost.
6.2. Approximation of the Installation Effects by Interpolation 107

Figure 6.9: Comparison of a. imposed, and b. equalised interpolation distribution of σzz .

The discretisation coarseness is therefore a limitation to the transfer of the install-


ation effects. The coarsest mesh, in which the imposed installation state would not
be lost (or influenced), should have at least 2 elements below the pile tip (preferably
using higher order elements, e.g. 15-noded triangular elements).
The number of calculation steps can be increased at the initial loading levels
(ūy < 0.2) for example by decreasing tolerance and as a result smoother curves can
be obtained. This would enable the comparison of initial stiffnesses of different cases.
Nevertheless, in general, if there is no locking effect, quite similar load displacement
curves were obtained for both coarse and fine mesh variations.
Another way of checking the effect of mesh on the distribution of interpolated
and imposed fields is to visualise the original (PR) surface data (e.g. σzz ) together
108 6. Generalisation of Installation Effects

with the interpolated data points. Fig.6.9 shows the difference between imposed and
resulting (equalised) vertical stress distributions compared to the original PR field,
indicated by black dots. In Figure 6.9, a clear reduction of the imposed stresses can
be seen for the fine mesh cases (F1-F3) indicated by red markers. On the contrary,
the very coarse mesh (C4) shows an increase under the pile tip where the stresses
are concentrated.

Figure 6.10: Deviations from imposed (interpolated) state.

Similarly, in Fig.6.10, the comparison of the imposed and equalised vertical stress
(σzz ) distribution is plotted in a normalised form. This allows a clear visualisation of
the deviation from the imposed state. It can be seen that the deviation is pronounced
near the pile soil boundary. This is a consequence of the high stress gradients near
the pile. Besides, the pile region was at geostatic stress level when the installation
effects are imposed around it. As a results of having low stresses in the pile and
high stresses in the soil just next to it, the pile deforms very slightly to get into
equilibrium. As a consequence of this deformation, a reduction in the stresses around
the pile can be observed. Similarly, the interfaces at the pile soil boundary cause also
slight reduction/alteration of the stresses to get into equilibrium with surrounding
stress state.

6.2.3 Conclusions
In this section the possibility of transferring the installation effects resulting from
a PR simulation to a different FE mesh has been investigated. It has been shown
that the mesh discretisation had an influence on the imposed and equalised states as
6.3. Approximation of the Installation Effects by Bivariate Nonlinear Regression 109

well as the load displacement behaviour of the pile, around which the interpolated
installation effects were imposed and equalised. Interpolation from a fine mesh to
a finer mesh, results in smoother distribution of transferred installation effects (e.g.
Cartesian stresses and void ratio). Load displacement behaviour of the pile after
interpolating the installation effects on to a finer mesh is also smoother than the
interpolation to the coarser discretisation.
It can be concluded that the transfer of the installation effects to meshes with
different discretisation is possible, provided that the new mesh does not diminish
the distribution of the field due to its coarseness.

6.3 Approximation of the Installation Effects by


Bivariate Nonlinear Regression
In Section 6.2 the possibility of transferring the installation effects to different FE
meshes was shown. In order to represent the installation effects an interpolation
method was employed. It is more practical and desirable to find a surface that
approximates the data rather than requiring the interpolation of data points. This
section presents a methodology to represent the installation effects by approxima-
tions defined by closed form mathematical functions. The idea is to simplify the
representation of the installation effects and provide a basis for the representation
of the installation effects including the effects of pile diameter, penetration length,
and relative density of the soil.
First the theory of multiple non-linear regression analysis is given shortly (Section
6.3.1). Then Section 6.3.2 presents the methodology of representing the surfaces by
approximation (fitting). Finally, the results of surface fittings obtained for different
cases in which pile diameter, penetration length and the soil density were varied, are
presented and discussed.

6.3.1 Theoretical Background


The theoretical background summarised in this section is adopted from [86]. If the
surface data points are given as:

z = f (x, y) (6.20)

Then an approximate surface can be reconstructed from these data points based on
a model function to represent the surface as:

z = f (x, y; a1 , a2 , a3 ,··· , am ) (6.21)

having m parameters (coefficients). Then the surface approximation problem reduces


to a regression problem of determining the model parameters of the approximated
surface function by minimisation of the sum of squared errors, χ2 :

n
χ2 = (zi − f (x, y; a1 , a2 , a3 ,··· , am )) (6.22)
i=1
110 6. Generalisation of Installation Effects

However this is an ill-posed problem since there are infinite number of functions that
can represent the data some of which fit equally well. [86] suggests the augmentation
of the regression form to constrain the selection of approximation to a single choice
by forcing the solution to be smooth by the following criterion:
n ˆ ˆ 2 
∂ f ∂f ∂f ∂2f
2
χ = (zi − f (xi , yi )) + α
2
+2 + 2 dxdy (6.23)
i
∂x2 ∂x ∂y ∂y

With the condition, α > 0. The second part in the above equation with the coef-
ficient α, i.e. the regularizing term, assures smoothness of the fit. The coefficient
α regulates how close the data will be represented in the approximation (for small
α) and how smooth the solution will be (by a large α). The modified form assures
that the approximation has not only a unique solution, but also definitely better
than other solutions. Depending on the model function, the regularizing term can
be difficult to calculate. Therefore, the partial derivatives can be replaced with finite
difference approximations.

6.3.2 Describing Installation Effects by Surface Fitting


Based on the theory summarised in Section 6.3.1, the stress (σ) and void ratio (e)
fields, ψσ and ψe , around a displacement pile can be represented by surfaces of form:
ψ σ = ψ σ (r, z) (6.24)
ψ e = ψ e (r, z) (6.25)
in axisymmetric coordinate system. Chapter 5 presented the stress and void ratio
fields of equalisation state of several PR simulations. In Section 6.2, interpolation
functions were used to represent these surfaces. As mentioned in the introduction
of this section, the idea is to represent these surfaces by a general form, which does
not require data point as in the case of interpolation. Therefore, forms given in
Eqns.6.24 and 6.25 were investigated by the surface regression technique.
In order to perform a regression analysis, a model function is required. However,
it is not straightforward to find model functions considering the complex shapes of
the installation effects data.
After a thorough investigation of possible functions that could represent these
surface data, it was found that a summation of a number of Gaussian (exponential)
functions could represent the installation effects. Therefore the following form is
proposed as model function:

n
2
+ci (z−zi )2 )
ψ̂ = ai e−(bi (r−ri ) (6.26)
i=1

where, ψ̂ is the approximate representation of the exact field ψ.


The number of exponential forms were selected based on the peaks observed in
the installation effects data. Ideally n should be as large as possible since:
lim ψ̂ = ψ (6.27)
n→∞
6.3. Approximation of the Installation Effects by Bivariate Nonlinear Regression 111

However, a compromise between having the surface representation as close to


the installation effects data and a form as simple and smooth as possible has to
be made. Unfortunately, an augmented form, as in the form of Eqn.6.23 could not
be employed since the regularisation term of the proposed model function does not
have an analytical solution (i.e. the integral of second derivatives of the proposed
model function does not exist). Therefore the regression was performed on the model
function directly.
First the model functions were selected as only single form. Then the number of
terms (n) was increased until a reasonable representation of the surface was obtained.
In order to ensure that the solution converges to a good approximation, the upper
and lower bounds of the coefficients were selected considering the magnitude and
decay rate in radial and vertical directions as well as the location of the peaks
observed in the surface data.
As discussed in Section 5.3, for a better representation of the installation effects,
  
each Cartesian stress component (σrr , σzz , σθθ and σrz ) was normalised by the
geostatic effective mean stress:
1 
p0 = 
(σ + σzz 
+ σθθ ) (6.28)
3 rr
and the void ratio (e) field was normalised by the geostatic void ratio, which is
determined based on the formula proposed by [21]:
 n
3·p0

eK0 = e0 · e
hs
(6.29)

where, eK0 is the void ratio at K0 state; e0 is the void ratio at zero stress; hs and
n are the Hypoplasticity model parameters representing the granulate hardness and
exponent controlling the curvature of the 1D compression curve, respectively. More
information on the hypoplastic constitutive model can be found in Chapter 3.
Therefore, the fitting was performed on the normalised Cartesian stresses, Rrr ,
Rzz , Rθθ and Rrz and void ratios, Re , which were presented in Chapter 5 by the
Eqns.5.9 -5.13.
Based on the preliminary study the following model functions, in which the length
scales are also normalised by the pile diameter D, are proposed for the representation
of each normalised Cartesian stress and void field:

3 · K0  3
2 2
ψ̂rr = + ai1 · e−(bi1 (r/D−ri1 ) +ci1 (z/D+L/D+zi1 ) ) (6.30)
1 + 2 · K0 i=1

3 · K0  4
2 2
ψ̂zz = + ai2 · e−(bi2 (r/D−ri2 ) +ci2 (z/D+L/D+zi2 ) ) (6.31)
1 + 2 · K0 i=1

3 · K0  3
2 2
ψ̂θθ = + ai3 · e−(bi3 (r/D−ri3 ) +ci3 (z/D+L/D+zi3 ) ) (6.32)
1 + 2 · K0 i=1
112 6. Generalisation of Installation Effects


6
2
+ci4 (z/D+L/D+zi4 )2 )
ψ̂rz = ai4 e−(bi4 (r/D−ri4 ) (6.33)
i=1


4
2
+ci5 (z/D+L/D+zi5 )2 )
ψ̂e = 1 + ai5 e−(bi5 (r/D−ri5 ) (6.34)
i=1

Where, aim , bim and cim are the fitting coefficients representing the magnitude and
spread of the exponential forms and rim and zim are similarly the fitting coefficients
representing the location of the Gaussian forms of corresponding variable m (m=
1, 2,··· ,5; where m = 1 represents radial, m = 2, the vertical, m = 3,the tangential,
m = 4, the shear stress components and m = 5 to represent the void ratio). For
the sake of compatibility for different L/D analyses, the vertical position of the
Gaussian form is given relative to the location of pile tip (e.g. the term in the
exponent cim (z/D + L/D + zim )2 ).
The model functions have in common two spatial variables, r and z, one constant,
K0 and five fitting coefficients, aim , bim , cim , rim and zim . Since the model functions
are not linear in r and z, the resulting regression analysis applied on the surface
data is a bivariate non-linear regression. The minimisation of χ2 can be performed
iteratively based on some initial guesses as well as upper and lower limits.
The PR results of each case have a different discretisation level. Most of the
effects are pronounced near the pile, as expected. To have a more reliable R2 value
and a consistent data distribution, the PR field data were interpolated onto a uniform
grid. The grid spacing is smaller near the pile, in the influence region, where the
installation effects shows clear deviation from the K0 state, and larger otherwise. As
a result of using the new grid the regression analysis is preconditioned to fit better
in the influence region.
The spacing in the grid can be selected as fine as possible near the pile, in the
influence region. For the sake of practicality, the grid is created based on the pile
dimensions D and L using the following discrete set of equations representing the
grid regions (Fig.6.11). The grid can be formed using the discrete set of equations
given in 6.36 and 6.37.
A generic form of the regions next to the pile shaft can be given as:

lis : Ri−1 ≤ r ≤ Ri ; Z0 ≤ z ≤ Z1 |rj = Ri−1 +(j−1)·Δsi , zj = Z0 −(j−1)·Δsi (6.35)

for i = 1.7 and j = 1, Ri −R


Δsi
i−1
where rj and zj indicate the location of each grid
point. Similarly, the generic form of the regions below the level of pile base can be
given as:

lit : Ri−1 ≤ r ≤ Ri+1 ; Zk+1 ≤ z ≤ Zk+2 |rj = Ri−1 + (j − 1) · Δsi , zj = Zk+1 − (j − 1) · Δsj (6.36)

for i = 1, 2,··· , 7; k = 1, 2,··· , 7 and


 *
Rk+1 −R0
i = 1; 1, Δs1
j= ;
i > 1; 1, Rk+1 −R Δsi
k+i−1
6.3. Approximation of the Installation Effects by Bivariate Nonlinear Regression 113

Figure 6.11: Geometrical properties of the grid prepared to precondition the distribution of
surface data for bivariate nonlinear regression calculations.

Table 6.1: Geometrical properties of the regions defined for the interpolation grid

i 0 1 2 3 4 5 6
Ri 0 0.5 · D 2.5 · D 5·D 10 · D 20 · D 50 · D
Zi 0 −L −L − 2.5 · D −L − 5 · D −L − 10 · D −L − 20 · D - 50 · D
Δsi - 0.02 · D 0.1 · D 0.5 · D 2.5 · D 10 · D −

where the boundaries of the regions Ri and Zi and the spacing in the corresponding
region, Δsi are given in Table 6.1.
The results of each PR analysis (equalisation states) were interpolated to the
grids prepared using the interpolation technique explained in Section 6.1. Based on
the grid data, bivariate nonlinear regression technique was applied for each norm-
alised Cartesian stress and void ratio field using the model functions ψ̂rr , ψ̂zz , ψ̂θθ ,
ψ̂rz and ψ̂e given in Eqns. 6.31 to 6.35, respectively.
As mentioned in Section 6.2, there are infinite number of fitting results that may
be equally good (e.g. having the same R2 ). Furthermore, the nonlinear regression
algorithm requires initial guesses bounded by upper and lower limits for each fitting
coefficient. Otherwise, a good or converged solution cannot be found. Therefore,
each fitting coefficient of each model function was preconditioned by an initial guess
together with the upper and lower limits. The upper and lower limits were estimated
based on the distribution of the surface data. For example the rim and zim values
were bound by the regions where a peak is observed in the surface data. Similarly,
the aim values were bound by the magnitude of these peaks. The difficult part is the
determination of the bim and cim limits as these values cannot be extracted directly
from the surface data, at least visually. To be able to visualise the meaning of the
coefficients bim and cim the following generalised Gaussian form can be used:

f (r, z) = a · e−{b(r−r0 ) +c·(z−z0 )2 +2d·(r−r0 )·(z−z0 )}


2
(6.37)
114 6. Generalisation of Installation Effects

where,

cos2 θ sin2 θ
b= + (6.38)
2σr2 2σr2

sin2 θ cos2 θ
c= + (6.39)
2σz2 2σz2
sin 2θ cos 2θ
d= + (6.40)
4σr2 4σz2
and, θ is the rotation of the Gaussian surface (the angle between the global r axis
and the local r axis of the Gaussian function); σr and σz are the first inflection
points in r and z directions, respectively.
In the model functions, we have used a simplified version of the Gaussian form,
where the cross-term (d = 0) and rotation θ = 0 rad were neglected. The aim was
to obtain a versatile form for the generalisation of the model functions (Section 6.4)
given in Eqns 6.30 to 6.34.
Without the cross-term, the general form and the coefficients reduce to:

f (r, z) = a · e−{b(r−r0 ) +c·(z−z0 )2 }


2
(6.41)
1
b= (6.42)
2σr2
1
c= (6.43)
2σz2
Therefore, bim and cim limits can be estimated by the Equations 6.42 and 6.43
considering the radial and vertical span of a patch of the surface data, which is rep-
resented by an exponential form. For example σr can be selected as the radial span
of the function, where the model function reduces, approximately, to its geostatic
value.
In order to eliminate the spurious concentrations and to obtain the underlying
functions that describe the installation effects, median filtering was applied on the
interpolated PR analysis results (uniform grid).
The optimum filtered outputs are determined by applying 2N+1 by 2N+1 win-
dow sizes, where N = 1, 2,··· , 10. Figure 6.12 shows the results of one example case
σrz of this sensitivity study.
Based on the sensitivity study, a minimum window size of 11 x 11 for σrr and 15
x 15 for the rest of the variables has been found. Figure 6.13 compares the original
and filtered data of the D = 0.30m; L = 2.5D; Id = 0.40 case. It can be seen
that the spurious data points have been filtered out and smooth surfaces have been
obtained.
6.3. Approximation of the Installation Effects by Bivariate Nonlinear Regression 115

Median Filters
The median filters are signal processing tools that are used to filter out noise. Me-
dian filters can preserve signal edges while filtering out impulses [69]. It considers
neighbouring values to decide the current data has a representative value of its sur-
rounding. The median filter requires a predefined window size of odd integer (2N+1,
N=0,1,2,...). For the edges the values are repeated depending on the window size.
The following 1D example presents how the median filter works.
Assume a 1D signal, s = [5 75 15 7 10]. The median filtered output, f (s) of a window
size of 3, will be:
f [1] = M edian[5 5 75] = 5
f [2] = M edian[5 75 15] = M edian[5 15 75] = 15
f [3] = M edian[75 15 7] = M edian[7 15 75] = 15
f [4] = M edian[15 7 10] = M edian[7 10 15] = 10
f [5] = M edian[7 10 10] = M edian[7 10 10] = 10
Similarly, median filtering can be applied on a 2D signal, i.e. surface data, considering
a window size of 2N+1 by 2N+1.
 

Figure 6.12: Comparison of window size effect of the median filter on the σrz distribution of
D = 0.30 m; L = 2.5D; Id = 0.40

6.3.3 Results
In this section, the results of the bivariate nonlinear regression analyses described
in Section 6.3.2 are presented. Each of the normalised Cartesian stress, Rrr , Rzz ,
Rθθ and Rrz , and void ratio, Re field is presented together with the approximated
116 6. Generalisation of Installation Effects

Figure 6.13: Comparison of filtered data with PR data of σrz distribution of D = 0.30 m;
L = 2.5D; Id = 0.40 a. σrr , b. σzz , c. σθθ , d. σrz ,and e. e

surfaces ψ̂rr , ψ̂zz , ψ̂θθ , ψ̂rz , and ψ̂e in Figures 6.14 to 6.18, respectively. The analysis
results of the D = 0.30m; L = 10 · D; Id = 0.60 and the D = 0.40m; L = 10 · D;
Id = 0.80 cases are presented side by side.
The goodness of the surface fits are summarised in Table 6.2. In general, the
fitting results have R2 values larger than 0.9. However, these values alone do not
evaluate how good the fits are. [10] showed that nearly identical simple statistical
properties can be obtained from quite different data sets and pointed out the import-
ance of visual inspection for the fitting result. Therefore, the results summarised in
Table 6.2 should be evaluated together with the graphical representations presented
in Figures 6.14-6.18.
The R2 is the ratio of sum of squares of the regression, SSR and the total sum of
squares, STT. SSE is the sum of squared errors. The general form of these properties
are defined as follows:

n
2
SSR = wi (ŷi − ȳ) (6.44)
i=1


n
2
SST = wi (yi − ȳ) (6.45)
i=1


n
2
SSE = wi (yi − y¯i ) (6.46)
i=1
6.3. Approximation of the Installation Effects by Bivariate Nonlinear Regression 117

'n 2 'n 2
SSR SSE i=1 wi (ŷi − ȳ) i=1 wi (yi − ŷi )
2
R = =1− = 'n 2 = ' n 2 (6.47)
i=1 wi (yi − ȳ) i=1 wi (yi − ȳ)
SST SST
where wi ’s are the weights, yi are the data points, ȳ is the mean and ŷi are the
predictions by the fit. In this study the data was not weighted, therefore wi = 1.
The root mean squared error, RMSE, is an average measure of the spread of
the of data points from the predicted surface (residuals). It is a good measure of
the accuracy of the fitting as it combines the residuals in a single predictive value.
The RMSE value provides information on the magnitude of the error, which is also
known as the standard error of the regression, s is determined using the residual
degrees of freedom, nrdof . The nrdof is the difference between the total number of
data points, n and the number of fitting coefficients, m as:
√ SSE
RM SE = s = M SE = ; nrdof = n − m (6.48)
nrdof

Table 6.2: Summary of regression analysis of a)D = 0.30m; L = 10D: Id = 0.60 and b)D = 0.40m;
L = 10D: Id = 0.80

SSE R2 RMSE
a b a b a b
Rrr 2.1188 · 104 1.2942 · 104 0.9801 0.9916 0.5447 0.4257
Rzz 7.9626 · 102 1.2159 · 104 0.9822 0.9841 0.3339 0.4127
Rθθ 1.6485 · 104 1.1126 · 104 0.9760 0.9889 0.4805 0.3947
Rrz 1.0695 · 104 1.1572 · 104 0.6720 0.7820 0.3870 0.4026
Re 87.6398 109.7207 0.5920 0.7648 0.0350 0.0392

Figure 6.14: Fitting results of normalised vertical stresses ψ̂rr of cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

  
After obtaining the fitting coefficients, Cartesian stresses, σrr , σzz , σθθ , σrz , and
the void ratio e can be determined using the Eqns. 6.49 - 6.53. The forms presented
118 6. Generalisation of Installation Effects

Figure 6.15: Fitting results of normalised vertical stresses ψ̂zz of cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

Figure 6.16: Fitting results of normalised hoop stresses ψ̂θθ of cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

in Eqns. 6.49 - 6.53 were obtained by substituting Eqns. 6.30 - 6.34 back into Eqns.
5.9 - 5.14 (Section 5.3).
+ ,
3 · K0 3
  −(bi1 (r/D−ri1 )2 +ci1 (z/D+L/D+zi1 )2 )
σrr = p0 · + ai1 · e (6.49)
1 + 2 · K0 i=1
+ ,
4 · K0  4
2 2

