Download as pdf or txt
Download as pdf or txt
You are on page 1of 261

Design of Piles

in Europe

How did Eurocode 7


change daily practice ?

International Symposium
28 and 29 April 2016
Leuven, Belgium

Volume I

Organised by :

Belgian Member Society


ETC3 - Piles of ISSMGE

Edited by :

M. De Vos, N. Huybrechts and M. Bottiau


Design of Piles in Europe
How did Eurocode 7 change daily practice ?

International Symposium
28 and 29 April 2016
Leuven, Belgium

Volume I

Edited by

Monika De Vos
Belgian Building Research Institute (BBRI-CSTC-WTCB), Belgium

Noël Huybrechts
Belgian Building Research Institute (BBRI-CSTC-WTCB), Belgium
KU Leuven, Belgium

Maurice Bottiau
Franki Foundations Belgium
Chairman ISSMGE-European Technical Committee 3 – Piles
Organised by

Belgian Member Society


ETC3 - Piles of ISSMGE
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table of contents of VOLUME I

Introductory address 1
M. Bottiau, N. Huybrechts, M. De Vos

Organisation 3

Contributions

Towards the second generation of Eurocode 7 7


A. Bond

Design of piles according to Eurocode 7 – Expected evolutions 15


C. Moormann

Discussion Session 1 – Pile design based on calculation methods


General Report – Calculation methods based on soil parameters 29

General Report – Calculation methods based on direct derivation from in situ tests 31
A. Van Seters

Keynote lecture – Design examples: Comparison of different national practices 47


T. Orr

Keynote lecture – Recent experiences with static pile load testing on real job sites 63
J. Verstraelen, W. Maekelberg, M. Medaets

Discussion Session 2 – Pile design based on pile load tests

General Report – Design methods based on static pile load tests 87


S. Burlon

General Report – Design methods based on rapid pile load tests 97


M. Brown
General Report – Design methods based on dynamic pile load tests 105
O. Klingmüller, M.Schallert

Special lecture – Accounting for cyclic loadings on piles with Eurocodes 109
A. Puech, S. Burlon
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Discussion Session 3 – Structural capacity, material and execution related aspects and other
loadings

Keynote lecture – Execution of piles in Europe, standardisation issues and 129


future perspectives
C. Gilbert

General Report – Execution related aspects and structural codes and previous 131
experiences: the need for calibration
D. Selemetas

Keynote lecture – Concrete for deep foundations 147


K. Beckhaus

Special lecture – Ground improvement vs. Pile Foundations ? 157


S. Varaksin, B. Hamidi, N. Huybrechts, N. Denies

Discussion Session 4 – Monitoring, quality control and testing practice

General Report 207


J. Powell

Keynote lecture – New advances and innovations in measurement techniques and 209
quality control tools
N. Huybrechts, M. De Vos, G. Van Lysebetten

Special lecture – Research and Development activities on pile foundations in Europe 235
K. Gavin, D. Igoe, K. Kataoka Sorensen, N. Huybrechts
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Introductory address

In April 1997, under the chairmanship of ir. Flor De Cock, ERTC3 organised a seminar on
the “Design of Axially Loaded Piles, European Practice”. This seminar gave a
comprehensive overview of the daily practice in 17 different countries, through National
reports including national geology, piling practice and design methods. For many
engineers, the proceedings of this seminar were used as a real reference book over many
years. The seminar also shed new light on the differences and the similarities between the
different approaches, helping the geotechnical engineers to better understand and compare
their respective methods.

Nearly twenty years have passed and since that time, Eurocode 7 has been introduced. As
one knows, Eurocode 7 gives a framework where design methods should fit in, but does not
provide any precise rules. For this reason, National Annexes have been drafted and are now
(being) introduced in most European countries defining the detailed calculation rules to be
applied. In some cases, these were in direct continuity with the existing rules. In other
countries, Eurocode 7 was a real occasion to question the practice. A lot of research work,
including field and model tests as well as pile-soil models have been undertaken in order to
adapt the national rules. For all these reasons, ETC3 members thought that the moment was
opportune to present an update on how pile design codes have evolved in the different
European countries, on the basis of EC7. This is why this symposium was put in place.

The first objective of the symposium is thus to give the European geotechnical engineers
the opportunity to familiarise themselves with those (new) various national design methods.
As in 1997, each country represented in the ETC3 committee was invited to elaborate a
detailed report concerning their national design methods. This includes axially or laterally
loaded, individual piles or pile groups,… These national reports are published in Vol. II of
the proceedings. During the symposium, we have decided to present a thematic analysis of
the National Reports instead of an exhaustive presentation of the National reports, enabling
a better comparison. Further to this, we have been working on design examples, in order to
try to identify the differences between the methods using practical cases. This exercise,
summarized by Trevor Orr, promises to be an exciting challenge.

In addition to that, the Organising Committee wanted to present an overview of the future
evolution of pile design according to Eurocode 7 (CEN/TC 250/SC 7), the current status of
standardisation on pile execution, testing in Europe (CEN/TC 288 & 341/ISO TC 182) and
of the practical application of static, dynamic and rapid load testing in pile design.

1
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Additional topics such as quality control, monitoring and new measuring techniques are
also covered. Finally, special attention will be paid to the requirements regarding concrete
for deep foundations and recent research and analysis of the working group of EFFC-DFI
on this topic will be highlighted.

Exchange of knowledge and better understanding of the various national design methods
will help geotechnical engineers to interpret the results obtained, enlarge their view on pile
behaviour and encourage, to some extent, a convergence of the European design rules. We
also formulate the hope that the debate between many practitioners and specialists during
this week will also address the challenges of drafting better codes for the future, codes
which would encourage design based on reliability as opposed to design based on
prescriptive codes obviously generalizing systems and soils, and leading to (relatively) high
safety factors reflecting the low reliability of execution. As we all know, the final global
performance of deep foundation largely depends on local conditions and installation related
parameters determining the soil response to the pile installation. In order to reflect this, our
codes will need to request an adequate understanding and documentation of the local soil
conditions, a correct specification of the expected performance (capacity, deformations, …),
the proper in-situ documentation of pile installation and pile performance control and
monitoring, including testing if necessary. Our codes could, if all this is achieved, evolve
towards lower safety factors reflecting this increase in safety. This challenge lies ahead
of us.

Leuven, April 2016

ir. Maurice Bottiau, Franki Foundations


ETC3 Chairman, President of the Belgian Member Society of ISSMGE

ir. Noel Huybrechts, Belgian Building Research Institute


ETC3 Secretary, Secretary of the Belgian Member Society of ISSMGE

ir. Monika De Vos, Belgian Building Research Institute


On behalf of the Organisation Committee

2
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Organisation

ETC3 MEMBERS & LIAISONS

Chairman :
M. Bottiau, Franki Foundations Belgium

Secretary :
N. Huybrechts, Belgian Building Research Institute & KU Leuven

Members :
R. Hofmann, Austria
C. Kolev, Bulgaria
J. Kos, Czech Republic
O. Møller, Denmark
M. Mets, Estonia
P. Tolla, Finland
S. Burlon, A. Le Kouby, J.P. Volcke, France
C. Moormann, Germany
F. Scheuring, Hungary
D. Igoe, Ireland
A. Mandolini, Italy
K. Gavin, the Netherlands
A. Simonsen, Norway
K. Gwizdala, Poland
J. Santos, Portugal
J. Josifovski, Republic of Macedonia
A. Ponomaryov, Russia
L. Prieto, C. Fernandez-Tadeo, Spain
G. Axelsson, Sweden
M. England, United Kingdom

Liaisons :
T. Orr (ETC10)
A. Mandolini (TC 212)
A. Bond (CEN TC 250/SC 7)
C. Moormann (CEN TC 250/SC 7/WG 3/TG 3)
C. Gilbert (CEN TC 288)
J. Powell (CEN TC 341/ISO TC 182)

3
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

INTERNATIONAL ADVISORY COMMITTEE

The international advisory committee is composed of :

The ETC3 Members and Liaisons

A. Gens
European Vice President of ISSMGE

R. Frank
President of ISSMGE

F. De Cock
Former ETC3 Chairman

SYMPOSIUM ORGANISING COMMITTEE

The organisation of the Symposium is under the leadership of the ETC3 Chairman and
composed of :

N. Huybrechts & M. De Vos


Belgian Building Research Institute (BBRI-CSTC-WTCB)

W. Maekelberg & J.F. Vanden Berghe


Vice Presidents of the Belgian Member Society of ISSMGE (BGGG-GBMS)

M. Roovers
President of the Belgian Association of Foundation Contractors (ABEF)

4
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Contributions

5
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

6
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Towards the second generation of Eurocode 7


Dr Andrew Bond, Geocentrix Ltd, UK, Chairman of TC250/SC7, chairman@eurocode7.com

ABSTRACT

Work on the second generation of Eurocodes got underway in October 2015, with the planned delivery
date for new versions of ENs 1990 to 1999 in 2020 or shortly afterwards. This paper summarizes the
work that is now being done to create an improved version of Eurocode 7 that will be fit for the 21 st
century.

1. EUROPEAN COMMISSION MANDATE M/515


In December 2012, the European Commission issued a Mandate (M/515) for “amending existing
Eurocodes and extending the scope of Structural Eurocodes”, which would involve:

• development of new standards or new parts of existing standards;


• incorporation of new performance requirements and design methods;
• introduction of a more user-friendly approach in several existing standards; and
• a technical report on how to adapt the ... Eurocodes ... to take into account the relevant impacts of
future climate change.

Under this Mandate, EN 1997 (Eurocode 7 – Geotechnical design) was classified as a


‘framework/assessement’ Eurocode and placed in Work Package I, alongside ENs 1990 (The Eurocode –
Basis of structural design), 1991 (Eurocode 1 – Actions on structures), and 1998 (Eurocode 8 – Design of
structures for earthquake resistance).

The outcomes that the Commission would like to see from this work package are:

a) Significant reduction in the number of Nationally Determined Parameters (NDPs).


b) Refinement to improve the ‘ease of use’ of Eurocodes by practical users.
c) Incorporation of recent results of international studies and practical experience from scientific and
technical associations and results from research programmes relevant to innovation (including the
performance based and sustainability concepts in design and construction).
d) Incorporation of recent results ... relevant to contribution of structural design to sustainability.
e) Adoption of ISO standards to supplement the Eurocodes family.
f) Developing auxiliary guidance documents to facilitate feedback from stakeholders and the practical
local implementation wherever necessary.
g) Developing information on the determination of material and resistance factors, serviceability for
buildings and bridges; fatigue verification; improving the fire safety engineering approach (EN 1990).
h) Incorporating new developments in the field of traffic loads and climatic actions; atmospheric icing;
waves and currents (EN 1991).
i) Providing a clear and complete list of background documents used during the standardisation process.
j) Developing a technical report analysing and providing guidance for ... relevant impacts of future
climate change (general and material specific)

CEN, the European Standards Organization, submitted a technical and financial proposal to undertake this
work and, in December 2014, the European Commission confirmed funding of €4.5M to enable Phase 1
of the project to get underway in January 2015. The National Standards Body for the Netherlands (NEN)
won the contract to manage the project on behalf of CEN.

A further two Phases of work are anticipated in response to Mandate M/515, which will – subject to
approval by and funding from the European Commission – bring the total cost of satisfying Mandate
M/515 to over €10M by the end of 2019. This figure does not include the value of all the time volunteered
by members of technical committees, working groups, and task groups that have been formed to oversee
and review the contracted work.

Figure 1 shows the timeline for development of the second generation of Eurocodes over the peiod 2010
to 2020.

7
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 1: Timeline for development of the second generation of Eurocodes

2. INITIAL PREPARATION FOR THE NEXT VERSION OF EN 1997


Planning for the next version of Eurocode 7 began in 2011, when TC250/SC7 – the sub-committee that is
responsible for maintenance and development of Eurocode 7 – decided to setup a total of fourteen
‘Evolution Groups’. These groups were asked to undertake ‘blue-sky thinking’ with regards to the
changes that could be made to improve the practice of geotechnical design according to Eurocode 7.
Figure 2 shows the titles of the fourteen Evolution Groups (EG12 Tunnelling was planned but never
brought into existence).

Figure 2: SC7's Evolution Groups, 2011-15

The Evolutions Groups submitted their final reports in December 2015 and these documents have formed
the starting point for the development work that has since got underway, as explained next.

8
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3. CURRENT PLANS FOR DEVELOPING EN 1997

3.1. SC7’s six Tasks


As part of CEN’s response to Mandate M/515, in 2014, SC7 proposed undertaking six tasks in order to
achieve the desired outcomes from Work Package I:

1. Harmonization and ease-of-use of Eurocode 7


2. Improvements to Eurocode 7 Part 1 – General rules
3. Improvements to Eurocode 7 Part 2 – Ground investigation
4. Creation of Eurocode 7 Part 3 – Geotechnical constructions (slopes and foundations)
5. Creation of Eurocode 7 Part 3 – Geotechnical constructions (retaining structures)
6. Improved treatment of rock mechanics and dynamic design

Tasks 1 and 2 are included in Phase 1 of the overall programme of work; Tasks 3-5 in Phase 2; and Task
6 in Phase 3.

Two Project Teams (PTs) were appointed in August 2015 to deliver the outcomes from Tasks 1 and 2.
The members of the Project Teams were selected following an open tendering process organized by NEN
and overseen by TC250 and the European Commission. SC7 appointed a ‘Pre-Selection’ Panel –
comprising Roger Frank (FRA), Giuseppe Scarpelli (ITA), Jorgen Steenfelt (DNK), Ivan Vanicek (CZE),
Norbert Vogt (DEU), and myself (GBR, SC7 Chairman, ex-officio) – to choose the best candidates to
serve as members of the Project Teams.

3.2. Task 1: Harmonization and ease-of-use of Eurocode 7


The Project Team for Task 1 (PT1) comprises Sébastien Burlon (Team Leader, FRA), Loretta Batali
(ROM), Bernd Schuppener (DEU), Brian Simpson (GBR), Carsten Steen Sørensen (DNK), Vincenzo
Pane (ITA), Mark Lurvink (SC7 Secretary, ex-officio), and myself (SC7 Chairman, ex-officio).

The major part of Task 1 is to propose a new structure for EN 1997 – one that is easier to navigate; is
more consistent externally with the other (structural) Eurocodes and internally between its own Parts; and
provides more space for detailed design rules covering a wider range of topics.

The most visible element of this Task is the division of Eurocode 7 into three Parts:

1. EN 1997-1 Eurocode 7 – Geotechnical design: Part 1 – General rules


2. EN 1997-2 Eurocode 7 – Geotechnical design: Part 2 – Ground investigation*
3. EN 1997-3 Eurocode 7 – Geotechnical design: Part 3 – Geotechnical constructions

*Note the title will be shortened from Ground investigation and testing.

One proposal for the new structure of Part 1 is shown (in part) in Figure 3.

9
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 3: One proposal for the new structure of Eurocoode 7: Part 1 – General rules

The current proposal for the structure of the new Part 3 is shown (again, in part) in Figure 4.

Figure 4: Proposed structure of new Eurocode 7: Part 3 - Geotechnical constructions

A second major element of Task 1 is to dispense with the infamous Design Approaches from EN 1997-
1:2004 and to propose an alternative method of selecting partial factors for design that is clearer for
practicising engineers to use. The current proposal (based on the work of Evolution Group 8) is to
introduce different ‘Design Combinations’ that depend on the geotechnical construction being designed.

3.3. Task 2: Eurocode 7 – Part 1: General rules


The Project Team for Task 2 (PT2) comprises Gunilla Franzén (Team Leader, SWE), Marcos Arroyo
(ESP), Michael Kavvadas (GRC), Andrew Lees (CYP), Adriaan van Seters (NDL), Herbert Walter
(AUT), Mark Lurvink (SC7 Secretary, ex-officio), and myself (SC7 Chairman, ex-officio).

The major elements of Task 2 are to propose a method that will allow for reliability discrimination,
depending on the consequences of failure in a particular design situation (persistent, transient, etc.); to

10
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

provide a better definition of the characteristic value of a geotechnical parameter, taking into account the
amount of data available and the size of the foundation; to identify the best methods of establishing
design water pressures, whether through the application of partial factors, the introduction of safety
margins, or by some other means; and to provide clear rules for using advanced numerical models in day-
to-day geotechnical practice.

3.4. Task 3: Eurocode 7 – Part 2: Ground investigation


The Project Team for Task 3 (PT3) will be appointed before the end of 2016.

One early decision that PT3 will need to make is how to re-structure Part 2, if at all. SC7’s Evolution
Group 2 has proposed a radical slimming down of the contents of Part 2, without altering its existing
structure.

An alternative proposal that is currently under consideration by SC7 is to completely re-organize the
contents of Part 2 in terms of the parameters that a ground investigation can provide to the designer. The
initial idea for this proposal came from the Nordic Mirror Group to SC7 and has been developed further
by a small group within SC7. The possible new structure is shown in Figure 5.

Figure 5: Possible new structure of Eurocode 7: Part 2 – Ground investigation

3.5. Tasks 4 and 5: Eurocode 7 – Part 3: Geotechnical constuctions


The Project Teams for Tasks 4 and 5 (PTs 4 and 5) will be appointed before the end of 2016.

These two Tasks together are intended to create a new European standard entitled Eurocode 7:
Geotechnical design – Part 3: Geotechnical constructions. The current proposal for the structure of the
new Part 3 is shown in Figure 4. A key aim of this work will be to ensure consistency between the various
sub-sections of the standard, by adopting a common sub-structure, currently expected to be similar to that
shown in Figure 6.

11
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 6: Proposed structure of each section within EN 1997-3

3.6. Task 6: Rock mechanics and dynamic design


The Project Team for Task 6 (PT6) is expected to be appointed sometime during 2017.

This Task was established to ensure that the treatment of rock mechanics and dynamic design in
Eurocode 7 is greatly improved – to such an extent that rock mechanics is seen as an equal bedfellow to
soil mechanics in the new code and the interface between Eurocode 8 (seismic design) and Eurocode 7 is
seamless.

4. REVIEWING/APPROVING THE WORK OF THE PROJECT TEAMS


The previous section of this paper outlines the very ambitious plans that SC7 has made for the next
generation of Eurocode 7. To ensure that the Project Teams deliver what is wanted, arrangements have
been made to provide technical support to the PTs as and when they need it. In addition, since SC7 retains
overall resposnsibility for the contents of EN 1997, it is important that it reviews the work of the PTs on a
regular basis and approves the final outcomes.

To this end, SC7 has now created three Working Groups to look after the separate Parts of the new
Eurocode. Task Groups have been created within those Working Groups, to help focus attention on
specific technical topics. The new organizational structure of SC7 is shown in Figure 7 and discussed
below.

Individual Task Groups have been established to align with the specific Tasks given in CEN’s response to
Mandate M/515. The Task Groups therefore ‘shadow’ the Project Teams that are contracted to do the
work. So, for example, Working Group 1’s Task Group 1 (WG/TG1) is responsible for supporting PT1
and reviewing its work on behalf of SC7.

12
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Each country within CEN has been asked to appoint delegates to serve on SC7’s Task Groups, on a
voluntary basis. As of the time of writing, over 200 geotechnical engineers have put their names forward
to help SC7 deliver on its promise of a better Eurocode 7, fit for the 21st Century.

Figure 7: Organizational structure of SC7 from 2016 onwards

5. CONCLUSION
A very ambitious programme of work is now underway that aims to deliver a vastly improved European
standard for geotechnical design. Major changes to the structure and content of EN 1997 are envisaged
that should lead to improved ease-of-use, greater consistency, and better harmonization between the
geotechnical practice in different European countries. A large body of experts have volunteered to assist
in this work, which will ensure a wide range of opinons will be heard on how to proceed.

My hope is that the second generation of Eurocode 7 will be a worthy successor to the orginal EN 1997,
which itself was a major achievement. As Isaac Newton said in 1676:

“If [we] have seen further, it is by standing on the shoulders of giants.”

REFERENCES

European Commission Enterprise and Industry Directorate-General, Brussels, 12 December 2012,


M/515 EN Mandate for amending existing Eurocodes and extending the scope of Structural Eurocodes.

CEN/TC 250, May 2013, Response to Mandate M/515, ‘Towards a second generation of EN Eurocodes’,
CEN/TC250, document N993

13
14
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Design of piles according to Eurocode 7 – Expected evolutions


Christian Moormann, University of Stuttgart, Institute for Geotechnical Engineering, Germany,
christian.moormann@igs.uni-stuttgart.de, christian.moormann@moormann-geotechnik.de

ABSTRACT

In progress of preparing the next generation of Eurocode 7 the Evolution Group EG 7 ´Pile Design´ of
TC250/SC7 has focused on the evoultion of Design of piles according to Eurocode 7 in the period 2011 to
2015. Based on the results of this period of collaboration and bringing together European experiences in
pile design the expected evolutions for the section ´Pile Foundations´ of the upcoming EC7-3are
presented in this contribution whereby several issues are still subject of ongoing dicussions and
developments. It is obvious that especially the design of pile foundations is still related to national
experiences resulting from diffferent geological conditions, different methods of soil investigation,
different type of piles commonly used as well as a wide field of different calculations methods often
related to the soil conditions/investigations and the pile type used. Nevertheless there is epecially for pile
design a realistic chance to achieve a further harmonizaton of the design principles even if the
exploration and calculation methods might differ still in future.

1. GENERAL
From 2011 to 2015 Evolution Group EG 7 ´Pile Design´ of TC250/SC7 had focused on the evoultion of
the existing design principles according to Eurocode 7 in order to prepare proposals for the further
evolution and improvement of the present section 7 of EC 7-1 called ´Pile Foundation´ for the next
generation of EC 7. Participants of this Evolution Group had been fifteen European experts representing
consulting, industry and universities from overall Europe.
The work of EG 7 was focused on:
a) Editing, changing and improving text of EN 1997-1, section 7 (mainly editorial task, e.g. cancel
doublings, resetting mistakes). Aim was to make the document much clearer and more user-
friendly.
b) Review of values for correlation factors, partial factors, model factors and action factors and
their combination in each country as this was considered to be of major interest for further
revision of EN 1997-1 in order to reduce the number of nationally determined parameters
(NDP).
c) Comparing and identification of proposals for harmonizing different design approaches for pile
design.
d) Comparing calculation methods for pile foundations for selected applications like axially loaded
piles, downdrag etc.
e) elaborate proposals for new aspects being relevant for engineering practice
In this context it was general understanding that the next generation of Eurocode 7 should focus on the
principles of design and safety concepts whereas recommendations might provide more detailed support
for engineering practice e.g. with different calculation methods, background information, continuative
literature etc.
After a preparatory two-year-period of more general and fundamental discussions and considerations EG
7 had worked on the present text of section 7 “Pile Foundations” since summer 2013 in order to make the
outcome of EG 7´s efforts as effective and valuable as possible and to result in a proposal of a revised
section for pile design.
The most relevant aspects of this working period of EG 7 are presented in this paper combined with some
general consideration for the further evolution of pile design according to Eurocode 7.

15
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2. HARMONISATION
Beside others activities EG 7 had focussed on harmonizing different design approaches and nationally
determined parameters like correlation factors, partial factors, model factors or action factors as this issue
is considered to be probably the most relevant one.
An analysis by Bond (2013) shows that the predominant number of countries is using the same design
approach (DA2 or DA2*) at least for the design of axially loaded piles (Figure 1). Therefore a general
harmonization in this regard is likely to be achieved.

Figure 1: Design approaches used for (axial) pile design in Europe (Bond 2013).

Nevertheless an analysis of the partial safety factors, model factors and correlation factors indicate
relevant differences even concerning the principal. As an example many countries do not follow the
proposal of EC 7-1 to apply different sets of partial safety factors on the pile resistance depending on the
type of pile. As well the approach for the correlation factors differs from country to country as discussed
in section 5.2 of this paper.
A comparison of the differing national approaches is reliably possible only on the basis of design
examples as all nationally determined parameters as well as the calculation methods used in engineering
practice influence the final pile design.
Whereas for axially loaded piles a harmonization of the design approach used might be achieved the need
for further discussion is greater in cases where the pile design is influenced by the displacement behavior
and therefore by the stiffness of the soil layer and the pile foundation. Typical examples are pile subjected
by negative skin friction, by horizontal soil movements (see Figure 2) or buckling. In such cases either
DA 2 or DA 3 is applied in different countries.

16
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 2: Design situations for piles stressed by lateral soil movements: a) piled slopes (dowels), b)
heavy surcharge loads, c) excavations, d) piled bridge abutments with inclined piles.

Figure 3: Numerical calculation of a pile foundation for a bridge abutment influenced by horizontal soil
movements with application of Design Approach DA2* (simulation with characteristic values).

17
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

DA 2 provides the benefit that with the characteristic values used for soil stiffness and strength and for
the pile stiffness as well as for the loads/stresses the characteristic displacements as well as the
characteristic inner forces of the pile foundation is calculated, but does involve the shortcoming that the
variability of soil conditions can only be covered when a range is considered for the input parameter. DA
2 has also advantages when numerical simulations are applied (Figure 3).
DA 3 on the other hand has the benefit that the influence of the variability of the soil parameter is
considered when calculating the displacements and inner forces, but is related to the shortcoming that
there are actually no partial safety factors available to be applied on soil or pile stiffness. The application
of DA 3 might also be difficult in cases where numerical approaches are used with advances soil models
depending on stress and strain rates.
In summary major efforts is requested in evaluating, harmonizing and further developing these issues for
the next generation of EC 7.

3. NEW STRUCTURE FOR SECTION ´PILE FOUNDATIONS´


Within the framework of EG 7 a new structure for the section ´Pile Foundations´ of EC 7 was elaborated
which is documented in Figure 4 in comparison to the structure of present Section 7.
The revised structure is adjusted on the first level to the principal structure of future EC 7-3 dealing with
geotechnical structures. The new structure is characterized by the basic idea to distinguish between single
standing piles, pile groups and piled rafts on the second level and on a third level between axially and
transversely loaded piles. The present differentiation between compressive and tensile resistance of
axially loaded piles will be relinquished and integrated in the subsections for axially loaded piles; this
helps to shorten the text of the standard.
Several new subsections as e.g. for ´Pile resistance due to cyclic, dynamic and impact loads´ are planned
and will have to be elaborated and incorporated into the new structure.

4. EASE OF USE: REVISION OF SECTION ´PILE FOUNDATIONS


From August 2013 on the work of EG 7 had been focused on preparing a revised section 7 “Pile
Foundations” with concrete and detailed recommendations for making this section more user-friendly as
well as for the further evolution of this chapter. In this context stepwise the present text of section 7 was
reviewed and complemented analysing the four questions:
• Which clauses should remain unchanged in the next edition?
• Which clauses should be deleted from the next edition 7? And why?
• Which of those clauses should be changed in the next edition? What changes should be made?
And why?
• What new clauses or even subsections should be added in the next edition? And why?
The review has followed the idea that a code - in sense of improving ease-of-use - should focus on
principles for design and safety concepts as well as basics rules for pile design avoiding textbook-like
explanations and description. Figure 5 shows an excerpt as example.
It looks quite promising that the aim to come up with a more user-friendly version of the section “Pile
Foundations” could be reached as the clauses revised could be focused on the essential principals.
These proposals and recommendations will later be used by the Project Team which will be mandated
with writing this section of the new code.

18
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 4: Present (left) and proposal for new (right) structure of section ´Pile Foundations´ of EC 7

5. FURTHER EVOLUTION OF PILE DESIGN ACCORDING TO EC 7

5.1. Proposals for new subjects of section ´Pile Foundations´


By preparing a revised section ´Pile Foundations´ stepwise the present text of section 7 of EC 7-1 is
reviewed and concrete and detailed recommendations for the evolution of this section are formulated. As
outcome of this evolution process it was agreed that for some aspects of pile design more guidance is
needed for engineering practice.
These aspects which are presently not covered appropriately should be included in more detail or even for
the first time in a future version of EC 7. Beside others such relevant aspects have been identified as
follows:
• calculation and design methods for transversally loaded piles,
• consideration of downdrag / negative skin friction (proposal elaborated, see section 6),
• ground heave,
• piles effected by lateral ground movements,
• calculation methods and design of pile groups for axial and horizontal loads,
• design of piled rafts including cases where piles are used as ´settlement-reducer´,
• pile resistance and design for dynamic, cyclic and impact loads,

19
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 5: Example for revised clauses of earlier section 7 ´Pile Foundations´ as part of review work of
TC 250/SC 7/EG 7: focusing on design basics in sense of improving ease-of-use.

• seismic design of pile foundations,


• proof of the serviceability of pile foundations,
• structural design of piles for lateral loading and buckling.
As an example section 6 of this paper presents a new approach for designing piles in case of downdrag /
negative skin friction.
Some considerations concerning the issue whether (most commonly used) calculation methods for axially
and transversally loaded single piles should be included in the next generation of EC 7-3 are addressed in
the following section 5.3.
Additionally to these issues which are not covered appropriately by the present EC 7-1 other issues were
identified which needs major revision. An example is the present approach for the correlation factors ξ i as
actually requested by EC 7-1.

5.2. Revised approach for correlation factors ξ i


In determining the design values of a pile resistance from the results of static and dynamic pile load
testing as well as from results from ground investigation the measured resp. calculated values R c;m of the
pile capacity have to be initially mitigated by correlation factors ξ i to derive the characteristic pile
resistance R c;k . With the correlations factors, the influence of the spatial variability of the subsoil, but also
of the uncertainties arising from the execution and evaluation of the pile load tests on the derivation of the
characteristic pile resistance from the measured values of a single or limited number of test loads is taken
into account. An analysis of the existing approach implemented in EC 7-1 indicates shortcomings
particularly with respect to the sole dependency of the correlation factors on the number of performed
load tests. Whereas correlation factors ξ i to derive characteristic pile resistances from static and dynamic
pile load tests are used quite frequently, correlation factors to derive characteristic pile resistances from
ground tests results (Table A.10 of EC7-1) are used only in a few countries (Table 1).
In this regard approaches used in different European countries were reviewed and proposals for an
evolution of the present regulations were discussed.
Beside countries using just the correlation factors proposed by EC 7-1 some countries have adaptions
which base on this EC 7-1 approach but consider additional aspects like the German approach that
specifies the correlation factors ξ 5,6 for dynamic pile load tests in dependency of the calibration of the
dynamic pile load tests with static pile load tests (DGGT 2012).

20
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 1: Correlation factors ξ i to derive characteristic pile resistances from static pile load tests (Table
A.9 of EC7-1) and from ground tests results (Table A.10 of EC7-1).

Other national standards have approaches for the correlation factors which do not only consider the
number of test piles or of ground tests but also the area (“surface”) for which this tests have been
executes. Thus the French code for pile foundations NF P 94-262 does refer the correlation factors to a
reference area of S ref = 2,500 m² and allows to reduce the correlation factors for smaller areas (Figure 6).
Such an approach does consider as appropriate that the spatial variability of pile resistances decrease for
smaller pile foundations when the same number of piles is tested.
The Dutch standard NEN 6743:1991 and the Belgian BBRI-Report 12 propose approaches for the
correlation factors ξ 3,4 which consider both the intensity of soil improvement as well as the number of
foundation piles for which this evaluation is applied (Table 2).

French Standard for piles (NF P 94-262):

with S = 2500 m² and 625 m² < S < S


ref ref


ξ1,2 (N, S) = 1 + �ξ ´1,2 (𝑁𝑁) − 1� ∙ �𝑆𝑆�𝑆𝑆
𝑟𝑟𝑟𝑟𝑟𝑟

Figure 6: Correlation factors ξ 1 and ξ 2 for deriving the characteristic pile resistance from static pile
load tests depending according to the French standard NF P 94-262 depending on the number of tested
piles (n) and on the ground area S of the pile foundation.

21
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 2: Values of the correlation factors ξ 3 and ξ 4 to derive characteristic values from CPT-tests
according to Belgian guidelines (BBRI-Report 12).
a) correlation factor ξ 3

Number of 1 CPT / 1 CPT / 1 CPT / 1 CPT / 1 CPT /


piles 10 m² 50 m² 100 m² 300 m² 1000 m²
1-3 1.25 1.29 1.32 1.36 1.40
4-10 1.15 1.19 1.21 1.25 1.29
> 10 1.14 1.17 1.20 1.24 1.27

b) correlation factor ξ 4

Number of 1 CPT / 1 CPT / 1 CPT / 1 CPT / 1 CPT /


piles l 10 m² 50 m² 100 m² 300 m² 1000 m²
1-3 1.08 1.17 1.23 1.31 1.40
4-10 1.00 1.07 1.13 1.21 1.29
> 10 1.00 1.06 1.12 1.20 1.27

Based on this experiences the improvement planned for the approach of correlation factors ξ i can be
summarized as follows:
• relate ξ 3/4 (ground test results) to investigations per area
• relate ξ 1/2 and ξ 5/6 (load tests) to percentage of piles tested ./. total number of piles or to the area
of the pile foundation
It has also to be discussed whether correlation factors should vary in dependency of the pile type.
Generally the reliability of dynamic load tests is considered to be lower than for static pile load tests and
therefore higher correlation factors have to be applied for dynamic pile load tests, a fact that is covered by
the present regulation of EC 7-1 already with ξ 1,2 being significantly smaller than ξ 5,6 already. But results
of a round robin test with dynamic pile load tests executed recently on bored piles conditions in sandy soil
near Berlin (Baeßler et al. 2012; Herten et al. 2013) indicated also that in this case a quite large scattering
due to the analysis of the results of dynamic pile load tests may occur, which makes dynamic pile load
tests more reliable for prefabricated driven piles than for cast-in-situ driven or bored piles.
Generally the basis concept of correlation factors which consider the extent and intensity of soil
investigation when defining the requested safety factor is considered to be a motivation for adequate soil
investigation programs. Therefore it might be an example for other applications (slope stability …) as
well.