σzz = p0 · + ai2 · e−(bi2 (r/D−ri2 ) +ci2 (z/D+L/D+zi2 ) ) (6.50)
1 + 2 · K0 i=1
6.3. Approximation of the Installation Effects by Bivariate Nonlinear Regression 119

Figure 6.17: Fitting results of normalised shear stresses ψ̂rz of cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

Figure 6.18: Fitting results of normalised void ratios ψ̂e of cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

+ ,
3 · K0  3
2 2

σθθ = p0 · + ai3 · e−(bi3 (r/D−ri3 ) +ci3 (z/D+L/D+zi3 ) ) (6.51)
1 + 2 · K0 i=1


6
2
+ci4 (z/D+L/D+zi4 )2 )
σrz = p0 · ai4 · e−(bi4 (r/D−ri4 ) (6.52)
i=1
+ ,

4
−(bi5 (r/D−ri5 )2 +ci5 (z/D+L/D+zi5 )2 )
e = eK0 1 + ai5 · e (6.53)
i=1
120 6. Generalisation of Installation Effects

Or considering the hypoplastic relation of [21]:


 n + ,
3·p0

4
−(bi5 (r/D−ri5 )2 +ci5 (z/D+L/D+zi5 )2 )
e = e0 · e ai5 · e
hs
1+ (6.54)
i=1

Where e0 is the void ratio at zero stress level, hs and n are hypoplasticity model
parameters representing granulate hardness and exponent of the logarithmic decay
function. Figures 6.19 - 6.23 compare the surface fits of the D = 0.30m; L = 10D;
Id = 0.60 and D = 0.40m; L = 10D; Id = 0.80 cases with the corresponding PR
results.

 with PR data of cases: a. D = 0.30m; L = 10D;


Figure 6.19: Comparison of the surface fit of σrr
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

 with PR data of cases: a. D = 0.30m; L = 10D;


Figure 6.20: Comparison of the surface fit of σzz
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80
6.3. Approximation of the Installation Effects by Bivariate Nonlinear Regression 121

 of cases: a. D = 0.30m; L = 10D; I = 0.60,


Figure 6.21: Comparison of the surface fit of σθθ d
b. D = 0.40m; L = 10D; Id = 0.80

 with PR data of cases: a. D = 0.30m; L = 10D;


Figure 6.22: Comparison of the surface fit of σrz
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

As one of the main goals of the study is to improve the numerical load test
curves of the displacement piles, i.e. increased stiffness and strength, the accuracy
of fits were evaluated by the load - displacement behaviour of the pile after the
installation effects are imposed using the forms given in Eqns 6.49 - 6.54. In order
to obtain a smooth distribution of stresses after equilibrium, the effects were imposed
to the whole mesh including the pile region with same soil properties and without
activating the interfaces around the pile. After imposing the installation effects and
obtaining equilibrium, the pile was activated by switching the material properties of
the corresponding elements from soil (hypoplasticity) to pile (linear elasticity). The
interfaces around the pile were also activated at the same time. The load tests were
122 6. Generalisation of Installation Effects

Figure 6.23: Comparison of the surface fit of e with PR data of cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

performed on this state to compare the force displacement curves with the reference
curves obtained from load tests performed after the PR simulations.

Figure 6.24: Comparison of numerical load test curves of the imposed fields using the fitting
results with PR result and K0 state of cases: a. D = 0.30m; L = 10D; Id = 0.60, b. D = 0.40m;
L = 10D; Id = 0.80

Figure 6.24 shows a general trend that the imposed states using the proposed
functions of stress and void ratio underpredict the PR results. The deviation from
the PR result increases for the cases with lower densities (e.g. Id = 0.40). One of the
reasons for obtaining reduced capacity for the imposed states is the approximation
error introduced during the regression analyses. More importantly, the surface peaks
were smoothed by the median filtering. Figure 6.24 also presents three different cases
6.3. Approximation of the Installation Effects by Bivariate Nonlinear Regression 123

in which the installation effects are imposed on each Cartesian stress component
and the void ratio (i.e., σ̂rr (r, z); σ̂zz (r, z); σ̂θθ (r, z); σ̂rz (r, z) and ê(r, z)), only on
Cartesian stress components (i.e.,σ̂rr (r, z); σ̂zz (r, z); σ̂θθ (r, z); σ̂rz (r, z); void ratio
follows Bauer’s exponential relation at current stress level) and only on orthogonal
Cartesian stress components (i.e., σ̂rr (r, z); σ̂zz (r, z); σ̂θθ (r, z); shear stresses are
assumed to be zero, σ̂rz (r, z) = 0). It is being realized that the latter situation will
certainly introduce an unbalance into the model. However, this unbalance will be
removed by the NiL step after the stresses have been imposed.
It can be seen from the same figure that the second case in which only Cartesian
stresses are imposed based on the approximated installation effects gives the closest
load displacement curve compared with the PR results. The PR curve has numerical
oscillation at the beginning of loading, whereas the approximations give smooth
curves.

Figure 6.25: Change of approximated field imposed (red), after equilibrium (soil, pink) and after
pile and interfaces were activated (pile+NiL, green) a. radial and b. vertical directions of case
D = 0.30m; L = 10D; Id = 0.60

Before performing the load tests, a NiL step is applied as soon as the approx-
imated state was imposed around the pile to ensure equilibrium. At the end of the
NiL step, reductions in the imposed stresses were observed (Figure 6.25). These
reductions caused the underestimated load displacement behaviour of the imposed
approximate state. The reduction in stresses is also an indication of unbalanced
forces resilient from the approximated stress field. Since each Cartesian stress com-
ponent was approximated separately, the equilibrium of the stress field should be
checked. The equilibrium of stresses can be checked using the force - equilibrium
equations for a continuum in axisymmetric conditions for radial and vertical direc-
tions:
1 ∂  ∂ σ
(rσrr )+ (σrz ) θθ = 0 (6.55)
r ∂r ∂z r
1 ∂  ∂ 
(rσzz )+ (σzz ) − γsoil =0 (6.56)
r ∂r ∂z
124 6. Generalisation of Installation Effects

The open forms of the unbalance in radial and vertical directions can be obtained
by substituting Eqns. 6.49 - 6.52 into Eqns. 6.55 and 6.56 (See Appendix A).

Figure 6.26: Distribution of radial unbalanced stresses for cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

Figure 6.27: Distribution of vertical unbalanced stresses for cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

Figures 6.26 and 6.27 present the radial and vertical unbalance stress surfaces
for the D = 0.30 m; L = 10 · D; Id = 0.60 and the D = 0.40 m; L = 10 · D; Id = 0.80
cases. The unbalance is high near the pile tip and close to the symmetry axis.
In order to satisfy the equilibrium conditions in radial and vertical directions, one
  
can assume three major stress components (σrr , σzz , σθθ ) are successfully represent-
ing the stress field and then try to calculate the shear component (σrz ) analytically
by satisfying both of the equilibrium equations (Eqns. A.5 & A.6). However the

integral of the ∂z (σrz ) term in the vertical equilibrium equation is unknown. As the
6.4. Generalisation of the Installation Effects 125

goal is to express the installation effects in geometric functional form, the numer-
ical integration was not employed to determine the shear stresses from equilibrium
equations using the geometric functions proposed for the major stress components.
Another way of satisfying the equilibrium conditions is performing the regression
analysis to minimise the unbalance in horizontal and vertical directions; in other
words, employing the Eqns.A.5 & A.6 as model functions. However, a converging
solution for these regression analyses could not be obtained. Therefore, performing
the NiL step after imposing the installation effects was employed in order to remove
the unbalance.

6.3.4 Discussion of Results and Conclusions


In this part of the study, the Cartesian stress and void ratio fields resulting from
pile jacking simulations by the PR technique, have been approximated by model
functions of Gaussian form. The fitting results have shown that the approximated
surfaces successfully represent the field data. Nevertheless, due to the approxima-
tion errors introduced in each fit, some unbalance was observed when the separate fit
surfaces were combined to form the stress field. The unbalance caused some reduc-
tions at the NiL step, after the stresses were imposed and equilibrium was satisfied
within the FE analysis. Several ways were investigated to reduce the unbalance and
none could reduce the unbalance as good as the FE NiL step.
It was shown that the equilibrium state obtained by the NiL step after imposing
the proposed functions from separate regression analyses yields a conservative load
displacement behaviour (Figure 6.24). Besides, the goodness of the separate surface
fits on normalised Cartesian stresses were better than those using the equilibrium
equations. Furthermore, the final goal is to represent the stress and void ratio fields
of the geometry and density variations in a unique form. Obtaining the unique form
requires a second regression over the fitting results of each variation. The method
will be explained in Section 6.4 in more details. It is obvious that the second fitting
will also introduce a certain level of approximation. Therefore, we believe that the
results of each surface fitting presented in this part are good enough to represent
the stress and void ratio fields obtained from the PR simulations.

6.4 Generalisation of the Installation Effects


In this section the possibility of combining the separate functional forms fitted for
different pile diameter, penetration length and sand density variations, into a ver-
satile and comprehensive form us described in which the coefficients are defined in
terms of D, L/D and Id . In Section 6.4.1 the multiple non-linear regression of each
coefficient is presented. The performance of the proposed forms are checked against
the PR analysis results based on load displacement curves after the installation
effects are imposed (Section 6.4.2).
126 6. Generalisation of Installation Effects

6.4.1 Determination of the Coefficients of Generalised Func-


tions
Several functional forms (surface equations) representing the installation effects of
each pile diameter, penetration length and soil density variations were derived in
Section 6.2. In order to introduce a more generalised representation of the installa-
tion effects, the approximated forms of each variation described in Section 6.2 should
be unified. One way of doing so is introducing the altered variables such as L/D
and Id in the model function, besides the spatial variables r and z.
An easy way to obtain a generalised form is to define the change of the fitting
coefficients, aim , bim , cim , rim and zim (Section 6.3.2) in terms of L/D and Id .
Since the model functions, ψ̂rr , ψ̂zz , ψ̂θθ , ψ̂rz and ψ̂e employed to approximate
surface data of each pile geometry, L/D and soil density, Id variation were the same,
it is expected that approximating the coefficients in the D, L/D and Id space would
yield a generalised form. This generalised form, then represents all variations of pile
diameter, penetration length and soil densities.
The length measures have already been scaled by the pile diameter D in Section
6.3 and the approximation of the surface data was performed on the scaled spatial
variables r and z. Considering the low variations in the coefficients for the two pile
diameter size variations (D = 0.30 m and D = 0.40 m), and for the sake of simplicity,
the approximation of the coefficients was performed on the model functions of type:

ψ̂coef f = f (L/d, Id ) (6.57)

Based on the fitting results of the variations presented in Section6.3 and surface plots
of each coefficient in the L/d, Id space, the following model function is proposed to
approximate the surface data of each coefficient, aim , bim , cim , rim and zim :
 q
L
ψ̂aim = ψ̂bim = ψ̂cim = ψ̂rim = ψ̂zim = k · Id p · (6.58)
D

6.4.2 Results
In this section, the results of the regression analyses on the coefficients, aim , bim ,
cim , rim and zim are presented. Figure 6.28 shows an example case (the coefficients
a31 and a21 of ψ̂rr ) of the results of the fitting applied on the coefficients of the
model functions (ψ̂rr , ψ̂zz , ψ̂θθ , ψ̂rz and ψˆe ) proposed in Section 6.3. The fitting
applied on these approximated fields (fitting results) introduce additional errors.
Therefore, the approximations, ψ̂aim , ψ̂bim , ψ̂cim , ψ̂rim and ψ̂zim in this part are
not as good as the fitting results of each variation separately. Yet, the generalised
forms provide a versatile representation of the Cartesian stress and void ratio fields
around a displacement pile of any diameter, penetration length or density within
the variation boundaries. In order to extract a representative surface, the outliers
were excluded in the fitting process. At the beginning a confidence interval of 0.65
(≈ σ, the standard deviation) was chosen. In an iterative scheme the confidence
interval was reduced (0.05 in each iteration) until an R2 value not less than 0.85 is
6.4. Generalisation of the Installation Effects 127

obtained or the number of excluded data points were not more than half of the total
data points. The goodness of the regression analyses are summarised in Table 6.3.
As conditioned, the fitting results have R2 values mostly higher than 0.8. There are
few cases with low R2 values (i.e. b31 , c21 , r31 , z21 ). Figures 6.28 & 6.29 show some
example fitting results having strong and weak correlations, respectively.

Figure 6.28: Approximation result for the coefficient a11 and a31 (of ψ̂rr ) in L/D, Id space.

Figure 6.29: Approximation result for the coefficient b31 and r31 (of ψ̂rr ) in L/D, Id space.

Using the generalised functions, the initial state around a jacked pile can be
updated so that the installation effects are adequately represented. Figures 6.30
-6.34 present the surface data of the installation effects variables, and corresponding
surface fit for the D = 30 m; L = 10 · D; Id = 60 and the D = 40 m; L = 10 · D;
Id = 80 cases. In general the approximations represent the surfaces reasonably well.
128 6. Generalisation of Installation Effects

Table 6.3: Summary of regression analysis on the generalisation of the normalised radial stresses

k m n SSE R2 RMSE
a11 12.6500 0.42041 0.62771 14.6125 0.9494 1.0602
a21 6.7982 0.20536 1.1027 1.6766 0.9602 0.3904
a31 3.2202 0.26622 0.59298 0.4360 0.9478 0.2088
b11 0.24033 0.1403 −0.2664 3.7143 · 10−4 0.9832 0.0061
b21 0.089398 −0.63766 −2.4166 0.0141 0.9496 0.0343
b31 0.16667 0.14443 −0.50662 0.0882 0.4825 0.0681
c11 2.0000 0.0000 0.0000 3.8157 · 10−18 0.9999 5.8896 · 10−10
c21 1.1202 −0.42806 −0.19518 0.1077 0.6474 0.1094
c31 0.64216 −0.11266 −0.61411 0.0147 0.9250 0.0384
r11 −0.49999 −3.676 · 10−6 −3.304 · 10−5 2.6628 · 10−10 0.9999 5.4393 · 10−6
r21 0.0011251 1.3995 −5.000 1.8777 0.8919 0.3426
r31 3.125 −0.079436 0.76681 13.9002 0.3593 0.8553
z11 0.59488 −0.077483 0.11228 0.0018 0.8926 0.0108
z21 2.7177 −0.29554 0.3013 0.4518 0.6431 0.2241
z31 0.0089438 0.38163 −4.7903 0.0773 0.9872 0.087

Furthermore, the approximate surfaces have a smooth representation of the results


of the PR analyses.

 distribution of generalised form with PR results of


Figure 6.30: Comparison of radial stress, σrr
the cases: a. D = 0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

A numerical pile load test can be employed to obtain the load-displacement be-
haviour of the pile with imposed installation effects and to assess the performance
of the proposed generalised functions. Figure 6.37 presents the load displacement
curves of a model pile with different stress and void ratio states. Both figures suggest
that the approximations yield improved capacities as well as stiffnesses compared to
geostatic states. Due to additional approximation errors as well as the unbalance
stresses (i.e. Figure 6.35 & 6.36), lower capacities and stiffnesses have been obtained
for the load displacement curves of the generalised forms, ψ̂(r, z, L/D, Id ) than those
for the individual fits , ψ̂(r, z). The load displacement curves of the reference cases
(the interpolated PR results) were obtained based on the equalisation state, after
penetration simulations. It can be seen that for the dense case a better approx-
6.4. Generalisation of the Installation Effects 129

 distribution of generalised form with PR results


Figure 6.31: Comparison of vertical stress, σzz
of the cases: a. D = 0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

 distribution of generalised form with PR results


Figure 6.32: Comparison of tangential stress, σθθ
of the cases: a. D = 0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

imation is obtained. In the generalised formulas and PR results a dilation zone


is evident around the pile. Employing the generalised formulas only for Cartesian
stresses and using a void ratio based on the current stress state result in a com-
paction zone around the pile. As a result of this density increase, the stiffness and
strength of the soil increase and yield improved load displacement behaviour. How-
ever, the improvement was based on the increase in stress and strength of the sand
around the pile, which was assumed to be compacted rather than dilated. However,
the dilation around a pile had been observed in previous experimental studies (e.g.
[56]) and numerical results (e.g. [112], PR results). The load displacement curve
of generalised case for the dense soil is an indirect indication of the accuracy of ap-
proximation.Therefore, the improvement in the global load-displacement behaviour
130 6. Generalisation of Installation Effects

Figure 6.33: Comparison of shear stress, σrz distribution of generalised form with PR results of
the cases: a. D = 0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

Figure 6.34: Comparison of void ratio, e distribution of generalised form with PR results of the
cases: a. D = 0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

should be evaluated with caution.

6.5 Conclusions
In this Chapter the possibility of representing the installation effects by generalised
geometrical functions has been investigated. First a sensitivity analysis was per-
formed to check the limitations of transferring the installation effects resulting from
a PR simulation to a different FE mesh. Furthermore, it was shown that the mesh
discretisation had an influence on the imposed and equalised states as well as the
6.5. Conclusions 131

Figure 6.35: Comparison of radial unbalanced stresses for cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

Figure 6.36: Distribution of vertical unbalanced stresses for cases: a. D = 0.30m; L = 10D;
Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

load displacement behaviour of the pile. Interpolation from a fine mesh to a finer
mesh, resulted in smoother distribution of transferred installation effects. Smoother
load displacement curves were obtained for the cases with interpolation to a finer
mesh than those to a coarser discretisation. The first conclusion is that the transfer
of the installation effects to meshes with different discretisation is possible, provided
that by the new mesh the distribution of the installation effects are not disturbed
due to the mesh coarseness.
Secondly, the Cartesian stress and void ratio fields resulting from pile jacking
simulations by PR technique, have been approximated by model functions of Gaus-
sian form. The fitting results have shown that the approximated surfaces successfully
132 6. Generalisation of Installation Effects

Figure 6.37: Comparison of the numerical load test curves based on the equalised approximate
installation effects with the state obtained using PR and the reference K0 state of cases: a. D =
0.30m; L = 10D; Id = 0.60, b. D = 0.40m; L = 10D; Id = 0.80

represented the field data. Nevertheless, due to the approximation errors introduced
in each fit, some unbalance was observed when the separate fit surfaces were com-
bined to form the stress field. The unbalance caused some reductions at the NiL
step, where the stresses were imposed and equilibrium was satisfied within the FE
analysis. Several ways were investigated to reduce the unbalance and none could
reduce the unbalance as good as the FE NiL step. It was shown that the equilib-
rium state obtained by the NiL step after imposing the proposed functions found by
separate regression analyses yields a conservative load displacement behaviour.
Finally the fitting results obtained in Section 6.3 were generalised by representing
the fitting coefficients of these functions in terms of the varied parameters, i.e. pile
diameter, D, penetration length, L and soil density, Id . As a result of second order
approximation, a lower level of approximation has been obtained. Nevertheless,
graphical inspection showed that the approximations were reasonable. The load
displacement curves support the goodness of the generalised forms indirectly. The
results also have shown that the deviation from the reference case reduced with
increasing soil density. One of the reason for the density dependent behaviour is
the level of reductions in the NiL step is lower for stiff material. The equilibrium
states can be obtained by minor deformations, whereas soft material (loose soil)
deforms much and therefore the stresses reduce more. In the following chapter,
the proposed generalised formulas will be employed first in different density and
geometry variations for verification purposes. Then, the results of a geocentrifuge
test in Baskarp Sand [56] sand will be compared for the validation of the formulas.
Chapter 7

Application

7.1 Introduction
In the previous chapters the concept of installation effects, modelling and incorpora-
tion with different techniques were presented. Necessary steps for the determination
of model parameters (Chapter 4) and the techniques employed (Chapters 3, 4 & 6)
have been explained in details. Chapter 7 is devoted to the verification (7.1) and
validation (7.2) of the generalised method proposed in the previous chapter. An
instrumented centrifuge model pile is selected as the validation case.

7.2 Verification Cases


In Section 6.4 a generalised form of installation effects has been proposed based on
the PR analyses summarised in Table 5.2. This section verifies the proposed formulas
by geometry and density variations within the calculation grid (i.e. 0.30 m ≤ D ≤
0.40 m; 2.5D ≤ L ≤ 10D and 0.40 ≤ Id ≤ 0.80). The performance of the proposed
functions is investigated by two different diameter and density cases, i.e. D = 0.30 m;
Id = 0.50 and D = 0.40 m; Id = 0.70, of two different penetration lengths, i.e.
L = 6D and 8D. Since these cases were not included in the fitting process, they are
used as objective testing cases.
Using Equation 6.58, one can obtain the coefficients of the corresponding approx-
imations (Equations 6.49 - 6.54). Employing these formulas to get the approximate
installation effects around the wished-in-place pile and performing a numerical pile
load test, the accuracy of the approximations can be determined.
The FE meshes used in this section for D = 0.30 m and D = 0.40 m are given
in Figure 7.1. As can be seen from Figure 7.1 the FE mesh is different than the
ones used in PR simulations, i.e. there are no interfaces in the pile region. In
order to have a more objective comparison of the results, the reference load test was
performed on a state interpolated from the equalisation state obtained at the end of
the PR simulation. One load test was performed for the wished-in-place pile case, at
geostatic stress and density state. Finally for the formulas proposed in Section 6.4

133
134 7. Application

were tested after imposing the new state and obtaining equilibrium. The formulas
were employed in the whole mesh (excluding the clay region) in order to obtain a
smooth distribution of the imposed stresses. Finally, the pile was activated and after
having the equilibrium, the load test was performed.