5.3. Calculation methods for pile design


In process of future evolution of the section ´Pile Foundation´ for the next generation of Eurocode 7 it
will have to be discussed in more detail whether calculation methods might be added as an informative
annex to EC 7-3.
Such information would enable engineers in practice under ideal circumstances to calculate and design
pile foundations just on basis of using EC7-3. On the other hand especially for pile design it has to be
considered that due to different geological conditions, different type of piles used and different long-term
design practice even for frequently used types of pile like bored piles or precast driven piles many
different calculations methods are presently used in Europe, which differ concerning their basic
approaches, their range of proven application, the input parameters needed etc.
Therefore – if possible at all - the documentation of calculation methods for piles in an informative annex
to EC 7-3 could be realized only for simple tasks like calculation the axial pile resistance for usual pile
systems (to be identified) and for most frequently calculation methods used in Europe (and not only in a
single country).
A benefit of selected calculation methods documented in an informative annex to EC 7-3 could be that
model factors related to these methods could be stated as well (at least in a range) what would be a
contribution to a further harmonization of pile design in Europe. As already mentioned this issue will
have to be discussed in more detail in future. In any case joint efforts on a European basis might become
necessary to identify and to validate such calculation methods once a positive decision for such an
informative annex is taken.

22
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

6. EXAMPLE: NEW SECTION FOR NEGATIVE SKIN FRICTION


An example for a new subject identified and subsicently elaborated by EG 7 in preparation of the next
generation of the section `Pile Foundation´ of Eurocode 7 is the new subsection on negative skin friction
resp. downdrag. The new approach considers both serviceability and ultimate limit state design and is part
of the proposal of EG 7 for a revised section ´Pile Foundations´. The subsection has been drafted by EG 7
as follows:
(1) Negative skin friction has to be regarded as a permanent action F n , originating from relative axial
movement between the ground and the pile, when the ground settles more than the pile.
(2) The pile continues to settle until the actions from negative skin friction t n , together with the actions
imposed on the pile by the superstructure, and the pile resistances resulting from the pile end
bearing capacity and supporting skin friction qs, are in equilibrium.
(3) Negative skin friction should be taken into account for the justification at SLS and ULS.
(4) Both for SLS and ULS, the characteristic values of negative skin friction should be estimated by
considering the pile loading level with appropriate calculation models, see (5), accounting for
relevant strain mechanisms between the pile and the soil surrounding. For simple cases
approximates approach can be used, see (6).
(5) An appropriate model to calculate negative skin friction is to take into account up to the neutral
point marking the boundary between positive and negative skin friction. Two separate neutral points
should be considered for the ULS and SLS (Figure 7). In the case of SLS the neutral point is the
point of the theoretical zero relative movement between the pile and the settling soil considering
total possible settlement (e.g. primary and secondary consolidation). In the case of the ULS the
neutral point should be moved up relative to the SLS neutral point.
To calculate the negative skin friction t n,k information are required on:
- pile settlements with depth;
- soil strata settlements with depth;
- the resulting relative movements and
- any mobilisation functions of t n,k and q s,k .

Figure 7: Evaluation of negative skin friction for ultimate limit state (ULS) and serviceability limit state
(SLS).

Comparing the relative displacements from pile settlement spile and the settlement of the
surrounding soil sground gives the location of the neutral point and thus the value of the
characteristic action Fn,k in the serviceability limit state and in the ultimate limit state.
(6) For simple cases the following approximate approaches can be used alternatively to (5):

23
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

a) In case that at ultimate limit state the settlements of the pile can be assumed to be greater than
the settlement of the surrounding soil the neutral point can be assumed to be located at the
ground surface resp. at top of the pile. Thus the pile resistance can be proven for the ultimate
limit state without calculating a negative skin friction.
b) In case that at ultimate limit state the settlements of the pile has to be considered to be smaller
than the settlement of the surrounding soil (example: “rock socket piles”) the neutral point can
be assumed to be located at the bottom of the settling soil layer. Thus the pile resistance can be
proven for the ultimate limit state by assuming characteristic values of negative skin friction for
the settling soil layers.
c) For serviceability limit state the calculation of the neutral point can be substituted by assuming
characteristic values of negative skin friction for the settling soil layers, thus considering the
neutral point to be located at the bottom of the settling soil layers. This value t n,k should be an
upper bound value and be determined on the safe side.
(7) Approaches for deriving the characteristic negative skin friction t n,k are given in Annex A.
(8) Normally, downdrag and transient loading need not be considered simultaneously in load
combinations.
(9) The pile resistance has to be proven analysing the ultimate and serviceability limit state:
a) Serviceability limit state (SLS): the characteristic action F n,k (SLS) and the location of the neutral
point have to be calculated by the deformation behaviour associated with the pile settlement
spile and the settlements in the soft stratum sground. The design value of the effects is:
F d = F k = F G,k + F n,k (SLS) + F Q,rep (1)
Additionally it should be proven that the pile displacements are compatible with the supported
structures.
b) Ultimate limit state (ULS): the characteristic action Fn,k(ULS) and the location of the neutral
point have to be calculated by comparing the deformations associated with the pile settlement
spile = sultin the ultimate limit state and the settlements in the soft stratum sground. The
location of the neutral point is normally higher than in the serviceability limit state, because the
pile settlement sult is greater than s(SLS). The design value of the effects is:
F d = (F G,k + F n,k (ULS)) ⋅ γ G + F Q,rep ⋅ γ Q (2)
(10) For the structural analysis of the pile shaft the action resulting from negative skin friction at
serviceability limit state, i.e. F n,k (SLS), has to be considered as permanent characteristic action
beside permanent and transient loading from the structure. The design value of the action relevant
for the internal pile design has to be calculated as
F d = (F G,k + F n,k (SLS)) ⋅ γ G + F Q,k ⋅ γ Q . (3)
(Note: It might be further discussed whether the same partial factor should be applied for structural
loads and negative skin friction.)

Annex A (Informative)

Two principle approaches for deriving the characteristic negative skin friction t n,k are given in the
literature dealing with negative skin friction:
• Using total stresses for cohesive soils
τ n,k = α ⋅ c u,k (A.1)

where:
α factor for specifying the value of the characteristic negative skin friction for cohesive soils;
cu,k characteristic value of the shear strength of the undrained soil.
Depending on the soil type and pile type the factor α generally ranges between 0.15 and 1.60,
whereby α = 1 is often adopted in approximation, which is generally recommended for cohesive soils.
More detailed information on the value of α can be taken from [Literature].

24
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

• Using effective stresses for non-cohesive and cohesive soils:


t n,k = K 0 ⋅ tanδ k ⋅ σ ′v = β ⋅ σ ′v (A.2)

where:
σ′ v effective vertical stress;
K0 coefficient of at-rest earth pressure;
δk characteristic value of the interface friction angle, whereby
δ k = ϕ 'k for concrete cast in situ piles,
δ k = 0.75 ϕ 'k for concrete precast and steel piles,
but K 0 ⋅ tanδ k ≥ 0.25
ϕ′ k characteristic value of the friction angle;
β factor for specifying the value of the characteristic negative skin friction for non-cohesive
and cohesive soils.
According to [relevant literature] the factor β generally ranges between 0.1 and 1.0, depending
on soil type. For non-cohesive soils β = 0.20 to 0.30 is often used.

7. OUTLOOK
For elaborating the Eurocode text for the next generation of Eurocode 7 the working structure of
TC250/SC7 was restructured at autumn of 2015. The new structure with three working groups (WG) is
shown in Figure 8. The preparation of revised Eurocode text will be the task of so called Project Teams
(PT) whose members are contracted to NEN and who will deliver the Eurocode text. The Project Teams
will receive assistance and guidance by Task Groups (TG) who will also review the work of the Project
Teams on behalf of SC7.

Figure 8: Evaluation of negative skin friction for ultimate limit state (ULS) and serviceability limit state
(SLS).

The elaboration of a final proposal for the section ´Pile Foundation´ will be the task of Project Team PT4.
They must develop a new proposal implementing new subjects and changes based on:
• M/515 Mandate,
• Evolution Group Reports,
• Systematic Review comments approved by SC7.

25
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Project Team PT4 will start working in the second phase of the program for revising Eurocode 7, i.e.
probably from end of 2016 on.
Task group TG 3 of Working Group WG 3 (WG3/TG3) called ´Pile Foundations´ has already started
working in autumn 2015 and will focus on the following tasks:
• Review comments received through Systematic Review,
• Respond to the Project Team PT4 request for assistance/guidance,
• Review work of Project Team PT4 on behalf of SC7,
• Recommend acceptance/rejection of Project Team PT4’s work.
Beside others probably the following subjects might be addressed by WG 3/TG 3 in a first approach:
• Proposal for evolution and harmonisation of NDP´s (Nationally Determined Parameters) and
DA´s (Design Approaches),
• Elaboration of an improved approach for correlation factors ξ,
• Selection and documentation of calculation methods for axially and transversely loaded piles and
for other applications as input for a possible informative annex,
• Elaboration of basic design rules and proof concept for pile groups and for piled rafts,
• Guidance for interpretation of pile load tests (e.g. assessment of limit resistance from static pile
load tests).

ACKNOWLEDGMENT
The progress gained in elaborating proposals for the evolution of the section ´pile foundations´ of EC 7
into the next generation bases on the collaboration and valuable contribution of all delegates (Table 3). of
Evolution Group EG 7 “Pile Design” of TC250/SC7 working from 2011 to 2015 on these issues.
The support of these colleagues in developing the next generation of the section ´Pile Foundation´ of
EC 7 is highly appreciated.

Table 3: Delegates of TC 250/SC 7/EG 7 ´Pile Design´ of the period 2011-2015


Name Country Name Country
Christian Moormann (Convenor) Germany Ole Møller Denmark
Chris Raison (Secretary) UK Panicos Papadopoulos Cyprus
Gary Axelsson Sweden Krzystof Sahajda Poland
Ioan Boldurean Romania Arne Schram Simonsen Norway
Sébastien Burlon France Frits van Tol Netherlands
Josif Josioivski Rep. of Mazedonia Veli-Matti Uotinen Finland
Boleslaw Klosinski Poland Monika de Vos Belgium
Jan Kos Czech Republic

REFERENCES

Baeßler, M., Niederleithinger, E., Georgi, S., Herten, M. (2012): Evaluation of the dynamic load test on
bored piles in sandy soil. 9th Int. Conf. on testing and design methods for deep foundations - IS
Kanazawa 2012, (Matsumoto ed.), 2012, 155-162.
Belgium Building Research Institute (BBRI): „Guidelines for the application of Eurcode 7 in Belgium
according to NBN EN 1997-1 ANB – Part 1: Geotechnical design in ULS of axially loaded piles, based
on CPT“. BBRI-Report 12
Bond, A. (2013): Implementation and evolution of Eurocode 7, Modern Geotechnical Design Codes of
Practice, P Arnold et al. (Eds), IOS Press.
De Cock, F., Legrand, C.(eds.) (1997): Design of axially loaded piles ⋅ European practice. International
Seminar, ISSMFE ⋅ ERTC3, Brussel, 17.-18. April 1997, Balkema, Rotterdam, 377 pp.
DGGT (2012), AK 2.1: Recommendation on Piling (EA-Pfähle) 2. Auflage, Ernst & Sohn Verlag, Berlin,
2012.

26
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Herten, M.; Baeßler, M.; Niederleithinter, E.; Georgi, S. (2013) Bewertung dynamischer Pfahlprobe-
belastungen an Bohrpfählen. Pfahl-Symposium 2013, Mitteilungen des Instituts für Grundbau und
Bodenmechanik Technische Universität Braunschweig, Heft 96, 2013, 79–98.
Moormann, Ch. (2014): Proposal for a Revised Section „Pile Foundations“ of Eurocode 7 by Evolution
Group 7 „Pile Design“ and DIN Committee on Piles. Proc. of the First PRB-Workshop on Ease of Use of
the Eurocodes, 04.-05. December 2014, Berlin, Nußbaumer, Hertle, Meyer (eds.), Beuth Verlag, Berlin,
51-52
NEN 6743:1991-12: Geotechniek - Berekeningsmethode voor funderingen op palen - Drukpalen.
Nederlandse Normalisatie-Instituut, Delft, 1991

27
28
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

General report - Calculation methods based on soil parameters

29
30
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

General Report – Calculation methods based on direct derivation from


in situ tests
Adriaan van Seters, Fugro GeoServices BV, a.vanseters@fugro.com

ABSTRACT
In Europe many ground investigation techniques are used for the analysis of geotechnical problems. The
analysis can be performed directly by application of the test results to the analysis In this paper the
practice of European countries to analyze pile design using in situ test results is addressed.
An overview is given of the various tests and their application in Europe. The quantity of ground
investigation is also considered in this paper. Furthermore the application of three tests: CPT,
Pressuremeter and Dynamic Probing in European pile design is described. Comparisons between the
practices of the various countries have been made.

1. INTRODUCTION
In Europe various types of in situ testing are in operation. Most of them are developed in a certain country
to test the ground types of that particular country. Thus in France the Pressuremeter was developed to
cope with e.g. limestone, marl, sands and clays. In the Netherlands the Cone Penetration Test came into
existence as a model test pile in soft clays and sands. Other countries developed the Dynamic Probing,
SPT, Vane test, Dilatometer probe.
These in situ tests were developed to measure the ground properties, but also to predict the bearing
capacity of a pile directly. Especially for the Cone Penetration Test, many national calculation methods
were developed to directly assess the pile bearing capacity.
For this symposium, representatives of many countries have published a National Report on Pile design.
This paper summarises the design methods, which directly derive the pile bearing capacity form in situ
tests, without a prior assessment of soil parameters. Three methods were mentioned: Cone Penetration
Test, Pressuremeter Test and Dynamic Probing Test.

2. IN SITU TESTS

2.1. Overview of the in situ tests


In Table 2.1 an overview of all in situ tests used in the various countries is given.

Table 2.1 – In situ tests used in EU-countries


CPT/CP CPT-
Country SPT PMT DPT WST Vane MWD(A) DMT
TU M
Austria No information specified
Belgium D D I
Denmark I I
Estonia D
Finland I D I I I
France D I D I I I
Hungary D I I
Ireland I I
Italy* I I I I
Macedonia I I I I
Netherlands D
Poland D I I
Russia D D I I I
Spain I I I I I
Sweden D I I I I
Legend: D – Test is used to directly assess pile bearing capacity
I – Test is used to derive soil parameters, indirect assessment pile bearing capacity

31
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

CPT/CPTU Cone Penetration Test with Electrical Cone, U = with porewater measurement
CPT-M Cone Penetration Test with Mechanical Cone
SPT Standard Penetration Test
PMT Pressuremeter Test
DPT Dynamic Probing Test
WST Weight Sounding Test
MWD Monitoring While Drilling – continuous measuring during percussion drilling
DMT Dilatometer Test

2.2. Quantity of in situ tests executed on a project


From the National Reports, the quantity of insitu tests executed on a project was investigated. From this
survey it appeared that some countries specified minimum distances between investigation points. Other
countries rewarded more investigation points by decreasing a correlation factor, model factor or partial
resistance factor.

The practice in each country is discussed below:

Belgium
In Belgium, the density of the investigation points is related to correlation factors ξ 3 and ξ 4 , as shown in
Table 2.2.

Table 2.2 - Values of the correlation factor ξ 3 and ξ 4 versus density of CPT’s
ξ 3 / ξ 4 versus CPT DENSITY
NUMBER
of PILES 1 CPT 1 CPT 1 CPT 1 CPT 1 CPT
10 m² 50 m² 100 m² 300 m² 1000 m²
1-3 1,25/1,08 1,29/1,17 1,32/1,23 1,36/1,31 1,40/1,40
4-10 1,15/1,00 1,19/1,07 1,21/1,13 1,25/1,21 1,29/1,29
> 10 1,14/1,00 1,17/1,06 1,20/1,12 1,24/1,20 1,27/1,27

Finland
In Finland the quantity of Dynamic Probing Tests - DPSH is specified: Soundings must be performed at
every corner of the building and every 5 to 15 m along the wall line.

France:
In France, the quantity of the investigation points (boreholes, Pressuremeter tests and CPT’s) is related to
the value of the correlation factors ξ 3 and ξ 4 .

The values of the correlation factors ξ 3 et ξ 4 are obtained from the following equations:

ξ i (N , S ) = 1 + [ξ ' i (N ) − 1]
S
with 100 m2≤ S ≤ 2500 m2.
S réf
S ref = 2500 m2 and N = number of investigation points.

The values of the correlation factors are given in Table 2.3.

Table 2.3 – correlation factors in France


ξ’ for N= 1 2 3 4 5 7 10
ξ3’ 1.40 1.35 1.33 1.31 1.29 1.27 1.25
ξ4’ 1.40 1.27 1.23 1.20 1.15 1.12 1.08

Thus, for a 50 x 50 m grid of 5 investigation points, a value of ξ 3 of 1.29 applies. However for a 10 x 10
m grid, this value is reduced to value of ξ 3 of 1,06.

Ireland
In Ireland guidelines for the spacing of the investigation points are given as shown in table 2.4.

32
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 2.4 - Suggested intensity of ground investigation points according to EuroCode 7 - Ireland
Structure Spacing Arrangement
High-rise and industrial 15 – 40m Grid
Large area ≤60m Grid
Linear – Roads, railways, channels, pipelines, dikes, tunnels, retaining walls 20 – 200m

Dams and weirs 25m – 75m Vertical


sections
Special – Bridges, stacks, machinery foundations 2 – 6 per foundation

Italy
In Italy in preliminary design stage of buildings, for instance, an average number of 2 boreholes with SPT
are carried out for a plan area of the building of about 1.000 m2; the number increases to about 5 if the
plan area approaches 10.000 m2. At final design stage, the number is approximately doubled.

Netherlands
Generally, CPT’s are performed in a relatively dense grid to calculate a pile foundation. A maximum
centre to centre distance of 25 m between CPT’s is required in building regulations for limited variability
of the subsoil to 15m in case of high variability.

The depth of the CPT should be at least 5 m below the pile tip level. At least 1 CPT should reach 10 times
the pile tip diameter below the pile tip level. In case of CC3 buildings (higher than 70 m), CPT’s should
reach to at least 3 times the smallest width of the building (maximum 25 m) below the pile tip level.

In the majority of the cases CPT’s are performed with electrical cones (10 cm2 and 15 cm2), while
mechanical cones are very occasionally used in the eastern and southern part of the country. Measurement
of the pore pressures is included in about 30% of the cases.

Russia
In Russia the specifications of the in situ testing depends on the importance of the structure, as shown
below:
Low importance level of structure:
• CPT: grid between 35x35m (>2CPT’s), 25x25m (>3CPT’s), 15x15m (>6CPT’s), depending on
variability of soil conditions
Normal importance level of the structure:
• CPT: grid between 25x25m (>6CPT’s), 20x20m(>7CPT’s), 15x15m(>10CPT’s), depending on
variability of soil conditions
• PMT: > 6 tests within one geological-geotechnical layer
High importance level of the structure:
• CPT: grid between 25x25m (>6CPT’s), 15x15m (>8CPT’s), 10x10m (>10CPT’s), depending on
variability of soil conditions
• PMT: > 6 tests within one geological-geotechnical layer

3. DESIGN BASED ON CONE PENETRATION TEST

3.1. Piles in compression

3.1.1. Pile capacity according to Eurocode 7


Eight countries have reported CPT-based methods when analysing the pile bearing capacity in
compression, i.e. Belgium, Estonia, France, Hungary, Netherlands, Poland, Russia, Sweden.

33
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

According to Eurocode 7 – 7.6.2.3 eq (7.6) – (7.8) is the design value of the pile resistance in
compression R c;d is determined from:
R c;d = R b;d + R s;d
where:
R b;d = R b;k / γ b
R s;d = R s;k / γ s
The characteristic value of the pile resistance in compression is

Rb; cal + Rs ; cal Rc; cal (R ) (R ) 


Rc; k = ( Rb; k + Rs ; k ) = = = Min c; cal average ; c; cal min 
ξ ξ  ξ3 ξ4 
In some countries a model factor γ R;d is introduced according
R c;cal = R c / γ R;d
Figure 3.1 (Huybrechts et al., 2016) gives a schematic overview of the different steps to calculate the
design value of the compressive resistance of the pile R c,d .

Figure 3.1 – Schematic overview of steps to arrive at a design value of the pile axial capacity
(Huybrechts et al., 2016)

3.1.2. Step 1 – compressive resistance from the CPT-result.

General
The pile base resistance is generally calculated according:

Rb = α b × Ab × qc
where α b is an empirical factor taking into account the pile type/installation method and the soil type.

In some countries like Belgium and the Netherlands the cone resistance is reduced in case of an
excavation after performing the CPT’s. In case of driving or vibrating piles in overconsolidated soils, in

34
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

the Netherlands, q c must be reduced by √(1/OCR) to account for reduction of the overconsolidation
effects due to the driving process.

Weighted cone resistance at the pile tip


The value of the cone resistance q c is generally an average value of the cone resistance at the pile tip.
This is due to the fact that the base resistance of the pile extends over a certain height above and below
the pile tip. Most countries use some averaging algorithm to obtain a weighted value of cone resistance q c,
as shown in Table 3.1. The averaging is performed over a height extending Y times the pile diameter
above the pile tip and X times the pile diameter below the pile tip.

Table 3.1: averaging the cone resistance at pile tip level (D = pile diameter)
Country Method X (below pile tip) Y (above pile tip)
Belgium De Beer Method, reducing the cone 0 0
resistance diagram by exchanging cone
resistance with the resistance of a rod
with the diameter of the actual pile.
Estonia No data
France Bustamante CPT2012-model 1.5 D, < 1.5 m 0.5 D, < 0.5 m
Hungary See Netherlands 0.7 D to 4 D 8D
Netherlands Taking the average q c below the pile tip 0.7 D to 4 D 8D
downward to 0.7 – 4 D (q c;I ) , from there
taking the average of the lowest q c values
upward to tip level (q c;II ) and taking the
average of the lowest q c values upward to
8 D above tip level (q c;III ).
q c = 0.25 q c;I + 0.25 q c;II + 0.5 q c;III
Poland Three soil conditions exist:
I –Uniform soil 1D 4D
II – Soil below tip is weaker than above 4D 2D
III – Tip is strong soil and weak soils 4D 4D
within range of X and Y
Russia No data
Sweden Bustamante – see France

Factor αb – pile type / installation and soil type


The factor α b has been tabulated for various pile types and various soil types.

In Belgium differentiation is made between Tertiary clays and other soil types.
France distinguishes: Siliceous Silt and Clays (CaCO 3 < 30 %), Heterogeneous soils, Sand and gravel,
Chalk, Marl and calcareous marl, Weathered rock.
In the Netherlands cohesive and cohesionless soils lead to different α b -values. The same applies for
Hungary.

In Poland the CPT-method is used for the derivation of the pile bearing capacity of two types of piles:
- Vibro and large diameter piles
- Screw displacement piles in non-cohesive soils

The data for Russia were not provided. In Sweden, the CPT-method is only used in sand for friction piles.
It is assumed that similar factors as in France are used.

The factor α b for sands has been tabulated in table 3.2 for typical pile types.

35
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 3.2: Factors for α b in sands


Cast in
Open Bored pile
Precast situ,
ended Steel using Micro
Country concrete driven CFA-pile
pipe pile thixotropic pile
pile casing,
driven fluid
retrieved
Belgium 1.0 1.0 1.0 0.5 0.5 n.a.
France/Sweden 0.4 0.4 0.25 0.25 0.2 0.2
Netherlands 1.0 1.0 1.0 0.8 0.5 0.5
Hungary 1.0 1.0 1.0 0.46*** 0.33*** n.a.
Poland (depending 1.0* to 0.173* to
on q c ) 0.35** 0.137**
*q c = 4 MPa ** q c = 40 MPa ***including a reduction factor of 0.65

It is noted that the Dutch α b values will be reduced by 2017 to 70 % of the presented values. Furthermore,
in the Netherlands the pile tip bearing capacity is limited to 15 MPa. For CFA-piles, in the Netherlands
the cone resistance above the pile tip is in the bearing capacity calculation reduced to 2 MPa.

For screw displacement piles in non cohesive soils in Poland, the following expression is used based on
pile load tests:
0 ,16
q 
qb;ult = 2475 ×  cb  in kPa
p 
 ref 
Where: q cs is the cone resistance at pile tip in MPa (average over 1 D above and 2 D below the pile tip)
P ref is reference stress equal to 1 MPa.

In Estonia, the cone resistance is related to the “yield point”, i.e. the point on the load-settlement curve,
where linear dependency between load and settlement is ended.
The pile tip resistance can then be approximated as:
q b;yield = 0.2 x q c
Conclusion

When the values of Table 3.2 are compared, the following can be concluded:
• The α b -values cannot be taken at face value, as the calculation is very different per country.
• France/Sweden have a straight forward approach, where the cone resistance is evaluated at pile tip
level (large upper deviations are neglected) and the bearing capacity is derived as product of α b x q c .
• In Belgium and the Netherlands, the cone resistance is reduced, resulting in higher α b -values than in
France. In Belgium by the method De Beer, in the Netherlands by a limiting value of 15 MPa.
• In Poland, the α b -values is a function of the cone resistance. The α b -value decreases with increasing
qc.

Other parameters applied to the base resistance

In Belgium and the Netherlands the base resistance is reduced by:


- factor for Tertiary stiff fissured Clays, taking into account the soil strength dependent characteristics
(B). This factor is between 0.476 and 1.0.
- shape factor for non circular/square pile bases. This factor is 1.0 for square/circular piles, but reduces
to 0.77 (B) and 0.60 (NL) for walls
- reduction factor for piles with a prefabricated enlarged pile tip.

36
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Shaft resistance

The shaft resistance is calculated according to:

Rs = As × ∑ (α s ;i × hi × qs ;i )

The values of α s in sands are given in Table 3.3.

Table 3.3: Factors for α s in sands


Cast in
Bored Micro
Precast situ,
Steel pipe pile using pile,
Country concrete driven CFA-pile
pile driven Bentonite High
pile casing,
fluid pressure
retrieved
Belgium*** 0.011 0.0067 0.0044 0.0056 n.a.
France/Sweden κ 1.0 1.0 0.5 1.25 1.0 2.0
Netherlands 0.01 0.014 0.006 0.006 0.006 0.017
Hungary 0.009 0.011 0.075 0.0055 0.0055
Poland (depending 0,009* to 0,008* to
on q c ) 0,003** 0,002**
*q c = 4 MPa ** q c = 40 MPa ***These values include a reduction factor η

In Hungary, the shaft friction is according to the Szepesházi method defined as:

q s = α s × qc
This formulation illustrates less increase of shaft friction with increasing q c . Here the factor α s is not
dimensionless.

In France the shaft friction is defined as:

q s = κ × g i ( qc )
Where the value of κ is depending on the soil type and the type of pile, and whereas the function g i (q c )
shows the non-lineair development of friction as function of q c in figure 3.2.

Figure 3.2 – Definition of functions g i for cone resistance values q c (Burlon et al, 2016)

37
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

For screw displacement piles in non cohesive soils in Poland, the following expression is used based on
pile load tests:
0 , 23
 q 
qs ;ult = 65 ×  cs  in kPa
p 
 ref 
Where: q cs is the cone resistance along the shaft in MPa
P ref is reference stress equal to 1 MPa.

In Belgium and in the Netherlands, the friction is limited (see Table 3.4).

Table 3.4 – Limiting friction values[kPa] :


Cast in
Bored
Precast situ,
Steel pipe pile using
Country concrete driven CFA-pile
pile driven Bentonite
pile casing,
fluid
retrieved
Belgium 150 90 60 75
Netherlands 150 210 90 90 90

In Estonia, the cone resistance is related to the “yield point”, i.e. the point on the load-settlement curve,
where linear dependency between load and settlement is ended.
The shaft friction of a pile can then be approximated as:
q s;yield = 0.8 x f s
With f s is the local sleeve friction from the CPT.

Conclusion
An evaluation of the α s values show a reasonable agreement between the values from various countries
for various pile systems.
All countries have some limitation on the shaft friction with increasing q c :
- Belgium and the Netherlands have limiting values
- France, Poland and Hungary show a less increase in friction with increasing q c France achieves this
phenomenon by using a function g i (q c ), where Poland reduces α s with increasing q c . Hungary relates
the friction with the square root of q c .

3.1.3. Step 2 – Application of a model factor


In 4 countries model factors γ Rd are introduced: Belgium, France, Poland and Sweden. The model factors
are greater than unity and the calculated pile capacity is reduced by multiplication with 1/γ Rd .

In Belgium the derived value from the CPT-result is reduced as follows:

Rc;cal = Rc / γ Rd
Where γ Rd is a model factor depending, whether a pile system is tested in comparable ground or at the job
location itself. The value varies between 1.0 and 1.3. The modelfactor only applies to cast insitu pile
systems: the value for driven and jacked piles is always 1.0.

In France two methods of analysis exist, i.e. the “model pile” procedure and the “ground model”
procedure. The former is based on calculating N values of the bearing capacity from N-test profiles (CPT,
PMT-test or boreholes). The latter procedure is based on dividing the ground in layers with characteristic
properties.
For the “model pile” procedure a model factor γ R;d1 is used. For the “ground model” procedure, the model
factor consists of the product of γ R;d1 x γ R;d2 . The factor γ R;d1 ranges between 1.18 and 2.0 depending on
soil and pile type. The factor γ R;d2 is always 1.1.

In Hungary a model factor γ Rd of 1.1 is used, when the bearing capacity is directly derived for CPT-
results, without back-calculation to soil properties.

38
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

In Poland a model factor γ Rd is applied of 1.25 for a single pile foundation, 1.12 for a two-pile foundation
and 1.00 for a foundation on more than two piles.

In Sweden in combination with design based on CPT (Bustamante, French method), a model factor γ Rd of
1.4 is used. When characteristic value of the bearing capacity is derived according an “alternative
procedure, not using ξ-values) an extra model factor of γ Rd,e of 1.4 is applied.

3.1.4. Step 3 – Application of the ξ-value


The correlation factor ξ is related to the uncertainty of the ground conditions.

When using CPT’s, the correlation factor is related to the number of CPT’s executed for a structure.
The countries use different methodologies to determine the value of ξ 3 and ξ 4

In Belgium, the value of ξ is related to the number of CPT’s per area. Also the number of piles below a
stiff structure may reduce the ξ-value due to redistribution of loads. The values of ξ 3 for a not-rigid
structure range from 1.40 (1 CPT per 1000 m2) to 1.25 (1 CPT per 10 m2).

The ξ-values used in France and tabulated below are valid for an area covered by 1 CPT of 2500 m2, i.e. a
grid of 50 x 50 m. When the CPT´s or other field tests are performed at a smaller grid, the ξ-values may
be reduced, leading to a more economic design (see section 2.2).

The values for ξ 3 for the other countries are tabulated in Tables 3.5 and 3.6.

Table 3.5 – Correlation factor ξ 3 for number of tests N


Country 1 2 3 5 10
France 1.40 1.35 1.33 1.29 1.25
Netherlands 1.39 1.32 1.30 1.28 1.25
Poland 1.30 1.25 1.21 1.17 1.13

Table 3.6 – Correlation factor ξ 4 for number of tests N

Country 1 2 3 5 10
France 1.40 1.27 1.23 1.15 1.08
Netherlands 1.39 1.32 1.30 1.03 1.00
Poland 1.25 1.15 1.12 1.05 1.00

The French values correspond with the Eurocode 7 – default values.

3.1.5. Step 4 – Application of a partial factor


The partial factors can be applied on the base resistance (γ b ) and the shaft friction (γ s ) separately or on the
total capacity as a whole (γ t ). The partial factor for the various countries are tabulated in Table 3.7.

Table 3.7 – Partial resistance factors – pile bearing capacity in compression:


Country γb γs γt
Belgium With QA* : 1.0 1.0
Without QA: 1.07 – 1.2
France 1.1
Hungary Driven pile: 1.1 Driven pile: 1.1 Driven pile: 1.1
Bored pile: 1.25 Bored pile: 1.1 Bored pile: 1.2
Netherlands 1.2 1.2 1.2
Poland 1.1 1.1 1.1
Sweden Driven pile: 1,2 Driven pile: 1,2 Driven pile: 1,2
Bored, CFA pile: 1.3 Bored, CFA pile: 1.3 Bored, CFA pile: 1.3
* Contractor has to prove that the pile installation takes place according a well-established quality plan.

39
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3.2. Piles in tension

3.2.1. Pile capacity according to Eurocode 7


According to Eurocode 7 – 7.6.3.3 eq (7.15) – (7.17) is the design value of the pile resistance in
compression R t;d is determined from:
R t;d = R t;k / γ s;t
The characteristic value of the pile resistance in compression is

 (R ) (R ) 
Rt ; k = Min t ;cal average ; t ;cal min 
 ξ3 ξ4 
In some countries a model factor γ R;d is introduced according
R t;cal = R t / γ R;d

3.2.2. Step 1 – tensile resistance from the CPT-result.

Shaft resistance

The shaft resistance for a single pile is calculated according to:

Rt = As × ∑ (α t ;i × hi × qt ;i )

The values of α t in sands are given in Table 3.8 as a comparison.

Table 3.8: Factors for α t in sands


Cast in
Bored Micro
Precast situ,
Steel pipe pile using pile,
Country concrete driven CFA-pile
pile driven Bentonite High
pile casing,
fluid pressure
retrieved
Belgium* 0.009 0.0053 0.0035 0.0044 n.a.
France/Sweden κ 1.0 1.0 0.5 1.25 1.0 2.0
Netherlands 0.007 0.012 0.004 0.0045 0.0045 0.017
*These values include a reduction factor η. For static tensile load

In France the shaft friction is determined according to the procedure for compression piles. No correction
factor is applied in this stage.

In Sweden shaft friction is calculated corresponding to the friction in compression. A correction factor on
the shaft friction in compression for tension piles in sand of 0.7 to 0.9 is applied

In Belgium and in the Netherlands, the friction is limited (see Table 3.9).