Figure 7.1: FE mesh used for the verification cases a. D = 0.30 m and b. D = 0.40 m

Since the shear stress and the void ratio have lower accuracy compared to the
Cartesian normal stress approximations, two extra variations were also investigated.
In the first variation, instead of using the proposed approximation for the void
ratio, Bauers [21] formula, which follows the current mean stress, was employed (i.e.
σ̂rr (r, z); σ̂zz (r, z); σ̂θθ (r, z) and σ̂rz (r, z)). Therefore, in contrast to the proposed
void ratio approximation, a compacted region around the pile was obtained. As a
result, a stiffer and stronger response was obtained for these cases (Figure 7.2). The
second variation excluded the shear stress component, i.e. σrz = 0 for the full FE
domain. Besides, similar to the first variation, Bauer’s [21] formula was employed
for the void ratio.
Figures 7.2 and 7.3 show the numerical pile load test curves. In these figures the
predictive capability of the formulas are presented with the reference PR and K0 sim-
ulations. In all variations employing only σ̂rr (r, z), σ̂zz (r, z), σ̂θθ (r, z) and σ̂rz (r, z)
formulas and Bauer’s [21] formula for the void ratio, e resulted in stiffer and stronger
load displacement response. For the dense cases, almost equal responses have been
obtained for variations in which all of the proposed formulas (i.e. σ̂rr (r, z), σ̂zz (r, z),
σ̂θθ (r, z), σ̂rz (r, z) and ê(r, z)). The medium dense analyses have indicated slightly
stiffer and stronger load displacement curves for the cases using all approximations
compared to the second variation.
7.2. Verification Cases 135

Figure 7.2: Comparison of the numerical load test curves of the PR results (interpolated on the
reference mesh), the equalised approximate installation effects and the reference K0 results of the
verification cases a. D = 0.30 m; Id = 70; and b. D = 0.40 m; Id = 50, both having a pile length,
L = 6D

Figure 7.3: Comparison of the numerical load test curves of the PR results (interpolated on the
reference mesh), the equalised approximate installation effects and the reference K0 results of the
verification cases a. D = 0.30 m; Id = 70; and b. D = 0.40m; Id = 50, both having a pile length,
L = 8D.

F P R −F ψ̂
For the quantification of the results, percentage error (%err = t F P R t ) and the
t
improvement level, η can be calculated for the total reaction force obtained from
approximations, Ftψ̂ at any pile displacement, uy using the total reaction force of
the PR simulations, FtP R as the reference and the K0 simulations, FtK0 as the lowest
level using:

Ftψ̂ − FtK0
η= (7.1)
FtP R − FtK0

Using Equation 7.1, the improvement level of each variation at uy = 0.10D have
been found and presented in Table 7.1. In general a level of, η ≈ 25% has been
136 7. Application

obtained for all variations. High improvement levels, i.e. η ≈ 37 and 40%, have
been obtained for the dense cases (Id = 0.70). On the other hand, the deviation
from PR curves can be considered as error. The % error of each approximation is
also given Table 7.1. It can be seen that the PR curves could be estimated with an
error range of ≈ 35 − 46%. The difference is both due to approximation (regression)
and the reduction of imposed stresses by the unbalance.
Table 7.1: Summary of improvement levels of the verification cases at uy = 0.10D

Cases η at uy = 0.10D %err at uy = 0.10D


D(m) L/D Id ψ̂ 1 ψ̂ 2 ψ̂ 3 ψ̂ 1 ψ̂ 2 ψ̂ 3
0.30 6 0.70 0.2771 0.3710 0.2771 42.09 36.62 42.09
0.30 8 0.70 0.2809 0.4009 0.3091 42.51 35.42 40.84
0.40 6 0.50 0.2570 0.3129 0.3129 43.36 40.10 45.60
0.40 8 0.50 0.2319 0.2885 0.1860 43.27 40.08 45.85
1
σ̂rr (r, z); σ̂zz (r, z); σ̂θθ (r, z); σ̂rz (r, z) and ê(r, z)
2
σ̂rr (r, z); σ̂zz (r, z); σ̂θθ (r, z) and σ̂rz (r, z)
3
σ̂rr (r, z); σ̂zz (r, z) and σ̂θθ (r, z)

7.3 Validation Case - Centrifuge Pile Jacking in


Baskarp Sand
As summarised in Chapter 2, Dijkstra (2009) investigated the effect of pile jacking
on the stress and density change by a centrifuge model test. In this section, the
proposed formulas are employed and validated by the pile load tests of Dijkstra [56].
The pile is pre-embedded 7.175 m in Baskarp sand and further jacked by 7 m
on prototype scale. The centrifuge model setup and the FE model employed in the
simulations are given in Figure 7.4 a & b, respectively. In the physical test, the
model pile is closer to one boundary. It is expected that the behaviour of the model
pile is dominated by the closest boundary, therefore in the axisymmetric FE model, a
radial boundary of equal to this distance was used (R = 175 · Ng = 6125 mm). Since
the validation is based on the pile load test at final penetration level (L = 405 · Ng =
14175 mm), the proposed formulas were employed for this configuration. At the
physical model test it is expected to have friction at the side boundaries. In order
to represent this boundary condition (BC) a thin stiff column, which was fixed in
horizontal and vertical directions was placed. Frictional boundary was represented
by the interface placed between the soil and the stiff column. The peak friction
angle, φpeak of sand was assigned to the interface since a full mobilisation at the
boundary is not expected.
The hypoplastic model parameters of the Baskarp sand (Table 4.3) were adopted
from Anaraki [8] and Pham et al. [121], for the basic model and the intergranu-
lar strain parameters, respectively. The minimum and maximum void ratios were
emin = 0.548 and emax = 0.929 [56]. Three density variations, namely, e0 = 0.783
/ Id = 0.38 (loose), e0 = 0.709/Id = 0.58 (medium), and e0 = 0.637 / Id = 0.77
(dense) were analysed. Figures 7.5 - 7.7 compare the numerical load test results of
the states based on approximated and equalised installation effects with the centri-
fuge model tests for the dense, medium dense and loose cases, respectively. In order
7.3. Validation Case - Centrifuge Pile Jacking in Baskarp Sand 137

Figure 7.4: a. Physical model test setup [56] (all units are in mm; acceleration level, Ng = 35),
and b. FE model of the validation case

to determine the improvement level of the load displacement behaviour of the FE


model piles, numerical load tests were simulated starting from the corresponding K0
states. These results were included as reference.
Table 7.2: Summary of improvement levels of the validation cases at uy = 0.10D

Cases η at uy = 0.10D %err at uy = 0.10D


e0 n Id ψ̂ 1 ψ̂ 2 ψ̂ 3 ψ̂ 1 ψ̂ 2 ψ̂ 3
0.637 0.389 0.77 0.296 0.293 0.263 51.07 51.26 53.44
0.709 0.415 0.58 0.616 0.730 0.584 20.08. 14.13 21.76
0.783 0.439 0.38 0.734 0.755 0.573 12.74 11.74 20.41
1
σ̂rr (r, z); σ̂zz (r, z); σ̂θθ (r, z); σ̂rz (r, z) and ê(r, z)
2
σ̂rr (r, z); σ̂zz (r, z); σ̂θθ (r, z) and σ̂rz (r, z)
3
σ̂rr (r, z); σ̂zz (r, z) and σ̂θθ (r, z)

For all three cases the improvement levels have been summarised in Table 7.2.
The highest improvement level, i.e. η ≈ 75%, has been obtained for the loose case
(Id = 0.38). The improvement levels reduce with increasing density. Similar to the
discussion for the verification cases (7.2), but this time the deviation from the meas-
urements can be considered as error, since there are no reference PR simulations.
The % error of each approximation is also given in Table 7.2. It can be seen that the
measurements could be estimated with an error range of ≈ 12 − 53%. The difference
is again due to approximation (regression) and the reduction of imposed stresses
by the unbalance, as well as the uncertainties arising from the sample preparation
of the centrifuge test. The initial uy = 0.01D part of the centrifuge pile load test
curves are not accurate. During the centrifuge tests, Dijkstra [56] points out that
138 7. Application

the unloading of the pile head load after 7 m penetration does not guarantee a com-
plete relaxation. Due to these uncertainties, the initial stiffness of the numerical
simulations are not compared with the centrifuge test results.

Figure 7.5: Comparison of the numerical load test curves with the centrifuge test result [56] based
on the equalised approximate installation effects with the state obtained using formulas and the
reference K0 state of the dense case, e0 = 0.637; Id = 0.77.

Figure 7.6: Comparison of the numerical load test curves with the centrifuge test result [56] based
on the equalised approximate installation effects with the state obtained using formulas and the
reference K0 state of the medium dense case, e0 = 0.709; Id = 0.58.
7.3. Validation Case - Centrifuge Pile Jacking in Baskarp Sand 139

Figure 7.7: Comparison of the numerical load test curves with the centrifuge test result [56] based
on the equalised approximate installation effects with the state obtained using formulas and the
reference K0 state of the loose case, e0 = 0.783; Id = 0.38.

As mentioned earlier, the relatively low improvement amounts with respect to


the K0 situation, especially for the dense case, are mostly due to the reduction of
stress levels during the equalisation step after imposing the generalised functions
besides the approximation errors. It can be anticipated that the PR results would
give a stiffer response compared to the approximations (i.e. Figures 7.2 and 7.3).
One way to circumvent the reduction of imposed stresses is to introduce an up-
scale factor for the functions and obtain intended stress levels after equalisation.
Since the increase should not affect the areas outside the influence region, the up-
scaling factor, λ can be applied only on the exponential term (See Equations 7.2-7.6).
The factor can vary throughout the mesh or it can be simply estimated from the
global response, i.e. using 1/η. Then the equalised stress fields can be checked
and the scaling factor can be modified in an iterative scheme until the equalised
stress distributions are close enough to the stress levels imposed by the generalised
functions.
As a result of the increase in all stress components, the unbalanced forces are
also magnified. Hence the decrease of imposed stresses in equalisation phase is
pronounced. Besides, convergence issues have been arisen when factored stresses
were applied in the equalisation phase.

3 · K0 3
2 2
σ̂  rr = p0 · [ +λ· ai1 · e−(bi1 (r/D−ri1 ) +ci1 (z/D+L/D+zi1 ) ) ] (7.2)
1 + 2 · K0 i=1
140 7. Application

4 · K0 4
2 2
σ̂  zz = p0 · [ +λ· ai2 · e−(bi2 (r/D−ri2 ) +ci2 (z/D+L/D+zi2 ) ) ] (7.3)
1 + 2 · K0 i=1

3 · K0 3
2 2
σ̂  θθ = p0 ·[ +λ· ai3 · e−(bi3 (r/D−ri3 ) +ci3 (z/D+L/D+zi3 ) ) ] (7.4)
1 + 2 · K0 i=1


6
2
+ci4 (z/D+L/D+zi4 )2 )
σ̂rz = p0 · [λ · ai4 e−(bi4 (r/D−ri4 ) ] (7.5)
i=1


4
2
+ci5 (z/D+L/D+zi5 )2 )
ê = p0 · [1 + λ · ai5 e−(bi5 (r/D−ri5 ) ] (7.6)
i=1

The pile tip and shaft capacities (Qp and Qs at uy = 0.10D) obtained by the
FE simulations using proposed functions are also compared with the state of the
art empirical methods such as API [12], Dutch method [54], Fugro-05 [92], ICP-05
[88], NGI-05 [43], and UWA-05 [101], in Figure 7.9. The empirical methods uses qc
data to estimate the shaft and tip capacities. In order to provide qc data for these
methods, the correlation of Baldi et al. [16] was inverted by using Id as input:
 C1
qc = C0 · (σv0 ) · e(Id ·C2 ) (7.7)

where, C0 = 157, C1 = 0.55 and C2 = 2.41. Appendix C provides more information


about the calculation details.
In short, these methods differ in selecting representative qc value for tip, account-
ing for friction fatigue, accounting for pile diameter size and correlations relating
these parameters.
Table 7.3 summarizes the accuracy of predictions of total pile capacities for three
different densities for a pile head displacement of uy = 0.10D. Figure 7.8 shows
the trend of prediction accuracies of each method employed. As expected, the FE
calculations based on K0 constitute the lower bound followed by the conservative
empirical method, API-01. The loose case prediction of the ICP-05 is lower than the
API-05. The FE predictions based on the states obtained using proposed functions
and ICP-05 method were in the same order, i.e. underpredicting the total capacity
by ≈ 15 − 50% for loose to dense cases, respectively. The dense case was predicted
best by UWA-05 method with an error less than 8.1%. It is followed by Fugro-05
and NGI-05 methods with an error of ≈ 20%. The best prediction of the loose case
capacity was obtained by NGI-05, i.e. error less than 7.3%. Rest of the predictions
using empirical methods resulted in an overestimation of the total capacity between
13 − 74%, having an upper bound by the Dutch method.
7.3. Validation Case - Centrifuge Pile Jacking in Baskarp Sand 141

Figure 7.8: Comparison of total pile capacities

Figure 7.9: Comparison of pile a. base and b. shaft capacities

Table 7.4 summarizes the accuracy of predictions of pile base and shaft capacities
obtained using the FE simulations based on K0 state and states based on proposed
formulas, and the empirical methods for a pile head displacement of uy = 0.10D.
The base capacities were underpredicted for the dense case (Id = 0.77) by all
calculation methods. Nevertheless, Dutch method predicted the capacity within an
accuracy of ≈ 13% followed by UWA-05 method (≈ 22%). For the medium dense
(Id = 0.58) and loose (Id = 0.38) cases, Dutch, Fugro-05, and UWA-05 methods
overestimated the base capacity between ≈ 29 − 46%. The lowest base estimate
was obtained by the FE calculations based on K0 state, which had accuracy level
of ≈ 24 − 41%. FE calculations based on the state using proposed functions (ψ̂ 1 )
as well as the NGI-05 method estimated the measured capacity quite well (< 12%
error) for the medium dense and loose cases; with underprediction of the former and
142 7. Application

Table 7.3: Summary of pile total capacity prediction accuracies of proposed functions and empir-
ical methods

Total pile capacity


Method Qcalculated
t /Qmeasured
t
Id = 0.77 Id = 0.58 Id = 0.38
K0 0.2745 0.4769 0.5248
ψ̂ 1 0.4890 0.8000 0.8723
FE

ψ̂ 2 0.4870 0.8564 0.8794


ψ̂ 3 0.4649 0.7846 0.7943
API-05 [12] 0.4386 0.6291 0.7831
Empirical

Dutch Method [54] 1.1301 1.7404 1.4223


Fugro-05 [92] 0.8148 1.4606 1.4198
ICP-05 [88] 0.5657 0.8660 0.7031
NGI-05 [43] 0.8021 1.2738 0.9271
UWA-05 [101] 0.9189 1.4404 1.2092

Table 7.4: Summary of pile base and shaft capacity prediction accuracies of proposed functions
and empirical methods

Pile base capacity Pile shaft capacity


Method Qcalculated
p /Qmeasured
p Qcalculated
s /Qmeasured
s
Id = 0.77 Id = 0.58 Id = 0.38 Id = 0.77 Id = 0.58 Id = 0.38
K0 0.2432 0.3835 0.4125 0.3643 0.6774 0.6721
ψ̂ 1 0.5378 0.9098 1.1250 0.3488 0.5645 0.5410
FE

ψ̂ 2 0.5432 0.9699 1.1250 0.3256 0.6129 0.5574


ψ̂ 3 0.4676 0.7970 0.9625 0.4651 0.7581 0.5902
API-05 [12] 0.3649 0.4614 0.6904 0.6500 0.9889 0.9046
Empirical

Dutch Method [54] 0.8720 1.4598 1.4342 1.8706 2.3421 1.4068


Fugro-05 [92] 0.6311 1.3620 1.7406 1.3418 1.6721 0.9990
ICP-05 [88] 0.3939 0.6595 0.6479 1.0585 1.3090 0.7755
NGI-05 [43] 0.5004 0.9864 1.1099 1.6675 1.8903 0.6873
UWA-05 [101] 0.7848 1.3138 1.2908 1.3037 1.7118 1.1021

overprediction of the latter. For the loose case, FE calculations based on the state
using proposed function employing only normal stresses (ψ̂ 3 ) slightly underpredicted
the base capacity with an error less than 4%.
Contrary to the base capacities, a large band of overestimation was obtained
using empirical methods. For example, Dutch method predicted 2.5 times the meas-
ured total shaft capacity for the medium dense case. Best prediction of the shaft
capacity for medium dense case was obtained by API-01 (error ≈ 1.1%). Simil-
arly, Fugro-05 predicted the almost the measured shaft capacity of the loose case
(error < 0.1%). The loose case shaft capacity was overestimated by UWA-05 (er-
ror 10.2%) and underestimated by API-01, ICP-05 and NGI-05 methods by errors
of 9.5%, 22.5%, and 31.3%, respectively. The FE predictions based on the state
obtained using proposed functions (ψ̂ 1 and ψ̂ 2 ) predicted shaft capacities even less
than the calculations based on K0 state. However, the FE predictions based on the
state obtained using proposed function ψ̂ 3 predicted the dense and medium dense
cases slightly better than (i.e. ≈ 10% less error) the calculations based on K0 state.
API predicted the shaft capacities quite well for the medium dense and loose cases.
The main reason for shaft capacity accuracy trend is due to the tested pile being
an intermediate case between bored and jacked pile (i.e. 7 m of jacking of 7.175 m
pre-embedded model pile).
7.4. Discussion of Results and Conclusions 143

7.4 Discussion of Results and Conclusions


In the previous chapters the concept of installation effects, modelling and incorpor-
ating them with different techniques were presented. In this chapter the generalised
formulas proposed in 6.4 have been verified 7.1 and validated 7.2. Improvement
of stress and density state in all of the verification and validation cases have been
observed by the load displacement curves obtained from numerical pile load tests.
The level of improvement has been quantified by load displacement curves of the
reference load tests initiated at K0 stress and density states.
It has been shown that the proposed formulations given in Chapter 6 can be used
to impose the stress and void ratio changes around a wished-in-place pile to obtain
the installation effects approximately without simulating the installation process.
As a result of the imposed installation effects, different levels of improvement in
global load displacement behaviour of the pile can be obtained. A general trend of
underestimation of the PR results and measurements have been observed.
For the verification cases, the range improvement level with respect to K0 sim-
ulations, at 0.10 · D displacement is η = 0.18 − 0.40. The deviation from the PR
curves, in other words the percent error has a range of ≈ 35 − 46%. The difference is
both due to approximation (regression) and the reduction of imposed stresses by the
unbalance. The η value could be improved (or the error could be reduced) by using
the exponential formula proposed by Bauer [21] for the void ratio instead of the void
ratio approximations. Employing the formula for the void ratio resulted in compac-
tion around the pile. The improvement is therefore artificial due to the strength
and the stiffness increase by the compaction. However, dilation had been observed
in previous experimental studies. Therefore, the improvement in the global load-
displacement behaviour should be evaluated with caution. From this perspective, it
is suggested to employ all formulations, i.e. σ̂rr (r, z); σ̂zz (r, z); σ̂θθ (r, z); σ̂rz (r, z)
and ê(r, z), to obtain a reasonable stress and density state.
In the last part (Section 7.3), the centrifuge test of Dijkstra [56] was modelled
to validate the proposed formulations. After the equalisation phase, the hypoplastic
model with intergranular strain was employed for the load tests. One can assume
that there is enough time between the pile installation and loading under service
loads. Therefore, the equalised state can be considered as the new virgin state in
relation to the small strain behaviour. Hence, it is a reasonable approach to use the
intergranular strain in the numerical simulations after imposing and obtaining the
equalised installation effects.
There is no reference PR simulations of the validation case. Therefore, the er-
rors and improvement levels are calculated considering the FE simulations based
on K0 state and the centrifuge test results (measurements). In general the total
load capacities obtained from the centrifuge tests are underestimated except for the
loose case. However, an improvement level as high as η = 0.755, in other words an
error as low as 11.7%, is obtained for the loose case. The maximum deviation from
the measured response was obtained for the dense case (η = 0.296; err% = 53.4%)
employing ψ̂ 3 .
144 7. Application