Table 3.9 – Limiting friction values[kPa] :


Cast in
Bored
Precast situ,
Steel pipe pile using
Country concrete driven CFA-pile
pile driven Bentonite
pile casing,
fluid
retrieved
Belgium 120 72 48 60
Netherlands 105 180 60 68 68

Furthermore, Belgium and the Netherlands reduce the cone resistance further, if the pile is loaded by
alternating loading (tension-compression). The reduction is maximum 1.33 in Belgium and 1.5 in the
Netherlands.

40
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

In case of pile groups, all countries consider the sum of the tensile capacities of the individual piles versus
the total weight of the block of soil, enclosed by the piles.

3.2.3. Step 2 – Application of a model factor


In 4 countries model factors γ Rd are introduced: Belgium, France, Poland and Sweden. The model factors
are greater than unity and the calculated pile capacity is reduced by multiplication with 1/γ Rd .

For all countries except France, the same model factor applies as for compression loading (see 3.1.3).

In France, for the “model pile” procedure a model factor γ R;d1 is used. For the “ground model” procedure,
the model factor consists of the product of γ R;d1 x γ R;d2 . The factor γ R;d1 ranges for tension piles between
1.45 and 2.0 depending on soil and pile type. The factor γ R;d2 is always 1.1.

3.2.4. Step 3 – Application of the ξ-value


The same ξ/values apply for tension loading as for compression loading (see 3.1.4).

3.2.5. Step 4 – Application of a partial factor


The partial factors can be applied on the tensile capacity as a whole (γ s;t ). The partial factor for the
various countries are tabulated below in Table 3.10:

Table 3.10 – Partial resistance factors – Pile bearing capacity in tension


Country γ s;t
Belgium 1.0
France 1.15
Hungary 1.25
Netherlands 1.35
Sweden Driven pile: 1.3
Bored, CFA pile: 1.4

4. DESIGN BASED ON PRESSUREMETER TEST

4.1. General
Only in France the ultimate capacity of piles is designed on the basis of Pressuremeter Testing. The
calculation model “PMT2012” is described in the National Report to this conference by Burlon et al
(2016) and is based on the work by Bustamante et al. (2009).

The model is comparable with the CPT2012-model described in Chapter 3.


The equations in section 3.1.1 also apply.

The model is based on the net limit pressure p* LM , which is equal to the measured limit pressure p LM
minus the total horizontal stress p h0 (rough estimate).

4.2. Determination of the base resistance


The base resistance is written as:

Rb = k p × p *LM ;e × Ab
Where p* LM;e is the equivalent net limit pressure, which is an average value of p LM over 1.5 D (max 1.5
m) below the proposed pile tip and 0,5 D (max 0.5 m) above the pile tip.

The k p value is depending on the pile class and the ground type. France distinguishes: Siliceous Silt and
Clays (CaCO 3 < 30 %), Heterogeneous soils, Sand and gravel, Chalk, Marl and calcareous marl,
Weathered rock.

The value of k p is tabulated for sand in Table 4.1.

41
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 4.1 – Value of k p for sand - PMT


Cast in
Open Bored pile
Precast situ,
ended Steel using Micro
Country concrete driven CFA-pile
pipe pile thixotropic pile
pile casing,
driven fluid
retrieved
France 3.1 3.1 1.9 1.65 1.1 1.1

4.3. Determination of the shaft resistance


The shaft resistance is calculated according to:

Rs = As × ∑ (κ × g i ( p *LM ) × hi )

Where: A s shaft circumference


hi height

The values for κ are dependent on pile category and soil type and are tabulated for sand in Table 4.2.

Table 4.2 – Value of κ for sand - PMT


Cast in
Open Bored pile
Precast situ,
ended Steel using Micro
Country concrete driven CFA-pile
pipe pile thixotropic pile
pile casing,
driven fluid
retrieved
France 1.4 1.4 0.7 1.8 1.4 2.9

The function g i (p* LM ) is dependent form the soil type. Function g 2 is valid for sand and gravel. This
function is presented in figure 4.1.

Figure 4.1 – Functions g i depending on p* LM (Burlon et al., 2016)

In France the shaft friction for a tension pile is determined according to the procedure for compression
piles. No correction factor is applied in this stage.

4.4. Application of a model factor


The model factor in France for the calculation method using the pressuremeter is the same as for the CPT
method.

42
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

In France two methods of analysis exist, i.e. the “model pile” procedure and the “ground model”
procedure. The former is based on calculating N values of the bearing capacity from N-test profiles (CPT,
PMT-test or boreholes). The latter procedure is based on dividing the ground in layers with characteristic
properties.
For the “model pile” procedure a model factor γ R;d1 is used. For the “ground model” procedure, the model
factor consists of the product of γ R;d1 x γ R;d2 . The factor γ R;d1 ranges between 1.18 and 2.0 depending on
soil and pile type. The factor γ R;d1 ranges for tension piles between 1.45 and 2.0 depending on soil and
pile type. The factor γ R;d2 is always 1.1.

4.5. Application of the ξ-value


The correlation factor ξ is related to the uncertainty of the ground conditions and is the same as for the
CPT-method. Reference is made to section 3.1.4.

4.6. Application of partial factor


The partial factors are applied on the total capacity as a whole (γ t ) and on the tension capacity (γ s;t ) in
case of tension piles. The partial factors are tabulated in Table 4.3.

Table 4.3 – Partial resistance factors – PMT-method:


Country γt γ s;t
France 1.1 1.15

5. PILE DESIGN BASED ON DYNAMIC PROBING

5.1. Introduction
In Finland, the axial pile bearing capacity can be estimated using Dynamic Probing results.
These tests are performed using a heavy DPSH-A apparatus. Reference is made to EN1997-2 for a
description and correlations. The resulting parameter is the number of blows per 0.2 m penetration

The correlations are from the following Figures 5.1 and 5.2 for the National Report for Finland by
Kinnunen et al, 2016.

Figure 5.1- Base and shaft resistance estimated using dynamic probing resistance for displacement piles
(Kinnunen et al, 2016).

43
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 5.2 - Base and shaft resistance estimated using dynamic probing resistance for replacement piles
(Kinnunen et al., 2016).

From Figure 5.1 or Figure 5.2 the base resistance q b and the shaft resistance q s can be determined.

Using: Rb = Ab × qb

and Rs = As × ∑ (hi × qs )

the total bearing capacity of a pile can be determined.

A model factor is not applicable.

Thus, from more DP-tests a characteristic value for the pile bearing capacity in compression can be
determined according:

Rc;cal  (R ) (R ) 
Rc; k = = Min c;cal average ; c;cal min 
ξ  ξ3 ξ4 
The values for ξ 3 and ξ 4 are given in Table 6.1.

Table 6.1 – values of ξ 3 and ξ 4 versus number of tests - Finland


Finland 1 2 3 5 10
ξ3 1.85 1.77 1.73 1.65 1.60
ξ4 1.85 1.65 1.60 1.50 1.40

The partial resistance factors are shown in Table 6.2.

Table 6.2 – Partial resistance factors - Finland


Finland γb γs γt γ s;t
1.2 1.2 1.2 1.35 permanent load
1.5 variable load

44
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

6. SYMBOLS
With regard to general used symbols reference is also made to EN 1990 and EN 1997 – 1.

A b (m²) the pile base section


A s (m) the pile circumference
D b (m) the pile base diameter
D s (m) the diameter of the pile shaft
h i (m) thickness of soil layer i
q b (kPa) unit pile base resistance
q c (MPa) cone resistance
q s (kPa) unit pile shaft friction
R (kN) pile resistance
R b (kN) pile base resistance
R c (kN) compressive resistance of the pile
R cal (kN) calibrated pile resistance
R d (kN) design value of the pile resistance
R s (kN) shaft fricton
R t (kN) resistance op a pile subjected to tension loads

α b (-) installation factor for the pile base resistance


α s (-) installation factor for the shaft friction of a pile subjected to compressive loads
α t (-) installation factor for the shaft friction of a pile subjected to tension loads
γ b (-) partial factor for the pile base resistance
γ Rd (-) model factor
γ s (-) partial factor for the shaft friction of piles subjected to compressive loads
γ st (-) partial factor for the shaft friction of piles subjected to tension loads
ξ (-) correlation factor

REFERENCES
Burlon, S, Szymkiewicz, F., Le Kouby, A., Volcke, J.P., 2016, “Design of piles – French practice”,
ISSMGE-ETC3 International Symposium on Design of Piles in Europe, Leuven, Belgium

Bustamante, M., Gambin, M., Gianeselli, L., 2009, “Pile Design at Failure using the Ménard
Pressiometer: an Update, IFCEE 2009, Proc. Int. Foundation Congress & Equipment Expo, Orlando,
ASCE Geotechnical Publication, 186, 127 – 134.

Huybrechts, N., De Vos, M., Bottiau, M., Maerten, L., 2016, “Design of piles – Belgian practice”,
ISSMGE-ETC3 International Symposium on Design of Piles in Europe, Leuven, Belgium

Kinnunen, J., Riihimäki, T., Tolla, P., Uotinen, V-M., Länsivaara, T.,2016, “Design of piles – Finnish
practice”, ISSMGE-ETC3 International Symposium on Design of Piles in Europe, Leuven, Belgium

45
46
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Design Examples: Comparison of different national practices


Trevor L.L. Orr, Trinity College Dublin, Ireland, torr@tcd.ie

ABSTRACT
Three examples, which involved the design of driven, bored, screw and CFA piles in different ground
conditions, were distributed to the members of European Technical Committee 3 (ETC3): Piling of the
International Society for Soil Mechanics and Geotechnical Engineering (ISSMGE) to prepare solutions to
these examples using their national practices. The solutions received are compared in this paper. These
solutions showed that all the European countries are now carrying out pile designs according to, or so as
to be compatible with, Eurocode 7. Since no models to calculate the pile resistance are provided in
Eurocode 7, many countries provide calculation models in their standards or guidance documents for
pile design. As a result of this a great variation occurred in the calculated pile resistances and
consequently in the characteristic loads that the designed piles would support in each country. However,
since most countries have adopted Eurocode 7, the solutions received are quite closely harmonized with
regard to the overall factor of safety determined as the ratio of the characteristic resistance to the
characteristic load. It was also found that the ratio of the calculated resistance to the characteristic load,
which is effectively the traditional factor of safety, for almost all the solutions lay within the range of 2.0
to 3.0, which is the range that was normally adopted for factor of safety for piles before the introduction
of Eurocode 7. From this it is concluded that the introduction of Eurocode 7 has not had a great effect on
pile design in Europe.

1. INTRODUCTION
In 2015, the member of European Technical Committee 3 (ETC3) Piles of the International Society for
Soil Mechanics and Geotechnical Engineering (ISSMGE) were invited to prepare national reports on the
design of piles in their countries and to provide solutions to three pile design examples. These reports and
design examples were prepared for the ETC 3 International Symposium Design of Piles in Europe - How
did Eurocode 7 change daily practice? This paper presents an analysis and comparison of the solutions
received from the ETC3 members based on their national practices.

Eurocode 7, i.e. EN 1997: Geotechnical design, with Part 1: General rules and Part 2: Geotechnical
investigation and testing, was due to replace the former national standards for geotechnical design in all
the CEN member countries. Eurocode 7 Part 1 in 2010 and following publication in each country of its
National Annexes that included the values of the partial factors and other parameters to be used in that
country. CEN in 2016 has 33 member states, which includes most European countries. The primary
objective of this symposium was to give the European geotechnical engineers the opportunity to
familiarise themselves with the different national methods for pile design in Europe and discover how pile
design in Europe has been affected by the introduction of Eurocode 7. ETC3 has 26 members, all of
which, except for Russia, are also members of CEN and hence are obliged, to prepare national annexes
for Eurocode 7 and to implement it. National reports on pile design were received from 15 members of
ETC3 together with 4 solutions for Example 1 – Driven pile, 10 solutions for Example 2 – Bored pile, 8
solutions for Example 2 - Screw pile, and 11 solutions for Example 3, as shown in Table 1. The aim of
these examples was to focus on the approach rather than the final result. Tables were prepared for each
example and the following results were requested:
• Characteristic (unfactored) value and ULS design (factored) value of the pile resistance
• SLS capacity, but not the settlement
• Factors applied to the loads.
Examining the reports on national practice submitted by the ETC3 countries that also submitted solutions
to the design examples, it was found that the national practices in all those countries were either following
Eurocode 7, and many have been doing so for several years ever though some, Poland for example, does
not have a national annex, or, as in the case of Russia, the national documents are being harmonised with
and are not contrary to the requirements in Eurocode 7.

47
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 1: Countries from which national reports and solutions to the ETC3 examples were received
Solutions received for examples
National reports received
Example 1 Example 2 Example 3
from 15 ETC3 members
Driven pile Bored pile Screw pile CFA pile
Austria
Belgium x x x
Denmark x x
Estonia
Finland x x
France x x x
FYR Macedonia x x
Hungary x x x x
Ireland x
Italy x x x
Netherlands x x x
Poland x (3) x (2) x
Russia x x x
Spain
Sweden x x
Totals 4 10 8 11

2. DESIGN EXAMPLE 1

2.1. Details of Design Example 1


Example 1 is the design of 22 precast concrete driven piles, 270 x 270 mm with a 60mm diameter rock
shoe, to support the abutments of a bridge and is based on a real project in an urban area in Sweden. The
soil consists of 0.8m of fill (mostly sand and gravel) over 1.15m of stiff clay (dry clay above the
groundwater level) over 4.5m of very soft clay (minimum c u = 9kPa) over 1.5m of hard till (very dense
moraine) overlying granite bedrock which normally has an intact unconfined strength, q u of 200MPa.

Ground investigations (GI) were carried out at the site of one at abutment, Borehole No. 6, and at a
distance of 27m from the other abutment, Borehole No. 7. Soil samples were obtained at the Borehole No.
7, from which laboratory tests were carried out to determine c u from the fall cone, sensitivity, density and
water content values. Vane shear tests were carried out at Borehole No. 6 and soil-rock probing and
weight sounding tests (WST) were carried out at both the Boreholes No. 6 and 7. Following installation
of the piles, 4 piles were subjected to dynamic load tests. The results of the tests carried out at Borehole
No. 6 and 7 are shown in Figures 1 and 2.

Figure 1: GI No. 6 showing vane shear test, soil- Figure 2: GI No. 7 showing laboratory fall cone,
rock probing and weight sounding test results sensitivity, density water content and in situ soil-
rock probing and weight sounding test results

48
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Following installation of the 22 piles, dynamic pile load tests were carried out on four piles, P3, P5, P9
and P19. Analysis of these dynamic tests gave the following dynamic pile resistances:

P3 = 2838 kN
P5 = 4100 kN
P9 = 2660 kN
P18 = 2722 kN

The objective of this example was first to predict the pile resistance based on the results of the
geotechnical investigation and then to determine the pile resistance based on the dynamic pile load test
results.

2.2. Solutions to Example 1

2.2.1. General comment on solutions


As shown in Table 1, only 4 solutions were received for Example 1 and these were received from
Denmark, Finland, Hungary and Sweden. The reason why so few solutions were received for this
example was because it involved the design of driven piles in soft ground and bearing on rock. This was
not considered a typical pile design situation for most of the ETC3 member countries. A statement
demonstrating this is the following comment received from Poland that Example 1 was: “not applicable
because of not enough experiences with driven precast concrete piles embedded in bedrock in Poland”.

2.2.2. ULS design


The results of ULS design calculations for Example 1, including the various factors used to obtain the
design resistances, are presented in the following tables: Table 2 gives the prediction method and the
predicted base, shaft and total compressive resistances, R b , R s and R c;pred ; Table 3 gives calculated
resistances from the dynamic load tests, R c;cal , the correlation and model factors, ξ 3 , ξ 4 and γ R;d , and
hence the characteristic resistance from the dynamic tests, R c;k ; and Table 4 gives the partial factor on the
total resistance, γ t , the design compressive resistance, R c;d , the partial factors on permanent and variable
loads, γ G and γ Q , the characteristic load, F k , that the pile will support assuming the variable load is one
third (33%) of the permanent load, the overall factors of safety (OFS) defined as R c;k /F k , and the total
safety factors (TSF) defined as R c;k /F k .

It was commented that there is no tradition in Denmark of estimating the preliminary resistance of piles
on rock. It was also stated that design would be based on driving resistance or dynamic or static load tests
and it would be necessary to have the pile driving record, with hammer weight, fall height and efficiency
for pile design in Denmark in coarse soils using the Danish Driving Formula (DDF). Using this method
and the set for hard driving, a resistance of 1748kN was determined as shown in Table 2. When the only
data available are the data for the tested piles, then it is only possible to determine the resistance of the
tested piles and not the resistance of the remaining working piles.

In order to design according to the Finnish practice it was necessary to convert the provided soil-rock
probing test (MWD) and weight sounding test (WST) results to DPSH_A (dynamic penetration super
hard) values as the Finish standard provides graphs relating the dynamic probing resistance to the pile
base and shaft resistance. Calculated base and shaft resistance values of 1450kN and 300kN giving a total
of 1750kN are obtained the using this method as shown in Table 2. This value is very similar to the
Danish predicted resistance of 1748kN. However, it is noted that as the pile is equipped with a 60mm
rock shoe and driven directly onto rock, the shaft resistance is neglected and correlations factors of 1.57
and 1.45 are used giving the design resistance, R c;d = 1468kN.

It was commented that dynamic tests are quite rare in Hungary. Generally static tests are performed to
verify the calculation method and designs are hardly ever based on pile load test results. No resistances
were predicted based on the GI results.

49
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 2: Example 1 (Driven pile) - Prediction method and predicted resistance


Country Prediction method Rb Rs R c;pred
Denmark Danish Driving Formula - - 1748
Finland To apply Finnish guide MWD and WST results needed 1450 300 1750
to be converted to DPSH_A
Hungary No predicted resistance calculated
Sweden WST and percussion drilling to bedrock. End bearing 2262 01 2262
resistance is a function of rock dowel shoe area A.
R=q u xA. Unit end bearing resistance ; q b =4q u .
1
The shaft resistance was assumed to be negligible, hence R s = 0

Table 3: Example 1 (Driven pile) - Calculated resistance, correlation and model factors, and
characteristic resistance from the dynamic tests
R c;cal Correlation factors Model
Country Method Min. Mean ξ5 ξ6 factor R c;k
γ Rm
Denmark Pile dynamic tests 2561 2695 1.41 -2 1829
Finland Pile dynamic tests 2561 2695 1.36 1.23 - 1982
Hungary Pile dynamic tests. 2561 2695 1.55 1.42 1.05 1656
Sweden Pile dynamic tests 2561 2695 1.55/1.13 1.45/1.13 0.85 2250
1
Only 4 of 22 piles tested. It is assumed all have resistance at least equal to minimum measured. If test
pile is representative of other piles not tested, then in Denmark ξ = 1.4
2
No model factor is used in Denmark
2
Divided by 1.1 due to high stiffness of footing

Table 4: Example 1 (Driven pile) – Characteristic resistance, total partial resistance factor, design
resistance, characteristic load, OFS and TSF
Country R c;k γt R s;d γG γQ F k1 OFS TSF
Denmark 1829 1.3 1407 1.0 1.5 1251 1.46 2.15
Finland 1982 1.2 1652 1.15 1.5 1335 1.48 2.02
Hungary 1656 1.1 1506 1.35 1.5 1085 1.53 2.48
Sweden 2250 1.2 1875 1.35 1.5 1351 1.67 1.99
1
F k = 4R d /(3γ G + γ Q )

The graphs in Figure 3 show, starting from the same mean bearing resistance obtained from the dynamic
tests, the variations in the characteristic and design resistance obtained by each country and the resulting
predicted characteristic load assuming 33% variable load. It was found that the coefficient of variation of
the characteristic resistance is 11%. The calculated overall factors of safety, OFS = R k /R k given in Table
4 and plotted in Figure 4 are very consistent since the combination of partial resistance factors and partial
load (action) factors in the case of these four countries ranges from1.46 to 1.67. The OFS values, which
are the combination of the partial resistance and load factors plus the correlation factors, range from 1.99
to 2.45 which consistent with range of values of 2.0 to 3.0 for the overall factor of safety that was
traditionally used for pile design since the OFS is the mean calculated resistance divided by the
characteristic load.

50
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3000

2500
Resistances and Characteristic load, kN

2000

1500

1000
1. Denmark Rc;mean
2. Finland Rc;min
3. Hungary Rc,pred
500 4. Sweden Rc;k

0
1 2 Countries 3 4

Figure 3: Example 1 (Driven pile) –Resistances and characteristic load

3,00

2,50

2,00
OFS and TSF

1,50

1,00 1.
Denmark OFS
0,50 2. Finland TSF

0,00
1 2 3 4
Countries

Figure 4: Example 1 (Driven pile) – Overall factor of safety and total safety factor from dynamic tests

2.2.3. SLS design


Since the pile passes through soft clay it was assumed that this clay was not fully consolidated and hence
downdrag would occur. This downdrag needs to be calculated to assess the serviceability limit state.
However, it is pointed out in the Danish solution that the maximum strength of the clay is required to
predict the downdrag rather than the minimum value of 9kPa that was provided. Furthermore, in the
Danish solution the load required to limit the crack width in the piles for SLS conditions was analysed
and, assuming G k =- 1173kN, the limiting load was found to be given by 1225kN – Q k so that if Q k
exceeds 52kN, then the SLS load with respect to the maximum allowable crack width is critical.

2.2.4. Structural design


Since the piles were driven through soft ground onto rock, the possibility of crushing or buckling of the
piles should be considered. This situation was commented upon in the Danish, Finnish and Swedish
solutions.

In the Danish solution, the design (STR) crushing strength of the piles without considering buckling was
calculated to be 2604kN. However, if buckling in the very soft clay layer is considered, the design (STR)
buckling load is 1694kN, which is greater than geotechnical (GEO) design resistance of 1407kN given in
Table 4 so that buckling would not be critical.

51
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

It is noted in the Finnish solution that, as the piles are bearing on rock, it is necessary to consider the pile
compressive stress. For concrete piles with concrete having a compressive strength of 45MPa, the
maximum design compressive resistance is 2887kN. Also in Finland it is necessary to check for buckling
and, while the buckling load was not calculated, it was estimated to be in the range 1100 to 1300kN. In
view of this, buckling would control the design in Finland as this load is less than R c;d , which is 1652kN.

The Swedish solution used a similar analysis to the Danish solution for pile buckling in a very soft clay
and calculated the design (STR) buckling load for the pile to be 1757kN compared to the Swedish
geotechnical (GEO) design bearing resistance of 1875kN shown in Table 4, so that in this case buckling
rather than bearing resistance was found to be the controlling ultimate limit state.

3. DESIGN EXAMPLE 2

3.1. Details of Example 2


Example 2 is the design of 20 piles for a stiff building in Belgium as part of a Belgian Building Research
Institute (BBRI) research project on a site 36m x 18m. Two types of piles were to be considered: 900mm
bored piles with bentonite suspension and a temporary casing over the first 3m, and 410mm displacement
(no soil excavation) screw piles. The soil consists of 1.0m of fill over a deep deposit of stiff over-
consolidated Boom clay. No additional surface load was to be considered and hence no downdrag.

The ground investigation involved 3 CPT, 2 SPT and 2 pressuremeter (PMT) tests well distributed over
the site. Boring with undisturbed soil sampling and laboratory tests were carried out centrally on the site.
The CPT q c and SPT N values are plotted in Figures 5 and 6 respectively. The results of the triaxial tests
gave a c’ value of 22kPa and a φ’ value of 28.2o.The groundwater was at a depth of 1.0m

The objective of this example was to predict the compressive resistance of the bored pile and the screw
pile, stating what ground investigation information was used.

Figure 5: Example 2 (Bored and screw piles) – CPT plots Figure 6: Example 2 (Bored and
screw piles) – SPT plot

3.2. Solutions to Example 2 - Bored pile


The results of the ULS design calculations for the bored pile in Example 2 are presented in the following
tables: Table 5 gives the test data and models used to calculate the pile resistance; Table 6 gives the
calculated base, shaft and total resistances, R b , R s and R c ; Table 3 gives calculated resistances from the
dynamic load tests, R c;cal , the correlation and model factors, ξ 3 , ξ 4 and γ R;d , and hence the characteristic
resistance from the dynamic tests, R c;k ; and Table 4 gives the partial factor on the total resistance, γ t , the
design compressive resistance, R c;d , the partial factors on permanent and variable loads, γ G and γ Q , the
characteristic load, F k , that the pile will support assuming the variable load is 50% of the permanent load,

52
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

and overall factors of safety (OFS) defined as R c;k /F k , and the total safety factors (TSF) defined as
R c;k /F k .

Table 5: Example 2: (Bored pile) – Ground investigations and calculation models


Country GI Calculation models
Belgium CPT qb based on de Beer’s method. Factors to account for pile
installation, scale effect, pile shape, etc. are given in
Belgian standard to calculate R c and R s .
France PMT Calculation model based on Ménard limit pressure.
FYR Macedonia c’ and φ’ from lab tests Calculation model in Macedonian Rulebook
Hungary CPT Resistance determined from CPT results using one of three
methods: (1) Former Dutch standard method in EC7 Part
2, (2) Fellenius-Eslami calculation method and (3)
Szepesházi method using Excel spreadsheet
Italy CPT Free of method on basis of experience and skill
Netherlands CPT Based on NEN 9997-1 – ULS at 20% toe deformation
Method has factors to account for pile type and base shape
Poland (a) I L estimated from CPT Calculated according to Polish code PN-83/B-02482
Poland (b) CPT Calculated according to Polish code PN-83/B-02482 and
Gwizdala and Steczniewski calculation model for the
bored pile and Krasinski model for the screw pile
Poland (c) Lab c u Calculated according to Polish code PN-83/B-02482
Russia CPT Resistance calculated using equations and factors given in
Russian code GOST 19912-21202

In France there are two possible ways to calculate the pile resistance: the “model pile procedure” and the
“ground model procedure”. In this example the most frequent method in France was used, which is the
ground model procedure. This method, given in the French standard, involves dividing the ground into
layers and calculating the base resistance and the shaft reistance in each layer from the calculation model.
In this procedure the reprsentative values for the soil parameters in each layer, which are cautious
estimates, is left to enginering judgement. In the case of Hungary, it is commented that bored piles with
drilling mud or shelling technology (i.e. a casing) are rarely used. In Macedonia the pile resistance
traditionally and at present is mostly calcuculated using soil parameters and a bearing capacity equation.
The Macedonian Rulebook is used in conjunction with Eurocodde 7 and provides equations for pile base
and shaft rsistances that include bearing capacity factors whose values are a function of the mobilised
friction angle, φ’ m and are presented in a chart. In this example, c’ = 20kPa and φ’ = 26o were used.

The calculated resistances are presented in Table 6. Since 7 of the 10 solutions are based on the results of
3 CPT tests, the minimum and mean calculated resistances are presented as the correlation and model
factors in Table 7 are applied to these to obtain the characteristic resistances in Table 8. The Russian
resistance is much smaller than the other but no details were provided, nor the design resistance.

Table 6: Example 2 (Bored pile) – Calculation method and calculated resistances


R b;cal R s;cal R c;cal
Country Method
Min. Mean Min. Mean Min. Mean
Belgium CPT 770 978 1027 1046 1797 2024
France PMT 732 1422 2154
FYR Macedonia Lab c,φ’ 1600 339 1939
Hungary CPT 546 635 1373 1502 1983 2137
Italy Lab c u 1000 1370 2370
Netherlands CPT 647 752 1475 1551 2140 2302
Poland (a) I L from CPT 1521 1540 1457 1486 2995 3026
Poland (b) CPT 1422 1474 453 471 1881 1945
Poland (c) Lab c u 9511 13981 2349
Russia CPT 386 408 534 581 944 989
1
.Based on q b =9c u , q s from graph of q s vs c u in Polish standard

53
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 7: Example 2 (Bored pile) – Correlation, model and partial resistance factors
Model
Correlation factors Partial resistance factors
Country factor
ξ3 ξ3 γ R;d γb γs γt
Belgium 1.23 1.18 1.21 1.21 1.0 -
France - - 1.265 1.1 1.1 1.1
FYR Macedonia - - 1.8 1.1 1.1 1.1
Hungary 1.33 1.23 1.1 1.2 1.1 1.15
Italy 1.6 1.48 - 1.35 1.15 -
Netherlands 1.303 1.30 - 1.2 1.2 1.2
Poland (a) 1.33 1.23 - 1.1 1.1 1.1
Poland (b) 1.33 1.23 - 1.1 1.1 1.1
Poland (c) 1.33 1.23 - 1.1 1.1 1.1
Russia - - - - - -
1
For bored pile without static load test
3
1.39 for stiff building, 1.3 for flexible building

Table 8: Example 2 (Bored pile) – Characteristic and design resistances


Country R b;k R s;k R c;k R b;d R s;d R c;d
Belgium 544 725 1269 453 725 1178
France 579 1124 1703 526 1022 1548
FYR Macedonia 889 188 1077 808- 171 979
Hungary 404 1014 1418 336 923 1259
Italy 625 855 1480 462 743 1215
Netherlands - - 1646 - - 1372
Poland 1158 1117 2275 1053 1015 2068
Poland 1108 354 1462 1007 322 1329
Poland 715 1051 1766 650 955 1605
Russia - - - - - -

Table 9: Example 2 (Bored pile) –Resistances, partial action factors, characteristic load, OFS and TSF
R C;mean/ca
Country R c;k R c;d γG γQ Fk OFS TSF
l/
Belgium 2024 1269 1178 1.35 1.5 849 1.49 2.4
France 2154 1703 1548 1.35 1.5 1116 1.53 1.9
FYR Macedonia 1939 1077 979 1.35 1.5 706 1.53 2.7
Hungary 2137 1418 1259 1.35 1.5 907 1.56 2.4
Italy 2370 1480 1215 1.31 1.5 900 1.65 2.6
Netherlands 2140 1646 1372 1.35 1.5 1.51 2.3
Poland (a) 3026 2275 2068 1.35 1.5 1504 1.53 2.0
Poland (b) 1945 1462 1329 1.35 1.5 967 1.53 2.0
Poland (c) 2349 1766 1605 1.35 1.5 1167 1.53 2.0
Russia 989 - - - - - -
1
For structural permanent loads; 1.5 for non-structural permanent loads

54
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3500
1. Belgium 6. Netherlands Rc;cal/mean
3000 2. France 7.Poland (a) Rc;k
3. FYR Macedonia 8. Poland (b) Rc;d
4. Hungary 9. Poland (c) Fk
2500 5. Italy 10. Russia
Resistances and Characteristic load, kN

2000

1500

1000

500

0
1 2 3 4 5 6 7 8 9 10
Countries
Figure 7: Example 2 (Bored pile) –Resistances and characteristic load

3,0
2,5
2,0
OFS and TSF

1,5
1. Belgium 6. Netherlands
1,0 2. France 7.Poland (a) TSF
3. FYR Macedonia 8. Poland (b)
0,5 4. Hungary 9. Poland (c) OFS

0,0 5. Italy
1 2 3 4 5 6 7 8 9
Countries

Figure 8: Example 2 (Bored pile) – Overall factor of safety and total safety factor

The graphs in Figure 7 show that the calculated pile resistances and the resulting characteristic load that
the pile can support are very variable using the different national practices, with coefficients of variation
of 14% for the calculated resistance and a COV value of 22% for both the characteristic resistance and the
characteristic load. The highest resistance was obtained by Poland using laboratory test data while the
lowest value was obtained by Russia. As in Example 1, the OFS value is close to 1.5 in the case of all the
solutions and the TSF values range from 1.9 to 2.7.

3.3. Solutions to Example 2 – Screw pile


Eight solutions were obtained for the screw pile and these are presented in Table 10 to 14. The methods
used to calculate the resistance for this pile were the same as those given in Table 5 for the bored pile.

55
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 10: Example 2 (Screw pile) – Calculation method and calculated resistances
R b;cal R s;cal R c;cal
Country Method
Min. Mean Min. Mean Min. Mean
Belgium CPT 203 246 833 871 1045 1117
France PMT 205 985 1190
Hungary CPT 244 289 625 684 899 973
Italy Lab c u 891 209 1100
Netherlands CPT 244 288 672 707 924 995
Poland (a) I L from CPT 316 320 930 949 1250 1268
Poland (b) CPT 388 392 1045 1108 1439 1500
Russia CPT 84 87 243 265 331 351

Table 11: Example 2 (Screw pile) – Correlation, model and partial resistance factors
Country Correlation factors Model factor Partial resistance factors
ξ3 ξ3 γ R;d γb γs γt
Belgium 1.23 1.18 1.3 1.07 1.0 -
France - - 1.265 1.1 1.1 1.1
Hungary 1.33 1.23 1.1 1.2 1.1 1.15
Italy 1.6 1.0 1.15 1.15 -
Netherlands 1.301 1.301 - - - 1.2
Poland (a) 1.212 1.122 - 1.1 1.1 1.1
Poland (b) 1.21 1.12 - 1.1 1.1 1.1
Russia - - - - - -
1
1.39 for stiff building, 1.30 for flexible building
2
Different from values for a bored pile

Table 12: Example 2 (Screw pile) – Characteristic and design resistances


Country R b;k R s;k R c;t R b;d R s;d R c;d
Belgium 132 549 681 124 549 673
France 162 779 941 147 708 855
Hungary 180 462 642 150 420 570
Italy 557 131 688 484 114 598
Netherlands - - 924 - - 711
Poland (a) 264 783 1048 240 712 953
Poland (b) 324 916 1240 295 832 1127
Russia - - - - - -

Table 13: Example 2 (Screw pile) – Resistances, partial action factors, characteristic load, OFS and TSF
Country Method R c;d γG γQ F k1 OFS TSF
Belgium CPT 673 1.35 1.5 485 1.46 2.15
France PMT 855 1.35 1.5 616 1.40 2.30
Hungary CPT 570 1.35 1.5 411 1.56 2.37
Italy Lab c u 598 1.3 1.5 431 1.55 2.48
Netherlands R c;mean/cal/ R c;k 924 1.35 1.5 711 1.51 2.33
Poland (a) I L from CPT 953 1.35 1.5 770 1.53 1.85
Poland (b) CPT 1127 1.35 1.5 812 1.53 1.85
Russia CPT - - - - - -
1
F k = 4R d /(3γ G + γ Q )

56
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

1600
1. Belgium 5. Netherlands

Resistances and Characteristic load, kN 1400 2. France 6.Poland (a)


3. Hungary 7. Poland (b)
1200 4. Italy 8. Russia

1000

800

600
Rc;cal/mean
400 Rc;k
Rc;d
200

0
1 2 3 4 5 6 7 8 9
Countries

Figure 9: Example 2 (Screw pile) –Resistances and characteristic load

3,0

2,5

2,0
OFS and TSF

1,5

1,0
OFS
TFS
0,5

0,0
1 2 3 4 5 6 7
Countries

Figure 10: Example 2 (Screw pile) – Overall factor of safety and total safety factor

The graphs of the resistances plotted in Figure 9 are similar to those for the bored pile and show that the
highest resistances were obtained by Poland using both the I L values and CPT values. The Russian
solution only provided calculated resistances. The OFS and TSF values were also similar.

4. DESIGN EXAMPLE 3

4.1. Details of Example 3


The third design example is the design of 6 x 12m long 520mm diameter CFA piles for a stiff building in
northern France. The soil consists of 0.8m unspecified material over 2.4m of quaternary silty sand and
sandy silt over 5.4m of dense quaternary sand over a deep deposit of very dense quaternary sand. No
additional surface loads are to be considered so that there is no downdrag. The groundwater level is at a
depth of 1.6m.

The ground investigation involved 1 CPT and 1 PMT plus 1 boring, all located centrally on the site. The
CPT was performed before the installation of the piles while the CPT was performed after the installation
of the piles. The results of the CPT are shown in Figure 7 and the soil properties obtained from laboratory
tests are given in Table 8. The c' and φ' values obtained from shear box tests are presented in Table 8.

57
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 8: Example 3: Soil properties

The ground investigation involved 1 CPT, 1 PMT and 1


boring, all located centrally on the site. The CPT was
performed before the installation of the piles while the
CPT was performed after the installation of the piles.
The CPT results are shown in Figure 7 and the soil
properties obtained from laboratory tests are given in
Table 8. The c' and φ' values were obtained from shear
box tests.
Figure 11: Example 3 - CPT

4.2. Design Solutions to Example 3


As shown in Table 1, 11 solutions were received for this example from Belgium, Denmark, Finland,
Hungary, Italy, Netherlands, Poland (3 solutions), Russia and Sweden. Seven of the design solutions are
based directly on the CPT results, one is based on ID values based on the CPT and one is based on the
angle of friction and density obtained from the laboratory tests. The individual calculated bearing
resistances submitted with the different solutions vary greatly as shown by the 11 sets of values for the
Table 14. The total individual pile resistances range from 1280 to 5093kN with an average value of
2978kN as shown at the bottom of Table 14. Some information about national practices used to calculate
the resistance values is given in Table 15. The Russian solution only provided the calculated resistances.

Table 14: Example 3 (CFA pile) –Calculated resistances


Calculated resistances, kN
Country Method
Base resistance Shaft resistance Total resistance
Belgium CPT 2900 1568 4468
Denmark φ’ and γ 1026 254 1280
Finland CPT 2548 2545 5093
France PMT 1226 2067 3293
Hungary CPT 1402 1650 3052
Ireland CPT 639 2064 2703
Italy CPT 460 1460 1920
Netherlands CPT 1868 1279 3147
Poland I D estimated from CPT 860 1690 2550
Russia CPT 311 1349 1660
Sweden CPT 1790 1530 3320
Average/max/min 1439 1539 2978
Min - max 311 - 2900 254-2545 1290-5093

58
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 15: Example 3 (CFA pile) – Methods used to calculate resistances


Country Method Details
Belgium CPT q b based on de Beer’s method. Factors to obtain R c and R s in
Belgian standard to account for pile installation, scale effect,
and pile shape.
φ’ and γ From lab tests in Table 3.1: φ’ =30 , γ = 18kN/m3 above GWL
o
Denmark
and 10kN/m3 below GWL
Finland CPT Finnish design guide – based on design methods in EC7
France PMT Design according to French standard which is compatible with
EC7. Calculation model based on Ménard limit pressure.
Hungary CPT CPT software package used
Ireland CPT q b = 0.1q c q c = 30.1MPa lower bound value averaged over
1.5D above and below pile base; q s = q c /150 – typical for
bored piles in sand with q c = 17MPa over shaft
Italy CPT Free choice of method on basis of experience and skill
Netherlands CPT Based on NEN 9997-1 – ULS at 20% toe deformation
Poland I D estimated from CPT Calculated according to Polish code PN-83/B-02482
Russia CPT -
Sweden CPT LCPC method - Bustamente & Gianesellis (1982)

Table 16: Example 3 (CFA pile) – Correlation, model and partial resistance factors
Model
Correlation factors Partial resistance factors
Country Method factor
ξ3 ξ4 γ R;d γb γs γt
Belgium CPT 1.211 1.131 1.352 1.13 1.0 1.1
Denmark φ’ and γ 1.5 1.5 -4 1.3 1.3 1.3
Finland CPT 1.68 1.68 - 1.2 1.2 1.2
France PMT - - 1.625 1.1 1.1 1.1
Hungary CPT 1.4 1.4 1.1 1.2 1.1 1.15
Ireland CPT 2.1 2.1 1.0 1.6 1.3
Italy CPT 1.4 1.4 1.0 1.3 1.15 1.1
Netherlands CPT 1.39 1.39 1.0 1.2 1.2 1.2
Poland I D from CPT 1.4/1.1=1.27 1.27 1.0 1.1 1.1 1.1
Russia CPT - - - - - -
Sweden CPT 1.4/1.1=1.27 1.4/1.1=1.27 1.4 1.3 1.3 1.3
1
ξ values from tables in Belgian code for 1 CPT and 6 piles over an area of 104m 2
2
γ Rd from table in Belgian code when there is no pile load test; γ Rd = 1.0 when piles on site tested
3
γ b for CFA pile when there is no quality assurance; γ b = 1.0 for CFA pile when there is quality
assurance
4
γ R;d Model factor not used in Denmark

Table 17: Example 3 (CFA pile) – Characteristic and design resistances


Country Method R b;k R s;k R c;k R b;d R s;d R c;d
Belgium CPT 1775 960 2735 1614 960 2574
Denmark φ’ and γ 684 169 853 210 130 340
Finland CPT 1517 1515 3032 1264 1262 2526
France PMT 969 1634 2603 881 1485 2367
Hungary CPT 910 1071 1982 759 974 1723
Ireland CPT 304 983 1287 190 756 946
Italy CPT 329 1042 1371 253 906 1159
Netherlands CPT 1344 920 2264 1120 767 1887
Poland I D from CPT 677 1331 2008 615 1210 1825
Russia CPT - - - - - -
Sweden CPT 1006 861 1867 774 662 1436

59
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 18: Example 3 (CFA pile) – Resistances, partial action factors, characteristic load, OFS and TSF

OFS = TFS =
Country Method R c;d CC ξ1 K FI γG γQ F k2
R c;k /F k Rc; av;cal /F k
Belgium CPT 2574 1.35 1.5 1855 1.5 2.4
Denmark Lab tests 340 1.35 1.5 245 3.5 5.2
Finland CPT 2526 2 1.0 1.15 1.5 2014 1.5 1.9
France PMT 2367 - - 1.35 1.5 1706 1.6 2.5
Hungary CPT 1723 - - 1.35 1.5 1242 1.5 3.1
Ireland CPT 946 - - 1.0 1.3 880 1.6 2.3
Italy CPT 1159 - - 1.3 1.5 859 1.7 2.3
Netherlands CPT 1887 1.35 1.5 1360 1.5 1.9
Poland I L from CPT 1825 1.35 1.5 1315 1.5 2.4
Russia CPT - - - - - - - -
Sweden CPT 1436 2 0.89 0.91 1.35 1.5 1237 1.1 2.7
1
ξ reduction factor from EN 1990 in Swedish NA used in load combination Equation 6.10b of EN 1990
2
F k = 4R d /(K FI (3ξγ G + γ Q ))

6000
1. Belgium 6. Ireland
2. Denmark 7. Italy Rc;cal
3. Finland 8.Netherlands Rc;k
Resistances and Characteristoc load, kN

5000
4. France 9. Poland Rc;d
5. Hungary 10. Sweden Fk
11. Russia
4000

3000

2000

1000

0
1 2 3 4 5 6 7 8 9 10 11
Countries
Figure 12: Example 3 (CFA) – Resistances and characteristic load

6,0
1. Belgium 6. Ireland
2. Denmark 7. Italy OFS
5,0
3. Finland 8.Netherlands TFS
4. France 9. Poland
4,0
5. Hungary 10. Sweden
OFS and TSF

3,0

2,0

1,0

0,0
1 2 3 4 5 6 7 8 9 10
Countries
Figure 13: Example 3 (CFA pile) – Overall factor of safety and total safety factor

60
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

The graphs in Figure 13 show that the lowest resistances for the CFA pile were obtained by Denmark,
which was the only country to use the laboratory test results to calculate the pile resistance. The low
resistance occurred because Denmark restricts both the base and shaft resistance in the case of a CFA
pile. According to the Danish report, since the CFA pile is a bored pile, its design base resistance is
limited to 1000kPa unless load tests are carried out and its shaft resistance is limited to 30% of the shaft
resistance of a corresponding displacement pile. Using φ’ = 30 gives N q = eπtanφ’tan2(45+φ’/2) = 18.4 and
o

using γ = 18kN/m above and 20kN/m below the groundwater table and γ w = 10kN/m3 to calculate the
3 3

effective vertical stress at the base of the pile, the base resistance is given by R b;cal = 2A b q b N b =
2x0.21x133x18.4 = 1028kN. To determine the characteristic base resistance, R b;k the calculated base
resistance is divided by the correlation factor ξ = 1.5, regardless of the number of investigations. The
design base resistance is obtained by the dividing R b;k by the partial resistance factor, γ R = 1.3, hence R b;d
= R b;cal /(1.5x1.3) = 527kN. Since the area of the base is 0.21m2, the design base bearing pressure σ b;d =
2509kPa, which greatly exceeds the maximum allowable design bearing pressure of 1000kPa according
to the Danish practice. Assuming the maximum design bearing pressure occurs, the maximum design
base resistance is 210kN. The calculated shaft resistance is given by R b;cal = 0.3N m ∑A s,i q m,i =
0.3x0.6x(πx0.52x1.6x0.8x18 +πx0.52(1.6x18+5.2x10)) = 254kN, where Nm is the friction coefficient for
a friction pile. As with the base resistance, the design shaft resistance is obtained by dividing calculated
shaft resistance by the correlation factor ξ = 1.5 and the partial resistance factor, γ R = 1.3 so that R s;d =
R s;cal /(1.5x1.3) = 130kN.

The following comments are made on the Danish solution:


1) The given values of φ’ range from 30.2 to 32.1 for the soil which is described as dense or very dense
o o

sand compared to the φ’ values of 40 to 42 given in Table D.1 of EN 1997-1.


o o

2) Even if a higher value of φ’ were used in the calculations, it would have no effect on the design
resistance as it is already at the maximum value allowed in Denmark, unless load tests are carried out,
and the friction coefficient is constant regardless of the φ’ value.
3) Rather than using model factors to correct the partial factors γ b and γ s in accordance with the note to
7.6.2.3(8), in Denmark a correlation factor ξ = 1.5 is applied to the calculated resistances.

As the graphs in Figure 12 show, there was a great variation in the calculated resistances for the CFA
piles, more than for the other piles. The coefficient of variation of the calculated resistances was 37%
resulting in a COV of both the characteristic resistance and characteristic load of 41%. In spite of this as
may be seen from Figure 13, all the solutions, except for the Danish solution, had OFS values within the
range 1.5 to 1.8 and TSF values in the range 1.9 to 3.2.

5. CONCLUSIONS
Examination and comparison of the solutions to the design examples submitted by the different countries
has shown that:
1. The design of pile foundations in Europe is being carried out according to, or so as to be compatible
with, Eurocode 7.
2. Harmonisation has occurred in the design method and consistent overall factors of safety (OFS) are
used.
3. However, consistent overall factors of safety does not meant that the outcome of the designs are
similar.
4. Since Eurocode 7 does not provide any calculation models for pile design, different methods, mostly
presented in national standards or guides, are used to calculate the pile resistances.
5. Most of the national practices are based on successful national experiences involving particular testing
methods, soil conditions and calculation models.
6. Many of the national methods to calculate pile resistance involve additional factors or requirements to
those in Eurocode 7, for example factors relating to pile type or pile shape.
7. The consequence of the different methods for calculating pile resistances is that there is a great
variation in the calculated resistances and hence in the calculated characteristic loads that the piles
could support with these having a COV value of 41% in the case of the CFA pile.
8. The TSF values were found to lie mostly within the range 2.0 to 3.0 and hence are consistent with
previous practice before the introduction of Eurocode 7. In response to the question, “How did
Eurocode 7 changes daily practice?”, the comment in the Danish report is “In Denmark, not much, if
at all.”

61
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

9. For more harmonisation to occur, there must be more agreement on the models to calculate pile
resistance, but is that possible considering the different soil conditions, testing methods, pile types,
installation methods and experiences that exist in Europe.
10. There was almost no consideration of serviceability in the solutions received. It appears the existing
practices for ULS design also cover SLS design.

Not many comparisons of the results national practice for pile design to Eurocode 7 have been published
to date. One example is the comparison by Frank (2005) of a pile designed from static load test results by
engineers from different countries for the International Workshop on the Evaluation of Eurocode 7 which
held in Dublin in 2005 (Orr, 2005). Similar to the author, Frank found that a great variety of calculation
models, which he showed were related to national practice and experience, were used to calculate the pile
reistance and hence a very large range of pile lengths was obtained for the pile designs. Simpson (2005),
who reported at this workshop on the solution of retaining wall design examples carried out to Eurocode
7, noted that the solutions to the retaining wall examples showed considerable scatter, even when
calculated by different contributors using nominally the same method. These results represented a
considerable range of safety and hence he considered that some would have been unsafe if used in
practice and that factors of safety have an important role in geotechnical design in covering for human
error.

The author concludes from this review of the solutions received for the pile examples that for more
harmonisation with regard to pile design to occur, there must be more agreement on the models to
calculate pile resistance. However, that raises the question if that is possible considering the different soil
conditions, testing methods, pile types, installation methods and experiences that exist in Europe.

ACKNOWLEDGEMENTS

The author wishes to acknowledge and thank the members of ETC3 and their colleagues for all their work
in preparing the national reports and the solutions to the design examples upon which this keynote paper
is based.

REFERENCES

Frank, R (2005) Evaluation of Eurocode 7 – Two pile foundation examples, Proceedings the
International Workshop on the Evaluation of Eurocode 7, Dublin, Department of Civil, Structural and
Environmental Enginering, Trinity College Dublin, 117-125

Orr, T.L.L. (2005) Proceedings the International Workshop on the Evaluation of Eurocode 7, Dublin,
Department of Civil, Structural and Environmental Enginering, Trinity College Dublin, edited by T Orr,
pp320

Simpson, B (2004) Eurocode 7 Workshop – Retaining wall examples 5-7, Proceedings the International
Workshop on the Evaluation of Eurocode 7, Dublin, Department of Civil, Structural and Environmental
Enginering, Trinity College, Dublin, 127-146

62
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Recent experiences with static pile load testing on real job sites
Jan Verstraelen, TUC Rail, Belgium, jve@tucrail.be

Wim Maekelberg, TUC Rail, Belgium, maw@tucrail.be

Marine Medaets, TUC Rail, Belgium, mmd@tucrail.be

ABSTRACT
Static pile load tests are sometimes carried out to verify a design, but are rarely used as design tools to
optimize or change the final design. Cases studies on driven cast-in-place piles, displacement screw piles
and CFA piles will be presented. In all presented cases, the SLT’s led to new insights regarding the
influence of pile installation or local soil conditions and led to a change in the initial foundation design.
The results of the SLT’s should not only be used to verify the pile’s bearing capacity, but can also be used
to estimate single pile and also group pile settlements.

1. HISTORICAL BACKGROUND
Up until the publication of the Belgian national guidelines in 2009, several owners and engineers used
different standards and calculation methods for the design of axially loaded bearing piles. Some of these
were based on the already published and normalised standards from neighbouring countries, mainly the
Dutch and French standards. Others were based on the locally developed the De Beer method for
calculating the ultimate pile base capacity.

In the Belgian pile design guideline of 2009, the original De Beer method has been retained. Compared to
others standards, the De Beer method distinguishes itself by using an analytical approach based on the
scale effect to calculate the ultimate pile base capacity (N. Huybrechts et al, 2016). The method and later
modifications of the De Beer method have also been reported in ECSMFE and ICSMFE (BGGG-GBMS,
1985) proceedings by De Beer and Van Impe among others.

Figure 1 compares the results of different analytical methods to determine the pile base unit capacity
based on CPT results. As can be seen, differences of up to 35% from the average value can be obtained
depending on which method was used.

Figure 1: comparison between unit bearing resistance (for driven piles-Db=0.6m) based on CPT results,
using different national standards or guidelines

63
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Part of these differences can be explained by the historical background of the different methods.

The Koppejan method as used in the NEN standard pre 2016 is an empirical method based on the
resemblance between a CPT test and a statically jacked pile (Plantema, 1948). Results of the De Beer
method, which also uses the scale effect, show great resemblance to the Koppejan method. Recently, and
based on the analysis of several pile load tests on driven piles (Van Tol, 2012), the results of the
Koppejan method should be reduced by 30%, shown as the NEN - post 2016 line in figure 1.

The French standard NF P94 262 relies more on the results of pressiometer tests but also includes a
method for calculating the unit bearing capacity based on CPT results. The method is also empirical and
based on the results of instrumented static pile load tests. Since more data is available for pile load tests
correlated to pressiometer tests, it is stated that the CPT based method is “conservative”.

The ICP method or Imperial college method as published by Jardine et al, 2005, is an empirical method
developed for the bearing capacity of open end and closed end tubular piles and aimed at the design of
offshore wind farms. Although being one of the simplest of all considered methods, predictions lie within
the range of the NF and NEN – post 2016 methods as can be seen in figure 1.

After the post 2016 reduction, the NEN, NF and ICP methods lead to predictions within a range of +-
15%. The De Beer method leads to the largest bearing capacity sometimes up to 45% higher than the
most conservative method.

The great differences between the national standards and guidelines show that there must remain a great
uncertainty and spread in the prediction of pile unit base resistance.

Since most of the techniques used to install piles rely on driving, screwing or boring, installation effects
are introduced to make the transition from the statically pushed CPT sounding rod to the equivalent pile
behaviour. For the De Beer method in the Belgian national guideline, the proposed installation factor
equals to 1 for driven, vibrated or jacked piles.

All of the standards and guidelines put emphasis on the calculation of ultimate bearing capacity and less
on pile settlement calculations:

• NEN uses normalised pile base and shaft load settlement graphs for 3 different pile categories,
• NF uses a t-z approach with (bilinear) transfer functions based on the E m -modulus measured in
pressiometer tests,
• The Belgian guideline does not give methods for calculating pile settlements although De Cock, 2008,
proposed parameters for hyperbolic transfer functions based on the analysis of several instrumented
pile load tests.

In the past, a safety factor of 2 between serviceability and ultimate bearing capacity was considered
sufficient to disregard pile settlements at working load. For critical structures, a safety of 3 on the pile
base bearing capacity was applied to limit working load settlements to less than 10 mm (Tomlinson,
2008).

In practice, also after using Eurocode pratice, it is usually assumed that settlements of piled foundations
are negligible and should only be considered in the case of underlying weak compressible layers, in which
case Terzaghi’s equivalent raft method is often used.

2. EUROCODE‘S PARTIAL SAFETY FACTORING

2.1. Design based on soil tests (CPT’s)


Since the use of Eurocode 7, a deterministic safety factor is no longer used but a combination of different
partial safety factors needs to be applied (probabilistic approach). The design load F c,d is calculated as 1 x
permanent load + 1,1 x variable load. Although F c,d should strictly be called an ULS load, it corresponds
closer to an SLS load and is further designated as SLS load.

64
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

The ratio between this design load F c,d and the ultimate load Q l consists of the following different safety
factors:
• A partial load factor γ F = 1,35 for permanent loads and = 1,5 for variable loads,
• A model factor γ Rd between 1 and 1,35 to calculate characteristic values based on the calibration
between CPT results and static pile load tests. The value depends on whether or not pile load tests
have been carried out (on site or not) and the pile type. The model factor thus reflects the spread in the
calibration between predictions and measured values for different pile types (probabilistic),

• A correlation factor ξ 3 and ξ 4 on the average and the minimum characteristic value to take the
variation on the soil characteristics and the uncertainty on this variation into account (probabilistic).
The values of ξ 3 and ξ 4 are determined by the number of CPT tests carried out, the number of piles
underneath the foundation and the possibility of the foundation to redistribute pile load forces.

As a result, the total safety factor (Q l /F c,d ), which is the product of all of the above mentioned partial
safety factors, can vary greatly between a theoretical minimum of 1,35 and maximum of 2,55. In many
cases, the designers are not aware of the total safety factor due to the different partial factors,
combinations and possibilities. Typically used total safety factor lie around 1,5 and are thus substantially
lower than those used in the previous deterministic approach.

It must be reminded that in the case of a lower total safety factor, pile settlements can no longer be
disregarded and should be taken into account in the design. Whether or not pile load tests have been
carried out (γ Rd ) or the number of CPT’s that were executed (ξ 3 and ξ 4 ), they both have no influence on
the pile settlement and only influence the allowable working load.

2.2. Design based on static load tests (SLT)


When pile load tests are used to predict pile bearing capacity according to Eurocode, similar partial safety
factors can be applied with the exception of:

• A model factor γ Rd of 1 since bearing capacities were measured directly and are not based on
predictions and correlations.

• A correlation factor ξ 1 and ξ 2 on the average and the minimum measured bearing capacities to take
the variation on the soil characteristics and the uncertainty on this variation into account
(probabilistic). This value is determined by the number of static load tests carried out on site, and the
possibility of the foundation to redistribute pile load forces.

As a result, the total safety factor, which is the product of the above mentioned partial safety factors, can
vary between a theoretical minimum of 1,35 and maximum of 2,1. Typically, a smaller number of pile
load tests are available so the total safety factor lies around 1,65 when 2 SLT’s are available and pile load
redistribution is allowed.

The design based on the results of pile load tests is based on the results of the ultimate or pile capacity at
10%D pile base displacement. The creep load is also measured but is not taken into account in
determining the allowable pile load according to the general Eurocode formulation or the Belgian
guideline, contrary to f.e. the NF P 94 270 standard.

Unlike the design based on CPT tests, a design based on static pile load tests allows determining the
single pile settlements at working loads. However, these settlements will most often not be in agreement
with the pile group settlements.

3. AN ALTERNATIVE APPROACH FOR DESIGNING BASED ON PILE


LOAD TESTS
In most cases, pile load tests are carried out to verify the contractor’s execution or check the designers’
assumptions. Only very rarely, it is reported that the final design was based on the results of the pile load
tests.

The pile load tests presented in this article were all used to alter the preliminary design based on CPT
tests. The SLT’s were carried out before the start of the production piling, which poses logistical and
planning difficulties. The cases in which pile load testing was used as a design tool consist typically of

65
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

larger projects in which >1000 piles were installed. When the testing is carried out at the start of the
project, this does not have to lead to planning difficulties.

Alternatively, sometimes special pile types or deep foundation techniques are used that are not covered by
the national guidelines, such as jet grout piles, post-grouted bored piles or grout injection screw piles.

3.1. Testing and instrumentation


All tests presented in this article are fully instrumented tests in which strains are measured every meter
along the pile shaft. Also extensometer, vibrating wire and optic fibre instrumentation were used with
success in other projects. The instrumentation, loading and interpretation of the raw data was handled by
specialised companies such as BBRI and APTS. For all SLT’s described in this article, optic fibre
measurements were used.

The optic fibres were installed only very shortly before the test itself in reservation tubes placed during
pile installation. Strains developed shortly after pile installation, such as those leading to residual loads,
were thus not measured.

The pile loading procedure follows the NF P94 150 standard with constant time steps of 60’ and
continuous load/displacement/strain measurements. Only one load cycle was applied, with loading carried
out till a predetermined theoretical failure load, 10%D pile base settlements or ultimate load which could
not be maintained. Subsequently, an unloading cycle was carried out. As reaction device, mostly tension
piles were used but also crown load testing and dead load testing were performed.

Typically, 2 or more SLT’s were carried out in the centre of previously executed CPT’s. Theoretical
failure loads were calculated based on the previously executed CPT, and an ultimate load of 120% of that
theoretical ultimate load was used to design the reaction device and load system.

3.2. Design of bearing capacity based on SLT results


Due to the large number of required SLT’s to base a design only on SLT’s, a combination of SLT’s and
CPT’s was used to design the pile’s bearing capacity.

Since the piles are fully instrumented, the results allow to determine the pile base installation factor α b as
well as the shaft installation factor α s , and if required, different values for each identified layer can be
used. The final design uses these installation factors (usually the minimum value) together with additional
CPT tests to determine the production pile’s length and pile tip level.

Shaft friction in upper compressible layers was ignored and subtracted from the tests results presented
further.

A model factor γ Rd of 1 is assumed since bearing capacities are measured directly and are not based on
predictions and correlations, such as is the case when designing with CPT’s.

A correlation factor ξ on the minimum characteristic value was applied to take the variation on the soil
characteristics and the uncertainty on this variation into account. This value depends on the number of
CPT and SLT tests carried out.

Additionally, the SLS design load was checked with the measured creep load and the partial safety factors
according to NF P 94 270: F c,d <= R c,cr,k /0.9 for permanent and variable load combination and F c,d <=
R c,cr,k /1.1 for permanent load combination. This last requirement is translated in a check on the ratio
between ultimate load Q l and creep load Q c which should be sufficiently small.

This leads to the following expressions:


𝑄𝑄𝑙𝑙
𝐹𝐹𝑐𝑐,𝑑𝑑 ≤ (1)
1,35 𝑥𝑥 𝜉𝜉

And
𝑄𝑄𝑐𝑐
𝐹𝐹𝑐𝑐,𝑑𝑑 ≤ (2)
0,9 𝑜𝑜𝑜𝑜 1,1

66
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

A typical value of 1,1 is proposed as a first approximation and minimum value for ξ.

In a case of predominantly permanent loads, this leads to a ratio of Q l /Q c that should be smaller than
1,35. When loads are predominantly variable loads, this ratio may rise to 1,65.
When the measured ratio is higher than can be accepted, the creep criterion determines the allowable load
and the coefficient ξ is raised to have the same limit as when the allowable load is based on the ultimate
load.

Figure 2 shows the measured load displacement curve for a driven cast-in-place pile with D b =0,550 and
D s =0,508 together with the ultimate load Q l (@10%Db) and creep load Q c . The ratio Q l /Q c is 1,17 so the
creep load is not used to adapt the SLS load (F c,d ) based on the measured Q l . The allowable SLS load F c,d
was calculated as Q l /(1,35x1,1). Measured single pile base displacements are 12mm at F c,d load.

Figure 2: results of SLT on driven cast in place pile, Db = 0.550m

Figure 3 shows the measured load displacement curve for a displacement screw pile with D b =0,56 and
D s =0,56 together with the ultimate load Q l (@10%Db) and creep load Q c . The ratio Q l /Q c is 1,61 so
distinction is made between two different load cases:
• When SLS loads are predominantly variable loads, and since Q l /Q c is smaller than 1,65, the allowable
SLS load F c,d can be calculated as Q l /(1,35x1,1) = F c,d characteristic.
• When SLS loads are predominantly permanent loads, and since Q l /Q c is smaller than 1,65 but higher
than 1,35, the value of ξ was raised to 1,31 to satisfy F c,d quasi-permanent <= Q c /1.1.

Measured single pile base displacements are 7,4 mm at F c,d – quasi-permanent and 13,5 mm at F c,d -
characteristic. One can notice the rather larger difference in settlement, due to the fact that both loads lie
around the measured creep load and although the loads vary only by a ratio of 1,18, the measured
settlements vary by a ratio of 1,82! This large difference in settlement is the raison why the creep criterion
was used in the pile design as presented here.

67
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 3: results of SLT on displacement screw pile, Db = 0.56m

This methodology was carried out for all instrumented piles on a job site and the minimum measured
installation factors were used for the design. The creep limit was transformed into a ratio between Q l /Q c
in order to be able to use this ratio for further design calculations. The creep limit is only measured during
pile load tests and cannot be calculated using CPT based methods.

The measured/calibrated installation factors together with the additional safety factor ξ were used to
design the production piles based on additional CPT’s within the foundation footprint and using analytical
methods (i.e. Belgian guideline).

Production piles were usually the same or smaller diameter than the test piles. Pile base level was
designed based on additional CPT’s and using the calibrated installation factors from the SLT’s.

3.3. Design of pile group settlements based on SLT results


Single pile settlement measured at SLS load cannot be assumed to be identical to the pile group
settlement at similar loads. As still common practice, Terzaghi’s equivalent raft method is used to
determine the pile group settlement. However, this method has the following limitations:
• Much of the success of the method relies on the chosen stress distribution along the pile group as well
as the depth of the equivalent raft,
• Centre-to-centre distances between piles cannot be taken into account, only the pile group
circumference is considered,
• The type of pile cannot be taken into account: f.e. driven and screw displacement piles can both be
end-bearing but will lead to different group settlements.

As explained by Poulos, 2006, the equivalent raft method should be limited for pile groups with more
than 16 piles and typical centre-to-centre distances of 3 pile diameters.

Due to advances in analytical techniques and experimental research, the equivalent raft approach is
becoming more and more replaced by more advanced techniques such as the interaction method and finite
element calculations. With the interaction factor method, it is proposed that the final settlement profile is
a combination of the deformation profile of each pile within its range of influence. Figure 4, taken from
Fleming’s Piling Engineering handbook, 2008, shows how the deformation profile of each pile should be
combined to create the combined settlement profile in the case of equivalently loaded piles without a pile
cap.

68
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 4: combined settlement of profile through combination of individually equivalently loaded piles
(from K. Fleming, Piling Engineering, 2008)

In the presence of a very stiff pile cap, which imposes equal pile and raft settlements, this leads to stiffer
piles around the edges and thus greater piles loads for corner piles. This is similar to the contact stresses
underneath a stiff direct foundation.

Randolph and Roth (1979) describe how the deformation filed around the pile shaft varies logarithmically
with distance while that of the pile base is related to 1/distance. The evolution of the interaction factor
with increasing centre-to-centre spacing is shown in figure 5. For typical centre-to-centre distances of 3
diameters and more, ignoring pile base interaction leads only to a smaller error. The shaft interaction
continuous to have an influence up to 10D spacing and more.

69
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 5: evolution of shaft and base interaction factor with increasing centre-to-centre distances,
according to Randolph and Wroth, 1979

Mylonakis and Gazetas (1998) modified the above approach to take into account the effect of the pile
stiffness of neighbouring piles and modified the shaft interaction factor through a diffraction factor. This
leads to an analytical approach in which a set of equations can be solved with chosen boundary
conditions. In most cases, this boundary condition is a rigid pile cap. A result of such an analysis is shown
in figure 6. On the left hand side, the results of solving the analytical equations using the interaction
factors is shown for a pile group of 9 piles, diameter 0.9 m and centre-to-centre distances of 3D. Since the
technique is rather laborious, commercially available software such as DPile is used in further analysis
presented here. The results of the DPile solution of the same geometry but based on a different calculation
model (Poulos model) are shown on the right hand side of figure 6. The results of both solutions not only
represent the distribution of pile loads but also calculate the pile group settlement.

Figure 6: comparison between analytical solution with interaction factor method and DPile

70
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

A similar interaction factor method exists to take into account the group effect of lateral loading. In DPile,
usually the “cap soil” or “cap-layered soil” models are used. These models also allow taking into
consideration inclined piles, different pile types and lengths, ... The single pile’s response is modelled
through t-z (shaft) and pile tip curves together with lateral p-y curves. The interaction between the piles
through the soil is modelled with the E-modulus and Poisson ratio of the surrounding soil. Multiple layers
can be used with the cap-layered soil model.

When SLT’s are carried out, the measured pile tip and pile shaft load displacements curves are used in the
design calculation.

Most importantly, the above mentioned methods show that:


• For a pile within a pile group, relatively more load is carried by the pile base than for a single pile,
• The settlement of closely space pile groups is determined mostly by the load-displacement behaviour
of the pile base.

The ASIRI research, which led to an experimentally deduced method for designing rigid inclusions, noted
that the settlement of the pile group corresponded very well and could be estimated as the settlement of
the pile base as measured in a single pile load test.

As a result, a lot of attention was paid to the pile’s base capacity and load displacement behaviour during
the analysis of the SLT’s.

4. CASE 1: DRIVEN CAST-IN-PLACE PILES


The first case discusses the results of 2 SLT’s on driven cast-in-place piles. The piles were installed on a
job site at 300 m distance from each other and in similar ground conditions. Penetration depth within the
bearing layer was more than 5 m. The casing diameter was 508 mm, the pile base was 550 and 600 mm
respectively.

The piles were driven with a hydraulic hammer at 72 kNm energy output and a blowcount between 8 to
10 blows/25cm along the final 2 metres. After reinforcement installation and pouring of the concrete, the
casing was driven back.