Depending on the initial density and penetration lengths, different levels of re-
ductions on the imposed stress field can be observed. The reductions in the imposed
stresses are due to the unbalance introduced by the approximations. One can cir-
cumvent this issue by overshooting the level of the stress field determined by the
generalised functions such that the equalised state is approximately at the same
level as the stress field using the generalised functions. Nevertheless, increasing
the level of imposed stresses also magnified the unbalance, resulting in convergence
issues at the equalisation phase. A more rigorous way would be to find the equi-
librium shear stresses using force equilibrium equations and the proposed Cartesian
normal stresses. However, as discussed in Chapter 6, a closed form solution of the
shear stresses could not be obtained. The option of determining these stresses in
a numerical integration was discarded to preserve the closed form of the proposed
functions.
Finally, total pile capacities as well as the base and shaft capacities obtained from
the simulations using the proposed formulas and state-of-the-art empirical methods
such as API, Dutch Method, Fugro, ICP, NGI and UWA, were compared with the
centrifuge measurements. The FE predictions based on the states obtained using
proposed functions and ICP-05 method were in the same order, i.e. underpredicting
the total capacity by ≈ 15−50% for loose to dense cases, respectively. The prediction
of total pile capacity for the dense case was more accurate (i.e. less than ≈ 20%
error) using CPT based direct methods except ICP-05. These methods (except for
loose case prediction of NGI-05), on the other hand, overestimated the total capacity
by ≈ 21 − 74%.
The proposed formulas predicted the base capacities of the medium dense and
loose cases at reasonable accuracy (within 12.5% confidence interval) except for
the ψ̂ 3 prediction of the medium dense case (≈ 20% error). Using the proposed
formulas, only ≈ 46 − 53% of the measured base capacity could be obtained for the
dense case. The shaft capacities were underpredicted for all density variations using
the proposed formulas (≈ 33 − 61% of the measured shaft capacities, excluding ψ̂ 3
prediction). The shaft capacities were even less than the K0 and API simulations,
except for the ψ̂ 3 formulation (75.8% of the measured shaft capacity), in which the
shear and dilation around the shaft that result in lower stress and stiffness levels at
the shaft compared to the K0 state was excluded.
The numerical load tests even based on a K0 would differ from the measurements,
due to the limitation of constitutive model, the parameters of which are determined
for element tests. Although the model was tuned for the element tests, the accuracy
at which BVP’s such as the particular case (which involve complicated stress paths)
can be simulated, is not guaranteed. This is one of the reasons for the difference
between the results of the numerical simulations and the measurements, besides the
approximations and the reductions in imposed stresses due to unbalance.
Chapter 8

Conclusions

8.1 Conclusions
In the geotechnical engineering practice, the installation effects are mostly discarded
in the numerical analysis of displacement piles. The information on the installation
effects is essential, not only to make reliable predictions of the pile bearing capacity
and its behaviour in the soil under different conditions, but also to predict the effects
of pile installation on surrounding structures in the ground.
Pile installation processes and their effects on the surrounding soil have been
investigated by several researchers. An important aspect of the numerical modelling
of displacement piles is to analyse its service behaviour, i.e. load displacement
behaviour after the installation process. Previous studies mostly focused on the
installation process rather than the equalisation stage corresponding to the situation
after the removal of the force on the pile.
The main goal of this research is to assess the installation effects of displacement
piles in a numerical framework and describe these effects in a flexible geometric func-
tion such that their load-displacement behaviour in the numerical simulations can
be improved compared to the one obtained from geostatic state. The approximated
stress and density state can be imposed by a subroutine that uses these functions.
The flowchart of the algorithm is given in Appendix B.
The installation effects around one pile can be analysed by several numerical
techniques. In this thesis, a simplified numerical technique, which is called the
’Press - Replace’ technique, was introduced and employed to model penetration of a
jacked pile in sand. The PR technique involves a step-wise updated geometry, which
consists of a straining phase followed by a geometry update. The geometry update
is to model the penetrated part of the pile, which can be achieved by changing the
material properties at the beginning of each phase.
The PR technique was validated with more advanced large strain methods for
a cone penetration simulation in an undrained clay. The PR technique allows the
use of non-smooth geometry (e.g. sharp corners at the pile or cone tip or shoulder).
However, the results made clear that a concentration of stresses at these locations
could not be avoided. It was shown that the PR technique could be used in a stand-

145
146 8. Conclusions

ard FE program to model large deformation problems, like pile jacking, benefiting
from stability without losing much accuracy of the stress field. The model has a
limitation in the sense of not fully capturing the continuous velocity field near the
tip of the pile being pushed. Other numerical techniques like Eulerian, ALE, CEL
and MPM are more suitable to obtain the continuous velocity field around the pile
tip during penetration. However, these methods are not yet generally available or
common for geotechnical engineering applications.
During pile penetration, the surrounding soil undergoes significant changes in
stress level and density. A hypoplastic constitutive model was chosen to model
the behaviour of the soil around the pile due to its state dependent stiffness and
strength properties. It was concluded that the hypoplastic model published by
[158] provides a flexible formulation, by which a large range of density and stress
levels can be represented by one set of parameters. This model has been further
developed by several researchers and it was calibrated for low stress levels (< 250
kPa). At very high stress levels (> 5 - 10 MPa) that are commonly occur at the
pile tip during pile penetration, grain crushing becomes evident. The soil properties
change as a result of crushing. Therefore a modified version of the hypoplasticity
was introduced, in an attempt to take grain crushing and its consequences for the
soil stiffness into account. It was shown that the modified model could be applied
at single stress point level. However, the model had convergence issues during FE
simulations of boundary value problems. Therefore it was concluded that the model
needed a thorough investigation and further improvement in order to get a stable
and robust formulation, which was beyond the scope of this study. In their 2D
DEM model, [105] showed that the stresses at the pile tip and shaft decreased
during penetration. Furthermore, the density around the pile also decreased due
to crushed material. Therefore, as a result of discarding the grain crushing effects,
overshooting of mobilised stresses around the pile might be expected during the
penetration simulations using PR technique. It is difficult to estimate the behaviour
after equalisation stage. The study was continued with the original hypoplastic
model.
The PR technique was employed to investigate the pile behaviour during and after
the jacking phase and corresponding installation effects for various pile geometries
and soil densities. The results were validated by well-established design charts, in
which load-displacement behaviour of the pile base and shaft are given separately. A
general stiff trend was observed for the PR simulations. However, the design charts
present the lower bound of the field observations. The actual field measurement
curves are more stiffer. Therefore, it was concluded that the results can be used as
a quantitative assessment of the installation effects.
The results of analyses of the pile geometry variations were plotted by normal-
ising length scales. These results indicated similar installation effects (i.e. stress and
density change) for both of the pile diameter variations. The stress change around
the pile increased with increasing penetration length and soil density. In all vari-
ations dilation was found around the pile extending up to approximately a distance
of 4D for dense, 3.5D for medium dense, and 2D for loose case. As expected, the
stresses around the pile had a clear influence on the capacity and stiffness of the pile
8.1. Conclusions 147

since the stiffness of the hypoplastic model is calculated based on the current stress
and void ratio. In order to have an objective evaluation of the stress and the void
ratio change due to the pile installation, the Cartesian stress components and the
void ratio were normalised with the geostatic mean effective stress and void ratio,
respectively. The normalised results set basis for assessment and continuous geo-
metric description of the installation effects. This way, the installation effects can
be transferred and imposed to different FE models with different discretisations. As
a result an improved load displacement behaviour of the model pile compared to
the one obtained for the geostatic conditions can be achieved without simulating the
installation process.
The main purpose of this study is to develop a method to incorporate the in-
stallation effects around a wished-in-place pile to account for these effects without
simulating the whole penetration process. As a result, there would be a consider-
able reduction in computational demand, especially when there is more than one
pile (e.g. a pile group or a piled raft model). This is particularly efficient and useful
for practical applications.
The installation effects of various pile jacking cases (i.e. pile geometry, soil density
and soil types) were investigated in Chapter 5. The results of these cases are in
discrete form (i.e. values at stress points). To be able to transfer these results to
arbitrary FE meshes, it is necessary to represent these results as a continuous field
so that the values can be determined at any location (e.g. at the stress points of the
new FE mesh). The continuous fields can be represented by geometric functions.
One way of determining these fields is using interpolation functions.
Section 6.2 presented the interpolation of the calculated installation effects (stress
and density) to FE models having different meshes to check the applicability and
possible mesh dependency of incorporating the installation effects. After imposing
the installation effects it was necessary to perform a NIL step in which the unbal-
ance forces are removed to obtain an equalised stress state. A sensitivity analysis
performed to check the limitations of transferring the installation effects resulting
from a PR simulation to a different FE mesh showed that the mesh discretisation
had minor influence on the imposed and equalised states as well as the load displace-
ment behaviour of the pile. However, the coarsest mesh, in which one element below
the pile base and two elements on the shaft were used, showed much stiffer load dis-
placement behaviour and much higher capacity compared to the behaviour obtained
for the rest of the mesh variations. Interpolation from a fine mesh to a finer mesh
resulted in smoother distribution of transferred installation field variables. Besides,
smoother load test curves were obtained for the cases with interpolation to a finer
mesh than those to a coarser discretisation. The interpolation functions provide
quite good approximations of the surface formed by guiding stress points. On the
other hand, these functions require information of all stress points (i.e. location and
state).
In order to represent the installation effects in a more practical way, model func-
tions of Gaussian form for each variable (i.e. normalised stresses and void ratio)
were selected. The stresses and the void ratio were normalised with their corres-
ponding geostatic values. Furthermore, the length scales were normalised with pile
148 8. Conclusions

diameter, D. The normalised stress and void ratio surfaces were successfully repres-
ented by model functions of Gaussian form. By regression analyses the coefficients
of these model functions were determined. Once having these fits, less data, i.e.
only the fitting coefficients, are required to represent a continuous field of install-
ation effects. Hence, it was shown that the installation effects can be extended to
functional forms simpler than the interpolation functions. Nevertheless, due to the
approximation errors introduced in each fit, some unbalance was observed when the
separate fit surfaces were combined to form the stress tensor field.
In order to satisfy the equilibrium conditions in radial and vertical directions,
  
three major stress components (σrr , σzz , σθθ ) were assumed to successfully repres-
ent the stress field and the analytical solution satisfying both of the vertical and
horizontal equilibrium equations for the shear component (σrz ) was investigated.

However the integral of the ∂z (σrz ) term in the vertical equilibrium equation is
unknown. As the goal is to express the installation effects in geometric functional
form, the numerical integration was not employed to determine the shear stresses
from equilibrium equations using the geometric functions proposed for the major
stress components.
Another way of satisfying the equilibrium conditions is performing the regression
analysis to minimise the unbalance in horizontal and vertical directions. In other
words, employing the unbalance equations as model functions. However, a conver-
ging solution for these regression analyses could not be obtained. Therefore, the
unbalance was solved by a numerical NIL step, where the stresses were imposed and
equilibrium was satisfied within the FE analysis. It was shown that the equilibrium
state obtained after imposing the proposed functions found by separate regression
analyses by the NIL step results in a reduction of the imposed stresses leading to a
conservative load displacement behaviour.
Finally, the fitting results obtained in Section 6.3 were generalised by express-
ing the fitting coefficients of these functions in terms of the varied parameters, i.e.
pile diameter, D, penetration length, L and soil density, Id . As a result of second
order approximation (i.e. performing regression analyses on the coefficients of previ-
ous approximations that is an approximation of an approximation), a less accurate
approximation was obtained. Nevertheless, graphical inspection on the maximum
values and distribution of the stress and void ratio surfaces of the approximation and
the reference PR results showed that the approximations were qualitatively reason-
able. The load displacement curves support the accuracy of the generalised forms
indirectly.
The load-displacement curves obtained from numerical pile load tests were used
to evaluate the improvement due to the imposed stress and density states using
direct interpolation, approximated and generalised formulas. The quantification of
the improvement level was performed using the load tests curves of reference PR
simulation and the geostatic state. It was shown that the FE calculations employ-
ing the direct interpolation of PR simulation results gave nearly 100% improvement
(η = 1.0). The level of improvement reduced from a range of minimum η = 0.50 and
maximum η = 0.85 obtained by the approximate functions (fitting each geometry
and density variation separately) to minimum η = 0.25 and maximum η = 0.55 ob-
8.1. Conclusions 149

tained by employing the generalised functions for loose and dense cases, respectively.
The verification cases were based on the PR simulations of geometry and density
variations, i.e. Id = 0.50 & 0.70 and L = 6D & 8D, the results of which were
not considered in the generalisation of geometric functions. For these testing cases,
the range of improvement level at 0.10D displacement was minimum η = 0.18 and
maximum η = 0.40 for the cases with density indices of Id = 0.50 and Id = 0.70,
respectively. The deviation from PR curves can be considered as error. An error
range of ≈ 35 − 46% was obtained. The difference is both due to approximation
(regression) and the reduction of imposed stresses by the unbalance.
It was possible to increase the improvement level, η by using the exponential
formula proposed by [21] for the void ratio instead of the void ratio approximations.
However, the improvement was based on the increase in stress and strength of the
sand around the pile as a result of the compaction obtained by substituting the
imposed stresses in Bauer’s equation [21]. However, besides the results of the PR
simulations, dilation was observed around a jacked pile also in previous experimental
studies. Therefore, the improvement in the global load-displacement behaviour by
using the Bauer s formula instead of the proposed exponential form for the void
ratio, should be evaluated with caution.
The centrifuge test results of [56] were used for validation. Using the intergranu-
lar strain during load tests improves the load displacement curves further. One can
assume that there is enough time after the pile installation and loading under service
loads. Therefore, the equalised state can be considered as the new virgin state in
relation to the small strain behaviour. Hence, the intergranular strain could be used
in the numerical simulations after imposing and obtaining the equalised installation
effects.
There is no reference PR simulations of the validation case. Therefore, the errors
and improvement levels are calculated considering the FE simulations based on K0
state and the centrifuge test results (measurements). The centrifuge load test result
was assumed as η = 1.0. Based on this interpretation, the range of improvement
level at 0.10D displacement was η = 0.26 & η = 0.75 for the dense and loose cases,
respectively. The improvement levels reduce with increasing density. The measure-
ments could be estimated with and error range of 11.7 − 53.4%. The difference is
again due to approximation (regression) and the reduction of imposed stresses by
the unbalance, as well as the uncertainties in the measurements arising from the
sample preparation of the centrifuge test.
The total pile capacities (Qt ) as well as the base and shaft capacities (Qp and Qs
at uy = 0.10D) obtained by the FE simulations using proposed functions were also
compared with the state of the art empirical methods such as API [12], Dutch method
[54], Fugro-05 [92], ICP-05 [88], NGI-05 [43], and UWA-05 [101]. The FE predictions
based on the states obtained using proposed functions and ICP-05 method were in
the same order, i.e. underpredicting the total capacity by ≈ 15 − 50% for loose to
dense cases, respectively. The prediction of total pile capacity for the dense case
was more accurate (i.e. less than ≈ 20% error) using CPT based empirical methods
except ICP-05. These methods (except for loose case prediction of NGI-05), on the
other hand, overestimated the total capacity by ≈ 21 − 74%.
150 8. Conclusions

The proposed formulas predicted the base capacities of the medium dense and
loose cases at reasonable accuracy (within 12.5% confidence interval) except for
the ψ̂ 3 prediction of the medium dense case (≈ 20% error). Using the proposed
formulas, only ≈ 46 − 53% of the measured base capacity could be obtained for the
dense case. The shaft capacities were underpredicted for all density variations using
the proposed formulas (≈ 33 − 61% of the measured shaft capacities, excluding ψ̂ 3
prediction). The shaft capacities were even less than the K0 and API simulations,
except for the ψ̂ 3 formulation (75.8% of the measured shaft capacity), in which the
shear and dilation around the shaft that result in lower stress and stiffness levels at
the shaft compared to the K0 state was excluded.
The numerical load tests even based on a K0 would differ from the measurements,
due to the limitation of constitutive model, the parameters of which are determined
for element tests. Although the model was tuned for the element tests, the accuracy
at which BVP’s such as the particular case (which involve complicated stress paths)
can be simulated, is not guaranteed. This is one of the reasons for the difference
between the results of the numerical simulations and the measurements, besides the
approximations and the reductions in imposed stresses due to unbalance.
It can be concluded that using the proposed formulations given in Chapter 6, one
can impose an approximate installation effects field around a wished-in-place pile,
and can obtain reasonable pile base capacities (estimation within 20%) for relative
densities, Id < 0.70. Conservative shaft capacities (estimation 50 − 75%) can be
obtained for relative densities, Id < 0.70.

8.2 Outlook for Further Research


It is clear that the method of incorporating the installation effects by generalised
geometric functions is an important step towards a more tangible representation and
incorporation of these effects in a virgin mesh without simulating the penetration
process.
The importance of incorporating the crushing effects in the constitutive model
was shown and some modifications were employed on the hypoplastic model to ac-
count for grain crushing. The modified model requires extra input parameters, some
of which are difficult to determine. The model can be improved by:

ˆ Using more physical input parameters,

ˆ Increasing stability by using a global predictor-corrector scheme with a suf-


ficient stiff elastic stiffness matrix (i.e. initial stiffness method) that is only
updated when required to avoid non-converging behaviour [89]),

ˆ incorporating the changes in all hypoplastic parameters (i.e. hs , n, α, β, χ)


due to grain crushing.

It should be noted that the constitutive models have limitations for complex
stress/strain paths observed in BVP’s, such as pile penetration. These limitations
8.2. Outlook for Further Research 151

can be investigated by modelling the stress/strain paths observed in physical test


utilizing photoelastic measurements (e.g. [6] and [57]).
In this study, it was shown that the representation of each Cartesian stress sep-
arately might result in force equilibrium issues. Selecting model functions that
∂  ∂ ∂
are integrable considering the ∂r (rσrr ), ∂z (σrz ) and ∂r (rσrz ) terms would enable
resolving the equilibrium issues by satisfying equilibrium equations intrinsically, i.e.
irrespective of the coefficients.
Recent simulations have shown that the reduction of imposed stresses during the
NIL step can be minimised by using stiff material (i.e. having a stiffness 100 times
of the original hs value) when the functions are employed. As a consequence, the
pile load-settlement curve will improve (Figure 8.1).

Figure 8.1: Comparison of the numerical load test curves using formulas employed with using
original stiffness (hs ) and high stiffness (100 · hs ) for the medium dense case, e0 = 0.709;Id = 0.58.

The range of variations in pile penetration length, diameter and soil density can
be extended to make more general geometric functions to describe the installation
effects. Furthermore, the variables used to describe the installation effects, i.e. D,
L and Id , can be increased to account for other effects, i.e. set-up, ageing. The
representation of installation effects can be employed for other materials such as silt
and clay.
To the authors knowledge, there is no physical test by which the installation
effects, i.e. stress and density field is obtained as a continuous field with a reliable
level of accuracy. If the installation effects can be quantified by a field test or
a physical test at a reasonable accuracy, the functions proposed should then be
updated accordingly.
152 8. Conclusions

It was shown that the level of improvement decreased by each approximation.


However, the PR results can be imposed around wished-in-place piles to repres-
ent the installation effects with high accuracy (η ≈ 1.0). The interaction between
piles during the installation of piles is not considered in the PR simulations. Nev-
ertheless, the methods proposed in this thesis can be a good starting point to take
the installation effects for the numerical analysis of pile groups by superposition of
the installation effects. If more than one displacement pile will be modelled, the
interaction between the piles during installation should be investigated for an ac-
curate representation of the installation effects for pile groups, rather than simply
superimposing the installation effects obtained for a single pile.
The interaction between the newly installed pile and the existing piles can be
analysed by simulating penetration processes of both piles at a cost of large compu-
tational demand. Alternatively, rather than simulating the penetration of all piles,
the geometric functions can be employed to account for the installation effects of the
first pile installed and as a result the computational effort can be reduced. Based on
the investigation of interaction effects, the proposed formulations can be then exten-
ded to account for these effects and can then be used to represent the installation
effects of pile groups.
Another foreseen benefit of incorporating the installation effects using the func-
tions is the possibility of modelling the displacement piles using embedded piles [60].
The embedded piles are beam elements enhanced with proper interaction with the
continuum. These elements increase the calculation efficiency. At present, the em-
bedded piles do not incorporate the installation effects and only non-displacement
piles can be modelled. Use of the functions will therefore improve the embedded pile
behaviour to a more realistic level and the FE simulations would be beneficial to
the calculation efficiency. As mentioned in the previous paragraph, if more than one
displacement pile will be modelled, the interaction between the piles during install-
ation should be investigated for an accurate representation of the installation effects
for pile groups, rather than simply superimposing the installation effects obtained
for a single pile.
Bibliography

[1] Règles techniques de conception et de calcul des fondations des ouvrages de


gènie civil, 1993.

[2] NBN EN 1997-1:2005. Eurocode 7: Geotechnisch ontwerp - deel 1: Algemene


regels, 2005.

[3] NEN 9997-2010. Geotechniek - berekeningsmethode voor funderingen op palen


- drukpalen, 2010.
[4] M. Abu-Farsakh, M. Tumay, and G. Voyiadjis. Numerical parametric study
of.piezocone penetration test in clays. International Journal of Geomechanics,
3(2):170–181, 2003.

[5] Y. B. Acar and M. T. Tumay. Strain field around cones in steady state pen-
etration. Journal of Geotechnical Engineering Division, 112:207–213, 1986.

[6] H. G. B. Allersma. Determination of the stress distribution in assemblies of


photoelastic particles. Experimental Mechanics, 9:336–341, 1982.