Figure 7: CPT in axis of test pile before installation together with CPT at 30cm from pile after
installation

71
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 7 shows the comparison between the CPT’s prior to installing the pile and after installation of the
pile. The soil underneath the pile base seems compacted by the pile driving and the pile base level can be
clearly identified. This was evident in all 4 of such comparisons (2 axial and 2 lateral SLT’s were carried
out). Along the pile shaft, a greater difference was noted between all 4 CPT tests. In all of them, the CPT
values in the upper layer were hardly affected by the pile driving. In the deeper more sandy layers, an
increase in cone resistance could be noted. The analysis presented in figure 7 is an average result which
shows a smaller increase within some layers of the bearing sand and a sharp increase underneath the pile
base level. In some tests, the cone resistance along the shaft seemed barely increased while others showed
a larger increase.

The piles were tested 26 and 28 days after installation. Figure 8 shows the ratio of measured to calculated
unit bearing resistance (Q b,meas /R bu,calc ) versus the ratio of pile base settlement with pile diameter (s b /D b ).

Figure 8: normalised results of measured/calculated unit bearing resistance (De Beer method)

Since the De Beer method and Belgian guideline assumes an installation factor of 1, the curves should
converge to a value of 1 at 10%D pile base settlement. This is clearly not the case. Also, it can be noticed
that full bearing capacity is not yet reached at a pile base displacement of 10%D and continuous to rise till
15 and even 20%D displacement.

Comparisons of the ratio between measured and calculated values according to different methods are
summarized in table 1. In the NEN method, a reduction of 30% of the Koppejan value has already been
applied.

Table 1: comparison of measured unit bearing resistance with calculated unit bearing resistance
according to different standards (q bmeas /q b,calc )
De Beer NF P 94 270 ICP NEN post 2016
driven cast in place (D b =0.55) 0.54 0.84 1.03 0.74
driven cast in place (D b =0.60) 0.56 1.05 1.10 0.82

Only the ICP method seems slightly conservative, all other methods are too optimistic with the De Beer
method being overoptimistic by almost 50%.

These results are in agreement with the analysis by Van Tol, 2010 and 2012, which showed that the ratio
of measured to calculated unit bearing resistance decreases with increasing pile penetration ratios. The

72
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

piles tested in this case study have penetration depths of 10xD b and more and should therefore lead to
smaller installation factors.

Figure 9 shows the normalised measured to calculated shaft bearing resistance. Although shaft
dimensions are identical, the results are very different for both piles.

Figure 9: normalised results of measured/calculated shaft bearing resistance (Belgian guideline)

When looking in detail at both results, it appeared that this difference can be attributed completely to the
difference in friction along the bottom 3 to 4 m of the pile (figure 10). In the other SLT, this is where
most of the shaft’s bearing capacity is generated, hence the rather large difference between both test
results.

73
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 10: pile load along shaft, Db=0.600,Ds=0.508

The negative effect of larger base plates is known but, in most guidelines, only for larger dimensions of
the enlarged base plate. In the Belgian guideline, a reduction on the pile base capacity is only introduced
for larger ratios of base plate to casing diameter and also the reduction is applied on the base capacity and
not on the shaft capacity. This value equals to λ=0.9 for the larger base plate dimensions but has not been
applied in figure 8 and table 1.

In Tomlinson (2008), it is reported that shaft friction should be disregarded along a length of 2D above
the enlarged pile base for belly reamed bored piles since the soil moves downward with the pile base
when the pile is loaded. This would be in agreement with the observation in figure 10 that for small pile
base displacement, friction is generated along the bottom part of the pile but that disappears again for
greater pile base settlements. However, underreamed bored piles have an enlargement which is created
after the pile has been bored till final depth and with pile base diameters of 2 times that of the shaft. This
is not the case here since the base plate has an oversize of only 5 cm.

However, unlike the belly reamed bored piles, the base plate is present during driving and pushes the soil
underneath the plate down and outwards. It is assumed that the concrete fills the created void during
pullback of the casing or that the soil moves inwards after the base plate has passed. This could only be
confirmed by excavating the pile, which was not in the scope of the project. Instead, production piles
were all installed with the smaller diameter base plate.

The design was based on the measured installation factors together with additional CPT’s within the
foundation footprint. The interpretation of the SLT for the pile with small base plate was presented in
figure 2. Together with the results of lateral load tests (J. Verstraelen, 2015), both the lateral and axial
curves where introduced in the calculation model.

74
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 11: foundation design based on axial and lateral SLT’s

Figure 11 shows the final design with the pile location and rake and axial load for each pile in a 149° pile
group.

5. CASE 2: DISPLACEMENT SCREW PILES


A second case shows the results of 2 SLT’s on displacement screw piles. The piles have a 560 mm
diameter and are installed with a tapered displacement screw and counterrotating screw above the soil-
displacement body. The test piles are situated on a job site at 150 m distance from each other and in
similar ground conditions. Penetration depth within the bearing layer was limited to 1.5 m or thus
approximately 3xD b .

The piles were drilled with a fixed leader machine with heavy torque rotary couple of 45 tonm and pull-
down capacity of 10 ton.

Figure 12: CPT in axis of test pile before installation together with CPT at 30cm from pile after
installation

75
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 12 shows the comparison between the CPT’s prior to installing the pile and after installation of the
pile. It appears that soil relaxation has occurred in the sandy layers. This is especially true in the vicinity
of the pile base. The Tertiary clay layer from 4 to -1 mTAW (member of Merelbeke) is hardly affected by
the pile installation process. During installation of the pile, nothing seemed out of the ordinary and no
difference could be noted on the pile-registration or by the present site engineer.

The piles were tested 35 and 37 days after installation. Figure 13 shows the ratio of measured to
calculated unit bearing resistance (Q b,meas /R bu,calc ) versus the ratio of pile base settlement with pile
diameter (s b /D b ).

Figure 13: normalised results of measured/calculated unit bearing resistance (De Beer method)

Since the Belgian guideline assumes an installation factor of 0,7, the curves should converge to a value of
0,7 at 10%D pile base settlement. It can be noticed that full bearing capacity is not yet reached at a pile
base displacement of 10%D and continuous to rise till 15 and even 20%D displacement. Only at these
higher levels of pile base settlement, the installation factor of 0,7 is reached.

When comparing the results with that of the driven piles (figure 8), it can be seen that the pile base
behaves less stiff. The ultimate ratio of measured to calculated unit bearing capacity seems higher than
that of the driven piles. However, it must be reminded that the penetration ratio of the screw piles is far
less than that of the driven piles.

Comparisons of the ratio between measured and calculated values according to different methods are
summarized in table 2. In the NEN method, a reduction of 30% of the Koppejan value was already
applied, but the installation factor α p = 0,9 has not yet been applied.

Table 2: comparison of measured unit bearing resistance with calculated unit bearing resistance
according to different standards (q bmeas /q b,calc )
De Beer NF P 94 270 ICP NEN post 2016
displacement screw pile (Db=0,56) 0.49 0.75 0.68 0.81
displacement screw pile (Db=0,56) 0.51 0.73 0.67 0.78

All of the methods seem too optimistic. For the ICP method, it must be noted that no guidelines are given
for screw piles. However, the spread in the results is very small for all of the considered methods.

Figure 14 shows the normalised measured to calculated shaft bearing resistance. Although shaft
dimensions are identical, the results are very different for both piles.

76
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 14: normalised results of measured/calculated shaft bearing resistance (Belgian guideline)

In Figure 14, only friction is considered in the Tertiary clay and the bearing sand underneath it.
Theoretical installation factors are 0,9 for the clay and 1 for the sand with equal shaft bearing
capacities/meter pile. One of both tests confirms these installation factors, but the other tests shows
considerably lower values.

As shown in figure 12, one of the CPT tests showed a decrease in cone resistance around the pile tip.
Whether or not this decrease is linked to the smaller shaft resistance as measured in the test is not known.
Either case, it is remarkable that the pile base capacity seems unaffected, while the cone resistance
beneath the pile base was also reduced in one of both CPT tests.

As described by Fleming, 2008, soil relaxation may occur if the pile’s rotational speed is greater than the
pile’s advance rate. This will lead to overflighting, as shown in figure 15.

Figure 15: overflighting of CFA-piles (from K. Fleming, Piling Engineering, 2008)

Although this is a typical problem for CFA-piles, this effect is also possible for displacement augers
especially underneath the soil displacement body.

As can be seen in figure 16, along the upper meter the advance rate per rotational cycle remains constant
and corresponds to the maximum pitch of the displacement auger (the auger has a variable pitch to limit
soil relaxation). When the drill reaches the harder bearing sand layer, the advance rate quickly decelerates
but rotation speeds remain constant. The rotational couple rises but remains lower than the drill’s

77
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

capacity. While drilling in those layers, soil is transported upwards until it is pushed aside at the
displacement body. The soil relaxation is most noticeable along the length underneath the displacement
body (also shown schematically but to scale in figure 16).

Figure 16: CPT before and after pile installation together with advance rate/cycle during pile installation

However, this decline in advancement per rotation is not that different from that noted in the other test
pile, as can be seen in figure 17.

Figure 17: CPT before and after pile installation together with advance rate/cycle during pile installation

The results of the CPT tests after installation are also presented, and show no decrease in cone resistance.
It can only be concluded that registration alone is not sufficient to establish if whether or not this
relaxation has occurred. Carrying out CPT tests next to every production is not feasible either.

Since the cause of the lower bearing capacity could not be identified, the results of the SLT with the
lower shaft friction could not be excluded and had to be taken into account in the design.

78
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 3 summarizes the results of both tests and compares them to the design calculation according to the
Belgian guideline. In the Q b /R b ratio, the installation factor was not taken into account. In the Q t /R t ratio,
all installation factors according to the Belgian guideline were taken into account. Finally, the ratio of
creep to ultimate load was also presented. This appeared also to be very different for both tests. The
consequences of this higher ratio were already explained in figure 3.

Table 3:summary of the results of two SLT’s on displacement screw pile – according to Belgian guideline
Q b /R b Q s /R s Q t /R t Q l /Q c
displacement screw pile 0.49 0.51 0.62 1.61
displacement screw pile 0.51 0.96 0.83 1.34

Both SLT’s showed bearing capacities considerably lower than those that should be theoretically
obtained. The results led to a change in design with longer piles.

6. CASE 3: HOLLOW STEM CFA PILES


A third case presents the results of 2 SLT’s on hollow stem CFA piles with a diameter of 600 mm. the
piles are bored with a hollow stem CFA-auger with a 324 mm diameter hollow stem and 600 mm
diameter outer flanges. The test piles are situated on a job site at very large distance from each other
(several km) and in different ground conditions. Test pile 1 is a friction/floating pile in very
heterogeneous loose to medium silt and sand layers, test pile 2 is an end bearing pile situated in dense
sand layers.

Figures 18 and 19 show the comparison between the CPT’s prior and after pile installation. Both show a
decrease in cone resistance in the sandy layers. Independent of the original cone resistance, the cone
resistance after pile installation seems limited to a maximum of 4 to 6 MPa. The decrease in cone
resistance is limited to a short distance beneath the pile tip.

Figure 18: CPT in axis of test pile before installation together with CPT at 30cm from pile after
installation – friction pile

79
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 19: CPT in axis of test pile before installation together with CPT at 30cm from pile after
installation – end bearing pile

This soil relaxation is probably caused by overflighting, as illustrated in figure 15. Due to the presence of
denser sand layers, this cannot be avoided. During installation of the pile, only limited soil was
transported to the surface and nothing seemed out of the ordinary and no difference could be noted on the
registration or by the present site engineer.
The piles were tested 16 and 15 days after installation. Figure 20 shows the ratio of measured to
calculated unit bearing resistance (Q b,meas /R bu,calc ) versus the ratio of pile base settlement with pile
diameter (s b /D b ).

Figure 20: normalised results of measured/calculated unit bearing resistance (De Beer method)

Since the Belgian guideline assumes an installation factor of 0,7, the curves should converge to a value of
0,7 at 10%D pile base settlement. It can be noticed that full bearing capacity is not yet reached at a pile

80
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

base displacement of 10%D and continuous to rise till 15 and even 20%D displacement and more. Even
at these higher levels of pile base settlement, the installation factor of 0,7 is not reached.

When comparing the results with that of the driven piles (figure 8) and displacement screw piles, it can be
seen that the CFA’s pile base behaves even less stiff.

Comparisons of the ratio between measured and calculated values according to different methods are
summarized in table 4. In the NEN method, a reduction of 30% of the Koppejan value was applied, as
well as the other specific guidelines and factors for CFA piles. For the NF method, also the relevant
factors were applied.

Table 4 : comparison of measured unit bearing resistance with calculated unit bearing resistance
according to different standards (q bmeas /q b,calc )
De Beer NF P 94 270 ICP NEN post 2016
CFA with hollow stem 0.40 0.96 0.67 0.95
CFA with hollow stem 0.31 0.82 0.64 0.80

All of the methods seem too optimistic. For the ICP method, it must be noted that no guidelines are given
for CFA-piles. Compared to the results on driven and displacement screw piles, the spread in the results is
larger for all of the considered methods.

Figure 21 shows the normalised measured to calculated shaft bearing resistance. Although shaft
dimensions are identical, the results are very different for both piles.

Figure 21: normalised results of measured/calculated shaft bearing resistance (Belgian guideline)

The Belgian guideline proposes and installation factor of 0.7 for CFA piles with hollow stem, so the
measurements should converge to a value of 0.7 at 10%D.

The pile installed through the heterogeneous layers presented in figure 18 shows higher than expected
shaft capacity, while the shaft friction in the denser sand layer (below a level of +5 mTAW) in figure 19
is lower than expected. Also, it is remarkable that shaft capacities are reasonably high given the reduction
in measured cone resistance after pile installation.

Table 5 summarizes the results of both tests and compares them to the calculation according to the
Belgian guideline. In the Q b /R b ratio, the installation factor was not taken into account. In the Q t /R t ratio,
all proper installation factors according to the Belgian guideline were taken into account. Finally, the ratio
of creep to ultimate load was also presented. This appeared also to be very different for both tests.

81
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 5: summary of the results of two SLT’s on hollow stem CFA piles – according to Belgian guideline
Q b /R b Q s /R s Q t /R t Q l /Q c
CFA with hollow stem 0.40 1.36 1.05 1.51
CFA with hollow stem 0.31 0.70 0.63 1.15

The SLT for the friction pile meets the expected calculated value. This is caused by the higher than
anticipated shaft friction which compensates for the lack in end bearing capacity.

The SLT for the end bearing pile showed dramatically lower bearing capacity. As a result of this test, all
end bearing hollow stem CFA piles were redesigned as driven cast-in-place piles, the results of which
were already presented as case 1. The original design tried to limit nuisance related to vibration and noise
for constructions that where closer to domestic housing. Additional precautions had to be taken to take
into account these nuisances (noise barriers, vibration measurements, …).

The CFA-friction piles needed additional settlement calculations to verify the design. Since the piles
behave even more as a friction pile than originally anticipated, the calculated settlements will be larger
than in the design.

Figure 22 shows the resulting load distribution for a pile group calculated suing the cap-soil model in
DPile calibrated with measured load displacements curves. The load on the centre pile is about 50% of
that of the corner pile.

Figure 22: design of CFA pile group under vertical load

The settlement of the pile group depends on the interaction between the piles. Since the pile centre-to-
centre distance is around 3D, this is especially true for these floating/friction piles. Since the load
displacements curves are based on measurement, this interaction is the greatest uncertainty. As mentioned
earlier, the interaction is determined on the E-modulus and Poisson coefficient of the soil between the
piles. Figure 23 shows the calculated load settlement curve for the pile group using 2 different E moduli
for the soil between the piles.

82
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 23: pile group settlement at different load stages, based on measured tip and t-z curves, with
different E-modulus for soil between piles

For a ratio in E of 3, settlements vary by factor of 2. The lower bound value was estimated as 3 à 4 q c
with q c limited to 6 à 8 MPa due to pile installation (figure 18). However, as indicated by Poulos, 2006,
this stiffness varies with load level and can estimated as a small strain modulus. This is represented by the
higher E-modulus of 72 MPa = 3 times the original E-modulus based on the reduced cone resistances.
Finally, it can also be noted that the estimate based on the reduced cone resistance from the CPT after pile
installation is conservative, since the influence of the pile installation should limited to the direct vicinity
of the pile.

7. CONCLUSION
The great differences in pile bearing capacity calculated with different standards and methods indicate
clearly the uncertainty encountered in estimating this capacity. This capacity is also defined by the
maximum settlement at ultimate pile load, which differs in each standard. Comparing the current
Eurocode calculation method with partial safety factor to the older calculation method with deterministic
safety factor usually shows a lower ratio between pile bearing capacity and allowable pile load. This can
only be accepted if also pile group settlement calculations are integrated in the foundation design. It can
be anticipated that in the future, similar to the design of direct foundations, attention will shift more from
the calculation of bearing capacity to settlement calculations. But in order to carry out more accurate
settlement predictions, one should be able to accurately predict pile bearing capacity.

In order to obtain a correct estimation on the pile bearing capacity and pile settlement, instrumented static
load tests should be carried out. When such tests are carried out, they can be used to change the design
according to the results. When higher pile loads are permitted in the design, essentially by allowing a
smaller total safety factor on bearing capacity, pile group settlements should also be checked.

Three different cases were presented where pile load tests were performed and the design was altered
based on the results. The measured pile base capacity showed little variation between different tests at the
same jobsite and for the studied pile types. Measured values are considerably lower than those in the
current Belgian design guideline, but are more consistent with other methods (ICP) or recent updates to
standards (NEN). Alternatively, the ultimate pile base capacity should be measured at higher settlement
ratios, e.g. 20%D instead of 10%D. However, this last solution or “quick fix” will shift emphasis even
more to settlement calculations.

83
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

When comparing the results of shaft bearing capacity, a greater variation was found between tests at the
same site. In some cases, a possible cause was identified and could sometimes be resolved, such as the
large base plate for driven piles. In other cases, this variation had to be taken into account in the design
since it seemed it could not be avoided.

When designing piled foundations, the allowable settlements of the structure should be considered and
reflected in the ratio of allowable to ultimate bearing capacity. When pile bearing capacity is not
accurately predicted, this will lead in a first instance to higher than expected settlements and not
catastrophic failure. Structures with a considerable amount of permanent and less variable load should be
designed with care, especially when they are also sensitive to settlements.

REFERENCES
AFNOR. 1999. NF P94-150-1. Sols : Reconnaissance et essais - Essai statique de pieu isolé sous un effort
axial Partie 1 : En compression

ASIRI national project, Recommendations for the design, construction and control of rigid inclusion
ground improvements, IREX

De Beer, E. 1971-1972. Méthodes de déduction de la capacité portante d’un pieu à partir des résultats
des essais de pénétration. Annales des Travaux Publics de Belgique, No 4 (p 191-268), 5 (p321-353) & 6
(p 351-405), Brussels.

De Cock, F. 2008. Sense and sensitivity of pile load-deformation behaviour. Proceedings of the 5th
international conference on Bored an Auger Piles – BAP V, Ghent

Fellenius, B.H., 2001. From strain measurements to load in an instrumented pile. Geotechnical news
magazine, Vol. 19, No. 1, pp 35-38

Fleming K., 2001. Piling Engineering. Taylor & Francis, London

Huybrechts, N. 2016, design of Piles – Belgian Practice, ISSMGE - ETC 3 International Symposium on
Design of Piles in Europe. Leuven, Belgium

ISSMFE Subcommittee on Field and Laboratory testing. 1985. Axial Pile loading test – Part 1: Static
Loading. ASTM geotechnical Testing Journal, pp. 79-90.

Jardine, R, Chow F., Overy R., Standing J., 2005, ICP design methods for driven piles in sands and clays,
Thomas telford, London

Mylonakis, G. & Gazetas G., 1998. Settlement and additional internal forces of gropued piles in layered
soil. Géotechnique Vol 48. No.1, pp. 55-72

Plantema G, 1948, Results of a special loading tests on a reinforced concrete pile, a so-called pile
sounding, interpretation of the results of deep soundings, permissible pile loads an extended settlement
obervations. Second International Conference on Soil Mechanics and Foundation Engineering,
Rotterdam

Poulos H., 2006, Pile Group Settlement estimation – research to practice, Geotechnical Special
Publication, ASCE
Randolp, MF & Wroth C.P., 1979. An analysis of the vertical deformation of pile groups, Géotechnique,
29(4): 423-439

Reese, L.C. & Van Impe, W.F. 2001. Single piles and pile groups under lateral loading. Rotterdam:
Balkema.

Tomlinson, M. & Woodward, J. 2008. Pile design and construction practice. Taylor & Francis, London

Van Tol A.F., 2010, Draagvermogen van geheide palen in internationale context, Vakblad geotechniek,
Funderingsdag 2010 Special

84
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Van Tol A.F., 2012, Draagkracht funderingspalen, een up-date, Vakblad geotechniek, Funderingsdag
2012 Special

Verstraelen, J. & Maekelberg, W. 2015. Ontwerpen op basis van resultaten van laterale
paalbelastingsproeven. Geotechiek, jaargang 19, nr. 5, december 2015 - Speciale uitgave n.a.l.v. de
Geotechniekdag ‘Proef op de Som”, 03.11.2015, Breda.

WTCB-CSTC, 2009. WTCB Rapport nr. 12. Richtlijnen voor de toepassing van de Eurocode 7 in België –
Deel 1 : het grondmechanische ontwerp in de uiterste grenstoestand van axiaal op druk belaste
funderingspalen/ CSTC Rapport n° 12. Directives pour l’application de l’Eurocode 7 en Belgique –
Partie 1 : Dimensionnement Géotechnique à l’état limite ultime de pieux sous charge axiale de
compression. WTCB-CSTC, Brussels.

ACKNOWLEDGEMENTS

The authors wish to thank the ABEF – Belgian community of piling contractors and BBRI – research
institute for their contribution for this paper.

85
86
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Brussels 28 & 29 April 2016

General Report – Design methods based on static pile load tests


S.Burlon, University Paris-Est, IFSTTAR, GERS Department, France, sebastien.burlon@ifsttar.fr

ABSTRACT
This paper presents the ETC3 general report dealing with the design methods based on static pile load
tests. It aims to summarise the practices used in Europe for the interpretation and the analysis of static
pile load tests. National reports considered in this paper has been provided by fifteen countries: Austria,
Belgium, Denmark, Estonia, Finland, France, Hungary, Ireland, Italy, Macedonia, Netherlands, Poland,
Russia, Spain and Sweden. Three main issues are addressed in this paper: the materials and the
equipments used for performing a static pile load test, the analysis of the results and the procedures
available in the Eurocode 7 to design a pile considering these results.

1. INTRODUCTION
Eurocode 7 allows the design of piles from the analysis of measurements obtained during static pile load
tests. A specific relationship including correlation factors is proposed for the calculation of the
characteristic value of the bearing capacity or the tensile resistance from measured values. Moreover,
Eurocode 7 mentions that “methods for assessing the compressive resistance of a pile foundation from
ground test results shall have been established from pile load tests and from comparable experience”.
In this framework, this paper presents the ETC3 general report dealing with the design methods based on
static pile load tests. It aims to summarise the practices used in Europe for the interpretation and the
analysis of static pile load tests. National reports considered in this paper are the following: Austria,
Belgium, Denmark, Estonia, Finland, France, Hungary, Ireland, Italy, Macedonia, Netherlands, Poland,
Russia, Spain and Sweden.
Very limited information is available in each national report concerning the use of static pile load tests.
This situation emphasizes that static piles load tests are not currently used in daily practice. Reading the
national reports, it appears that the main difficulty is related to the differences that are obviously observed
between the measured values obtained during a pile load test and the calculated values obtained by a
calculation model. It seems that the scatter of the calculation models is not really known and consequently
it is really difficult to define the range of measured values that could be expected. Regarding this issue, no
clear guidance is given in Eurocode 7 for the pile design based on static load tests.
Despite this context, in order to have a global overview about the use of static load tests for pile design
and to better understand how their use could be improved, three main issues are addressed in this paper:
the materials and the equipments used for performing a static load test, the analysis of the results and the
procedures available in Eurocode 7 to design a pile considering these results.

2. MATERIALS AND EQUIPMENT


Very few information is available about the materials and the equipment used for a static pile load test. It
may be interesting to underline that static pile load tests can be performed with a jack or an Osterberg
cell. The jack is directly installed on the pile head and the load is applied with a load beam, two transfer
load beams and reaction piles (Figure 1). The Osterberg cell is installed within the foundation unit. As the
load is applied to the Osterberg cell, it allows to working in two directions; upward against upper side
shear and downward against base resistance and lower side shear. Among these two techniques, it seems
that static pile load tests with loading jack are the most used in Europe.
Measurements of the strains along the pile can be done. Many materials are now available (Figure 2):
removable extensometers with strain gauges or optic fibers, strain gauges installed on steel
reinforcements, vibrating string strain gauges, continuous optic fibers, etc. These materials aim at
measuring the variation of the normal force into the pile in order to analyse how the shear resistance is
mobilized at the soil-pile interface.

3. LOADING PROCEDURES
Limited information is presented concerning loading procedures. The future drafts of European standards
EN 22477-1&2 should provide more details. The maintained load procedure appears to be the most used
across Europe even if it seems that the displacement rate procedure can be used. Regarding the
maintained load procedure, the main differences concern the time interval of each loading step and the

87
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Brussels 28 & 29 April 2016

criterion for increasing the load. In the current draft of EN 22477-1, a time interval of at least 60 minutes
is recommended for the loading steps and the displacement rate of the pile has to be less than 0.1 mm/20
min. Other values are considered in other European countries: for example, 0.025mm/5min with a
minimum time of 30 minutes in Hungary, 1 mm between 10 and 100 minutes in Switzerland, etc. The
bearing capacity is obtained for a loss of stability or a displacement of the pile head corresponding to 10
% of its diameter. Other values can be considered: 5 % of the pile diameter if the diameter is greater than
0.8 m in Italy, a decreasing or constant deformation speed, not exceeding 20 mm per hour in Denmark.
Other considerations related to loading procedures are also mentioned:

− for large diameter piles, it is allowed to perform static load test on a “model” pile having the same
length but a smaller diameter (d model ≥ 50%⋅d) provided that the model pile is fully instrumented in
order to separately obtain the mobilization curves for the shaft and the base resistances;

− the waiting period between the execution of the pile and the load test is not really precised. For bored
piles, a delay comprised between 2 and 4 weeks corresponding to the curing time is considered but for
driven piles, the question is more complex. Dissipation of pore pressures and ageing effects are very
complex to take into account.

Figure 1: Jack, load cell, beam (until 10000 kN)

(a) (b)

(d)

(c)

Figure 2: Strain measurement


(a) removable extensometers – (b) optic fibers – (c) strain gauges on steel reinforcements – (d) vibrating
string strain gauges

88
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Brussels 28 & 29 April 2016

4. ANALYIS OF A STATIC LOAD TEST


For a static pile load test with measurements of axial strains, the main results concern: the load-settlement
curve (Figures 3a and 3b), the creep curves for each loading level and the creep rate curve (with many
definitions depending on the choice of the times t 1 and t 2 ) (Figure 4), the load distribution along the pile
(Figure 5) and the mobilization of shear strength at the soil-pile interface at several depths (Figure 6).

Load (kN)
Charge [kN]
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000 5500
0

8
Settlement (mm)

16 Loading #11
Chargement

24
Tassement [mm]

Loading #22
Chargement
32
Extrapolation
Extrapolation
40 (Chin)
Chin et Poulos

48

56

64
(a)

Load (kN)
0 1000 2000 3000 4000 5000 6000 7000 8000 9000
0

20
Settlement (mm)

40

60 Loading
Série1 #1

80

100

120

(b)
Figure 3: Load-settlement curve
(a) screw pile – D=16.59 m, B=0.62 m – (b) CFA piles – D=20.39 m, B=0.62 m

40
t2, s 2 10

35 t1, s 1 5000 kN 9
k
30 8 Yield resistance
(creep resistance) R Y
Creep rate (mm/h)
Vitesse de fluage [mm/h]
Enfoncement(mm)

7
[mm]

25 4500 kN
6
Settlement

20 5
4000 kN 4
15
3
10 3500 kN
2

5 1

0
0
0 1000 2000 3000 4000 5000 6000
1 10 100
Charge [kN]
Minutes Load (kN)
Minutes

(a) (b)
Figure 4: Creep curves and creep rate curve (screw pile D=16.59 m, B=0.62 m)

89
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Brussels 28 & 29 April 2016

Load (kN)
0 1000 2000 3000 4000 5000 6000
0

6
Depth (m)
8

10

12

14

16

18
Rb Rs
(a)
Load (kN)
0 1000 2000 3000 4000 5000 6000 7000 8000
0

8
Depth (m)

10

12

14

16

18

20

22
Rb Rs
(b)
Figure 5: Load distribution
(a) screw pile – D=16.59 m, B=0.62 m – (b) CFA piles – D=20.39 m, B=0.62 m

450

400
Level A – z=15.81 m

350 Level B – z=14.81 m


(kPa)

300 Level C – z=13.06 m


[kPa]

Level D – z=10.56 m
friction

250
unitaire

Level E – z=8.31 m
axialshaft

200
Level F – z=6.31 m
Frottement Unit

150 Level G – z=4.31 m

100 Level H – z=2.31 m

Level I – z=0.81 m
50

0
-8 0 8 16 24 32 40
Déplacement [mm]
Vertical displacement (mm)
(a)

90
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Brussels 28 & 29 April 2016

350

Level A – z=18.65 m
300
Level B – z=17.65 m

250 Level C – z=16.65 m

unitaire [kPa]
(kPa) Level D – z=15.08 m
200
Level E – z=12.05 m
friction

Level F – z=9.05 m
Unit shaftaxial

150
Level G – z=6.05 m
Frottement

100 Level H – z=3.75 m

Level I – z=2.0 m
50
Level J – z=1.0 m

0
-6 0 6 12 18 24 30 36 42 48 54 60
Déplacement [mm]
Vertical displacement (mm)
(b)
Figure 6: Unit friction mobilisation curves

(a) screw pile – D=16.59 m, B=0.62 m – (b) CFA piles – D=20.39 m, B=0.62 m

In Figure 3a, the procedure proposed by Chin (1970, 1972) is used to extrapolate the load-settlement
curve and assess the pile bearing capacity for a displacement criterion equal to B/10. In figure 3b, the
bearing capacity can be directly assessed since the displacement criterion of B/10 is reached.
The creep curves (Figure 4) can be used to assess the yield resistance (creep resistance), i.e., the
resistance of the pile where the pile head displacement does not increase according to the time. On the
load-settlement curve, the yield resistance corresponds more or less to the end of the linear relation
between the load and the displacement.
Regarding this creep curves, it is still surprising that the yield resistance as defined on Figure 4 is not
really considered in calculation and design standards. Figure 7 shows an example of a pile located in
chalk and submitted to various loading procedures. The bearing capacity is highly variable and depends
on the pile age and the loading procedure whereas the yield resistance seems to be more constant
according to the time (the beginning of each load-settlement curve until 600 kN is quite similar).
The creep curves can be used for the assessment of the bearing capacity in terms of displacements or in
terms of displacement rate. In many cases, the failure of the pile could be defined for a displacement
lower than B/10 since the pile displacement increases with the time due to creep effects. In these cases,
the accumulation of displacements would lead to too large disorders for the supported structure. This type
of analysis could be developed to introduce ultimate limit state verifications in terms of displacements
knowing that the bearing capacity can be very uncertain in some conditions. In the same way, it should be
interesting to develop more detailed procedures to assess a value of pile resistance for serviceability limit
states based on displacement rate criteria. For example, the creep rate k calculated in a semi-logarithmic
scale permits the extrapolation of pile displacements for several periods according to this relation
(assuming that the creep rate remains constant):
s2 − s1
k=
log(t 2 t1 )
The last issue presented in Figures 5 and 6 is related to the variation of the normal force along the pile and
the mobilization of unit shaft frictions. These results permit to identify the base resistance R b and the
shaft resistance R s . It is interesting to note the values of the unit shaft frictions are not reached for the
same pile head displacement due to the effects of the axial pile stiffness (see, Randolph, 2003). At the pile
top, as the displacements increase faster, the soil strength is mobilised faster and decreases due to too
large shear displacements. It is important to check if the calculation methods used for a design take into
account this aspect especially if the axial stiffness of the pile is very low.

91
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Brussels 28 & 29 April 2016

Load (kN)

Pre-cast driven pile 0.4x0.4 m


Chalk
Step 1 60’ (16-04-76)
Step 2 15’ (01-04-76)
Step 3 120’ (14-04-76)
Step 4 60’ (28-04-76)
Step 5 720’ (17/20-05-76)
Settlement (mm)

Step 6 30’ (11-1-77)

Figure 7: Pile submitted to several loading procedures

5. DESIGN OF PILES WITH STATIC PILE LOAD TEST

5.1. Basic relationships


The design of piles with static pile load tests can be performed according to the relationships proposed by
Eurocode 7-1. These equations seem to be very rarely used. The characteristic value is obtained as
follows:

R R 
Rc / t ;k = min  c / t ;moy ; c / t ;min 
 ξ1 ξ2 
where R c/t;k is the characteristic value of the bearing capacity or the tensile resistance and ξ 1 and ξ 2 are
correlation factors that should account for:

− the number of tests: the uncertainty related to the bearing capacity variation decreases when more tests
become available ;
− the scatter of the measured values of bearing capacity: when a large variability is observed, the lowest
value of the measured values should govern the foundation design; when the variability is small (small
variation coefficient), a value close to the mean value should govern the design.