[7] A. Altaee, B. H. Fellenius, and Evgin E. xial load transfer for piles in sand. i.
tests on an instrumented precast pile. Canadian Geotechnical Journal, 29:11–
20, 1992.

[8] K. E. Anaraki. Hypoplasticity investigated parameter determination and nu-


merical simulation. Master’s thesis, Delft University of Technology, The Neth-
erlands, 2008.

[9] K. H. Andersen, L. Andresen, H. P. Jostad, and E. C. Clukey. Effect of skirt-


tip geometry on set-up outside suction anchors in soft clay. In Proc. 23rd
Int. Conf. on Offshore Mechanics and Arctic Engineering, volume 1, pages
1035–1044, Vancouver, Canada, 2004.

[10] F. J. Anscombe. Graphs in statistical analysis. The American Statistican,


27(1):17–21, 1973.

[11] N. Aoki and D. A. Velloso. An approximate method to estimate the bearing


capacity of piles. In Proc. 5th Pan-American Conference of Soil Mechanics
and Foundation Engineering, pages 367–376, 1975.

153
154 BIBLIOGRAPHY

[12] API. Geotechnical and Foundation Design Considerations. Americal Petro-


leum Institute, Washington D.C., 2011.

[13] F. Aurenhammer and R. Klein. Voronoi diagrams. Technical report, FernUni-


versitt Hagen, Department of Computer Science, 1996.

[14] G. Axelsson. Long-term increase in shaft capacity of non-cohesive soils. In


Proc. 4th Int. Conf. on Case Histories in Geotech. Engng., 1998.

[15] G. Axelsson. Long-term set-up of driven piles in sand-Basic mechanisms and


methods of prediction. PhD thesis, Royal Institute of Technology, 2000.

[16] G. Baldi, R. Bellotti, V. Ghionna, M. Jamiolkowski, and E. Pasqualini. Inter-


pretation of cpts and cptus; 2nd part: drained penetration of sands. In 4th
International Geotechnical Seminar, pages 143–156, Singapore, 1998.

[17] M. M. Baligh. Theory of deep static cone penetration resistance. Technical


report, MIT Dept. of Civil Eng., 1975.
[18] M. M. Baligh. Strain path method. Journal of Geotechnical Engineering,
111(9), 1985.

[19] M. M. Baligh. Undrained deep penetration i:shear stresses. Géotechnique,


36(4):471–486, 1986.

[20] K. J. Bathe, E. Ramm, and Wilson E. L. Finite element formulations for large
deformation dynamic analysis. Int. J. Num. Meth. Engng., 9:353–386, 1975.

[21] E. Bauer. Calibration of a comprehensive hypoplastic model for granular ma-


terials. Soils and Foundations, 36:13–26, 1996.
[22] E. Bauer and I. Herle. Stationary states in hypoplasticity. Constitutive Mod-
elling of Granular Materials, pages 167–192, 2000.
[23] K. Been and M. G. Jefferies. A state parameter for sands. Géotechnique,
35(2):99–112, 1985.

[24] T. Belytschko, W. K. Liu, and B. Moran. Nonlinear finite elements for con-
tinua and structures. Wiley, Chichester ; New York, 2000.

[25] T. Benz. Small-strain stiffness of soils and its numerical consequences. PhD
thesis, Universitat Stuttgart, 2007.

[26] L. Beuth, T. Benz, and P. A. Vermeer. Formulation and validation of a quasi-


static material point method, 2007.

[27] J. Bigler, J. Giulkey, C. Gribble, C. Hansen, and S. Parker. A case study:


Visualizing material point method data. Stress The International Journal on
the Biology of Stress, pages 299–306, 2006.
BIBLIOGRAPHY 155

[28] R. F. Bishop, R. Hill, and N. F. Mott. The theory of indentation hardness


tests. Proc. Phys. Soc., 57:147–159, 1945.

[29] M. Boulon and P. Foray. Physical and numerical simulation of lateral shaft
friction along offshore piles in sand. In 3rd International Conference on Nu-
merical Methods in Offshore Piling, pages 127–147, 1986.

[30] N.G. Bourago and V.N. Kukudzhanov. A review of contact algorithms. Izv.
RAS, Mechanics of Solids, (1):44–85, 2005.

[31] E. T. Bowman and K. Soga. Mechanisms of setup of displacement piles in sand:


laboratory creep tests. Canadian Geotechnical Journal, 42(5):1391–1407, 2005.
[32] J. Brinch-Hansen. Discussion on hyperbolic stress-strain response. cohes-
ive soils. ASCE, Journal of Soil Mechanics and Foundation Engineering,
97(6):931–932, 1963.
[33] J. Brinch-Hansen. A revised and extended formula for bearing capacity. Danish
Geotechnical Institute, 1970.
[34] W. Broere and A.F. van Tol. Modelling the bearing capacity of displacement
piles in sand. Proceedings of the ICE - Geotechnical Engineering, 159(3):195–
206, 2006.

[35] H. J. Burd and G. T. Houlsby. Finite element analysis of two cylindrical


expansion problems involving near incompressible material behaviour. Inter-
national Journal Numerical and Analytical Methods in Geomechanics, 14:351–
366, 1990.

[36] M. Bustamante and L. Gianeselli. Pile bearing capacity by means of static pen-
etrometer. In In Proceedings of the 2nd European Symposium on Penetration
Testing, pages 493–500, 1982.
[37] J. P. Carter, J. R. Booker, and S.K. Yeung. Cavity expansion in cohesive
frictional soils. Géotechnique, 36(3):349–353, 1986.

[38] M. Chong. Density changes of sand on cone penetration resistance. In In


proceedings of the first international symposium on penetration testing ISOPT-
1, pages 707–714, 1988.

[39] F. C. Chow. Investigations into the behaviour of displacement piles for offshore
foundations. PhD thesis, University of London, 1996.

[40] F. C. Chow and Jardine R. J. Investigations into the behaviour of displacement


piles for offshore foundations. Ground engineering, 29:24–25, 1996.

[41] F. C. Chow, Jardine R. J., F. Brucy, and Nauroy J. F. Effects of time on


capacity of pipe piles in dense marine sand. Journal of Geotechnical and
Geoenvironmental Engineering. ASCE, 124(3):254–264, 1998.
156 BIBLIOGRAPHY

[42] A. Cividini and G. Gioda. A simplified analysis of pile penetration. In


Swoboda, editor, 6th International Conference on Numerical Methods in Geo-
mechanics, pages 1043–1049. Rotterdam:Balkema, 1988.

[43] C.J.F. Clausen, P.M. Aas, and K. Karlsrud. Bearing capacity of driven piles
in sand, the ngi approach. In 1st International Symposium on Frontiers in
Offshore Geotechnics, pages 677–681, Perth, Australia, 2005.

[44] I. F. Collins, Pender M. J., and Wang Y. Cavity expansion in sands under
drained loading conditions. International Journal for Numerical and Analytical
Methods in Geomechanics, 16(1):3–23, 1992.
[45] A. Colombi. Physical modeling of an isolated pile in coarse grained soils. PhD
thesis, University of Ferrara, 2005.

[46] M. R. Coop, E. U. Klotz, and Clinton L. The influence of the in situ


state of sands on the load-deflection behaviour of driven piles. Géotechnique,
55(10):721–730, 2005.
[47] R. Cudmani and V. A. Osinov. The cavity expansion problem for the inter-
pretation of cone penetration and pressuremeter tests. Canadian Geotechnical
Journal, 38(3):622–638, 2001.

[48] R.O. Cudmani. Statische, alternierende und dynamische Penetration in nicht-


bindigen Bden. PhD thesis, Universitat Fridericiana in Karlsruhe, 2001.

[49] J. L. Davidson and A. Boghrat. Displacements and strains around probes in


sand. In S.G. Wright, editor, In proceedings of the Conference on Geotechnical
Practice in Offshore Engineering, pages 181–202, 1983.
[50] M. T. Davisson. High capacity piles. In Proc. Lecture Series of Innovations
in Foundation Construnction, ASCE, Illinois Section, pages 81–112, 1972.
[51] E. E de Beer. The scale effect in the trans- portation of the results of deep-
sounding tests on the ultimate bearing capacity of piles and caisson founda-
tions. Géotechnique, 13(1):503–513, 1963.
[52] E. E de Beer. Different behavior of bored and driven piles. In Proc. of VI
Conf. on Soil Mechanics and Foundation Engineering, pages 307–318, 1984.

[53] R. de Borst and P. A. Vermeer. Finite element analysis of static penetration


test. In Proceedings of 2nd ESOPT, volume 2, pages 457–462, 1982.

[54] J. de Ruiter and F. L. Beringen. Pile foundations for large north sea structures.
Mar. Geotech., 3(3):267–14, 1979.

[55] A. D. Deeks, D. J. White, and M. D. Bolton. A comparison of jacked, driven


and bored piles in sand. In Proceedings of the 16th International Conference
on Soil Mechanics and Geotechnical Engineering, Vols 1-5, pages 2103–2106,
2005.
BIBLIOGRAPHY 157

[56] J. Dijkstra. On the Modelling of Pile Installation. PhD thesis, Delft University
of Technology, 2009.

[57] J. Dijkstra, W. Broere, and A. F. van Tol. Experimental investigation into the
stress and strain development around displacement pile, 2006.

[58] I. Einav and M. F. Randolph. Combining upper bound and strain path
methods for evaluating penetration resistance. Int.J. Num. Meth. Engng.,
63(14):1991–2016, 2005.

[59] H. K. Engin and L. Andresen. Comparison of zipper type techniques for finite
element analysis of pile penetration problem. In The 13th Int. Conf. of the
Int. Assoc. for Computer Methods and Advances in Geomechanics, volume 1,
pages 48–53, 2011.

[60] H. K. Engin, E. G. Septanika, and R. B. J. Brinkgreve. Improved embedded


beam elements for the modelling of piles. In G. Pande and S. Pietruszczak, ed-
itors, Numerical Models in Geomechanics: Numog X, pages 475–480, Rhodes,
Greece, 2007. Taylor & Francis Group.
[61] B. H. Fellenius. Determining the true distribution of load in piles. In Interna-
tional Deep Foundation Congress. ASCE, volume 2, pages 1455–1470, 2001.

[62] V. Fioravante, M. Jamiolkowski, and S. Pedroni. Modelling the behaviour of


piles in sand subjected to axial load. In Centrifuge 94, pages 455–460, 1994.

[63] K. Fischer and D. Sheng. Different aspects of large deformation contact formu-
lations applied to pile-soil fe-analysis. In Numerical Methods in Geotechnical
Engineering, pages 613–618. Taylor & Francis, 2006.
[64] W. G. K. Fleming. A new method for single pile settlement prediction and
analysis. Géotechnique, 42(3):411–425, 1992.
[65] R. Frank. Calcul des fondations superficielles et profondes. Presses de l’École
Nationale des Ponts et Chaussées, Paris, 1999.

[66] E. Franke. Co-report to discussion, session 13, on large diameter piles. In


Proc. 12th Int. Conf. Soil Mech. Found. Engrg., 1989.

[67] E. Franke. Measurements beneath piled rafts. In Keynote lecture to the ENPC
- Conference on Deep Foundations, pages 1–28, 1991.

[68] E. Franke. Design of bored piles, including negative skin friction and horizontal
loading. In Proc., 2nd Int. Geotechnical Seminar on Deep Foundations on
Bored and Auger Piles, pages 43–57, 1993.

[69] Jr. Gallagher, N. and G. Wise. A theoretical analysis of the properties of


median filters. Acoustics, Speech and Signal Processing, IEEE Transactions
on, 29(6):1136–1141, 1981.
158 BIBLIOGRAPHY

[70] R. E. Gibson and W. F. Anderson. In situ measurement of soil properties with


the pressuremeter. Civil Engineering and Public Works Review, 56(658):615–
618, 1961.

[71] D. R. Gill and B. M. Lehane. Extending the strain path method analogy for
modelling penetrometer installation. International Journal for Numerical and
Analytical Methods in Geomechanics, 24(5):477–489, 2000.

[72] D. V. Griffiths. Computation of bearing capacity factors using finite elements.


Géotechnique, 32(3):195–202, 1982.

[73] G. Gudehus. Code codicil. Ground Engineering, 29(7):8–8, 1996.


[74] G. Gudehus. A comprehensive constitutive equation for granular materials.
Journal of the Japanese Geotechnical Society : Soils and Foundation, 36:1–12,
1996.
[75] E. P. Heerema. Predicting pile driveability: Heather as an illustration of the
”friction fatigue” theory. In Proc. SPE European Petroleum Conference, 1978.
[76] S. Henke. Influence of pile installation on adjacent structures. International
Journal for Numerical and Analytical Methods in Geomechanics, 34(11):1191–
1210, 2010.

[77] S. Henke and J. Grabe. Simulation of pile driving by 3-dimensional finite


element analysis. In 17th European Young Geotechnical Engineers Conference,
pages 215–233, 2006.

[78] S. Henke and J. Grabe. Numerical modeling of pile installation. In 17th Inter-
national Conference on Soil Mechanics and Geotechnical Engineering (17IC-
SMGE), pages 1321–1324, 2009.

[79] I. Herle and G. Gudehus. Determination of parameters of a hypoplastic con-


stitutive model from properties of grain assemblies. Mechanics of Cohesive-
frictional Materials, 4(5):461–486, 1999.

[80] A. Hettler. Predicting pile driveability: heather as an illustration of the friction


fatigue theory. Ground Engineering, 13:15–37, 1980.

[81] H.D. Hibbit, P.V. Marcal, and J.R. Rice. A fe formulation for problems of large
strain and large displacement. International Journal of Solids and Structures,
6:1069–1086, 1970.

[82] R. Hill. The Mathematical Theory of Plasticity. Oxford University Press,


London, 1950.

[83] A. E. Holeyman. Soil behavior under vibratory driving. In Proceedings of the


1st International Conference on Vibratory Pile Driving and Deep Soil Com-
paction, pages 3–19, 2002.
BIBLIOGRAPHY 159

[84] T. Hughes and T. Belytschko. A Course of Nonlinear Finite Element Analysis.


Paris, 2008.
[85] White D. J. and M. D. Bolton. Comparing cpt and pile base resistance in
sand. In nstitution of Civil Engineers, Geotechnical Engineering, pages 3–14,
2005.
[86] Ramesh Jain, Rangachar Kasturi, and Brian G. Schunck. Machine Vision.
McGraw-Hill, 1995.
[87] R. J. Jardine and F. C. Chow. New design methods for offshore piles. Technical
report, Center for Petroleum and Marine Technology Department, 1996.
[88] R. J. Jardine, F. C. Chow, R. F. Overy, and J. R. Standing. Icp design methods
for driven piles in sands and clays, 2005.
[89] H.P. Jostad. Private communication, 2013.
[90] E. U. Klotz and M. R. Coop. An investigation of the effect of soil state on the
capacity of driven piles in sands. Géotechnique, 51(9):733–751, 2001.
[91] T. Kobayashi and R. Fukagawa. haracterization of deformation process of
cpt using x-ray tv imaging technique. In In 3rd International Conference on
Deformation Characteristics of Geomaterials, pages 43–47, 2003.
[92] H. J. Kolk, A. E. Baaijens, and M. Senders. Design criteria for pipe piles
in silica sands. In Proc., Int. Symp. on Frontiers in Offshore Geomechanics,
ISFOG, page 711716. Taylor & Francis, London, 2005.
[93] D. Kolymbas. A rate-dependent constitutive equation for soils. In Mech. Res.
Comm., volume 4, pages 367–372, 1977.
[94] D. Kolymbas. A novel constitutive law for soils. In 2nd Int. Conf. on Con-
stitutive Laws for Engineering Methods, volume 1, pages 319–326, 1987.
[95] D. Kolymbas. An outline of hypoplasticity. Archive of Applied Mechanics,
(61):143–151, 1991.
[96] F. H. Kulhawy. Limiting tip and side resistance: Fact or fallacy? In Analysis
and Design of Pile Foundations, ASCE, pages 80–98, 1984.
[97] K. Kuwajima, M. Hyodo, and A. F. L. Hyde. Pile bearing capacity factors and
soil crushabiity. Journal of Geotechnical and Geoenvironmental Engineering,
135(7):901–913, 2009.
[98] B. Lehane. Experimental investigations of pile behaviour using instrumented
field piles. PhD thesis, University of London, 1992.
[99] B. M. Lehane, R. J. Jardine, A. J. Bond, and R. Frank. Mechanisms of shaft
friction in sand from instrumented pile tests. J. Geotech. Engng Div. Am. Soc.
Civ. Engrs, 119(1):19–35, 1993.
160 BIBLIOGRAPHY

[100] B. M. Lehane, J. A. Schneider, and X. Xu. Evaluation of design methods for


displacement piles in sand. Technical report, UWA, 2005.
[101] B. M. Lehane, J. A. Schneider, and X. Xu. The uwa-05 method for prediction of
axial capacity of driven piles in sand. In International Symposium on Frontiers
in Offshore Geotechnics, pages 683–689, 2005.
[102] B. M. Lehane and D. J. White. Lateral stress changes and shaft friction for
model displacement piles in sand. Canadian Geotechnical Journal, 42(4):1039–
1052, 2005.
[103] W. Liu. Axisymmetric centrifuge modelling of deep penetration in sand. PhD
thesis, University of Nottingham, 2010.
[104] S. Lobo-Guerrero and L.E. Vallejo. Dem analysis of crushing around driven
piles in granular materials. Géotechnique, 55(8):617–623, 2005.
[105] S. Lobo-Guerrero and L.E. Vallejo. Influence of pile shape and pile interaction
on the crushable behavior of granular materials around driven piles: Dem
analyses. Granular Matter, 9(3-4):241–250, 2007.
[106] Q. Lu, M. F. Randolph, Y. Hu, and I.C. Bugarski. A numerical study of cone
penetration in clay. Géotechnique, 54(4):257–267, 2004.
[107] H. J. Luger, P. Lubking, and J. D. Nieuwenhuis. Aspects of penetrometer tests
in clay. In Proc. ESOPT II, pages 683–687, 1982.
[108] T. Lunne, P.K. Robertson, and J.J.M. Powell. Cone Penetration Testing in
Geotechnical Practice. Taylor & Francis Group, 1997.
[109] M. T. Luong and A. Touati. Sols grenus sous fortes contraintes. Revue Fran-
caise de Géotechnique, 24:51–63, 1983.
[110] Konrad J. M. Sand state from cone penetrometer tests. a framework consid-
ering grain crushing stress. Géotechnique, 48(2):201–216, 1998.
[111] S.P.G. Madabhushi and S.K. Haigh. Effect of superstructure stiffness on
liquefaction-induced failure mechanisms. International Journal of Geotech-
nical Earthquake Engineering, 1:70–87, 2009.
[112] K. P. Mahutka, F. König, and J. Grabe. Numerical modelling of pile jacking,
driving and vibratory driving. In Proceedings of International Conference on
Numerical Simulations of Construction Processes in Geotechnical Engineering
for Urban Environment (NSC06), pages 235–246, 2006.
[113] T. Marcher, P.A. Vermeer, and P.A. von Wolffersdorff. Hypoplastic and
elastoplastic modelling - a comparison with test data. In D. Kolymbas, editor,
Constitutive Modelling of Granular Materials, pages 353–374. Springer, 2000.
[114] D. Masin. Plaxis implementation of hypoplasticity. Technical report, Plaxis
bv, 2010.
BIBLIOGRAPHY 161

[115] G. G. Meyerhof. Some recent research on the bearing capacity of foundations.


Canadian Geotechnical Journal, 1:16–31, 1963.

[116] G. G. Meyerhof. Bearing capacity and settlement of pile foundations. ASCE


Journal of Geotechnical Engineering, 102(3):197–228, 1976.

[117] G. G. Meyerhof. Scale effects of ultimate pile capacity. ASCE Journal of


Geotechnical Engineering, 109(6):979–806, 1983.

[118] C. Mueller-Blenkle, P. K. McGregor, A.B. Gill, M. H. Andersson, J. Metcalfe,


V. Bendall, P. Sigray, D. Wood, and F. Thomsen. Effects of pile-driving noise
on the behaviour of marine fish. Technical report, COWRIE, 2010.
[119] Y. Nakata, M. Hyodo, and A.F.L. Hyde. Microscopic particle crushing of sand
subjected to high pressure one-dimensional compression. Soils and Founda-
tions, 41(1):69–83, 2001.
[120] A. Niemunis and I. Herle. Hypoplastic model for cohesionless soils with elastic
strain range. Mechanics of Cohesive-Frictional Materials, 2:279–299, 1997.
[121] H. D. Pham, H. K. Engin, R. B. J. Brinkgreve, and A. F. van Tol. Mod-
elling of installation effects of driven piles using hypoplasticity. In T. Benz
and S. Nordal, editors, Numerical Method in Geotechnical Engineering, Nu-
merical Methods in Geotechnical Engineering 2010 : Proceedings of the Sev-
enth European Conference on Numerical Methods in Geotechnical Engineer-
ing, pages 261–266. Taylor & Francis, 2010.

[122] D.M. Potts and L. Zdravković. Finite element analysis in geotechnical engin-
eering - Theory. Thomas Telford, London, 1999.
[123] H. G. Poulos and E. H. Davis. Pile foundation analysis and design. Wiley,
New York, 1980.
[124] G. Qiu, S. Henke, and J. Grabe. Application of a coupled eulerian-lagrangian
approach on geomechanical problems involving large deformation. Computers
and Geotechnics, 38(1):30–39, 2011.
[125] M. F. Randolph. 43rd rankine lecture: Science and empiricism in pile found-
ation design. Géotechnique, 53(10):847–875, 2003.

[126] L. C. Reese and M. W. O’Neill. New design method for drilled shaft from com-
mon soil and rock test. In Proceedings of Congress Foundation Engineering:
Current Principles and Practices, ASCE, volume 2, pages 1026 – 1039, 1989.

[127] P. K. Robertson. Soil classification using the cone penetration test. Canadian
Geotechnical Journal, 27(1):151–158, 1990.
[128] E. I. Robinsky and C. F. Morrison. Sand displacement and compaction around
model friction piles. Canadian Geotechnical Journal, 1(2):81–93, 1964.
162 BIBLIOGRAPHY

[129] A. Rohe. On the modelling of grain crushing in hypoplasticity. Technical


report, TUDelft, 2010.

[130] H. Rondón, T. Wichtmann, Th Triantafyllidis, and A. Lizcano. Hypoplastic


material constants for a well-graded granular material for base and subbase
layers of flexible pavements. Acta Geotechnica, 2(2):113–126, 2007.

[131] X. Rui, J. Hu, and Q. H. Liu. Higher order finite element method for inhomo-
geneous axisymmetric resonators. Progress In Electromagnetics Research B,
21:189–201, 1989.

[132] A. R. Russell and N. Khalili. Drained cavity expansion in sands exhibiting


particle crushing. International Journal of Numerical and Analytical Methods
in Geomechanics, 26(4):323–340, 2002.

[133] C. Sagaseta and G. T. Houlsby. Stresses near the shoulder of a cone penet-
rometer in clay. In 3rd International conference on Computational Plasticity:
Fumdamentals and Applications, volume 2, pages 895–906, 1992.
[134] C. Sagaseta, G. T. Houlsby, and H. J. Burd. Quasi-static undrained expansion
of a cylindrical cavity in clay in the presence of shaft friction and anisotropic
initial stresses. In K. J. Bathe, editor, Computational Fluid and Solid Mech-
anics 2003, pages 619–622, Oxford, 2003. Elsevier Science Ltd.
[135] I. Said, V. De Gennaro, and R. Frank. Axisymmetric finite element analysis
of pile loading tests. Computers and Geotechnics, 36(1-2):6–19, 2008.

[136] R. Salgado and J. Lee. Pile design based on cone penetration test results.
Technical report, Purdue University, 1999.
[137] J. H. Schmertmann. Guidelines for cone penetration test, performance and
design. Technical report, U.S. Department of Transportation, Washington,
D.C., 1978.
[138] J. A. Schneider and B. M. Lehane. End bearing formulation for cpt based
driven pile design methods in siliceous sands. In CPT’10 2nd International
Symposium on Cone Penetration Testing, volume 2, page 28, 2010.

[139] J. A. Schneider and B.M. Lehane. Correlations for shaft capacity of offshore
piles in sand. In International Symposium on Frontiers in Offshore Geotech-
nics, pages 757–764, 2005.

[140] D. Sheng, P. Wriggers, and S. W. Sloan. Application of frictional contact


in geotechnical engineering. International Journal of Geomechanics, 7(3):176,
2007.

[141] D. A. Shuttle and M. G. Jefferies. Dimensionless and unbiased cpt interpret-


ation in sand. International Journal for Numerical and Analytical Methods in
Geomechanics, 22:351–391, 1998.
BIBLIOGRAPHY 163

[142] A. W. Skempton. The bearing capacity of clays. In Proc. Building Research


Congress, pages 180–189. Rotterdam:Balkema, 1951.

[143] S. W. Sloan and M. F. Randolph. Numerical prediction of collapse loads using


finite element methods. International Journal for Numerical and Analytical
Methods in Geomechanics, 6(1):47–76, 1982.

[144] N. Sukumar. A note on natural neighbor interpolation and the natural ele-
ment method (nem). Technical report, Theoretical and Applied Mechanics of
Northwestern University, 1997.

[145] D. Sulsky, Z. Chen, and H. L. Schreyer. A particle method for history-


dependent materials. Computer Methods in Applied Mechanics and Engin-
eering, 118:179–196, 1994.

[146] E. Susila and R. D. Hryciw. Large displacement fem modelling of the cone
penetration test (cpt) in normally consolidated sand. nternational Journal for
Numerical and Analytical Methods in Geomechanics, 27:585–602, 2003.
[147] F. Tavenas and R. Audy. Limitation of the driving formulas for predicting the
bearing capacities of piles in sand. Canadian Geotechnical Journal, 9(1):47–62,
1972.

[148] C. I. Teh. An analytical study of the cone penetration test. PhD thesis, Uni-
versity of Oxford, 1987.

[149] C. Truesdell. Hypoelasticity. Journal of Ration. Mech. Anal., 4:83–133, 1955.

[150] M. Uesugi and H. Kishida. Influential factors of friction between steel and dry
sand. Soils and Foundations, 26(2):33–46, 1986.
[151] Lade P. V., J. A. Yamamuro, and P. A. Bopp. Significance of particle crushing
in granular materials. Journal of Geotechnical Engineering ASCE, 22(4):309–
316, 1996.

[152] P. van den Berg. Analysis of Soil Penetration. PhD thesis, Delft University of
Technology, 1994.

[153] H. van Langen. Numerical Analysis of Soil-Structure Interaction. PhD thesis,


Delft University of Technology, 1991.

[154] W. C. van Mierlo and A. W. Koppejan. Lengte en draagbermogen van


heilpalen. Bouw, 1952.
[155] A. S. Vesic. Tests on instrumented piles, ogeechee river site. J. Soil Mech.
Fnds. Div. ASCE, 96:561–584, 1970.

[156] A. S. Vesic. Expansion of cavities in infinite soil mass. J. Soil Mech. Fnds.
Div. ASCE, 98:265–290, 1972.
164 BIBLIOGRAPHY

[157] A. S. Vesic. Design of pile foundations. Technical report, Transportation


Research Board, Washington D.C., 1997.
[158] P. A. von Wolffersdorff. A hypoplastic relation for granular materials with
a predefined limit state surface. Mechanics of Cohesive-frictional Materials,
1(3):251–271, 1996.
[159] J. Walker. Adaptive Finite Element Analysis of The Cone Penetration Test.
PhD thesis, University of Nottingham, 2007.
[160] J. Walker and H.S. Yu. Adaptive finite element analysis of cone penetration
in clay. Acta Geotechnica, 1(1):43–57, 2006.
[161] R. G. Wan and P. J. Guo. A simple constitutive model for granular soils:
Modified stress-dilatancy approach. Computers and Geotechnics, 22(2):109–
133, 1998.
[162] J. Wang and Gadala M. S. Formulation and survey of ale method in nonlinear
solid mechanics. Finite Elem Anal Design, (24):253–269, 1997.
[163] E. Wernick. Skin friction of cylindrical anchors in non-cohesive soils. In Sym-
posium on Soil reinforcing and Stabilising Techniques, pages 201–209, 1978.
[164] D. J. White. An investigation into the behaviour of pressed-in piles. PhD
thesis, Univesity of Cambridge, Churchill College, 2002.
[165] D. J. White and M. D. Bolton. Displacement and strain paths during plane-
strain model pile installation in sand. Géotechnique, 54(6):375–397, 2004.
[166] D. J. White and A. D. Deeks. Recent research into the behaviour of jacked
foundation piles. In International Workshop on Recent Advances in Deep
Foundations, pages 3–26, 2007.
[167] D. J. White and B. M. Lehane. Friction fatigue on displacement piles in sand.
Géotechnique, 54(10):645–658, 2004.
[168] A. Wiȩckowski. A particle-in-cell solution to the silo discharging problem. Int.
J. Numer. Meth. Engng, 45:1203–1225, 1999.
[169] A. Wiȩckowski. The material point method in large strain engineering prob-
lems. Computer Methods in Applied Mechanics and Engineering, 193:4417–
4438, 2004.
[170] D. M. Wood, K. Belkheir, and D. F. Liu. Strain softening and state parameter
for sand modelling. Géotechnique, 44(2):335–339, 1994.
[171] R. D. Woods. Dynamic effects of pile installations on adjacent structures.
Technical report, Transportation Research Board, Washington D.C., 1997.
[172] R. D. Woods and V. M. Sharma. Dynamics effects of pile installations on
adjacent structures. Balkema Taylor & Francis, Leiden, 2004.
BIBLIOGRAPHY 165

[173] W. Wu, E. Bauer, and D. Kolymbas. Hypoplastic constitutive model with


critical state for granular materials. Mechanics of Materials, 23:45–69, 1996.

[174] W. Wu and D. Kolymbas. Numerical testing of the stability criterion for


hypoplastic constitutive equations. Mechanics of Materials, 9:245–253, 1990.

[175] W. Wu and A. Niemunis. Beyond failure in granular materials. International


Journal for Numerical and Analytical Methods in Geomechanics, 21:153–174,
1997.

[176] X. T. Xu, J. A. Schneider, and B. M. Lehane. Cone penetration test (cpt)


methods for end-bearing assessment of open- and closed-ended driven piles in
siliceous sand. Canadian Geotechnical Journal, 45(1):1130–1141, 2008.

[177] S. Yaghmai. Incremental analysis of large deformations in mechanics of solids


with applications to axisymmetric shells of revolution. Technical report, De-
partment of Civil Engineering. University of California. Berkeley, 1968.
[178] S. Yaghmai and E. P. Popov. Incremental analysis of large deflections of shells
of revolution. Int. J. Solids Struct., 7:1375–1393, 1971.
[179] L. Yang and R. Liang. Incorporating setup into load and resistance factor
design of driven piles in sand. Canadian Geotechnical Journal, 46:296–305,
2009.

[180] N. Yasufuku and A. F. L. Hyde. Pile end-bearing capacity in crushable sands.


Géotechnique, 45:663–676, 1995.
[181] A. G. Yetginer, D. J. White, and M. D. Bolton. Field measurements of the
stiffness of jacked piles and pile groups. Géotechnique, 56:349–354, 2006.
[182] T. L. Youd. Factors controlling maximum and minimum densities of sands.
In E. T. Selig and R. S. Ladd, editors, 75th Annual Meeting American Society
for Testing and Materials, pages 98–112, 1973.

[183] H. S. Yu, L. R. Herrmann, and R. W. Boulanger. Analysis of steady cone pen-


etration in clay. Journal of Geotechnical and Geoenvironmental Engineering,
126(7):594–605, 2000.

[184] H. S. Yu and G. T. Houlsby. A new finite element formulation for one-


dimensional analysis of elastic-plastic materials. Computers and Geotechnics,
9(4):241–256, 1990.

[185] H. S. Yu and G. T. Houlsby. Finite cavity expansion in dilatant soils: loading


analysis. Géotechnique, 41(2):173–183, 1991.

[186] H. S. Yu and A. J. Whittle. Combining strain path analysis and cavity expan-
sion theory to estimate cone resistance in clay. Unpublished notes, 1999.
166 BIBLIOGRAPHY
Appendix A

Calculation of Unbalance

In this appendix open form of the unbalance in radial and vertical directions will be
presented based on the general forms proposed for the stresses around the pile due
to installation. The new cartesian stresses were expressed in terms of pile diameter,
D and penetration length, L as:
+ ,
3 · K 
3
2 2

= p0 · ai1 · e−(bi1 (r/D−ri1 ) +ci1 (z/D+L/D+zi1 ) )
0
σrr + (A.1)
1 + 2 · K0 i=1
+ ,
3 · K0 
4
 −(bi2 (r/D−ri2 )2 +ci2 (z/D+L/D+zi2 )2 )
σzz = ·p0 + ai2 · e (A.2)
1 + 2 · K0 i=1
+ ,
3 · K 
3
2 2


= p0 · ai3 · e ( i3 )
0 b (r/D−r ) +c (z/D+L/D+z )
σθθ + i3 i3 i3
(A.3)
1 + 2 · K0 i=1


6
2
+ci4 (z/D+L/D+zi4 )2 )

σrz = p0 · ai4 · e−(bi4 (r/D−ri4 ) (A.4)
i=1

These stresses can be substituted into the equations:

1 ∂  ∂ σ
δrr = (rσrr )+ (σrz ) − θθ (A.5)
r ∂r ∂z r
1 ∂  ∂ 
δzz = (rσrr )+ (σrz ) − γsoil (A.6)
r ∂r ∂z
to find the unbalance in corresponding direction.
This system can be evaluated by the following MATLAB code:

167
168 A. Calculation of Unbalance

1 % symbolic check of functions for equilibrium:


2 syms r z
3 syms L D
4 syms Rrr Rtt Rtt Rrz
5 syms srr szz stt srz
6 syms srrK0 szzK0 sttK0
7
8 syms a11 a12 a13 a14
9 syms a21 a22 a23 a24
10 syms a31 a32 a33 a34
11 syms a42 a44
12 syms a54
13 syms a64
14 syms a74
15
16 syms b11 b12 b13 b14
17 syms b21 b22 b23 b24
18 syms b31 b32 b33 b34
19 syms b42 b44
20 syms b54
21 syms b64
22 syms b74
23
24 syms c11 c12 c13 c14
25 syms c21 c22 c23 c24
26 syms c31 c32 c33 c34
27 syms c42 c44
28 syms c54
29 syms c64
30 syms c74
31
32 syms r11 r12 r13 r14
33 syms r21 r22 r23 r24
34 syms r31 r32 r33 r34
35 syms r42 r44
36 syms r54
37 syms r64
38 syms r74
39
40 syms z11 z12 z13 z14
41 syms z21 z22 z23 z24
42 syms z31 z32 z33 z34
43 syms z42 z44
44 syms z54
45 syms z64
46 syms z74
47
48 syms S
49 syms gsoil
50 syms srz1 srz2
51 syms K0
52
53 Rrr = (3*K0)/(2*K0 + 1) + a11/exp(b11*(r11 − r/D)ˆ2 + ...
c11*(z11 + L/D + z/D)ˆ2) + a21/exp(b21*(r21 − r/D)ˆ2 + ...
c21*(z21 + L/D + z/D)ˆ2) + a31/exp(b31*(r31 − r/D)ˆ2 + ...
c31*(z31 + L/D + z/D)ˆ2)
169

54
55 Rtt = (3*K0)/(2*K0 + 1) + a13/exp(b13*(r13 − r/D)ˆ2 + ...
c13*(z13 + L/D + z/D)ˆ2) + a23/exp(b23*(r23 − r/D)ˆ2 + ...
c23*(z23 + L/D + z/D)ˆ2) + a33/exp(b33*(r33 − r/D)ˆ2 + ...
c33*(z33 + L/D + z/D)ˆ2)
56
57 Rzz = 3/(2*K0 + 1) + a12/exp(b12*(r12 − r/D)ˆ2 + c12*(z12 + L/D ...
+ z/D)ˆ2) + a22/exp(b22*(r22 − r/D)ˆ2 + c22*(z22 + L/D + ...
z/D)ˆ2) + a32/exp(b32*(r32 − r/D)ˆ2 + c32*(z32 + L/D + ...
z/D)ˆ2) + a42/exp(b42*(r42 − r/D)ˆ2 + c42*(z42 + L/D + z/D)ˆ2)
58
59 Rrz = a14/exp(b14*(r14 − r/D)ˆ2 + c14*(z14 + L/D + z/D)ˆ2) + ...
a24/exp(b24*(r24 − r/D)ˆ2 + c24*(z24 + L/D + z/D)ˆ2) + ...
a34/exp(b34*(r34 − r/D)ˆ2 + c34*(z34 + L/D + z/D)ˆ2) + ...
a44/exp(b44*(r44 − r/D)ˆ2 + c44*(z44 + L/D + z/D)ˆ2) + ...
a54/exp(b54*(r54 − r/D)ˆ2 + c54*(z54 + L/D + z/D)ˆ2) + ...
a64/exp(b64*(r64 − r/D)ˆ2 + c64*(z64 + L/D + z/D)ˆ2)
60
61 p0 = 1/3*(−100+gsoil*z)*(1+2*K0)
62 srr = Rrr*p0
63 szz = Rzz*p0
64 stt = Rtt*p0
65 srz = Rrz*p0
66 % Force equilibrium in radial direction:
67 Equil1 = 1/r*(srr+r*diff(srr,r))+diff(srz,z)−stt/r
68 % Force equilibrium in vertical direction:
69 Equil2 = 1/r*(srz+r*diff(srz,r))+diff(szz,z)−gsoil

and following unbalance forms can be obtained:

1 Equil1 = ((2*K0 + 1)*((gsoil*z)/3 − 100/3)*((3*K0)/(2*K0 + ...


1) + a11/exp(b11*(r11 − r/D)ˆ2 + c11*(z11 + L/D + ...
z/D)ˆ2) + a21/exp(b21*(r21 − r/D)ˆ2 + c21*(z21 + L/D + ...
z/D)ˆ2) + a31/exp(b31*(r31 − r/D)ˆ2 + c31*(z31 + L/D + ...
z/D)ˆ2)) + r*(2*K0 + 1)*((gsoil*z)/3 − ...
100/3)*((2*a11*b11*(r11 − r/D))/(D*exp(b11*(r11 − r/D)ˆ2 ...
+ c11*(z11 + L/D + z/D)ˆ2)) + (2*a21*b21*(r21 − ...
r/D))/(D*exp(b21*(r21 − r/D)ˆ2 + c21*(z21 + L/D + ...
z/D)ˆ2)) + (2*a31*b31*(r31 − r/D))/(D*exp(b31*(r31 − ...
r/D)ˆ2 + c31*(z31 + L/D + z/D)ˆ2))))/r + (gsoil*(2*K0 + ...
1)*(a14/exp(b14*(r14 − r/D)ˆ2 + c14*(z14 + L/D + z/D)ˆ2) ...
+ a24/exp(b24*(r24 − r/D)ˆ2 + c24*(z24 + L/D + z/D)ˆ2) + ...
a34/exp(b34*(r34 − r/D)ˆ2 + c34*(z34 + L/D + z/D)ˆ2) + ...
a44/exp(b44*(r44 − r/D)ˆ2 + c44*(z44 + L/D + z/D)ˆ2) + ...
a54/exp(b54*(r54 − r/D)ˆ2 + c54*(z54 + L/D + z/D)ˆ2) + ...
a64/exp(b64*(r64 − r/D)ˆ2 + c64*(z64 + L/D + z/D)ˆ2)))/3 ...
− (2*K0 + 1)*((gsoil*z)/3 − 100/3)*((2*a14*c14*(z14 + ...
L/D + z/D))/(D*exp(b14*(r14 − r/D)ˆ2 + c14*(z14 + L/D + ...
z/D)ˆ2)) + (2*a24*c24*(z24 + L/D + z/D))/(D*exp(b24*(r24 ...
− r/D)ˆ2 + c24*(z24 + L/D + z/D)ˆ2)) + (2*a34*c34*(z34 + ...
L/D + z/D))/(D*exp(b34*(r34 − r/D)ˆ2 + c34*(z34 + L/D + ...
z/D)ˆ2)) + (2*a44*c44*(z44 + L/D + z/D))/(D*exp(b44*(r44 ...
− r/D)ˆ2 + c44*(z44 + L/D + z/D)ˆ2)) + (2*a54*c54*(z54 + ...
L/D + z/D))/(D*exp(b54*(r54 − r/D)ˆ2 + c54*(z54 + L/D + ...
z/D)ˆ2)) + (2*a64*c64*(z64 + L/D + z/D))/(D*exp(b64*(r64 ...
− r/D)ˆ2 + c64*(z64 + L/D + z/D)ˆ2))) − ((2*K0 + ...
170 A. Calculation of Unbalance

1)*((gsoil*z)/3 − 100/3)*((3*K0)/(2*K0 + 1) + ...