Regarding this last aspect, it is interesting to assess the variability that correlation factors can take into
account. The following example present an ideal case including three sets of three piles (Table 2). The
characteristic value of bearing capacity is calculated following three methods: correlation factors, normal
distribution with k N assessed by Student distribution, log-normal distribution with k N assessed by Student
distribution with a 5% risk (see Annex D of EN1990, CEN, 2002). It can be concluded that correlations
factor values are calibrated for a low variability with a coefficient of variation (COV) close to 5 %. These
results show that the procedure proposed by Eurocode 7 for correlation factors could be improved with a
direct statistical analysis of the data obtained from the pile load tests. Another way could be to better
explain the role of these correlation factors.
Eurocode 7 allows to divide these correlation factors by 1.1 when the stiffness of the structure is
sufficient to transfer loads from weak to strong piles.
Other modifications are proposed in some countries:

− in Finland, in tension values given in table 6 are multiplied by another factor equal to 1.25 ;
− in France, in order to take into account the distance between the pile load test and the piles to be
executed, alternative correlation factors ξ 1 ' and ξ 2 ' have been introduced:

92
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Brussels 28 & 29 April 2016

ξ ' i (N , S ) = 1 + [ξ i (N ) − 1]
S
with 625 m2 ≤ S ≤ 2500 m2
S réf
For values S greater than 2500 m2, the following relationship may be considered if the site is
homogeneous:

ξ ' i (N , S ) = ξ i (N )

For one pile load test at the location of the pile to design, this equation leads to the following values: ξ 1 ’ =
ξ 2 ’ = 1.2 ;
− In United Kingdom, values of correlation factors are directly modified.

Table 1: Values of the correlation factors ξ 1 et ξ 2


ξ pour N=
1 2 3 4 ≥5
N number of pile load tests
ξ1 1.40 1.30 1.20 1.10 1.00
ξ2 1.40 1.20 1.05 1.00 1.00

Table 2: Variability taken into account by correlation factors


Set 1 Set 2 Set 3
Load test #1 3000 kN 3000 kN 3000 kN
Load test #2 2850 kN 2700 kN 2550 kN
Load test #3 3150 kN 3300 kN 3450 kN
Arithmetic mean ( X moy ) 3000 kN 3000 kN 3000 kN

Geometric mean ( X geom ) 2997 kN 2990 kN 2977 kN

Standard deviation ( σ ) 150 kN 300 kN 450 kN

Standard deviation ( σ ln ) 0.05 0.1 0.15

COV ( σ X moy ) 0.05 0.1 0.15


ξ factors 2500 kN (mean) 2500 (mean) 2429 kN (min)
Normal distribution
(5 % risk)
2494 kN 1988 kN 1483 kN
X k = X moy − k N σ
Log-normal distribution
(5 % risk)
2532 kN 2131 kN 1788 kN
X k = X geom e − k N σ ln

5.2. How to improve the use of static pile load tests?

5.2.1. Context
Pile design based on load tests are seldom in many European countries, and in practice, it is rare to do
more than one preliminary pile test on a site. Many reasons can explain this situation. The organization
and the cost of a static pile load test can be a first argument. Regarding only technical aspects, it seems
that the results obtained from a pile load test cannot be considered as isolated values and are
systematically compared to calculated values. The pile load test is used to confirm the design using
ground strength parameters. For example, in some countries such as Belgium (BBRI, 2009) or United
Kingdom, this procedure allows to considering a reduced value for the model factor in the calculation
model considered for the design. Only in exceptional cases the results of preliminary static pile load tests
are used to locally calibrate the semi-empirical design method and/or to adapt the original pile design.
Another difficulty is due to the absence of guidance on how to compare piles of different diameter or
different length. Furthermore, when a static pile load test is performed, it is expected to have some
differences between the measures and the calculations but it seems very difficult to quantify them. The
main problem seems to be linked to the scatter of the calculation model.

93
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Brussels 28 & 29 April 2016

5.2.2. Some ideas to use static pile load tests


The analysis of the results obtained from a pile load test lead to consider three types of parameters: the
bearing capacity R c , the contribution of the base resistance R b and the shaft resistance R s and the limit
shaft frictions q s .
These quantities cannot be measured with the same precision. The bearing capacity is easy to measure
since it only requires a load cell on the pile head and displacement sensors. As mentioned earlier, the
definition of the bearing capacity is not unique but the rule of B/10 seems to be accepted by a majority.
The base resistance, the shaft resistance and the limit shaft friction require the measure of the strains
along the pile. Several parameters may be a source of uncertainties during the interpretation of the
measures especially for bored piles: the geometry of the pile, the Young modulus of the pile, etc. These
parameters increase the scatter of the calculation of the normal force distribution into the pile, the
mobilized shaft frictions and finally the unit shaft frictions. The comparison between the measurements
and the calculations has to take into account these factors. Moreover, it is important to underline that
values of unit shaft frictions provided by calculation models do not systematically take into account the
effects due to the axial stiffness of the pile, which can increase the difference between measured and
calculated values. As noted earlier, during a static pile load test, the maximum values of unit shaft
frictions along the pile cannot be obtained at the same moment. Other aspects should be considered for
displacement piles like the dissipation of pore pressures or the ageing effects.
Pile load test databases can give very interesting information about the scatter of each specific calculation
model. They allow to take into account the systematic and random components of the variation in the
ground for the interpretation of static load tests. These pile databases can provide some information about
how a static pile load test can be representative.
Information concerning the assessment of resistance predictions from pile database is available for several
kinds of methods: for example, the ICP method (Jardine et al., 2005) (Table 3), the French pressuremeter
method PMT2012 (Burlon et al., 2014) (Table 4) or other methods (see Paikowski et al., 2004 for α-C u
methopd, β-σ' v method, etc.). Indeed, it is important to know if the calculation model is supposed to
provide values that are close to 1.0 on average. For the two methods presented in Tables 3 and 4, the
calibration has been made to ensure that the ratio of the calculated to the measured value is close to 1.0 on
average. For these two methods, the measured value of bearing capacity or tensile resistance should be
close to the calculated value but the results presented in Tables 3 and 4 show that a relative difference of
25 % between the measured and the calculated values can be expected.

Table 3: Assessment of resistance predictions for ICP method (all piles) (from Jardine et al., 2005)

Ratios of the calculated to the measured values

Arithmetic mean Standard deviation Coefficient of variation


Shaft resistance (sand) 0.99 0.28 0.28

Shaft resistance (clay) 1.03 0.21 0.2

Base resistance (sand) 1.01 0.19 0.19

Base resistance (clay) 0.85 0.26 0.30

Table 4: Assessment of resistance predictions for PMT2012 method (174 piles)

Ratios of the calculated to the measured values

Arithmetic Standard Geometric Coefficient of


mean deviation mean variation
Bearing capacity 0.96 0.24 0.92 0.25

Shaft resistance 0.95 0.35 0.88 0.37

Base resistance 1.04 0.48 0.97 0.46

Unit shaft friction 1.66 2.79 1.07 1.68

Figure 8 and Table 4 show other results for the French pressuremeter calculation: the scatter of the ratio
of the calculated values to the measured values for the bearing capacity R c , the contribution of the base
resistance R b and the shaft resistance R s and the unit shaft frictions q s (Burlon et al., 2014).These results
show that the uncertainty of the calculation models increases when the considered quantity concerns a

94
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Brussels 28 & 29 April 2016

low volume of ground. These results mean that the calculation model can provide “wrong” assessments at
a local scale, for example, for unit shaft frictions but “good” assessments at a large scale, for example, for
the shaft resistance and the bearing capacity. Errors on mean value (shaft resistance and bearing capacity)
are lower than errors on single values (base resistance and unit shaft frictions). This comment is very
important for the analysis of pile load tests in order to optimize a design during a project.

Figure 8: Assessment of resistance predictions (174 piles – French pile database PMT2012 method)

6. CONCLUSION
Static pile load tests can be considered as the base of the design of deep foundations since they are the
reference for the calibration of calculation models. Nevertheless, in Europe, their use in the daily practice
remains rare due probably to high costs and the need of a complex organization. The materials required
for the measure of strains can be sometimes complex to be installed.
Nevertheless, it seems that some technical issues restrict the use of static load tests. The approach
proposed by Eurocode 7-Part 1 and based on correlations factors could be improved. The links between
the results obtained during a load test and the reliability of the calculation models should be detailed.
During a project, it is very difficult to compare the uncertainty of the measures and the scatter of the
calculation models. No precise guidance exists for the analysis of pile load test results within the
framework of a project. Some ideas have been proposed in this paper but it would be interesting to
develop some aspects for including in the next Eurocode 7 more details about this issue.
Anyhow, European standards for pile design and pile execution should more underline the need of static
pile load tests in order to improve calculation models and increase reliability and robustness of piles. Each
year, new pile techniques are proposed by deep foundations contractors and design standards should
include these advancements. Pile database across Europe should be shared to compare calculation models
and enhance harmonised practices.

ACKNOWLEDGEMENTS
This paper is based on the National reports from Austria, Belgium, Denmark, Estonia, Finland, France,
Hungary, Ireland, Italy, Macedonia, Netherlands, Poland, Russia, Spain and Sweden. The authors of these
papers and the organisation committee of the International Symposium Design of Piles in Europe – How
did Eurocode 7 change daily practice? are gratefully thanked.

95
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Brussels 28 & 29 April 2016

REFERENCES

BBRI – Belgian Building Research Institute (2009) Directives pour l’application de l’Eurocode 7 en
Belgique – Partie 1 : dimensionnement géotechnique à l’état limite ultime de pieux sous charge axiale de
compression. Rapport n°12.

Burlon, S., Frank, R., Baguelin, F., Habert, J. and Legrand. S. (2014) Model factor for the bearing
capacity of piles from pressuremeter test results A Eurocode 7 approach. Géotechnique, 64(7), 513-525.

Chin, F.K. (1970) Estimation of the ultimate load of piles from tests not carried to failure. Proceedings of
the Second Southeast Asian Conference On Soil Engineering, pp83-91.

Chin, F.K. (1972) The inverse slope as a prediction of ultimate bearing capacity of pile. Proceedings 3rd
Southeast Asian Conference on Soil Engineering, pp83-91.

CEN (2002) EN 1990 Eurocode: Basis of structural design. European Committee for Standardization
(CEN), Brussels.

CEN (2004) Eurocode 7 – Part 1: Geotechnical design – Part 1: General rules, EN 1997-1. European
Committee for Standardization (CEN), Brussels.

CEN (2016) Geotechnical investigation and testing — Testing of geotechnical structures — Part 1: Pile
load test by static axial compression EN22471-1. European Committee for Standardization (CEN),
Brussels. (to be published, first draft)

Jardine R., Chow F., Overy R., Standing J. (2005). ICP design methods for driven piles in sands and
clays. Imperial College, London.

Paikowski, S., Birgisson, B., McVay, M., Nguyen, T., Baecher, G., Ayyub, B., Stenersen, K., O’Malley, K.,
Chernauskas, L. & O’Neill, M. (2004). LRFD design and construction for deep foundation, 1st edn.
Washington, DC, USA: Transportation Research Board.

Randolph, M.F. (2003) 43rd Rankine Lecture: Science and empiricism in pile foundation design.
Géotechnique, 53(10) 847-875.

96
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Design methods based upon rapid pile load tests


Michael Brown, University of Dundee, UK, m.j.z.brown@dundee.ac.uk

ABSTRACT
This report considers the development of guidance on rapid load pile testing within the European context
that has led to the development of FprEN ISO 22477-10, Geotechnical investigation and testing —
Testing of geotechnical structures — Part 10: Testing of piles: rapid load testing. This report summarises
current European national guidance on rapid load pile testing and goes on to outline an informative test
interpretation method included in FprEN ISO 22477-10 that goes someway to developing design methods
under EN 1997-1. The report finishes with considerations for EN 1997-1 evolution and future ETC3
reporting.

1. INTRODUCTION
Guidance on the design of piles based upon the use of rapid load pile testing (RLT) is generally limited
outside of Europe and currently lacks codification within CEN or ISO standards. The majority of existing
published guidance is limited to general information or execution and testing specification. Currently
guidance is lacking on how to interpret measured data captured during rapid testing so that it can be
considered a characterisitic measured resistance in EN 1997-1. Published national guidance in both
Holland and Germany does though offer some guidance on test interpretation and a way forward in
developing design to EN 1997-1 and its evolution. In order to improve European guidance on rapid load
testing CEN TC 341 WG7 have recently developed an execution code, FprEN ISO 22477-10,
Geotechnical investigation and testing — Testing of geotechnical structures — Part 10: Testing of piles:
rapid load testing. The execution code also includes an informative annex outlining a potential method of
test analysis but stops short of developing model and characteristic values for RLT. This report identifies
current existing European design and execution guidance, considers the national reports, outlines the
informative test interpretation guidance given in FprEN ISO 22477-10 and finishes with
recommendations for evolution of EN 1997-1 and future reporting of ETC3.

2. EXISTING GUIDANCE AND EXPERIENCE

2.1. National and international guidance

National guidance in the UK is limited to operational and specification type guidance as outlined in
SPERW (ICE, 2007). No specific mention is made of how to determine ultimate pile resistance from the
test but specific mention is made of how piles subjected to rapid load testing may experience significantly
higher loads than during static pile testing and should be designed and reinforced structurally to take these
additional loads (Brown & Hyde, 2006). SPERW states that rapid load testing can be considered as either
an alternative or complementary method to static load testing providing the results and the interpretation
of the results are fully understood. Rapid load testing may also be used for site specific quality and
consistency work where a maintained load test is available on an identical pile. Guidance on RLT in the
context of other pile testing techniques is also included in the ICE Manual of Geotechnical Engineering
(2012).
The most extensive European guidance exists in Dutch guidance in CUR Guideline 230 (Hoelscher et al,
2011) which outlines potential test interpretation and design approaches and also discusses the ethos and
development of the design and interpretation methodologies. Two interpretation methods are outlined in
detail in this document. The first is a method designed to be applicable in a variety of soil types and is
based upon a simplified version of the Unloading Point Method, UPM (Middendorp, 2000). This method
requires the use of a soil dependent factor which was developed in CUR Guideline 230 based upon a
database of published RLT undertaken in various soils in the USA, Belgium and Holland. CUR Guideline
230 also outlines and discusses the use and interpretation of single or multiple cycle testing for the first
time. The second interpretation method outlined (referred to as the Sheffield Method) is designed
currently for use in clay soils only and has not been extended to other soil types. This method was
initially developed based upon model pile testing and then verified against various clay deposits in the
UK (Brown & Powell, 2013, Brown et al, 2006). Current development allows it to be used in different

97
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

clay types through correlation of velocity dependant rate effects with simple laboratory consistency
measurements (e.g. plastic limit and liquid limit).
In Germany national guidance is covered in EAPfaehle – Recommendations on Piling, (GGS 2013). This
covers both execution and test interpretation for design. The interpretation approach outlined is as per that
outlined in CUR Guideline 230.
International guidance has been in existence for some time with formal codification in ASTM D7383 and
the Japanese Geotechnical Society (JGS). In both of these cases though guidance is limited to execution
rather than interpretation and design.

2.2. European experience and national reports


National reporting gives little mention of rapid load testing apart from mention as a possible verification
approach in Dutch practice. It is also mentioned in the Belgium national report (referred to as Kinematic
testing) as it was undertaken as part of the large screw pile field test study reported in Holeyman et al.
(2001) and Holeyman et al. (2003). This lack of inclusion in the national reports may be due to limited
experience, omission from reporting templates and current omission from EN 1997-1. It is noted later that
the national reporting templates for ETC3 did not include rapid load testing and that RLT now comes
under the heading of “non-static piles tests”. Therefore in future it is suggested that RLT is not grouped
under the heading of dynamic testing but as a separate entity as recognised by CEN/ISO and that
reporting on its use is also identified.

3. DESIGN OF PILES BASED UPON RAPID LOAD PILE TESTING

3.1. Informative guidance on tests interpretation in FprEN ISO 22477-10


As part of the development of FprEN ISO 22477-10 it was identified that as well as giving guidance on
execution of RLT it would also be useful to give informative information on the interpretation of RLT
especially where RLT is not mentioned in the current EN 1997-1. As FprEN ISO 22477-10 is an
execution code it was decided that it could not consider pile design directly but that there was a need to
give informative guidance on how to proceed from load and displacement measured during RLT to a
characteristic measured resistance (R c,m ) that could be used as a basis for design. The other reason that the
annex was made informative only was that it was felt that only one method of interpretation should be
included (due to space considerations) but that as other methods of interpretation do exist that their use
should not be precluded. It was also acknowledged that RLT interpetation is still the subject of research
and development and that codification should not stifle the development of improved analysis techniques
in the future.
The method currently included in the informative annex is based upon the CUR Guideline 230 for the
simplified UPM method for single and multiple cycle analysis. This method was used as a basis for the
informative annex as it was felt that this method was outlined in both Dutch and German national
guidance documents and that it’s development was well documented against a published data base of field
testing that can be verified by an end user.

3.2. Unloading Point Method (UPM) for determination of the compressive pile
behaviour

3.2.1. General
Rapid load test results can be analysed using the Unloading Point Method (UPM) to derive the measured
compressive resistance R c,m and the corresponding load deflection behaviour. The approach can be used
for either a single load cycle or multiple load cycles (undertaken on a single pile) as required. Only the
method for analysis of a single load cycle is outlined here although multiple cycle analysis is covered in
CUR Guideline 230 and FprEN ISO 22477-10.

3.2.2. Analysis of a single load cycle


The process of analysis of a single load cycle is outlined below with Figure 1 to Figure 3 included to
clarify the steps of the analysis process. The process is as follows:

a) Obtain the measured signals from the rapid load test


b) Calculation of the inertia corrected mobilised pile resistance at the unloading point:

98
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

• Obtain the pile properties required to calculate the mass of the pile, m (pile cross sectional area,
pile length, density of the pile material, records of the concrete volume used to form the pile) and
the mass of any other components contributing to the inertial resistance of the pile.
• Obtain the variation of force (F), velocity (v), displacement (w) and acceleration (a) with time (t)
for a single load cycle.
• Determine the time (t w-max ) during the test where the measured pile velocity becomes 0 m/s or a
value close to this. Referred to as the unloading point.
• Determine the magnitude of the force measured at t w-max , F c, tw-max.
• Determine the magnitude of the acceleration at t w-max , a tw-max.
• Solve equation (1) to determine the inertia corrected pile resistance at t w-max.

R c, ic, tw-max = F c, tw-max -(m × a tw-max ) (1)

The parameters in equation (1) are defined as:


a tw-max is the acceleration measured during the test at t w-max
F c, tw-max is the magnitude of the force measured during the test at t w-max
M is the mass of the pile and any other components contributing to inertial resistance of
the pile
R c, ic, tw-max is the inertia corrected pile resistance at t w-max
t w-max the time some time after the commencement and before the end of the load application
phase of the test where the pile velocity effectively becomes 0 m/s or a value close to
this.

c) Correction of the inertia corrected pile resistance for the soil dependant empirical parameter
• Determine the soil dependent empirical parameter based upon Table 1 recommendations and
calculate the corrected derived static pile resistance, R c,corrected as per equation (2).
• Determine the magnitude of the displacement at t w-max , w tw-max .

R c,corrected = η×(R c, ic, tw-max ) (2)

Where:
R c,corrected is the corrected mobilised resistance at w tw-max .
w tw-max magnitude of the pile displacement at t w-max
η soil dependant empirical parameter

Where the mobilised pile load deflection behaviour at t w-max can be shown to correspond to the ultimate
resistance of the pile then R c,corrected may be considered equivalent to the measured compressive pile
resistance R c,m.
The parameters shown in Table 1 should only be adopted where it can be demonstrated that rapid load
testing was undertaken on the same type of pile, of similar length and in comparable soil conditions as
those that were used to determine the parameters as outlined in CUR Guideline 230.

Table 1: Selection of soil dependant empirical parameter η


Soil type Clay Sand
Empirical factor, η 0,66 0,94
Standard deviation 0,32 0,15
Coefficient of variation 0,49 0,15
Number of cases 12 22
Number of sites 6 10
The values shown in Table A.1 are based upon data presented in Table C2 for clay soils and
Table C3 for sand of CUR Guideline 230.
NOTE 1 Only cases with pile head displacements greater than 5% of the equivalent
pile diameter are considered in Table 1.
NOTE 2 The variation of values in Table 1 is significant for clay soils particularly.
An alternative suggestion for a selection method for empirical factor for clay based upon
measured soil properties can be found in Brown & Powell (2013). The majority of the
values presented in Table A.1 are derived from pile tests on driven piles.
NOTE 3 Guidance on analysis in clays and layered soils is also given in Weaver &
Rollins (2010).

99
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure1: Flow chart showing the analysis process for a single cycle of rapid load testing

100
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 2: The analysis process for a single cycle using example data and calculation of R c,m

d) Determination of initial pile stiffness

If an approximation of the complete corrected load-displacement behaviour during the test is required
then further analysis steps are necessary to construct a hyperbolic based approximation of the load
displacement behaviour (Figure 3). Firstly it is necessary to determine the initial pile stiffness at low
displacement levels.

NOTE The hyperbola is described by F = w /(p + q × w); p is the inverse value of the initial pile
stiffness k c and q is the inverse value of the hyperbola asymptote. In a hyperbola diagram with

101
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

coordinates w / F and w, a pure hyperbola will be represented by a straight line. Background information
on this technique can be found in Chin (1972). The required steps are as follows:

• Plot the displacement of the pile (w) divided by the inertia corrected pile resistance (R c, ic ) for the complete test
versus the pile displacement up to w Fc-max where F c-max is the maximum measured force .
• Determine the magnitude of the w/R c, ic axis intersection with a best fit linear line to determine the value for the
hyperbola parameter p which represents the inverse of the initial pile stiffness, k c .

e) Determination of the inertia corrected load displacement curve

The procedure for determination of inertia corrected load displacement curve is as follows:
• Determine the initial stiffness parameter as outlined above.
• Calculate the remaining the hyperbola formula parameter q equation (3)

q = 1,0 / R c, ic, tw-max – p / w tw-max (3)

• Draw the resulting hyperbola load displacement curve using equation (4)

R c, ic = w / (p+(q × w)) (4)

f) Determination of the corrected compressive load displacement curve


The procedure for determination of the corrected compressive load displacement curve is as follows:
• Calculate the hyperbola parameter q corrected as per equation (3) with R c,ic,tw-max replaced by R c,corrected ;
• Draw the resulting corrected compressed load displacement curve using equation (4) with parameters
determined for R c,corrected .

Figure 3: Single cycle load displacement calculation using example data

3.2.3. Correlations for piled foundations


Although FprEN ISO 22477-10 outlines the potential analysis technique above it does not go as far as
giving model (if required) and correlation factors for RLT as this is considered “design” and outside of
the scope of an execution standard. FprEN ISO 22477-10 states in A.3 “For the verification of Structural
(STR) and Geotechnical (GEO) limit states, correlation and model factors should be applied to the R c,m
determined from Rapid Load testing to derive the characteristic resistance of axially loaded piles. These
should be determined based upon appropriate experience of comparison of Rapid Load testing analysis
results with static pile tests for similar pile types and ground conditions. Model factors should be used
that reflect the performance of the particular analysis technique used”. It is planned to develop such

102
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

factors based upon and independent data base of piles tests and publish these as an addendum to FprEN
ISO 22477-10 with a view to these being incorporated in the evolution of EN 1997-1.

4. CONSIDERATIONS FOR EVOLUTION OF EN 1997-1 & ETC3

4.1. Model and charateristic values for RLT and other pile tests
As mentioned in 3.2.3, model (if required) and characteristic values are required for RLT based design to
Eurocode. Therefore clear guidance is required on how these values should be derived for RLT but also
for the re-evaluation of the values for other pile testing techniques (previously published in EN 1997-1).
One shortcoming of the use of model and characteristic values is that they potentially prescribe pile test
analysis techniques and have the potential to stifle the use of newer and improved techniques. This could
be avoided by making the use of such factors less prescriptive and giving clear guidance on how these
values could be determined for example how many pile tests, comparison to what static equivalent test,
how many tests in each different soil type, transparency and publication of data, etc...

4.2. Definition of Rc,m and primary measurements


As RLT (and dynamic load tests, DLT) do not measure R c,m directly, and must first undergo some
interpretation, then definition of at what stage of the process of interpretation R c,m is defined is required.
As EN 1997-1 already includes model and characteristic values for DLT it would be expected that clear
definitions of how R c,m should be obtained would exist. Without such guidance it is difficult to maintain
consistent analysis and design based upon RLT or DLT. Adding this to an evolved Eurocode 7 would add
significant complexity and thus it would seem logical to include this in execution codes for testing
methodologies as outlined above in 3.2.2. As there is a requirement for primary measurements
(potentially several) and then interpretation there is a need for definition of these meaurements and
appropriate symbols within the execution codes.

4.3. The use of piles testing in Europe for verification not design?
Based upon the national reports submitted to ETC3 it would seem that pile load tests are very rarely used
(of any type) for stand alone design but are more commonly used as design verification. Thus new
guidance on the use for verification rather than just design may be required.

4.4. General considerations for Eurocode evolution main text


Although the presence of execution codes has the potential to reduce test specific text within an evolution
of EN 1997-1 there may still be specific requirements for individual test peculiarities to be mentioned.
For RLT these may include:

• Notes to the effect that all tests must be analsysed to some extent to determine R c,m and that primary
data obtained should not be used for design purposes (See 4.2 above).
• Piles should be designed structurally to withstand higher stresses than they may experience during
static loading.
• RLT may not be applied to all pile lengths without first checking that the duration of loading is
appropriate for the pile material and pile length being tested. Where the load duration criteria is not
met it may be necessary to pre-install appropriate embedded pile instrumentation and specialised
analysis will be required.

More generally the use of open format easily readable/transferable data should be encouraged for all pile
testing as this leads to greater transparency, opportunities to develop current practice and meet the
growing requirements for public accessibility of government funded research.

4.5. Future considerations for ETC3


It is noted that the national reporting templates for ETC3 did not include rapid load testing and that RLT
(and DLT) now comes under the heading of “non-static piles tests”. Therefore in future it is suggested
that RLT is not grouped under the heading of dynamic testing but as a separate entity as recognised by
CEN/ISO and that reporting on its use is also identified. Also, it would appear that pile load testing is not
frequently used for design purposes but for verification of design approaches. This may form a more
useful enquiry heading i.e. pile testing for design and verification of design.

103
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

REFERENCES

American Standard Testing Methods. Methods for axial compressive force pulse (rapid) testing of deep
foundations. ASTM D7383 - 08 Standard Test.

Brown, M.J., Hyde, A.F.L. & Anderson, W.F. (2006) Analysis of a rapid load test on an instrumented
bored pile in clay. Geotechnique. Vol. 56, No. 9. pp. 627-638. DOI: 10.1680/geot.2006.56.9.627.

Brown, M.J. & Hyde, A.F.L. (2006) Some observations of Statnamic pile testing. Proc. Inst. of Civil
Engineers: Geotechnical Engineering Journal, Vol 159, GE4. pp. 269-273. DOI:
10.1680/geng.2006.159.4.269.

Brown, M.J. (2012) Pile capacity testing. In: Burland, J., Chapman, T. Skinner, H. & Brown, M.J., ICE
Manual of Geotechnical Engineering. 1st ed. London, ICE Publishing Limited. pp 1451-1469. DOI:
10.1680/moge.57098.1451..

Brown, M.J. & Powell, J.J.M. (2013) Comparison of rapid load test analysis techniques in clay soils.
ASCE Journal of Geotechnical & Geoenvironmental Engineering. Vol 139, No. 1, pp. 152–161.

Chin, F.K. (1972) The inverse slope as a prediction of ultimate bearing capacity of piles. Lumb, P.L.
(eds) Proc. of the 3rd Southeast Asian Conf. on Soil Engineering, Hong Kong.

EAPfaehle – Recommendations on Piling, 2013. 1st Ed. German Geotechnical Society. Wiley

Hölscher, P. Brassinga, H., Brown, M.J. Middendorp, P. & Profittlich, M. & van Tol, F.A (2011) Rapid
load testing on piles: Interpretation guideline. CUR Building and Infrastructure. CUR Guideline 230. The
Netherlands, CRC Press/Balkema.

Holeyman, A., Couvreur, J.-M. & Charue, N . 2001. Results of dynamic and kinetic pile load tests and
outcome of an international prediction event. Screw Piles – Installation and Design in Stiff Clay,

Holeyman (ed.). Proceedings of the 1st symposium on screw piles, Brussels 15 March 2001. Swets &
Zeitlinger, Lisse.

Holeyman, A. & Charue, N. 2003. International pile capacity prediction event at Limelette. Belgian
screw pile technology – design and recent developments, Maertens, J. & Huybrechts, N. (eds).
Proceedings of the 2nd symposium on screw piles, Brussels 7 May 2003. Swets & Zeitlinger, Lisse.

Middendorp, P. (2000) Statnamic the engineering of art. Proc.6th Int. Conf. on the Application of Stress
Wave Theory to Piles, Balkema, Rotterdam, 551-562.

The Institution OF Civil Engineers (2007) ICE Specification for Piling and Embedded Retaining Walls
(SPERW). 2nd ed, Thomas Telford Publishing, London, UK.

Weaver, T.J. & Rollins, K.M. (2010) Reduction factor for the unloading point method at clay soil sites.
ASCE Journal of Geotechnical & Geoenvironmental Engineering. Vol 136, No. 4, pp. 643–646.

104
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Design methods based on dynamic pile load tests


Oswald Klingmueller, GSP mbH, Germany, ok@gsp-mannheim.de

Matthias Schallert, GSP mbH, Germany, ok@gsp-mannheim.de

ABSTRACT
Even as it is over 20 years that dynamic pile load tests have been introduced by EC7.7.5.3 as an
alternative load tests for the verification of pile capacity the tests were carried out in most countries
without any regulation. In the last 5 years EC7 has come into force in many European countries and also
National attachments have been installed. However the practice is still varying. There are countries were
dynamic pile load tests are used as a routine practice (like Sweden) to countries were dynamic tests are
not known (Like Macedonia). An overview over European practice on the basis of National reports is
given and recommendations for the evolution of EC7 is given. The development of a Standard 22477-4
for dynamic load tests is described.

1. INTRODUCTION
Dynamic pile load tests are considered as a tool for pile design in EC7 7.5.3 and 7.6.2.4 to 7.6.2.7. EC7
defines impact tests evaluated by driving formula and especially refers to dynamic pile load tests with
measurements of force and velocity at pile head. The latter methods of dynamic pile load tests are
available since 1970 and could be accepted as a state of the art when EC7 was first drafted in the 80ies.
EC7 was finally published in 1994 as ENV 1997-1 however it took until 2004 to turn it into a European
standard and it took more years that this standard was accepted in the different countries. Also fairly
recently the countries where EC7 came into force edited national annexes. Some countries where
Dynamic Pile Load Tests are accepted for verification of safety concepts addressed Dynamic Pile Tests in
their national annexes and some even provided their own national rules including national correlation
factors.
For 17 countries a report on pile design according to EC7 is available and has been evaluated with respect
to dynamic pile load tests.
A short review of the ongoing process of formulating an executional standard for dynamic pile load tests
is also given.

2. OVERVIEW
For the following countries information on the practice of dynamic pile load tests is available (see the
national reports of this :

2.1. Austria
Dynamic pile load tests are not mentioned in a National attachment. Not used in regular practice but in
special cases. Especially in the Western provinces.

2.2. Belgium
Dynamic pile load tests are named as means to verify the ULS provided the equivalence to static tests
have been demonstrated.

2.3. Denmark
Dynamic pile load tests are used as the standard to verify the ULS for driven piles. Also wave equation
based methods and a Danish driving formula are in use. Cast-in-place piles are rare and dynamic pile tests
on these piles the exception.
National regulations and safety factors are used.
In addition a calculation of setup is proposed with suitable parameters. This latter feature is unique in
comparison to all other countries reviewed.

105
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2.4. Estonia
Special driving formulas are in use as dynamic pile load tests. Primarily Gersevanov’s theoretical
considerations are followed.

2.5. Finland
Dynamic pile load tests according to EC7 regulations are widely in use. In some cases it is even possible
to base design on dynamic load tests without accompanying static tests. As an extension to the correlation
factors in table A11 the correlation factors are not only dependent on number of piles tested but also on
the percentage of piles tested in a foundation. The lowest correlation factor also applies to a situation
where 100% of the piles are tested.

Also national model factors are defined.

2.6. France
Dynamic pile load tests are not addressed in national regulations on pile design.
Description of integrity testing of piles also covers dynamic pile load tests.

2.7. Germany
Dynamic pile load tests are applied for driven piles mostly in the coastal areas and for cast-in-place piles
more in the middle and south. Germany national annex clarifies the implicit hierarchy of dynamic piles
tests as can be extracted from EC7 (see Klingmüller/Schallert 2012).
As a special regulatory feature a set of corrections for the correlation factors is proposed depending on the
demonstrated equivalence of dynamic to static pile load tests.
A correction to correlation factors with respect to the percentage of piles tested is proposed in an
addendum to the German recommendations for piles (see Moormann/Kempfert 2014).

2.8. Hungary
Dynamic pile load tests are regularly applied. No national annex to EC7 exists for Dynamic Pile Load
tests. Simple driving formulas are not used since at least 30 years.

2.9. Ireland
Dynamic pile load tests are regularly applied and understood to be related to quality control. No national
annex to EC7 exists for Dynamic Pile Load tests.

2.10. Italy
NTC 2008 encourages the design based on static or dynamic load test, with a clear preference for the
former since the validity of the results of dynamic load tests has to be demonstrated by comparing the
results against static load tests in comparable situations.
Correlation factors are taken from EC7.

2.11. Macedonia
No pile design by load tests. No comment on dynamic pile tests.

2.12. The Netherlands


Pile design based on load tests is seldom performed in the Netherlands because of the good relation
between CPT results and bearing capacity and the relative dense grid of CPT locations (25 m x 25 m).
It is expected that in the future, with the development of new cast in situ piles, the amount of pile load
tests will increase.
No special considerations in a National annex to EC7.
Dynamic pile tests including rapid load tests are seen in the context of quality control.