a13/exp(b13*(r13 − r/D)ˆ2 + c13*(z13 + L/D + z/D)ˆ2) + ...
a23/exp(b23*(r23 − r/D)ˆ2 + c23*(z23 + L/D + z/D)ˆ2) + ...
a33/exp(b33*(r33 − r/D)ˆ2 + c33*(z33 + L/D + z/D)ˆ2)))/r
2
3 Equil2 = ((2*K0 + 1)*((gsoil*z)/3 − 100/3)*(a14/exp(b14*(r14 ...
− r/D)ˆ2 + c14*(z14 + L/D + z/D)ˆ2) + a24/exp(b24*(r24 − ...
r/D)ˆ2 + c24*(z24 + L/D + z/D)ˆ2) + a34/exp(b34*(r34 − ...
r/D)ˆ2 + c34*(z34 + L/D + z/D)ˆ2) + a44/exp(b44*(r44 − ...
r/D)ˆ2 + c44*(z44 + L/D + z/D)ˆ2) + a54/exp(b54*(r54 − ...
r/D)ˆ2 + c54*(z54 + L/D + z/D)ˆ2) + a64/exp(b64*(r64 − ...
r/D)ˆ2 + c64*(z64 + L/D + z/D)ˆ2)) + r*(2*K0 + ...
1)*((gsoil*z)/3 − 100/3)*((2*a14*b14*(r14 − ...
r/D))/(D*exp(b14*(r14 − r/D)ˆ2 + c14*(z14 + L/D + ...
z/D)ˆ2)) + (2*a24*b24*(r24 − r/D))/(D*exp(b24*(r24 − ...
r/D)ˆ2 + c24*(z24 + L/D + z/D)ˆ2)) + (2*a34*b34*(r34 − ...
r/D))/(D*exp(b34*(r34 − r/D)ˆ2 + c34*(z34 + L/D + ...
z/D)ˆ2)) + (2*a44*b44*(r44 − r/D))/(D*exp(b44*(r44 − ...
r/D)ˆ2 + c44*(z44 + L/D + z/D)ˆ2)) + (2*a54*b54*(r54 − ...
r/D))/(D*exp(b54*(r54 − r/D)ˆ2 + c54*(z54 + L/D + ...
z/D)ˆ2)) + (2*a64*b64*(r64 − r/D))/(D*exp(b64*(r64 − ...
r/D)ˆ2 + c64*(z64 + L/D + z/D)ˆ2))))/r − gsoil + ...
(gsoil*(2*K0 + 1)*(3/(2*K0 + 1) + a12/exp(b12*(r12 − ...
r/D)ˆ2 + c12*(z12 + L/D + z/D)ˆ2) + a22/exp(b22*(r22 − ...
r/D)ˆ2 + c22*(z22 + L/D + z/D)ˆ2) + a32/exp(b32*(r32 − ...
r/D)ˆ2 + c32*(z32 + L/D + z/D)ˆ2) + a42/exp(b42*(r42 − ...
r/D)ˆ2 + c42*(z42 + L/D + z/D)ˆ2)))/3 − (2*K0 + ...
1)*((gsoil*z)/3 − 100/3)*((2*a12*c12*(z12 + L/D + ...
z/D))/(D*exp(b12*(r12 − r/D)ˆ2 + c12*(z12 + L/D + ...
z/D)ˆ2)) + (2*a22*c22*(z22 + L/D + z/D))/(D*exp(b22*(r22 ...
− r/D)ˆ2 + c22*(z22 + L/D + z/D)ˆ2)) + (2*a32*c32*(z32 + ...
L/D + z/D))/(D*exp(b32*(r32 − r/D)ˆ2 + c32*(z32 + L/D + ...
z/D)ˆ2)) + (2*a42*c42*(z42 + L/D + z/D))/(D*exp(b42*(r42 ...
− r/D)ˆ2 + c42*(z42 + L/D + z/D)ˆ2)))
Appendix B

Implementation of Functions

In this appendix the update of stresses and void ratio using proposed functions, if
activated by the switch (an input parameter) is described by the following flowchart:

Is this the yes Is the yes Read


beginning switch additional
of the step? ON? input data

no no
Calculate
Assign
Continue Is the model 3D rGP ,zGP , and
rGP = xGP
2D 2D or 3D? rotation
zGP = yGP
matrix, A

Apply
geometric
functions
to update
σij and e

∗ 2D Is the model 3D ∗ 2D T
σij = σij σij = Aσij A
2D 2D or 3D?

171
172 B. Implementation of Functions
Appendix C

Calculation of Pile Bearing


Capacity Using Empirical Methods

In this appendix the calculation of pile base, Qp and shaft, Qs based on the state-of-
the-art empirical methods API [12], Dutch method [54], Fugro-05 [92], ICP-05 [88],
NGI-05 [43], and UWA-05 [101] is presented in the MATLAB code provided herein.
The empirical methods uses qc data to estimate the shaft and tip capacities. In
order to provide qc data for these methods, the correlation of Baldi et al. [16]:
 
1 qc
Id = ln  ) C1 (C.1)
C2 C0 · (σv0

is inverted to have Id as input:


 C1
qc = C0 · (σv0 ) · e(Id ·C2 ) (C.2)

where, C0 = 157, C1 = 0.55 and C2 = 2.41.

173
174 C. Calculation of Pile Bearing Capacity Using Empirical Methods

1 tic
2 %formattedDate
3 dateF = strrep(date, '−', '');
4 dateF = strrep(dateF, '20', '');
5 % qc estimated from Baldi et al 1986
6 % qc=0.001*EXP(I d * C 2)* C 0 *C2ˆC 1
7 C0 = 157; C1=0.55; C2=2.41;
8 D = 0.001*15*35; %m
9 R = D/2; %m
10 L = 14.175; %m
11 Ap = pi*Rˆ2; %m2
12 As = pi*D*L; %m2
13 emin = 0.548;
14 emax = 0.929;
15 zphead = 7.175; %m
16 zptip = −7; %m
17 zmin = −7−1.5*D;
18 nseg = 100;
19 dz = (zphead−zptip)/nseg;
20 z = zeros(nseg,1);
21 delcv = tan(30*pi/180); % interface friction
22 for i=1:nseg+1
23 z(i) = zphead − dz*(i−1);
24 end
25 dz2 = 2*D/20; %below tip
26
27 for i=nseg+2:nseg+22
28 z(i) = zptip − dz2*(i−nseg−1);
29 end
30
31 for p = 1:3;
32 switch p
33 case 1 % e0=0.637;Id=0.77
34 e0=0.637;
35 gam = 11;
36 Id = (emax−e0)/(emax−emin);
37 betaAPI = 0.46;
38 limshAPI= 96;% kPa
39 Nq = 40;
40 qblimAPI = 10; %MPa
41 sigv = (z−zphead)*gam;
42 qc = C0*(−sigv).ˆC1*exp(C2*Id);
43
44 % Centrifuge test results at uy=0.10D
45 Qt DJK = 4.99; % MN
46 Qp DJK = 3.70; % MN
47 Qs DJK = 1.29; % MN
48
49 % FE results
50 Qt K0 = 1.37; % MN
51 Qp K0 = 0.90; % MN
52 Qs K0 = 0.47; % MN
53
54 Qt fall = 2.44; % MN
55 Qp fall = 1.99; % MN
56 Qs fall = 0.45; % MN
175

57
58 Qt f4s = 2.43; % MN
59 Qp f4s = 2.01; % MN
60 Qs f4s = 0.42; % MN
61
62 Qt f3s = 2.32; % MN
63 Qp f3s = 1.73; % MN
64 Qs f3s = 0.60; % MN
65 case 2 % e0=0.709;Id=0.58
66 e0=0.709;
67 gam = 10;
68 Id = (emax−e0)/(emax−emin);
69 betaAPI = 0.37;
70 limshAPI= 81;% kPa
71 Nq = 20;
72 qblimAPI = 5; %MPa
73 sigv = (z−zphead)*gam;
74 qc = C0*(−sigv).ˆC1*exp(C2*Id);
75
76 % Centrifuge test results at uy=0.10D
77 Qt DJK = 1.95; % MN
78 Qp DJK = 1.33; % MN
79 Qs DJK = 0.62; % MN
80
81 % FE results
82 Qt K0 = 0.93; % MN
83 Qp K0 = 0.51; % MN
84 Qs K0 = 0.42; % MN
85
86 Qt fall = 1.56; % MN
87 Qp fall = 1.21; % MN
88 Qs fall = 0.35; % MN
89
90 Qt f4s = 1.67; % MN
91 Qp f4s = 1.29; % MN
92 Qs f4s = 0.38; % MN
93
94 Qt f3s = 1.53; % MN
95 Qp f3s = 1.06; % MN
96 Qs f3s = 0.47; % MN
97 case 3 % e0=0.783;Id=0.38
98 e0=0.783;
99 gam = 9;
100 Id = (emax−e0)/(emax−emin);
101 betaAPI = 0.37;
102 limshAPI= 81; % kPa
103 Nq = 20;
104 qblimAPI = 5; %MPa
105 sigv = (z−zphead)*gam;
106 qc = C0*(−sigv).ˆC1*exp(C2*Id);
107
108 % Centrifuge test results at uy=0.10D
109 Qt DJK = 1.41; % MN
110 Qp DJK = 0.80; % MN
111 Qs DJK = 0.61; % MN
112
113 % FE results
176 C. Calculation of Pile Bearing Capacity Using Empirical Methods

114 Qt K0 = 0.74; % MN
115 Qp K0 = 0.33; % MN
116 Qs K0 = 0.41; % MN
117
118 Qt fall = 1.23; % MN
119 Qp fall = 0.90; % MN
120 Qs fall = 0.33; % MN
121
122 Qt f4s = 1.24; % MN
123 Qp f4s = 0.90; % MN
124 Qs f4s = 0.34; % MN
125
126 Qt f3s = 1.12; % MN
127 Qp f3s = 0.77; % MN
128 Qs f3s = 0.36; % MN
129 end
130
131 % API
132 qs API = zeros(nseg+1,1);
133 z API = zeros(nseg+1,1);
134 dFs = zeros(nseg,1);
135 for i=1:nseg+1
136 qs API(i) = min(−sigv(i)*betaAPI,limshAPI);
137 if i>1
138 dFs(i−1) = pi*D*dz*0.5*(qs API(i−1)+qs API(i));
139 end
140 z API(i) = zphead−(i−1)*dz;
141 end
142 Qs API = 0.001*sum(dFs);
143 Qp API = min(0.001*Nq*(−sigv(nseg+1)),qblimAPI)*Ap;
144 Qt API = Qs API+Qp API;
145
146 % Dutch Method
147 cnt = 0;
148 qsum = 0;
149 for i=1:length(z)
150 if z(i)<zptip+8*D
151 if z(i)>zptip−0.7*D
152 cnt = cnt+1;
153 qsum = qsum + qc(i);
154 end
155 end
156 end
157 qcav Dutch = 0.001*qsum/cnt; %MPa
158 qs Dutch = zeros(nseg+1,1);
159 dFs = zeros(nseg+1,1);
160 for i=1:nseg+1
161 qs Dutch(i) = 0.01*qc(i);
162 if i>1
163 dFs(i−1) = pi*D*dz*0.5*(qs Dutch(i−1)+qs Dutch(i));
164 end
165 end
166 Qs Dutch = 0.001*sum(dFs); % MN
167 Qp Dutch = 1*qcav Dutch*Ap; % MN
168 Qt Dutch = Qs Dutch+Qp Dutch; % MN
169
170 % Fugro−05
177

171 cnt = 0;
172 qsum = 0;
173 for i=1:length(z)
174 if z(i)<zptip+1.5*D
175 if z(i)>zptip−1.5*D
176 cnt = cnt+1;
177 qsum = qsum + qc(i);
178 end
179 end
180 end
181 qcav Fugro = 0.001*qsum/cnt; %MPa
182 hR = zeros(nseg+1,1);
183 tau Fugro = zeros(nseg+1,1);
184 for i=1:nseg+1
185 hR(i) = (z(i)−zptip)/R;
186 if hR(i)≥4
187 tau Fugro(i)= ...
0.08*qc(i)*(−sigv(i)/100)ˆ0.05*(hR(i))ˆ(−0.9); %kPa
188 else
189 tau Fugro(i)= ...
0.08*qc(i)*(−sigv(i)/100)ˆ0.05*4ˆ(−0.9)*(hR(i)/4); %kPa
190 end
191 if i>1
192 dFs(i−1) = pi*D*dz*0.5*(tau Fugro(i−1)+tau Fugro(i));
193 end
194 end
195 Qs Fugro = 0.001*sum(dFs); % MN
196 Qp Fugro = 0.001*8.5*100*(qcav Fugro*1000/100)ˆ0.5*1*Ap; % MN
197 Qt Fugro = Qs Fugro+Qp Fugro; % MN
198
199 % ICP
200 cnt = 0;
201 qsum = 0;
202 for i=1:length(z)
203 if z(i)<zptip+1.5*D
204 if z(i)>zptip−1.5*D
205 cnt = cnt+1;
206 qsum = qsum + qc(i);
207 end
208 end
209 end
210 qcav ICP = 0.001*qsum/cnt; %MPa
211 hR = zeros(nseg+1,1);
212 eta = zeros(nseg+1,1);
213 sigrc ICP = zeros(nseg+1,1);
214 Gm ICP = zeros(nseg+1,1);
215 qs ICP = zeros(nseg+1,1);
216 dsigr = zeros(nseg+1,1);
217 tau ICP = zeros(nseg+1,1);
218
219 for i=1:nseg+1
220 hR(i) = max((z(i)−zptip)/R,8);
221 sigrc ICP(i)= 0.029*qc(i)*(−sigv(i)/100)ˆ0.13*(hR(i))ˆ(−0.38);
222 if i>1
223 eta(i) = qc(i)*(100*(−sigv(i))ˆ(−0.5));
224 end
225 Gm ICP(i) = qc(i)*((0.0203+0.00125*eta(i)−...
178 C. Calculation of Pile Bearing Capacity Using Empirical Methods

226 1.216e−6*(eta(i))ˆ2)ˆ(−1));
227 dsigr(i) = 2* Gm ICP(i)*0.02*0.001*hR(i);
228 tau ICP(i) = delcv*(sigrc ICP(i)+dsigr(i));
229 if i>1
230 dFs(i−1) = pi*D*dz*0.5*(tau ICP(i−1)+tau ICP(i));
231 end
232 end
233 Qs ICP = 0.001*sum(dFs); % MN
234 Qp ICP = qcav ICP*max((1−0.5*log10(D/0.036)),0.3)*Ap; % MN
235 Qt ICP = Qs ICP+Qp ICP; % MN
236
237 % NGI−05 (Clausen et al., 2005)
238 qc NGI = 0.001*qc(nseg+1); %MPa
239 zi ztip = zeros(nseg+1,1);
240 tau NGI = zeros(nseg+1,1);
241 Dr = zeros(nseg+1,1); % According to NGI−05
242 FDr = zeros(nseg+1,1);
243 Fsig = zeros(nseg+1,1);
244 pa = 100; %kPa
245 Fload = 1.3; % for tension = 1.0
246 Ftip = 1.6; % closed end piles / for driven&open end=1.0
247 Fmat = 1.0; % for steel / 1.2 for concrete
248 for i=1:nseg+1
249 zi ztip(i) = (−z(i)+zphead)/(−zptip+zphead);
250 if i>1
251 Dr(i) = 0.4*log(qc(i)/(22*(−sigv(i)*100)ˆ0.5));
252 FDr(i) = 2.1*(Dr(i)−0.1)ˆ1.7; % According to NGI−05
253 Fsig(i) = (−sigv(i)/pa)ˆ(−0.25);
254 end
255 tau NGI(i) = ...
max(zi ztip(i)*pa*FDr(i)*Fload*Ftip*Fmat*Fsig(i),...
256 −0.1*sigv(i));
257 if i>1
258 dFs(i−1) = pi*D*dz*0.5*(tau NGI(i−1)+tau NGI(i));
259 end
260 end
261 Qs NGI = 0.001*sum(dFs); % MN
262 Qp NGI = (0.8* qc NGI/(1+Dr(nseg+1)ˆ2))*Ap; % MN
263 Qt NGI = Qs NGI+Qp NGI; % MN
264
265 % UWA
266 cnt = 0;
267 qsum = 0;
268 for i=1:length(z)
269 if z(i)<zptip+8*D
270 if z(i)>zptip−0.7*D % as Dutch method
271 cnt = cnt+1;
272 qsum = qsum + qc(i);
273 end
274 end
275 end
276 qcav UWA = 0.001*qsum/cnt; %MPa
277 hD = zeros(nseg+1,1);
278 sigrc UWA = zeros(nseg+1,1);
279 Gm UWA = zeros(nseg+1,1);
280 qc1N = zeros(nseg+1,1);
281 dsigr = zeros(nseg+1,1);
179

282 tau UWA = zeros(nseg+1,1);


283 for i=1:nseg+1
284 hD(i) = max((z(i)−zptip)/D,2);
285 sigrc UWA(i)= 0.03*qc(i)*(hD(i))ˆ(−0.5);
286 if i>1
287 qc1N(i) = (qc(i)/100)*(−sigv(i)/100)ˆ0.5;
288 Gm UWA(i) = qc(i)*185*(qc1N(i))ˆ(−0.7); %kN
289 dsigr(i) = 4* Gm UWA(i)*0.02*0.001/D;
290 end
291 tau UWA(i) = min(delcv,0.55)*(sigrc UWA(i)+dsigr(i));
292 if i>1
293 dFs(i−1) = pi*D*dz*0.5*(tau UWA(i−1)+tau UWA(i));
294 end
295 end
296 Qs UWA = 0.001*sum(dFs); % MN
297 Qp UWA = 0.9*qcav UWA*Ap; % MN
298 Qt UWA = Qs UWA+Qp UWA; % MN
299 % Collect data of each method in a vector for plotting purposes
300 switch p
301 case 1
302 vId(1,1) = Id;
303 % Total
304 vQt K0(1,1) = Qt K0;
305 vQt DJK(1,1) = Qt DJK;
306 vQt fall(1,1) = Qt fall;
307 vQt f4s(1,1) = Qt f4s;
308 vQt f3s(1,1) = Qt f3s;
309 vQt API(1,1) = Qt API;
310 vQt Dutch(1,1)= Qt Dutch;
311 vQt NGI(1,1) = Qt NGI;
312 vQt Fugro(1,1)= Qt Fugro;
313 vQt ICP(1,1) = Qt ICP;
314 vQt UWA(1,1) = Qt UWA;
315
316 % Shaft
317 vQs K0(1,1) = Qs K0;
318 vQs DJK(1,1) = Qs DJK;
319 vQs fall(1,1) = Qs fall;
320 vQs f4s(1,1) = Qs f4s;
321 vQs f3s(1,1) = Qs f3s;
322 vQs API(1,1) = Qs API;
323 vQs Dutch(1,1)= Qs Dutch;
324 vQs NGI(1,1) = Qs NGI;
325 vQs Fugro(1,1)= Qs Fugro;
326 vQs ICP(1,1) = Qs ICP;
327 vQs UWA(1,1) = Qs UWA;
328
329 % Tip
330 vQp K0(1,1) = Qp K0;
331 vQp DJK(1,1) = Qp DJK;
332 vQp fall(1,1) = Qp fall;
333 vQp f4s(1,1) = Qp f4s;
334 vQp f3s(1,1) = Qp f3s;
335 vQp API(1,1) = Qp API;
336 vQp Dutch(1,1)= Qp Dutch;
337 vQp NGI(1,1) = Qp NGI;
338 vQp Fugro(1,1)= Qp Fugro;
180 C. Calculation of Pile Bearing Capacity Using Empirical Methods

339 vQp ICP(1,1) = Qp ICP;


340 vQp UWA(1,1) = Qp UWA;
341 case 2
342 vId(2,1) = Id;
343 % Total
344 vQt K0(2,1) = Qt K0;
345 vQt DJK(2,1) = Qt DJK;
346 vQt fall(2,1) = Qt fall;
347 vQt f4s(2,1) = Qt f4s;
348 vQt f3s(2,1) = Qt f3s;
349 vQt API(2,1) = Qt API;
350 vQt Dutch(2,1)= Qt Dutch;
351 vQt NGI(2,1) = Qt NGI;
352 vQt Fugro(2,1)= Qt Fugro;
353 vQt ICP(2,1) = Qt ICP;
354 vQt UWA(2,1) = Qt UWA;
355
356 % Shaft
357 vQs K0(2,1) = Qs K0;
358 vQs DJK(2,1) = Qs DJK;
359 vQs fall(2,1) = Qs fall;
360 vQs f4s(2,1) = Qs f4s;
361 vQs f3s(2,1) = Qs f3s;
362 vQs API(2,1) = Qs API;
363 vQs Dutch(2,1)= Qs Dutch;
364 vQs NGI(2,1) = Qs NGI;
365 vQs Fugro(2,1)= Qs Fugro;
366 vQs ICP(2,1) = Qs ICP;
367 vQs UWA(2,1) = Qs UWA;
368
369 % Tip
370 vQp K0(2,1) = Qp K0;
371 vQp DJK(2,1) = Qp DJK;
372 vQp fall(2,1) = Qp fall;
373 vQp f4s(2,1) = Qp f4s;
374 vQp f3s(2,1) = Qp f3s;
375 vQp API(2,1) = Qp API;
376 vQp Dutch(2,1)= Qp Dutch;
377 vQp NGI(2,1) = Qp NGI;
378 vQp Fugro(2,1)= Qp Fugro;
379 vQp ICP(2,1) = Qp ICP;
380 vQp UWA(2,1) = Qp UWA;
381
382 case 3
383 vId(3,1) = Id;
384 % Total
385 vQt K0(3,1) = Qt K0;
386 vQt DJK(3,1) = Qt DJK;
387 vQt fall(3,1) = Qt fall;
388 vQt f4s(3,1) = Qt f4s;
389 vQt f3s(3,1) = Qt f3s;
390 vQt API(3,1) = Qt API;
391 vQt Dutch(3,1)= Qt Dutch;
392 vQt ICP(3,1) = Qt ICP;
393 vQt NGI(3,1) = Qt NGI;
394 vQt Fugro(3,1)= Qt Fugro;
395 vQt UWA(3,1) = Qt UWA;
181