2.13. Poland
Dynamic pile load tests are used. Static load tests on the same site are required for demonstration of
equivalence.

106
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2.14. Russia
Dynamic pile load tests are used in the form of refined driving formula derived by Gersevanov using a set
of adequate safety factors.

2.15. Spain
There is not a unique national annex to EC7 for application to all piles. There are different codes for
buildings and bridges.
Load test of piles is not common in Spain. Perhaps the reason is that many piles are bearing on rock. Just
some static load tests of piles in compression in projects with high loads. Dynamic load test in driven
piles are routinely done by drivers, only a few by specialized independent consultants. Dynamic load test
is not standard in bored piles.
Non-destructive or integrity testing of piles is common in Spain. Geotechnical codes recommend them in
many cases. Big piles are routinely tested with cross-hole sonic logging or sonic transparency inside
embedded tubes (CSL). Piles in building foundations are tested sometimes by sonic or echo test method
with a handheld hammer. There are few specialized consultants and many general testing laboratories
doing integrity testing of piles.

2.16. Sweden
Dynamic load testing is the main design method in Sweden since an overwhelming majority of piles are
either driven end-bearing piles or drilled end-bearing piles.
As dynamic load tests have a 40 years tradition also the EC7-National annex is advanced compared to
other countries. There is an extension to teh table of correlation factors as well as a refined set of model
factors.

2.17. United Kingdom


The National Annex provides a safety concept for piles including dynamic pile load tests. The correlation
factors of table A.NA.11 define a more conservative approach than in the original EC7 A.11 proposal.
However also the loading factors are evaluated by a different approach (see NA to BS EN 1997-1:2004)..

3. REVIEW OF CURRENT SITUATION


As can be learned by the reports of 16 countries the application of dynamic pile load tests is varying in a
wide range within the area where the EC7 is effective. On the one side a country like Sweden where
dynamic pile tests are regularly applied as a tool for pile design and on the other side a lot of countries
where dynamic pile load tests are not known. Between these two extremes there are all variations of
regular application sometimes according to EC7 regulations and sometimes without a special regulation.
However in most cases it is found necessary to proof equivalence of dynamic to static tests in a reliable
way.

4. INTEGRITY BY DYNAMIC TESTING


Some national regulations consider dynamic pile tests in context with quality control and with low strain
tests and other non destructrive tests. This may be due to the fact that also in EC7 quality control of piles
by non destrctive testing is not described in a consistent manner.

5. PERSPECTIVES
As was shown in Klingmueller/Schallert 2012 the EC7 regulations are not consistent. Several items need
clarification:
 In addition to dynamic pile load tests by measurements of force and velocities at pile top and
evaluation by wave equation models (CASE) and signal matching (CAPWAP) the use of driving
formulas with and without elastic deformation and also wave equation solutions are accepted.
However the latter are not described in a way that assures comparable and reliable results.

 Calibration. The verification of dynamic and static equivalence is defined with different wording.
Therefore there are no unique approaches. In some cases it is possible to relate ultimate capacities
however in many cases a static pile test is only carried out for 150% of the working load and a relation
to ultimate resistance from dynamic tests can not be calculated

107
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

 For larger pile foundations a combination of load tests is used: dynamic, static pile tests, wave
equation solutions and driving formula. A consistent safety concept should provide benefits for any
additional tests (see Klingmüller/Schaalert 2012).

A better and consistent application of dynamic pile tests can be achieved by the new standard
prEN 22477-4 Dynamic Pile Load Tests which is in the course of elaboration.

REFERENCES
Klingmüller, O.; Schallert, M. (2012): “Resistance factors for high-strain dynamic testing regarding
German application of Eurocode 7 and correlation of dynamic on static pile tests”,
Proceedings of IS-Kanazawa 2012: The 9th International Conference on Testing and Design Methods for
Deep Foundations, Kanazawa, Japan, 18.-20.09.2012 .

Moormann, Ch.; Kempfert, H.-G. 2014: Jahresbericht 2014 des Arbeitskreises „Pfähle“ der Deutschen
Gesellschaft für Geotechnik (DGGT). Bautechnik 91 (2014), Heft 12, S. 922–932

United Kingdom 2007: National Annex to BS EN 1997-1:2004, BSI 2007

108
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Accounting for cyclic loadings on piles with Eurocodes


Alain Puech, Fugro GeoConsulting, France, a.puech@fugro.com

Sébastien Burlon, IFSTTAR, France, sebastien.burlon@ifsttar.fr

ABSTRACT
Based on the results obtained in several international research projects related to the cyclic design of
piles and especially the French SOLCYP project, this paper proposes an original strategy to account for
cyclic loadings on piles within the Eurocodes framework. Amongst the questions to be considered, three
main topics are analysed in details. The first topic is related to a proper and complete definition of the
actions to take into account since very few elements are presented in the Eurocodes 0 and 1. The second
topic deals with the cyclic behaviour of soils and soil-structure interfaces and leads to propose a reliable
procedure for the assessment of the cyclic pile response for both axial and lateral loadings. Finally, a
strategy to analyse cyclic loadings and their effects on pile-soil systems is presented However
clarifications are required on how the verifications of the ultimate limit states (ULS) and the
serviceability limit states (SLS) should be conducted and more consensual values should be recommended
on the partial factors proposed by Eurocode 7 compatible codes and international offshore codes.

1. INTRODUCTION
Cyclic loadings are of major importance for the reliability of many structures such as offshore structures,
wind power plants or vibrating machines. Indeed, waves, wind or vibrations impose transient and cyclic
actions to these structures, which may induce excessive displacements and eventually lead to a loss of
stability. Specific analyses have to be developed especially for the design of the deep foundations and
piles supporting these structures.
The oil and gas industry has developed procedures for considering the effects of large wave cyclic loads
on foundations for offshore structures. The design guidelines include API RP 2GEO (2011), DnV
Foundations (1992) and ISO 19902 (2007). The offshore turbines industry is progressively adapting such
methodologies with, e.g. DnV-OS-J101 (2010); GL (2007); BSH (2007). DnV and GL past documents
are now being superseded by DNVGL-ST-0126 (2016). These documents are largely compatible in terms
of partial factor considerations with Load and Resistance Factor Design (LRFD) procedures (except for
the API RP2A series remaining in Working Stress Design -WSD), levels of safety and recommended
design methods and procedures.
The effects of cyclic loading on foundations are largely ignored in most civil engineering and building
activities. Eurocode 7 and national application standards reflect this poor level of consideration.
Addressing cyclic loads is not a straightforward exercise since these codes basically address the design of
buildings and bridges that are not submitted to large transient loadings.
In most countries offshore structures have been designed using offshore codes of international
applicability without any direct link with onshore standards and consequently Eurocodes. The design of
wind turbines offshore Germany is an exception. The Design Standard specifies that primarily national
and European construction standards shall be applied. A bridging document (BSH, 2011) recognizes that
some aspects of the design of offshore wind turbine foundations are not covered by these standards and
makes reference to the international offshore codes. ‘BSH’ note also provides specific guidance “on the
assessment of cyclic loads and on the verification of ultimate limit states for foundation elements taking
into account cyclic loading”.
Definitely, the use of Eurocodes and especially of Eurocode 7 for cyclic pile design requires the definition
a dedicated strategy. Amongst the questions to be considered, three main topics appear particularly
relevant:

− a proper and complete definition of the actions to take into account since very few elements are
presented in the Eurocodes 0 and 1 ;
− a reliable procedure for the assessment of the pile resistance considering the cyclic behaviour of soils
and soil-structure interfaces ;
− acceptance of precise criteria for the verification of the limit states (ULS and SLS).

109
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

This paper, largely based on the results of the French project SOLCYP, deals with these three main topics
in order to propose an original strategy to account for cyclic loadings on piles within the Eurocodes
framework.

2. MAIN RESEARCH PROJECTS RELATED TO CYCLIC LOADINGS

2.1. Main international research projects on piles


The main research projects including cyclic loading tests can be split into three groups (Jardine et al.,
2012; SOLCYP Recommendations, 2016):

− the first tests have been performed at the end of the years 70 and the beginning of the years 80 to
respond to the demand of the offshore oil and gas industry, which was moving to deeper waters,
involving new concepts of platforms (namely Tension Leg Platforms). The first cyclic loading tests
have been performed for research projects by the Institut Français du Pétrole in France and the
Building Research Establishment in United Kingdom or for specific projects (Empire, Aquatic Park,
WD58A in the USA). They were conducted on driven tubular piles or pile sections. Briaud and Felio
(1986) have established a database for the American Petroleum Institute related to the behaviour of
piles in marine soils. This database included laboratory tests, model tests and 16 axial load tests on
piles with a diameter greater than 150 mm ;
− the tests carried out by N.G.I. and by Imperial College of London (in the period 1980 - 2005) on
driven tubular piles are particularly significant. They include extensive testing campaigns on well
documented reference sites (NC clays, OC clays, sands) involving different loading modes and a very
high level of instrumentation both on the pile and in the soil. The analysis of the results has driven the
development of the first design methods ;
− more recently, two projects have been conducted: the SOLCYP project (see below), between 2008 and
2013, involving bored piles and driven piles, and the PISA project (2014-2015) related to transversal
static and cyclic loadings on driven piles (Byrne at al., 2015; results still unpublished).

2.2. SOLCYP Project

2.2.1. Presentation
The SOLCYP project (French acronym for Piles under CYclic SOLlicitations) was launched in 2008 with
the objectives of:
− (i) understanding the physical phenomena which affect the response of piles to vertical and horizontal
cyclic loads ;
− (ii) developing advanced design methods ;
− (iii) initiating pre-normative development of methodologies that may subsequently be included in
national and international codes or professional standards.

The project total budget approached 4.5 millions €, funded by Agence Nationale de la Recherche (ANR),
Ministère de l’Ecologie, du Développement Durable et de l’Energie (MEDDE), Fédération Nationale des
Travaux Publics (FNTP) and 14 private companies of the civil engineering and energy sectors. The
behaviour of bored and driven piles under cyclic loading in sands and clays was investigated through a
range of laboratory, field and theoretical approaches.
A detailed description of the scope of work can be found in Puech et al. (2012). Main experimental results
were presented at the TC209 Workshop organized during the 18th ICSMGE Conference in Paris and are
available on the French Committee of Soil Mechanics (CFMS) website (www.cfms-sols.org).
Interpretation of results and development of design procedures are now completed. The last step of
SOLCYP will be the publication of “Recommendations for the design of piles under cyclic loading”
(French version is expected mid-2016 and English version for end of 2016; SOLCYP Recommendations,
2016).

2.2.2. . Field tests


In situ axial pile tests were conducted on two sites. The first site is Merville, in northern France where
stiff to very stiff overconsolidated (OCR of about 10) plastic Eocene Flanders clay is encountered below
3 m depth. Ten test piles were installed in March 2011 (Figures 1 and 2), including four driven closed-
ended tubular steel piles, four bored (CFA) piles and two screw piles. The piles are 13.5 m long with
diameters of 406 mm (driven piles) or 420 mm (bored and screw piles). Piles were instrumented with

110
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

removable extensometers for axial strain measurements. Load tests were conducted from May to June
2011, including conventional incremental static tests, rapid load tests and cyclic load tests. The latter
covered both tests that failed after relatively small numbers of large cycles and those that extended to very
large numbers (N > 10 000) of relatively low amplitude cycles. All loading modes were investigated:
tension, compression, one-way and two-way (Benzaria et al., 2012, 2013a).

Figure 2: Two-way cyclic testing.

SOLCYP pile tests at Merville and Loon-plage


Figure 1: Installing driven and bored piles.

The second site is in Loon-Plage near Dunkirk where post-glacial Flandrian dense to very dense fine
sands are encountered. Given the availability of previous cyclic pile testing by Imperial College in the
same sands at a neighbouring site (see Jardine and Standing, 2000 and 2012 and Rimoy et al., 2013) only
two steel piles (13.5 m long, 406 mm diameter) were driven and tests conducted to complement and
extend the existing data set. Five bored piles (420 mm diameter, 8 m or 10.5 m long) were installed in
November 2011 and submitted in March 2012 to a load testing programme very similar to the Merville
piles (Benzaria et al., 2013b).

2.2.3. Model tests


Extensive centrifuge tests on model piles were performed in Fontainebleau sand and Speswhite clay using
the large centrifuge of IFSTTAR (formerly LCPC) in Nantes. The main objective under axial loading was
to derive cyclic stability diagrams for different soil conditions (medium dense or dense sands and soft
clays), installation modes (cast in place or driven) and loading modes (tension or compression, one-way
and two-way). Series of tests were also conducted under cyclic transversal loading aimed at capturing the
evolution of lateral displacements and moments in both sands and clays (Khemakem et al., 2012).
Highly instrumented axial displacement pile tests in sands were performed in the large calibration
chamber of the 3S-R Laboratory in Grenoble. These tests, carried out in collaboration with Imperial
College London, provide highly valuable data on the mobilisation of the interface pile-sand friction and
its evolution with number of cycles (e.g. Tsuha et al. 2012, Rimoy et al. 2012). Similar tests were also
performed in clay using the calibration chamber of the Navier-Cermes laboratory in Marne-la-Vallée.

2.2.4. Other aspects


The SOLCYP project also included extensive series of static and cyclic laboratory tests on soils and soil-
pile interfaces: triaxial tests, direct simple shear tests (DSS, CSS) and simple shear tests with constant
normal stiffness (CNS) or constant normal loading (CNL). The databases were used to support the
development of procedures to assess degradation effects for axial and transversal loadings.
Different ways were explored for modelling the response of piles under axial or transversal cyclic
loading. Positive issues are mentioned in Sections 4 and 5. Finally complete procedures for designing
piles under axial and transversal loading could be developed. These procedures are basically compatible
with the Eurocode 7 framework but may require some adaptations. The main steps of these procedures are
described in Section 6.

111
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3. CHARACTERIZATION OF ACTIONS
In standards such as Eurocode 7, loads are only characterized by their maximum intensity (characteristic
load). This approach is by far insufficient for characterizing a “cyclic” loading.
The term “cyclic” loading is used generically to characterise variable loads that have clearly repeated
patterns and a degree of regularity in amplitude and return period. Cyclic loads can be essentially of
environmental origin (waves, tide, current, wind, earthquakes, ice sheets) or anthropogenic (due to traffic,
blasting operations, plant operations or rotating machinery).

3.1. Regular cyclic loading


The most common practice is to conduct uniform cycling with load or displacement series’ that employ a
fixed frequency and regular amplitude. Such regular cycle tests can be defined by their number of cycles
N, period (T), mean (or average) load (Q m ), half-cyclic load amplitude (Q c ) (Figure 3).
Regular cyclic loading can be symmetrical (Q m = 0) but is more normally non-symmetrical due to a non-
zero average component. One-way loading imposes Q c < Q m while two-way loading leads to Q c > Q m .

Load Q

Cycle period T

Time

Figure 3: Definition of a regular cyclic loading.

3.2. Cyclic load histories


Cyclic load histories, as obtained from simulations of the design event, are expected to be irregular in
nature. Transforming “actual” cyclic loadings into “ideal” series of regular cycles is a requirement to
match with the outcomes of the uniform cycling imposed in laboratory or field testing tests performed in
order to facilitate the assessment of cyclic degradation of soils, interfaces or pile-soil systems. Cycle-
counting methods, most often based on ‘rainflow’ analyses (e.g., ASTM E1049-85), are available to this
purpose. The idealised loading is defined by uniform series of N i cycles of same period T characterized
by their mean value Q mi and their half-amplitude Q ci (Figure 4 – Step 1)

3.3. Equivalent cyclic loading


The equivalent number of cycles N eq of a cyclic event is defined as the number of cycles of maximum
cyclic amplitude causing the same degree of damage to the pile-soil system as the design cyclic event
(Figure 4 – Step 2). It must be realized that the equivalent cyclic loading depends not only on the cyclic
loading itself but also on the soil response (schematically the soil mass response for transversal loading or
the soil-pile interface response for axial loading). Several methods can be called for to assess N eq (e.g.,
Andersen, 1976, 2004, 2009; Andersen et al., 2013; Lin and Liao, 1999; Peralta, 2010). They require the
selection of a memory parameter to describe the degradation process: the shear strength in clay, the
excess pore pressure in sands or the displacement of the pile head. These methods follow a stepping
procedure to determine the number of cycles of series i+1 causing the same damage as the complete
series i, this number being added to the number of cycles of the series i+1 to calculate the damage at end
of series i+1. This type of procedure implicitly obeys the Miner’s law assuming that damage is
independent of the order of occurrence of the loads. It should be recognised that the Miner’s law is only
an approximation for soils.

112
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Load Q
True
cyclic
loading Time

Step 1 Cycle counting


Rainflow method

Load Q
Ideal
cyclic
loading Time

Step 2 Damage law


(Miner law)
Load Q

Equivalent
cyclic
loading Time

Figure 4: Procedure for transforming expected irregular cyclic loading into equivalent cyclic loading.

4. CYCLIC DEGRADATION OF PILE-SOIL SYSTEMS


This section is devoted to the effects of soils and soil-pile interfaces degradation on the response of piles.
The mechanisms controlling soils and interfaces degradation are a vast subject and are not discussed here.

4.1. Axially loaded piles


Typical experimental results obtained from static and cyclic pile load tests on axially loaded piles are
shown in Figure 5. Cyclic loads generate permanent cyclic displacements, associated to a degradation of
the shaft friction, even if the maximum cyclic levels are significantly lower than the static capacity. The
rate of displacement depends on the load combination (Q m , Q c ). Three modes of response are generally
considered (Figure 6):
- for severe loading combinations, the rate of displacement is high and the pile reaches failure at
typically less than 100 cycles; the pile is said unstable under the loading applied ;
- for low to moderate loading combinations, the initial rate of displacement decreases and the pile head
displacements stabilise: the pile is said stable ;
- for intermediate loading combinations, the rate of displacement tends to decrease over the first tens to
hundreds of cycles and then increases continuously. Pile failure may occur for typically a thousand
cycles or more. The pile is said metastable.

Stability diagrams (also called interaction diagrams), as introduced by Poulos (1988), are useful tools for
representing the response of piles to cyclic loading (Figures 7, 8, 9). The number of cycles to failure or
the number of cycles N applied to the piles without causing failure is plotted as a function of the
normalised average (Q m /R s ) and cyclic (Q c /R s ) load components. The diagonal line Q max =R s corresponds
to the static bearing capacity, whereas the second diagonal line indicates the separation between one-way
(Q c <Q m ) and two-way (Q c >Q m ) loading. For failed piles, the distance of the point to the first diagonal

113
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

indicates the amount of degradation in capacity. Degradation is much higher under two-way loading than
under one-way loading.
The stability diagram may have different aspects according to the type of soil. In sands a representation
with the three zones (stable, metastable, unstable) is considered as the more appropriate (Figure 7). In
normally consolidated or slightly overconsolidated clays, it may be feasible to draw lines of isovalues of
the number of cycles to failure N f (Figure 8). In heavily overconsolidated clays (Figure 9), it is observed
that the separation between the stable zone and the unstable zone is thin, so that no metastable zone can
be identified. In the one-way domain, threshold values for piles at Merville are in the range 0.8 to 0.9.

F4
Static
CC1
N=14
Load at pile head (kN)

Vuc=1000 kN at 42 mm

CC2
N=5000

Pile head displacement (mm)

Figure 5: Piles submitted to static and cyclic axial loading - Piles F4 and F5 are identical (402mm bored
piles) – F4 submitted to static test – F5 submitted to cyclic tests CC1 and CC2.

Failure in Failure in
tension compression
Cyclic displacements zm/B (%)

F: Failure after N cycles


Qc/Rc

Cycle number Qm/Rs


Rs/Rc
Figure 6: Typical responses of cyclically loaded Figure 7: Schematic cyclic stability diagram
piles (axial). showing stable (A), metastable (B) and unstable
(C) zones.

114
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Failure
No failure

Two-way
Loading

Qc/Rs

One-way
Loading

Qm/Rs

Figure 8: NGI tests in Haga clay (Karlsrud & Haugen, 1985) – Cyclic interaction diagram, tension tests.

Cyclic failure (Nf)


Instable zone
No failure after N cycles
Stable zone

Two-way
loading

One-way
Qc/Rc loading

Qm/Rc

Figure 9: SOLCYP - Merville field tests (Benzaria et al., 2013a) – Cyclic interaction diagram for bored
piles in compression – One-way and two-way loading modes.

4.2. Transversally loaded piles


For transversally loaded piles, typical experimental results obtained on flexible piles in the centrifuge are
shown in Figures 10a and 10b. An increase of the transversal pile displacements corresponding to the
accumulation of permanent cyclic displacements can be observed in Figure 10a. Concomitantly the
maximum bending moment measured in the pile increases slightly with the number of cycles and its point
of application tends to move downwards (Figure 10b).

115
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Final monotonic (a)

Transversal load (kN)


loading

Transversal pile displacement (mm)

Bending moment (kN.m)

(b)
Depth (m)

Figure 10: Pile submitted to cyclic transversal loading - Centrifuge tests – Prototype pile - (a) pile head
displacements and b) bending moments.

5. JUSTIFICATIONS FOR PILES UNDER CYCLIC LOADING

5.1. Introduction
In the framework of Eurocodes, the justifications for the design of geotechnical structures are linked to
the limit states concept: ultimate limit states (ULS) and serviceability limit states (SLS). ULS’ are
checked by introducing partial factors on loads and resistances: global resistances, such as the bearing
capacity, or shear strength resistances, such as the undrained shear strength or the friction. These two
approaches correspond to the resistance factor approach or the material factor approach. SLS’ refer to the
displacements that the supported structure can accept without considering any partial factors.
For pile design, a distinction is made between axial and transversal loads. For axial loading, it is possible
to define the limit resistance of a pile both in tension and in compression. For transversal loading, a limit
resistance can only be defined for short and rigid piles. For flexible piles, ULS concern only structural
(pile integrity) verifications, whereas displacements may be considered both for SLS and ULS.

5.2. Piles submitted to axial cyclic loads


For piles submitted to axial loads, the framework of ULS leads to make a comparison between loads or
actions Q and resistances R: Q < R. With the use of partial factors, the idea is to limit the load applied on
the pile to a fraction of the bearing capacity of the pile. The combination of all partial factors provides in
many cases the former global safety factor used in the past. Splitting the global safety factor in several
partial factors can be viewed as a way to optimize each individual factor. Regarding the pile design, even
if the national standards of European countries are allowed to adopt different global safety factors, at the
end of the calculation, the global safety factor related to bearing capacity of onshore piles is usually
between 1.9 and 2.5 (e.g., Vardanega et al., 2012). With these values, the aim is clearly to ensure the
stability of the pile and avoid that no failure mechanism may occur. Moreover, in a sense, it permits the
limitation of the pile displacements.
It is interesting and amazing to note at this stage that partial factors recommended by the international
standards in use in the offshore industry lead to lower global safety values in the range 1.5 to 2.0. This

116
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

discrepancy requires certainly clarifications where Eurocodes 7 would claim to become a reference
document for offshore foundation works.
For piles submitted to cyclic loadings, direct comparison between actions and resistances is no more
possible. Indeed, resistances are affected by actions and the bearing capacity of a pile varies according to
the characteristics of the cyclic loading (see section 4.1). Inversely, actions are in essence highly variable
with time and it is necessary to define the main features of these actions before any calculations (see
section 3). Moreover, the axial stiffness of the pile affects the degradation process at the soil-pile
interface.
A prerequisite to the verification process is to have a clear understanding of the concept of pile bearing
capacity. When a pile is submitted to a monotonic increasing load, the bearing capacity is defined by the
load that induces a displacement of the pile head greater than 0.1B (or any other criterion) or a loss of
equilibrium. The shape of the load-settlement curve depends on the type of soil: in clayey soils, the failure
criterion corresponds to a loss of equilibrium whereas, in sandy soils, the failure criterion is related to a
large displacement. Many calculation models are available to assess the static bearing capacity. They may
depend on intrinsic shear parameters or in situ resistance parameters such as cone resistance or limit
pressure. Well documented procedures for conducting static load test (CRL or incremental) allowing the
assessment of the static capacity of piles are available.
When a pile is submitted to a cyclic loading, two other definitions should be introduced: the cyclic
bearing capacity (Figure 11) and the post-cyclic bearing capacity (Figure 12).
The cyclic bearing capacity corresponds to a loss of equilibrium or a large displacement (0.1B) mobilized
during the cycling loading. It means that the pile resistance (base resistance and shaft resistance) has been
fully mobilized or the accumulated displacements during the cyclic loading are too large before the full
mobilization of the pile resistance. The cyclic bearing capacity depends on both the ground resistance and
the loading parameters: load average and load amplitude.
There are, for a particular soil-pile configuration, an infinite number of cyclic bearing capacities, each one
being defined by three parameter (Q m , Q c , N f ). A significant number of in situ cyclic tests have now been
conducted all over the world to analyse the evolution of the cyclic bearing capacity for various soil-pile
configurations (see section 2). Calculation models are available to estimate the cyclic bearing capacity of
a particular soil-pile configuration and for a defined cyclic loading event (design cyclic event). Some
methods will be shortly presented later in this paper. Procedures for conducting cyclic load tests on piles
aimed at assessing the cyclic capacity under calibrated design events are proposed in SOLCYP
Recommendations (2016).
The post-cyclic bearing capacity corresponds to the static bearing capacity of a pile which has been
subjected to a cyclic loading. The post-cyclic capacity of an unfailed pile, which has been submitted to
low to moderate cyclic loads, can be higher or lower than the static capacity of the equivalent virgin pile.
The post-cyclic capacity of a pile which has reached cyclic failure may be higher or lower than its
previous cyclic capacity. The post-cyclic capacity captures the effects of a particular cyclic loading event
on the subsequent response of the pile. It may be affected by the time elapsed between the end of the
application of the cyclic loads and the start of the post-cyclic test (pore pressure dissipation in clays,
ageing in sands, etc.).
Q

Static failure

C1
Load

Cyclic failure (Qm1, Qc1, N1)


Cyclic failure (Qm2, Qc2, N2)

δ
δc: failure displacement
C2 criterion

Figure 11: Definition of cyclic failure and cyclic capacity

117
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

δc
A
A : Static failure
δc
B B : Post-cyclic failure C1
C1 C : Post-cyclic failure C2
Load

C δc: failure displacement


δc criterion

C2

Figure 12: Definition of post-cyclic capacity

5.3. Piles submitted to transversal loads


For piles submitted to (static or cyclic) transversal loads, the framework of ULS and SLS may appear a
little bit confusing since, except for short and rigid piles (e.g., monopiles used for wind turbines
foundation, it is generally not possible to define clearly a limit resistance for the pile-soil system. For
flexible piles, failure of the pile structure is reached before loss of global stability. Transversal loading is
therefore limited by two criteria: lateral pile head displacements that can be accepted by the
superstructure and resistance of the pile structure to the resulting bending moments.
Concerning the structural resistance of the pile, several procedures are used: in some cases, values of
actions are multiplied by ULS partial factors at the beginning of the calculation and values of bending
moments directly correspond to ULS values (resistance factor approach), and, in other cases, values of
actions are not increased by partial factors and values of bending moments correspond to characteristic
values and have to be multiplied by ULS partial factors to be considered as ULS values (effect factor
approach).
For cyclic loads, the aim of the design should be to assess the evolution of the displacements at the pile
head under the cyclic design event and compare these values to the requirements of the supported
structure. In today practice, as long as reliable methods to assess permanent cyclic displacements were not
recognized, prudent design procedures consisted in limiting the yield development in the ground around
the pile. This can be done by using reduced p-y curves. For verification of the structural integrity of the
piles, values of bending moments into the piles may be obtained by considering the actions without any
partial factors. These values are then multiplied by appropriate partial factors to check ULS’ at the end of
the calculation.

6. SOLCYP STRATEGY
The full design of a pile subject to cyclic loads may call for relatively complex procedures which may not
be justified in every day’s practice. The relative success of standard engineering approaches (which
basically ignore potential effects of cyclic loads) suggests that, in most cases, cyclic loads can be
accommodated by expert engineering judgement and reasonably conservative assumptions. However this
is not always the case. Graduated approaches are thus proposed, corresponding to increasing complexity,
and criteria have been developed to determine which level of analysis is appropriate to the case
considered.
Criteria should integrate the following:
- Sensitivity of the structure to foundation displacements and rigidity;
- Relative importance of the cyclic load component (characterised by the ratio Q c /Q m of the cyclic
amplitude to the average load)
- The nature of the soil and its more or less degradation potential to cycles;
- Potential human, environmental or economic consequences of a poor foundation response.

The proposed design philosophy is illustrated by the flow charts in Figure 13 for axial loading and in
Figure 14 for transversal loading.

118
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 13: SOLCYP strategy for axially loaded piles (SOLCYP Recommendations, 2016)

119
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 14: SOLCYP strategy for transversally loaded piles (SOLCYP Recommendations, 2016)

In both cases the procedure proposes two steps:


- STEP1 is aimed at deciding whether specific cyclic design analyses are required or not ;
- STEP 2 describes how proceeding with detailed cyclic design analyses.

6.1. Axial loading

6.1.1. STEP1
The SOLCYP procedure assumes firstly that the sizing of the pile under the ELU and ELS static load
combinations is available according to standard Eurocode 7 compatible procedures. At this stage loads are
known by their characteristics values:
- G k the characteristic value of permanent loads (not factored) ;
- Q k the characteristic value of variable loads (not factored).

120
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

A preliminary analysis of the load cases is suggested, starting by considering all combinations (ULS and
SLS) and calculating:
- the minimum load applied to pile as:
∑Gi
k ;i

- the maximum load applied to pile as:


∑G i
k ;i + ∑ Qk ; j
j
The loading can be characterized by:
- its mean value:
∑Q k; j

Gm = ∑ Gk ;i +
j

i 2
- its cyclic component (half-amplitude):
∑Q k; j

∆Q =
j

2
The coordinates (G m ; ΔQ) can be normalized by the calculated static capacity of the pile R c and the
corresponding point plotted in a stability diagram representative of the soil-pile system considered (Figure
15). A diagram may be considered representative when the type of soil and the type of pile used to
establish the diagram are similar to the soil-pile system considered. A number of stability diagrams are
available covering medium dense to dense sands, normally consolidated to highly overconsolidated clays,
driven metallic piles and bored CFA or screw piles. The SOLCYP Recommendations contain guidance to
perform cyclic pile tests and procedures to adapt stability diagrams to a particular case.

0.8 Limit resistance


(compression/tension)
FOS=3

0.6 FOS=2
DQ/Rc
∆Q/R c

FOS=1.5
Nf=100
0.4 Bridge or Building

Zero Jacket for oil platform


damage
0.2 Jacket for wind turbine

One way/Two way


loading
0
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1

GmGm/Rc
/Rc

Figure 15: Checking cyclic loading severity and cyclic pile stability with the aid of a cyclic stability
diagram

In figure 15, three loading points are represented in a typical stability diagram. The severity of the cyclic
loading, for the soil-pile configuration considered, can be estimated by comparing the position of the
points to the zero damage line and to the failure line for N f = 100 , which can be assimilated to the line
separating the metastable and unstable zones.
The point representative of the “bridge or building” loading plots below the zero-damage line: detailed
cyclic analyses are not required and the static design can be validated.
The point representative of the “wind turbine jacket” loading plots above the N f =100 line in the unstable
zone: detailed cyclic analyses are required (or pile should be redesigned).

121
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

The point representative of the “oil jacket platform” loading plots in between the two lines. Different
alternatives can be envisaged: revisit the static design and increase the pile dimensions; question the
representativeness of the stability diagram and possibly perform cyclic pile tests to check/adapt the
position of the stability lines; proceed to detailed cyclic analyses.
It should be highlighted that the concept of stability diagrams has a major limitation. Existing stability
diagrams have been performed for (short) piles of high relative soil-pile axial rigidity. Long piles with
lower relative soil-pile axial rigidity may develop stress reversal in the upper part of the pile and more
severe degradation. Guidance is given in the SOLCYP recommendations on this aspect. In practice the
stability diagram concept may be relevant for most onshore piles and rigid offshore monopiles used for
wind turbines.

6.1.2. STEP 2
A prerequisite to start a detailed cyclic design is to have a full characterization of the cyclic loadings. The
way to properly characterize cyclic loadings is described in section 3.
The degradation laws to introduce in the models may be obtained by different ways:
- laboratory testing on soils using direct simple shear devices (DSS, CSS) or triaxial (CAUc, CAUe) ;
- laboratory testing on interfaces (e.g. constant normal stiffness CNS tests ; constant normal load CNL
test; constant volume CV tests) ;
- model pile testing in calibration chambers or in the centrifuge ;
- field testing.

These aspects will not be discussed here.


Several models are available to assess the degradation of the pile capacity as a result of the integration of
the local degradation of the interfaces. They include t-z envelope models; t-z cycle by cycle models and
finite element models. Some examples are briefly described in section 7.

6.1.3. Verifications for ULS and SLS


At the end of the calculation procedure the cyclic bearing capacity of the pile under the cyclic design
event (R t,cyc in tension or R c,cyc in compression) is available. It is reminded that the degradation has been
evaluated using characteristic (not factored) values of the loads (G k and Q k ).
The verifications for ULS or SLS are made using appropriate load factors and applying a resistance factor
on the cyclic resistance:
Rt ,cyc
γ G G k + γ Q Qk < in tension
γR

Rc ,cyc
γ G Gk + γ Q Qk < in compression
γR
As already mentioned in section 5.2 the partial factors may vary according to the contractual context.