396
397 % Shaft
398 vQs K0(3,1) = Qs K0;
399 vQs DJK(3,1) = Qs DJK;
400 vQs fall(3,1) = Qs fall;
401 vQs f4s(3,1) = Qs f4s;
402 vQs f3s(3,1) = Qs f3s;
403 vQs API(3,1) = Qs API;
404 vQs Dutch(3,1)= Qs Dutch;
405 vQs NGI(3,1) = Qs NGI;
406 vQs Fugro(3,1)= Qs Fugro;
407 vQs ICP(3,1) = Qs ICP;
408 vQs UWA(3,1) = Qs UWA;
409
410 % Tip
411 vQp K0(3,1) = Qp K0;
412 vQp DJK(3,1) = Qp DJK;
413 vQp fall(3,1) = Qp fall;
414 vQp f4s(3,1) = Qp f4s;
415 vQp f3s(3,1) = Qp f3s;
416 vQp API(3,1) = Qp API;
417 vQp Dutch(3,1)= Qp Dutch;
418 vQp NGI(3,1) = Qp NGI;
419 vQp Fugro(3,1)= Qp Fugro;
420 vQp ICP(3,1) = Qp ICP;
421 vQp UWA(3,1) = Qp UWA;
422 end
423 end
424 % colors of the different diameter and density cases
425 % Djk
426 pcolors(1,:) = [0 0 0]; % black
427 % Djk corrected
428 pcolors(2,:) = [0.1 0.1 0.1]; % gray
429 % fall
430 pcolors(3,:) = [1 0 0]; % red
431 % f4s
432 pcolors(4,:) = [1 0.5 0]; % orange
433 % f3s
434 pcolors(5,:) = [1 0.4 0.7]; % light soft red−pink
435 % K0
436 pcolors(6,:) = [0 0.25 1]; % light blue
437
438 % Qpˆcalculated/Qpˆmeasured
439 RQp K0 = vQp K0./vQp DJK;
440 RQp fall = vQp fall./vQp DJK;
441 RQp f4s = vQp f4s./vQp DJK;
442 RQp f3s = vQp f3s./vQp DJK;
443 RQp API = vQp API./vQp DJK;
444 RQp Dutch = vQp Dutch./vQp DJK;
445 RQp Fugro = vQp Fugro./vQp DJK;
446 RQp ICP = vQp ICP./vQp DJK;
447 RQp NGI = vQp NGI./vQp DJK;
448 RQp UWA = vQp UWA./vQp DJK;
449 RQp=[RQp K0';RQp fall';RQp f4s';RQp f3s';RQp API';...
450 RQp Dutch';RQp Fugro';RQp ICP';RQp NGI';RQp UWA'];
451
452
182 C. Calculation of Pile Bearing Capacity Using Empirical Methods

453 % Qsˆcalculated/Qsˆmeasured
454 RQs K0 = vQs K0./vQs DJK;
455 RQs fall = vQs fall./vQs DJK;
456 RQs f4s = vQs f4s./vQs DJK;
457 RQs f3s = vQs f3s./vQs DJK;
458 RQs API = vQs API./vQs DJK;
459 RQs Dutch = vQs Dutch./vQs DJK;
460 RQs Fugro = vQs Fugro./vQs DJK;
461 RQs ICP = vQs ICP./vQs DJK;
462 RQs NGI = vQs NGI./vQs DJK;
463 RQs UWA = vQs UWA./vQs DJK;
464 RQs=[RQs K0';RQs fall';RQs f4s';RQs f3s';RQs API';...
465 RQs Dutch';RQs Fugro';RQs ICP';RQs NGI';RQs UWA'];
466
467 % Qtˆcalculated/Qtˆmeasured
468 RQt K0 = vQt K0./vQt DJK;
469 RQt fall = vQt fall./vQt DJK;
470 RQt f4s = vQt f4s./vQt DJK;
471 RQt f3s = vQt f3s./vQt DJK;
472 RQt API = vQt API./vQt DJK;
473 RQt Dutch = vQt Dutch./vQt DJK;
474 RQt Fugro = vQt Fugro./vQt DJK;
475 RQt ICP = vQt ICP./vQt DJK;
476 RQt NGI = vQt NGI./vQt DJK;
477 RQt UWA = vQt UWA./vQt DJK;
478 RQt=[RQt K0';RQt fall';RQt f4s';RQt f3s';RQt API';...
479 RQt Dutch';RQt Fugro';RQt ICP';RQt NGI';RQt UWA'];
480
481 g = figure( 'Name', 'Comparison with DJK' );
482 subplot(1,2,1)
483 plot([0 0.5 1],vQp DJK./vQp DJK,'LineStyle','−','Color','k',...
484 'LineWidth',2,'DisplayName','Dijkstra (2009)');
485 axis([0.3,0.8,0.,2])
486 set(gca,'XDir','reverse')
487 axis square
488 set(gca,'fontsize',20)
489 title ( 'Calculated to measured tip capacities','fontsize',22);
490 xlabel( '$$I d$$','interpreter','latex','fontsize',30);
491 ylabel( '$$Q pˆ{calculated}/Q pˆ{measured} $$','interpreter','latex',...
492 'fontsize',30);
493 hold on
494 plot(vId,RQp K0,'Marker','.','MarkerSize',4,'Color','k','LineStyle',...
495 ':','LineWidth',2,'DisplayName','$K 0$');
496 plot(vId,RQp fall,'Marker','+','MarkerSize',25,'Color',pcolors(6,:),...
497 'LineWidth',2,'LineStyle','−−','DisplayName','$\hat{\psi}ˆ1$');
498 plot(vId,RQp f4s,'Marker','o','MarkerSize',25,'Color',pcolors(3,:),...
499 'LineWidth',2,'LineStyle','−.','DisplayName','$\hat{\psi}ˆ2$');
500 plot(vId,RQp f3s,'Marker','ˆ','MarkerSize',25,'Color',pcolors(4,:),...
501 'LineWidth',2,'LineStyle',':','DisplayName','$\hat{\psi}ˆ3$');
502 plot(vId,RQp API,'Marker','*','MarkerSize',20,'Color','k',...
503 'LineWidth',1,'LineStyle',':','DisplayName','$API−01$');
504 plot(vId,RQp Dutch,'Marker','s','MarkerSize',20,'Color','k',...
505 'LineWidth',1,'LineStyle',':','DisplayName','$Dutch \,M.$');
506 plot(vId,RQp Fugro,'Marker','>','MarkerSize',20,'Color','k',...
507 'LineWidth',1,'LineStyle',':','DisplayName','$Fugro−05$');
508 plot(vId,RQp ICP,'Marker','d','MarkerSize',20,'Color','k',...
509 'LineWidth',1,'LineStyle',':','DisplayName','$ICP−05$');
183

510 plot(vId,RQp NGI,'Marker','h','MarkerSize',20,'Color','k',...


511 'LineWidth',1,'LineStyle',':','DisplayName','$NGI−05$');
512 plot(vId,RQp UWA,'Marker','x','MarkerSize',20,'Color','k',...
513 'LineWidth',1,'LineStyle',':','DisplayName','$UWA−05$');
514 %
515 subplot(1,2,2)
516 plot([0 0.5 1],vQs DJK./vQs DJK,'LineStyle','−','Color','k',...
517 'LineWidth',2,'DisplayName','Dijkstra (2009)');
518 axis([0.3,0.8,0,2.5])
519 set(gca,'XDir','reverse')
520 axis square
521 set(gca,'fontsize',20)
522 title ( 'Calculated to measured shaft capacities','fontsize',22);
523 xlabel( '$$I d$$','interpreter','latex','fontsize',30);
524 ylabel( '$$Q sˆ{calculated}/Q sˆ{measured} $$','interpreter','latex',...
525 'fontsize',30);
526 hold on
527
528 plot(vId,RQs K0,'Marker','.','MarkerSize',40,'Color','k',...
529 'LineStyle',':','LineWidth',2,'DisplayName','$K 0$');
530 plot(vId,RQs fall,'Marker','+','MarkerSize',25,'Color',pcolors(6,:),...
531 'LineWidth',2,'LineStyle','−−','DisplayName','$\hat{\psi}ˆ1$');
532 plot(vId,RQs f4s,'Marker','o','MarkerSize',25,'Color',pcolors(3,:),...
533 'LineWidth',2,'LineStyle','−.','DisplayName','$\hat{\psi}ˆ2$');
534 plot(vId,RQs f3s,'Marker','ˆ','MarkerSize',25,'Color',pcolors(4,:),...
535 'LineWidth',2,'LineStyle',':','DisplayName','$\hat{\psi}ˆ3$');
536 plot(vId,RQs API,'Marker','*','MarkerSize',20,'Color','k',...
537 'LineWidth',1,'LineStyle',':','DisplayName','$API−01$');
538 plot(vId,RQs Dutch,'Marker','s','MarkerSize',20,'Color','k',...
539 'LineWidth',1,'LineStyle',':','DisplayName','$Dutch \,M.$');
540 plot(vId,RQs Fugro,'Marker','>','MarkerSize',20,'Color','k',...
541 'LineWidth',1,'LineStyle',':','DisplayName','$Fugro−05$');
542 plot(vId,RQs ICP,'Marker','d','MarkerSize',20,'Color','k',...
543 'LineWidth',1,'LineStyle',':','DisplayName','$ICP−05$');
544 plot(vId,RQs NGI,'Marker','h','MarkerSize',20,'Color','k',...
545 'LineWidth',1,'LineStyle',':','DisplayName','$NGI−05$');
546 plot(vId,RQs UWA,'Marker','x','MarkerSize',20,'Color','k',...
547 'LineWidth',1,'LineStyle',':','DisplayName','$UWA−05$');
548
549 if ishandle(g)
550 eval(strcat('saveas(g,''Comp DJK f Emp',dateF,''',''fig'');'));
551 close(g)
552 end
553
554 %total capacities
555 g1 = figure( 'Name', 'Comparison with DJK' );
556 plot([0 0.5 1],vQt DJK./vQt DJK,'LineStyle','−','Color','k',...
557 'LineWidth',2,'DisplayName','Dijkstra (2009)');
558 axis([0.3,0.8,0.,2])
559 set(gca,'XDir','reverse')
560 axis square
561 set(gca,'fontsize',20)
562 title ( 'Calculated to measured tip capacities','fontsize',22);
563 xlabel( '$$I d$$','interpreter','latex','fontsize',30);
564 ylabel( '$$Q tˆ{calculated}/Q tˆ{measured} $$','interpreter',...
565 'latex','fontsize',30);
566 hold on
184 C. Calculation of Pile Bearing Capacity Using Empirical Methods

567 plot(vId,RQt K0,'Marker','.','MarkerSize',40,'Color','k',...


568 'LineStyle',':','LineWidth',2,'DisplayName','$K 0$');
569 plot(vId,RQt fall,'Marker','+','MarkerSize',25,'Color',...
570 pcolors(6,:),'LineWidth',2,'LineStyle','−−','DisplayName',...
571 '$\hat{\psi}ˆ1$');
572 plot(vId,RQt f4s,'Marker','o','MarkerSize',25,'Color',...
573 pcolors(3,:),'LineWidth',2,'LineStyle','−.','DisplayName',...
574 '$\hat{\psi}ˆ2$');
575 plot(vId,RQt f3s,'Marker','ˆ','MarkerSize',25,'Color',...
576 pcolors(4,:),'LineWidth',2,'LineStyle',':','DisplayName',...
577 '$\hat{\psi}ˆ3$');
578 plot(vId,RQt API,'Marker','*','MarkerSize',20,'Color','k',...
579 'LineWidth',1,'LineStyle',':','DisplayName','API−01');
580 plot(vId,RQt Dutch,'Marker','s','MarkerSize',20,'Color','k',...
581 'LineWidth',1,'LineStyle',':','DisplayName','Dutch M.');
582 plot(vId,RQt Fugro,'Marker','>','MarkerSize',20,'Color','k',...
583 'LineWidth',1,'LineStyle',':','DisplayName','Fugro−05');
584 plot(vId,RQt ICP,'Marker','d','MarkerSize',20,'Color','k',...
585 'LineWidth',1,'LineStyle',':','DisplayName','ICP−05');
586 plot(vId,RQt NGI,'Marker','h','MarkerSize',20,'Color','k',...
587 'LineWidth',1,'LineStyle',':','DisplayName','NGI−05');
588 plot(vId,RQt UWA,'Marker','x','MarkerSize',20,'Color','k',...
589 'LineWidth',1,'LineStyle',':','DisplayName','UWA−05');
590 if ishandle(g1)
591 eval(strcat('saveas(g1,''Comp DJK f Emp Qt',dateF,''',''fig'');'));
592 close(g1)
593 end
594
595 toc %total calculation time
Acknowledgements

First and foremost I would like to thank my supervisors Frits van Tol and Ron-
ald Brinkgreve for their continuous support, enthusiasm and encouragement since
the very beginning of my research. Ronald, your translation of the Summary and
Propositions into Dutch is gratefully acknowledged.
The thesis received partial funding and support from Plaxis bv. I’m indebted to
many of my co-workers at Plaxis b.v. who made all the facilities available to me and
sincerely extended help with any issues related to the use of the Plaxis software. I
owe great debt of gratitude to Paul Bonnier for his continuous support throughout
my research. I also would like to thank Ed Hartman for his enormous time-saving
help, through which the numerical technique to model the pile penetration could be
automated.
Special thanks to David Mašı́n for his support and feedback with the use of the
hypoplastic model.
I’d like to thank all my colleagues at the Geo-Engineering Section, in particular:
Frans Molenkamp, Jelke, Wout, Dominique, Alex, Christiaan, Jon, Patrick, Nathan-
ael, Michel, Daniele and Anders who all contributed in several ways to this research.
Thanks to Arno and Han for their assistance during the use of laboratory facilities.
Furthermore, I would like to thank Neslihan Özmen Eryılmaz, Onur Pekcan,
Vahid Galavi, Nallathamby Sivasithamparam, Joris Vergeest, Gordon Fenton, Caner
Anaç and Maaike Belien for their help and suggestions.
I’d like to thank all colleagues at the Computational Geomechanics Division,
NGI Oslo who made my secondment period quite productive and pleasant.
Many thanks to my friends Ümit, Pelin, Alper, Burak and Fırat travelling all
the way to Delft and supporting me in various ways. Burak, your final touch to the
cover design made this thesis complete.
I’d also like to thank my family, who have kept supporting me remotely in many
ways.
My deepest gratitude goes out to Oğuzhan Çopuroğlu who has always inspired
me, and who, in a way, shifted my career path towards Delft. Together with my
sister Emine and my niece Elif, your sincere support, guidance and suggestions made
this period not only productive but also joyful.
Finally, my beloved wife, Jona, like all difficulties and challenges we have sur-
mounted together, without your continuous support and sincere help, this thesis
would not have been possible.

185
186 Acknowledgements
Curriculum Vitae

Harun Kürşat Engin was born on April 13, 1980, in Ankara, Turkey. In 1997 he
graduated from İzmir Salih Dede High School, Turkey. From 1997-2001 he studied
Civil Engineering at the Middle East Technical University (METU ), Turkey. In
2002 he started to work as research and teaching assistant in the Soil Mechanics
Laboratory at METU. During this period he studied the effect of soil nails under
footings as settlement reducers using a 1g model test. He obtained his Master of
Science degree, from same department (METU ) in 2005. In 2006, he visited Geo-
Engineering Section of the Delft University of Technology and Plaxis bv as a guest
researcher for 3 months. In this period he worked on the improvement and validation
of ‘Embedded Piles’. In February 2008 he started to work as a research assistant
at the Geo-Engineering Section of the Delft University of Technology to study the
effects of pile installation numerically, the results of this work are described in this
thesis. He presented this work in several international symposia and conferences. In
2009 a 3 months period of research at the Norwegian Geotechnical Institute (NGI )
was financially supported by the GeoINSTALL project (FP7/2007-2013 under grant
agreement PIAG-GA- 2009-230638). There he studied the effect of pile installation
in clay in the Computational Geomechanics Group, the results of which were also
presented in an international conference.

187
188 Curriculum Vitae
Publications

Conference and Journal Publications:

Engin, HK, Brinkgreve, RBJ & Tol, AF van (2013) Simplified Numerical Modelling of Pile Penetration
- the Press- Replace Technique, Int. J. An. and Num. Methods in Geomechanics, accepted for
publication. (Chapter 4)
Ngan-Tillard, DJM, Verwaal, W, Mulder, A, Engin, HK & Ulusay, R (2011). Application of the needle
penetration test to a calcarenite, Maastricht, the Netherlands. Engineering Geology, 123(3), 214-
224.
Engin, HK & Andresen, L (2011). Comparison of zipper type techniques for finite element analysis of
pile penetration problem. In N Khalili & M Oeser (Eds.), IACMAG 2011 (pp. 48-53). Sydney:
Centre for Infrastructure Engineering and Safety.
Engin, HK & Andresen, L (2011). Use of explicit solver for finite element analysis of a quasi-static
penetration problem. In N Khalili & M Oeser (Eds.), IACMAG 2011 (pp. 54-59). Sydney: Centre
for Infrastructure Engineering and Safety.
Engin, HK, Brinkgreve, RBJ & Tol, AF van (2011). Numerical analysis of installation effects of pile
jacking in sand. In S Pietruszczak & GN Pande (Eds.), Computational Geomechanics Comgeo II
(pp. 744-755). Rhodes, Greece: IC2E.
Ngan-Tillard, DJM, Engin, HK, Verwaal, W & Mulder, A (2011). UCS estimation for Maastricht
limestones by needle penetration testing. In A Anagnostopoulos, M Pachakis & C Tsatsanifos
(Eds.), Geotechnics of hard soils - Weak rocks (pp. 235-239). Amsterdam: IOS Press BV.
Pham, HD, Engin, HK, Brinkgreve, RBJ & Tol, AF van (2010). Modelling of installation effects of
driven piles using hypoplasticity. In T Benz & S Nordal (Eds.), Numerical methods in geotechnical
engineering Numge 2010 (pp. 261-266). Leiden: CRC Press.
Brinkgreve, RBJ, Engin, E & Engin, HK (2010). Validation of empirical formulas to derive model
parameters for sands. In T Benz & S Nordal (Eds.), Numerical methods in geotechnical engineering
Numge 2010 (pp. 137-142). Leiden: CRC Press.
Engin, HK & Brinkgreve, RBJ (2009). Investigation of pile behaviour using embedded piles. In M
Hamza, M Shahien & YE Mossallamy (Eds.), Proceedings of the 17th International Conference
on Soil Mechanics and Geotechnical Engineering ( pp. 1189-1192). Amsterdam: IOS press.
Engin, HK, Septanika, EG, Brinkgreve, RBJ and Bonnier, PG (2008). Modelling piled foundation by
means of embedded piles. 2nd Int. Workshop on Geotechnics of Soft Soils - Focus on Ground
Improvement. 3-5 Sep. 2008, Uni.of Strathclyde, Glasgow, Scotland.
Engin, HK, Brinkgreve, RBJ and (2008) Estimation of pile group behaviour using embedded piles,
Proc. of the 12th International Conference of International Association for Computer Methods
and Advances in Geomechanics (IACMAG), Goa, India.
Ergun, M.U., Kul, İ and Engin, HK(2008) Use of Inclusions under Footings and Rafts as Settlement
Reducers A Model Study, BGA International Conference on Foundations ICOF 2008, Dundee,
Scotland.
Ülgen, D and Engin, HK (2007) Predicting Earthquake -Induced Displacements of Slopes, 8th Pacific
Conference on Earthquake Engineering, Singapore.
Ülgen, D and Engin, HK (2007) A Study of CPT Based Liquefaction Assessment Using Artificial Neural
Networks, 4th International Conference on Earthquake Geotechnical Engineering, Thessaloniki,
Greece.
Engin, HK (2007) Çekme Kazıklarının Gömülü Kazık Elemanlarıyla Modellenmesi, TMMOB İnşaat
Mühendisleri Odası 2.Geoteknik Sempozyumu, Adana, Turkey.

189
190 Publications

Engin, HK, Septanika, EG & Brinkgreve, RBJ (2007). Improved embedded beam elements for the
modelling of piles. In G.N. Pande & S. Pietruszczak (Eds.), Numerical Models in Geomechanics
(pp. 475-480). London: Taylor & Francis Group.

Reports, Manuscripts and Other Publications:

Engin, HK (2009) NGI Internal report: TNO-T01 Running ABAQUS/Explicit for static problems,
October 2009.
Engin, HK (2006) Plaxis Internal Report on Embedded Piles, April 2006.
Engin, HK (2008) Plaxis 3D Foundation Validation Manual Version 2, Validation of Embedded Piles-The
Alzey Bridge Pile Load Test, 2008.
Engin, HK (2007) Plaxis Internal Report on Tension Pile Testing Using Embedded Piles, February 2007.
Engin, HK (2005) Use of Soil Nails under Footings as Settlement Reducers, MS Thesis, METU, Ankara,
Turkey.

You might also like