6.2. Transversal loading


The approach recommended for transversal loading is very similar in essence and will not be detailed
here.
For STEP1, specific criteria have been developed for checking the severity of the loading case and decide
whether a detailed cyclic analysis is required.
For STEP 2, global approaches have proved to be sufficient for most applications. A brief description of
this type of approach is given in section 7. The derivation of “true p-y curves” accounting explicitly for
the parameters of the loading (H m , H c , N) remains a challenge.
Verifications to be performed are described in section 6.3.

7. CALCULATION MODELS FOR CYCLIC LOADING

7.1. Axial cyclic loading


Several approaches are available for assessing pile degradation under axial cyclic loading and calculating
a cyclic bearing capacity. One can mention the NGI approach (codes PXCY/PAX2), the ICL approach
(ABC method), Poulos approach (code SCARP), UWA/Randolph approach (code RATZ/CYCLOP) and
the SOLCYP approach. A brief outline of the NGI, ICL and SOLCYP approaches is presented below.

122
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

7.1.1. NGI approach


The approach developed by NGI is based on a total stress approach and inspired from the methodology
proposed by Andersen (1976, 2004) for assessing the response of the huge gravity platforms installed in
the North Sea and adapted after for studying the stability of suction anchors and caissons. The NGI cyclic
axial pile codes PAXCY and PAX2 were developed by Karlsrud et al., (1986) and Nadim and Dahlberg
(1996). T-z springs are established by integrating the shear strains developed in the disk of soil
surrounding the pile segment. The cyclic soil response is based on shear strength contour diagrams
constructed from series of static and cyclic DSS laboratory tests assuming that the stress path in the soil
volume when the failure mechanism occurs is close to the one imposed in the DSS test. This approach has
been validated for driven piles in clayey soils.

7.1.2. Imperial College approach


The approach known as the ABC method (Jardine and Standing, 2000 and Jardine et al., 2005) links, for
both sands and clays, the decrease of the local shaft friction to the decrease of the effective radial stress
∆s’ r at the soil-pile interface. The decrease of the effective radial stress is induced by the cyclic shear
amplitude t cyc at the soil pile interface and can be estimated in conducting series of cyclic DSS tests at
constant volume. The main result can be expressed by the following equation:

∆s ' r  t cy  C
= A B +  N
s 'r 0  t max stat 

where: s’ r0 is the initial effective stress applied to the sample (assimilated to the radial effective stress
acting on the pile after installation), t max stat is the maximum shear stress under monotonic loading
(assimilated to the limit shaft friction), N is the cycle number and A, B, C are three parameters describing
the soil behaviour.
It is noted that results are assumed independent on the average shear stress.

The local shaft friction t s is linked to radial effective stress using the following equation:

τ σ = (σ ' r 0 −∆σ ' r ) τan δ

This approach allows calculating the cyclic bearing capacity of the pile under the design event.

7.1.3. SOLCYP approach


The approach developed within the SOLCYP project presents similarities with the Imperial College
approach but considers that both the cyclic shear amplitude t cyc and the mean shear stress t m at the soil
pile interface should be considered in assessing the decrease of the radial effective stress and
consequently of the shaft friction.
The decrease of the radial effective stress is obtained from series of interface tests (Pittos, 2014; Pra-aï et
al., 2016a) conducted at normal constant stiffness (CNS), this condition being considered well
representative of the stress conditions acting on the pile wall. A large database of CNS interface tests on
Fontainebleau and Loon-Plage sands has been established (Pra-aï, 2013). It is assumed that a low to
moderate cyclic loading induces a decrease of the normal stress at the soil-pile interface due to
contractance. The variation of the normal stress leads to a variation of the normal displacement u n at the
soil-pile interface. In the SOLCYP approach, the decrease of normal effective stress Δσ’ n depends on six
parameters: the cyclic number N, the initial normal effective stress σ’ n0 , the initial mean shear stress t cm ,
the cyclic shear amplitude t cyc , the normal stiffness K n and the soil relative density Id.
The decrease in radial effective stress derived from the CNS tests can be implemented in three models of
increasing complexity:

1- Using envelope t-z curves where the maximum value is degraded as a function of the loading
characteristics (Q m , Q c ) and the number of cycles N (Figure16).

2- Using the TZC model (Burlon et al., 2013). In this approach, degradation laws are coupled with an
explicit cyclic t-z curve (Figure 17). Pile displacements and pile resistance are calculated cycle by
cycle.

123
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3- Using the finite element method. An “explicit” approach in the sense of Wichtman (2005) is
implemented, where the volumetric strains induced by the cyclic loading are calculated according to
SOLCYP laws and imposed through special interface elements controlling shaft friction (Pra-aï et al.,
2016b).

After 5 cycles
After 10 cycles Static at
13.8 m depth
After 20 cycles
After 100 cycles
Shear stress (kPa)

Shear stress (kPa)


After 341 cycles

Static at
3.3 m depth

After 5 cycles
After 10 cycles
After 20 cycles
After 100 cycles
After 341 cycles

Local displacement (m) Local displacement (m)

Figure 16: Degradation of t-z curves as a function of cyclic loading characteristics and number of cycles
(SOLCYP Recommendations, 2016)

Figure 17: Concept of t-z cyclic curve in TSZ model (Burlon et al., 2013)

7.2. Transversal loading

7.2.1. “Cyclic” p-y curves


Offshore codes propose so-called “cyclic” p-y curves to assess the transversal response of offshore piles.
These envelope curves are to be applied at the end of a storm event to assess the transversal response of
the pile under the extreme environmental load. The degradation of the static response is based on data
obtained on relatively small diameter piles submitted to cyclic loadings simulating storm loadings. It is
noted that the characteristics of the loading and the number of cycles are not considered explicitly.

7.2.2. Transversal behaviour of monopiles


The extensive use of large diameter monopiles (typically 4-7m) for offshore wind turbines in northern
Europe has posed new design questions. Research has been conducted with model piles in sands
submitted to large number of cycles (typically 104 to 105) and full ranges of cyclic conditions at e.g.
Oxford University (Leblanc et al, 2010) or Leipnitz University, Hannover (Peralta, 2010).
There is consensus to consider that the accumulation of transversal displacements (or rotations) due to
cycles can be captured with exponential or logarithmic expressions such as:

124
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

yN  θ 
 or N  = k .N − m
y1  θ1 
 θ 
 or N  = 1 − t. ln ( N )
yN
y1  θ1 

where y N and y 1 (respectively θ N and θ 1 ) are the value of the pile head transversal displacement
(respectively pile rotation) at cycle N and 1 respectively; m is an exponent depending on the soil type; and
k or t are factors depending on the average and the cyclic amplitude of the load.
This approach is named “global approach” by opposition of “local” p-y approaches.
In SOLCYP, the global approach has been privileged: a first calculation is performed in monotonic
conditions and displacements and bending moments are extrapolated according to logarithmic equations.
These equations have been derived from extensive series of centrifuge tests in sands and clays.
As an example, the following equations are proposed for sands:

0.35
0.235  H c 
log( N )
yN
= 1+  
y1 CR  H max 
0.35
M max,N 0.094  H c 
= 1+   log(N )
M max,1 CR  H max 

where: CR is a rigidity coefficient introduced to account for pile flexibility and reconcile experimental
data from long piles and short monopiles.
Attempts have been made to develop “true” cyclic p-y curves which could explicitly take into account the
characteristics of the loading and the number of cycles. Although some progress has been noted, the
exercise is still challenging.

Static Static

Cyclic

a – Depth z < zR b – Depth z > zR

Figure 18: Example of cyclic p-y curve proposed in offshore codes (API RP2A – soft clay)

8. CONCLUSION
This paper addresses the feasibility and the associated potential difficulties of accounting for cyclic
loading in pile design with the Eurocodes and in particular with Eurocode 7.
A large experience with cyclic pile design is available in the offshore industry and recent progress is
being made in the field of monopiles for offshore wind turbines. In France, the SOLCYP project has
provided extensive and valuable data on the cyclic axial and transversal response of onshore piles. Based
on this amount of data, it is possible to propose avenues for designing piles under cyclic loadings.
Degradation effects of cyclic loads on pile capacities and pile head displacements are well documented.
Procedures to evaluate the degradation of soils and soil-pile interfaces are becoming mature and combine
laboratory testing, model pile testing and field testing.
Several models for calculating the response of piles under cyclic loading have been developed and
reasonably calibrated on field data. They integrate the local degradation of the soil-pile along the pile,
accounting for structural flexibility, in order to provide pile head displacements and capacity degradation

125
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

at the end of the design cyclic event. Models are of different complexity: they include global models,
local models based on cyclic load transfer curves (t-z or p-y) and finite element models.
This know-how may be capitalised and easily introduced in Eurocode 7 under the form of informative
annexes. The SOLCYP Recommendations, which will be published shortly, provide extensive and up to
date information on these aspects.
Three main issues were identified in the paper dealing with (1) the definition of actions, (2) the
verification procedures and (3) the harmonization of the partial factors to be applied.
A more complete definition is needed for cyclic loadings. This include the need for assessing load time
histories, transforming these time histories into idealized loadings and assessing equivalent number of
cycles for the design cyclic event. To this respect the SOLCYP Recommendations propose a procedure
similar in essence to the procedure recommended by the German BSH guidance note (2007) for offshore
wind turbines monopiles.
Regarding safety considerations and verification procedures, in depth discussions are required at the
European level but it is anticipated that consensus can be reached. The SOLCYP procedure is presented
in the paper as a starting point. Basically ULS cases are verified with respect to the cyclic pile resistance
(i.e., the reduced resistance of the pile at the end of the design cyclic event). Degradation effects due to
cyclic loadings are evaluated for both ULS and SLS cases by considering characteristic values. Partial
factors are applied at the end of the calculation both on actions and on resistances. This procedure is
similar to the ‘effect factor approach’ introduced in the next version of Eurocode 7. It is to a large extent
compatible with the German approach outlined in BSH and EA-Pfähle, Chapter 13.
Partial factors on loads and resistances may vary from one country to another according to national
application standards. In addition there are significant discrepancies on the use of partial factors between
Eurocode compatible codes and international offshore codes. Global safety factors in use in the offshore
industry are lower than those used in building and civil engineering. This point will require clarification
and possibly harmonization to avoid conflictual situations in the future.

ACKNOWLEDGEMENTS
Most of the results presented in this paper have been obtained during the French National Project
SOLCYP, managed by IREX and funded by: the French Research National Agency, forteen companies
from the civil engineering and energy sectors, the French Ministry for Sustainable Development
MEDDE), the Region “Pays de la Loire”.

REFERENCES
Andersen, K.H. (1976) Behaviour of clay subjected to undrained cyclic loading. Int. Conf. on Beh. of
Offsh. Struct., BOSS'76. Trondheim. Proc., Vol. 1, pp. 392 403. Also NGI Publ. 114, pp. 33 44.

Andersen K.H. (2004). Cyclic clay data for foundation design of structures subjected to wave loading.
Keynote Lecture; Proc. Int. Conf. on Cyclic Behaviour of Soils and Liquefaction Phenomena, CBS04,
Bochum, Germany. Proc. p. 371-387, A.A. Balkema, Ed Th. Triantafyllidis.

Andersen K.H. 2009. Bearing capacity of structures under cyclic loading; offshore, along the coast and
on land. 21st Bjerrum Lecture presented in Oslo 23 Nov. 2007. Can. Geotech. J. 46: 513-535. Also:
Norsk Geoteknisk Forening, Bjerrums Foredrag Nr. 21.

Andersen, K., Puech A. and Jardine R. (2013) Cyclic resistant geotechnical design and parameter
selection for offshore engineering and other applications, TC 209 Workshop, 18th ICSMGE, Paris
September 2013.

API RP2A (2000) Recommended practice for planning, designing and constructiong fixed offshore
platforms – Working stress design. RP2A-WSD, American Petroleum Institute.

API RP2GEO 2001 (2011) API Recommended Practice, Geotechnical and Foundation Design
Considerations, First Edition, American Petroleum Institute.

ASTM 1985 Standard practices for cycle counting in fatigue analysis. Designation E 1049-85, vol. 03. 01
of Metal Test Methods and Analytical Procedure. pp. 836-848.

126
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Benzaria O., Puech A. and Le Kouby A. (2012). Cyclic axial load-tests on driven piles in
overconsolidated clay, 7th Int. Conf. on Offshore Site Investigations and Geotechnics, OSIG, London.

Benzaria A., Puech A. et Le Kouby A. (2013a). Essais cycliques axiaux sur des pieux forés dans l’argile
surconsolidée des Flandres. Proceedings 18th ICSMGE, Paris.

Benzaria A., Puech A. et Le Kouby A. (2013b). Essais cycliques axiaux sur des pieux forés dans des
sables denses. Proceedings 18th ICSMGE, Paris.

Briaud, J.L. and Felio, G.Y. (1986) Cyclic axial loads on piles: analysis of existing data. Canadian
Geotechnical Journal, 23,362-371.

BSH - a) Standard: Design of Offshore Wind Turbines 2007, b) Guidance for use of the BSH standard
„Design of Offshore Wind Turbines“ 2011, c) Ground investigations for offshore wind farms 2008,
Federal Maritime and Hydrographic Agency (BSH), Hamburg and Rostock.

Burlon, S., Thorel, L., Mroueh, H. (2013) Proposition d’une loi t-z cyclqiue au moyen d’expérimentations
en centrifugeuse. Proceedings of the 18th International Conference on Soil Mechanics and Geotechnical
Engineering, Paris.

Byrne B.W., McAdam R., Hurd H.J., Houlsby G.T., Martin C.M., Zdravkovic L., Taborda D.M.G., Potts
D.M., Jardine R.J., Dideri M., Schroeder F.C., Gavin K., Doherty P., Igoe D., Muir Wood A., Kallehave
D. & Skov Gretlund (2015) New design methods for large diameter piles under lateral loading for
offshore wind applications, ISFOG III – Meyer (Ed.), pp. 705-710.

DNV. 1992 (1992) Classification Note 30-4, Foundations. Det Norske Veritas, Oslo.

DNV-OS-J101. (2010) Design of Offshore Wind Turbines Structures, DnV Offshore Standards, Det
Norske Veritas, Oslo. Dyvik R., Andersen K.H., Madshus C. and Amundsen T.

DNV-GL –ST (to be published, 2016 ?) Supports structures for windturbines – Proposal 2015-028
GL Rules and Guidelines (2005). IV Industrial Services. Part 2- Guidelines for the Certification of
Offshore Wind Turbines, Germanischer Lloyd, Reprint 2007.

Jardine, R.J. and Standing, R.J. (2000) Pile load testing performed for HSE cyclic loading study at
Dunkirk, France. Offshore Technology Report OTO 2000 007; Health and Safety Executive, London. Vol,
1, 160.p., Vol, 2, 200p.

Jardine R., Chow F., Overy R., Standing J. (2005). ICP design methods for driven piles in sands and
clays. Imperial College, London.

Jardine, R.J. and Standing, J.R. (2012) Field axial cyclic loading experiments on piles driven in sand.
Soils and Foundations. Vol. 52, No 4, pp 723-737.

Jardine, R.J., Andersen, K. and Puech, A. (2012). Cyclic loading of offshore piles: potential effects and
practical design. Keynote Paper. Proc 7th Int. Conf. on Offshore Site Investigations and Geotechnics,
SUT London, pp 59-100.

Karlsrud K and Haugen T. (1985). Behaviour of Piles in clay under cyclic axial loading-results of field
model tests. Proc. Behaviour of Offshore Structures, BOSS’85, 589–600.

Karlsrud, K., Nadim, F. and Haugen, T. (1986) Pile in clay under cyclic axial loading- Field tests and
computational modeling. Proc. 3rd International Conference on Numerical Methods in Offshore Piling,
Nantes, France. Paris, Editions Technip. Pp. 165-190.

Khemakhem M., Chenaf N., Garnier,J., Favraud C and Gaudicheau P. (2012). Development of
degradation laws for describing the cyclic lateral response of piles in clays. 7th Int. Conf. Offshore Site
Investigation and Geotechnics, SUT, London.
ISO 19902 (2007) International standard for the design of fixed steel offshore platforms, International
Standards Office, AFNOR, Saint Denis.

127
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

LeBlanc, C., Houlsby, G., T. and Byrne, B.W. (2010) Response of stiff piles in sand to long-term cyclic
lateral loading, Geotechnique 60, No. 2, 79–90.

Lin, S. and Liao, J. (1999) Permanent Strains of Piles in Sand due to Cyclic Lateral Loads. Journal of
Geotechnical and Geoenvironmental Engineering, Vol 125(9), 798-802.

Nadim, F and Dahlberg, R. (1996) Numerical modeling of cyclic pile capacity in clay. Proc. Offshore
Technology Conference, OTC paper 7994, Houston, USA.

Rimoy, S., Jardine, R.J and Standing, J.R. (2013) Displacement response to axial cycling of piles driven
in sand. Geotechnical Engineering, 116 (2): 131-146.

Peralta, P. (2010) Investigations on the behaviour of large diameter piles under long-term lateral cyclic
loading in cohesionless soil, PhD Thesis, Leibnitz University, Hannover, pp 201.

Pittos G. (2014) Contribution à la proposition de lois d’interface sol-pieu sous sollicitations cycliques.
Rapport Fugro. N°IREX LC/13/SOL/49.

Pra-Ai, S. (2013) Behaviour of soil-structure interfaces subjected to a large number of cycles. Application
to piles. Thèse de l’Université de Grenoble.

Pra-aï S., Pittos G., Boulon M. and Puech A. (2016a) Cisaillement direct cyclique sol-structure en vue du
calcul des pieux, JNGG 2016, Juillet 2016, Nancy (to be published)

Pra_aï S. and Boulon M. (2016b) Modélisation par éléments finis du comportement des pieux à
l’arrachement cyclique, JNGG 2016, Juillet 2016, Nancy (to be published)

Puech, A., Canou, J., Bernardini, C., Pecker, A., Jardine, R., and Holeyman, A. (2012) SOLCYP: a four
year JIP on the behaviour of piles under cyclic loading. Offshore Site Investigation and Geotechnics,
SUT, London.

Poulos, H.G. (1988) Cyclic Stability Diagram For Axially Loaded Piles Journal of Geotechnical
Engineering, Vol. 114, No. 8, pp. 877-895.

SOLCYP Recommendations (2016) Recommendations for planning and designing piles under cyclic
loading. To be published.

TC209 Workshop (2013) Proceedings of the TC 209 Workshop: Design for cyclic loading: piles and
other foundations, 18th ICSMGE, Paris, 4th September 2013, Ed. IREX / A. Puech (www.cfms-sols.org)

Tsuha, C.H.C, Foray, P.Y., Jardine, R.J., Yang, Z.X., Silva, M. and Rimoy, S.P. 2012. Behaviour of
displacement piles in sand under cyclic axial loading. Soils and Foundations, Vol. 52, No 3, pp 393-410.

Vardanega, P.J., Kolody, E., Pennington, S.H. and Simpson, B. (2012) Bored pile design in stiff clay I:
codes of practice. Proceedings of the Institution of Civil Engineers, Geotechnical Engineering, Vol. 165
Issue GE 4, 213–232.

Wichtmann T. (2005). Explicit accumulation model for non-cohesive soils under cyclic loading. In:
Triantafyllidis Th. (ed.). Schriftenreihe des Institutes für Grundbau und Bodenmechanik der Ruhr-
Universität Bochum, Heft 38.

128
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Execution of piles in Europe standardisation and future perspectives


C. Gilbert, Systra & Chairman CEN TC288 “Execution of special geotechnical works”

129
130
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Execution related aspects, structural codes and previous experience:


the need for calibration
Dr. Dimitrios Selemetas, Cementation Skanska, United Kingdom, Dimitrios.Selemetas@skanska.co.uk

ABSTRACT
This paper provides a brief review of some of the European practices on the structural design of piles to
Eurocodes. The information presented here is based on the available written contributions that were
submitted to the Symposium and other published literature. Comparisons are drawn between the current
implementation of Eurocode 2 in the UK on cast-in-place bored piles without permanent casing and
previous UK structural codes and executions standards making the case for the need for calibration of
structural codes in line with previous well established practice and constructability requirements.

1. INTRODUCTION
Geotechnical and structural design should all be part of one integrated approach to the design of piles.
Although significant emphasis is placed on the implementation of EN 1997 (EC7), which govers the
geotechical design of piles (i.e. pile length and typically pile diameter), the structural design of piles does
not always receive equal attention. Neverthelss, some of the structural requirements in EN 1992 (EC2)
may govern the final pile design solution and could have significant cost and constructability
implications. Despite considerable advances in sophisticated numerical analyses leading to more realistic
predictions of shear forces and bending moments in piles, some of the structural Code requirements may
render this level of sophistication impractical due to prescribed rules.

Prior to the implementation of a set of partial factors in a structural Code of Practice it is necessary to
carry out a series of calibration exercises to establish factors that strike a balance between (a) previous
well established practice and experience and (b) the state-of-the-art in terms of construction practice,
material technology and pile-soil interaction. The purpose of this paper is to provide a review of the
current practices in the structural design of piles from the the available information presented in this
Symposium and with reference to some of the recent UK experiences in the design of piles to EC2. Due
to limitations in scope and available information, this paper is mostly focused on cast-in place piles
without permanet casing.

2. EUROPEAN PRACTICE
The following review is based on the information provided on the structural design of piles within the
national reports submitted to this Symposium at the time of writing and other published work. Due to the
nature of this report and the limitations in the available time and resources, the author does not claim that
this report is exchaustive and the author wishes to apologise if some of the views of practising engineers
have not been captured as part of this appraisal.

2.1. Design Approaches to Structural Design


In attempting to capture the various methods with which practicing engineers carry out the structural
design of piles to Eurocodes across Europe, one must first recognise that different countries may follow
different Design Approaches, although Design Approach 2 (DA2) of EC7 appears to be the most popular,
with only one combination required for design verification (see clause 2.4.7.3.4 of EC7 Part 1 for a brief
explanation of all three Design Approaches). Table 1 below is an attempt to summarise the various
different Design Approaches used for the structural design of piles based on the information presented in
the National Reports to this Symposium. To allow this comparison, certain simplifying assumptions were
necessary, which are summarised in the list of notes below Table 1. In general, the Ultimate Limit State
(ULS) effect of actions (E d ) for the structural design is calculated by applying partial factors of 1.35 on
permanent actions (γ G ) and 1.50 on variable actions (γ Q ). Notable exceptions are Denmark, Finland and
Sweden, where the partial factors on actions are generally smaller than for the other countries.

131
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 1 :Summary of Design Approaches and partial factors used for the structural design of piles
Country Design Approach Combination γG γQ
Austria DA2 A1(1) + M1 + R2 1.35 1.50
Belgium DA1/1 A1 + M1 + R1 1.35 1.50
Denmark DA3(2) (A1 or A2) + M2 + R3 1.00(3) 1.50(3)
Finland DA2 A1 + M1 + R2 1.15(4) 1.50(4)
France DA2 A1 + M1 + R2 1.35 1.50
Germany DA2 A1 + M1 + R2 1.35 1.50
Hungary DA2(5) A1 + M1 + R2(5) 1.35 1.50
Ireland DA1 or DA2 or DA3(6) All combinations 1.35 1.50
Italy DA1 or DA2 A1(7) + M1 + R1 or R2 1.30 1.50
Rep. Macedonia DA2 A1 + M1 + R2 1.35 1.50
Netherlands DA3 A1 + M2 + R3 1.35 1.50
Poland DA2 A1 + M1 + R2 1.35 1.50
Russia SP 24.13330.2011 SP 24.13330.2011 NA(8) NA(8)
Spain DA2 A1 + M1 + R2 1.35 1.50
Sweden DA3 A1 + M2 + R3 1.23 1.37
UK DA1/1 A1 + M1 + R1 1.35 1.50
Notes:
1. Partial factos used in the Table above are for persistent actions(BS1)
2. For axially loaded piles in Demark a Resistance Factor Approach is actually used
3. Partial factors shown are for medium consequences (normal) class (CC2) i.e. K FI = 1.0
4. Partial factors shown are for consequence class of 2 i.e. K FI = 1.0
5. A new set of R2 factors is used to model the pile technique based on local experience
6. In Ireland, it is most common to apply DA1 or DA2 for pile design
7. Values shown are for structural permanent actions, as opposed to non structural permanent actions
8. Partial factors not available within the national report
9. Partial factors shown are for safety class 2, i.e. SK2: 𝛾𝛾𝑑𝑑 = 0.91

The Austrian practice follows Design Appropach 2 (DA2) of EC7, however Hofmann (2016) reports that
the A1 partial factors on the effects of actions are split into three design situations (BS1 to BS3), BS1
(persistent), BS2 (transient) and BS3 (accidental) to aligh with previous well establised practice by DIN
in Germany. In countries, where DA1 may be used (e.g. UK and Ireland) the designer must ensure that
the design satisfies two combinations of actions (DA1/1 and DA1/2). DA1/1 is referred to as the
Structural Limit State (STR) and DA1/2 is called the Geotechnical Limit State (GEO), but both limit
states affect the reinforcement design (typically, DA1/1 govers the area of reinforcement, while DA1/2
controls the length of steel required). In DA1/1 the actions are factored by higher partial factors (typically
1.35 on permanent actions and 1.50 on variable actions) while the components of pile resistance remain
un-factored. In comparison, in DA1/2 the actions are factored by lower partial factors (typically 1.00 on
permanent actions and 1.30 on variable actions) and the pile resistance components are factored by a set
of factors specfied by the National Annex (NA) of each country depending on the regime of testing and
the requirements for explicit verification of the serviceability limit state (SLS).

The merits of DA1 have been discussed by Simpson (2007) and in most cases it would be obvious which
combination is more critical, thus avoiding a dual calculation. However, where there are net tensile
actions, this requires special consideration. The example below shows the case of a pile subjected to net
tension under the STR limit state whereas the pile is in compression under the GEO limit state. For this
example, the minimum permanent action is G k = 1000kN and the variable uplift action is Q k = 700kN.
The permanent action is acting favourably when calculating the minimum ULS actions, hence a partial
factor of 1.0 is applied on permanent actions, as per UK NA to EC2 Part 1. The calculations below give
the derivation of the STR ULS and GEO ULS minimum actions respectively:

(DA1/1 STR-ULS): E d = 1.00 × Gk − 1.50 × Qk = 1.00 × 1000 − 1.50 × 700 = −50kN ( pile in tension)
(DA1/2 GEO-ULS): E d = 1.00 × Gk − 1.30 × Qk = 1.00 × 1000 − 1.30 × 700 = +90kN ( pile in compression)

Although, the net tensile action of -50kN, in the example above, appears insignificant, if this pile is also
subjected to high shear actions in combination with net tension, this pile may require significant amount
of shear reinforcement to comply with clause 6.2.3 of EC2 Part 1, depending on the relevant NA choices.
The design implications of these requirements are discussed in Section 3.8 of this report.

132
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2.2. Structural Capacity of Piles


The structural design of piles is often described as a grey area due to the unique nature of piles. Although
conventional structural analyses can be applied to the structural design of piles, consideration must be
given to factors unique to piling. Bearing in mind that a pile is not a column in air, it follows that for fully
embedded piles, nominal requirements for shear links, minimum number of main bars and crack control
reinforcement deserve special treatment. Nevertheless, structural codes continue to provide the reference
for the structural design of an element, which is not a column and in most cases the design is carried out
by structural engineers, who may not be familiar with the method of pile construction and the specific
constructability requirements.

From the information presented within the National Reports in this Symposium at the time of writing, it
would appear that, in general, the structural design is carried out in accordance with the principles of
EN1992 (EC2) and associated NA for reinforced concrete piles, EN1993-1-1 and EN 1993-5 (EC3) and
associated NA for steel piles, and EN 1994-1-1 (EC4) and associated NA for composite piles. Direct
reference to structural Eurocodes is made in the reports by Huybrechts et al. (2016), Møller et al. (2016),
Kinnunen et al. (2016), Burlon et al. (2016), Szepesházi and Scheuring (2016), Mandolini (2016),
Josifovski and Susinov (2016), Reinders et al. (2016), Tadeo and Estaire (2016) and Axelsson (2016).
Where there is no direct reference to the structural design of piles, Hofmann (2016), Mets and Leppik
(2016), Igoe (2016), Gwizdała and Krasiński (2016), it is assumed that the design is also carried out in
accordance with the general principles of the structural Eurocodes outlined above. Although, it would
appear that there is harmonisation, in that the structural design of piles follows the structural eurocodes, it
is the author’s experience, that the NA criteria, as set out by each country, are the key factors that govern
the structural design and these criteria may not always be in harmony. The notes below provide a
summary of some of the specific structural requirements that are presented in this Symposium.

Kinnunen et al. (2016) give a summary of the maximum allowed compressive and tensile stresses during
installation of driven piles, in accordance with the Finnish practice. Burlon et al. (2016) give supplemtary
guidelines for calculating the charactericitic concrete cylinder strength (f* ck ) and the design compressive
strength (f cd ) of cast in place reinforced concrete piles. In addition, there is specific guidance on limiting
values for the mean normal stress and the maximal normal stress in the pile section, as a function of the
pile construction technique, pile geometry and level of integrity control. Szepesházi and Scheuring
(2016) report that although the structural design in Hungary follows the principles of Eurocodes, there are
some cases where constructability issues occur due to high reinforcement density specified by ovelry
conservative designs carried out by structural engineers. Structural engineers may be knowledgeable of
the requirements of structural Eurocodes but they are not always aware of the execution standards for
special geotechnical works (e.g. EN 1536, 2015). Szepesházi and Scheuring (2016) also report that issues
with concrete quality arise due to lack of appropriate mix specifications to ensure flowability of concrete
and lack of experience with suitable concrete mix design for foundations.

Mandolini (2016) reports that piles in Italy are reinforced with a minimum reinforcement equal to 0.3% of
the nominal pile section along the entire length of the pile if the pile is cast-in-situ. The design intent is to
design piles to have a ductile response and minimum shear reinforcement of 8 mm helical is provided
along the pile portion potentially affected by plastic hinges. This portion of pile is assumed to extend to a
length equal to 3 times the pile diameter, where the minimum reinforcement is 1% of the pile area.

Josifovski and Susinov (2016) report that a simple structural check is usually carried out to confirm that
the concrete strength is not exceeded for a specific type of concrete. The National report describing the
Russian piling practice states that structural design of pile foundations is carried out in accordance with
Section 8 of SP 24.13330.2011. Tadeo and Estaire (2016) report that the structural design of piles must
satisfy the so called “structural limit”. This limit is defined as the maximum average compression stress
that can be applied to a pile at the quasi-permanent load combination. Recommended values of this
“structural limit” are given by Tadeo and Estaire (2016) ranging in value between 4.0MPa and 6.0MPa
depending on the piling technique and the level of control on site.

Axelsson (2016) reports that for steel design, the Swedish practice uses a partial coefficient of 1.0, when
the calculation is based on yield stress, see section 6.1 in EC3 Part 1-1. No partial coefficient is used for
the elastic modulus of steel but the modulus itself is defined to be 210 GPa, which would make the partial
safety coefficient equal to 1.0. The design compressive strength of concrete is calculated as per equation
3.15 of EC2 Part 1, where the partial factor for concrete (γ c ) is taken as 1.50 and the coefficient taking
account of the long term effects (α cc ) is taken as 1.0, as recommended in the main text of EC2 Part 1.

133
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3. UNITED KINGDOM PRACTICE

3.1. Introduction
EN 1992-1-1:2004 provides supplementary requirements for cast-in-place piles constructed without
permanent casing. These requirements are discussed below with reference to their design implications and
comparisons are drawn with previous UK practice.

3.2. Diameter for cast in place piles


In the absense of other provisions, clause 2.3.4.2(2) of EC2 Part 1 requires a reduction in the diameter of
cast in place piles without permanent casing for structural design calculations. Figure 1 shows that this
reduced diameter ranges between 80% and 98% of the nominal pile diameter depending on the size of the
pile. For large diameter piles, this reduction does not not have a noticeable effect on the pile bending
capacity, however this reduction could have a significant effect on the capacity of smaller diameter piles
(see section 3.4 of this report).

100
EC2 Reduced Diameter / Nominal Diameter (%)

95

90

85

80
0 500 1000 1500 2000 2500

Nominal Pile Diameter (mm)

Figure 1 :Reduction in pile diameter for cast in place piles as per EC2 Part 1 cl.2.3.4.2(2)

In the UK, BS 8002 (2015) considers that the reduction in pile diameter for structural calculations is not
necessary provided that the piles are constructed in accordance with the relevant execution standards (i.e.
BS EN 1536 or BS EN 14199). Indeed, nowadays it is common practice in the UK to implement
controlled construction processes, which are deemed to be other provisions as required by the Eurocode
Standard such that the diameter reductions in Clause 2.3.4.2 (2) of EC2 Part 1 do not need to be applied.
Details of these controls for two specific piling systems are given below.

3.2.1. Continuous Flight Auger (CFA) Piles


Where CFA pile installation is fully instrumented to control and record installation parameters, including
rate of boring volume of concrete and concreting pressure with depth these controls ensure repeatable
construction of piles with high integrity and excellent dimensional control. Additional measures include
measurements of drilling tools, checks that overbreak is positive i.e. the average diameter can be verified,
and the use of spacers in piles based on a nominal diameter. Where these provisions are in place during
construction the reduction in nominal pile diameter in clause 2.3.4.2 (2) of EC2 Part 1 is not necessary.

134
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3.2.2. Driven Cast In Situ (DCIS) Piles


Driven Cast In situ Piles are generally installed using an oversized shoe (typically 25mm greater than the
nominal pile shaft diameter to close the end of the temporary steel drive tube. The driving and concreting
of DCIS piles is increasingly instrumented. The oversize shoe creates a pile bore of greater diameter than
the drive tube, and as the concrete is effectively tremied into place by the drive tube, excellent
dimensional control is assured. These provisions repeatedly produce piles of good integrity and uniform
section. Where these provisions are in place during pile driving the reduction in nominal pile diameter
given in EC2 Part 1 clause 2.3.4.2 (2) is not necessary.

3.3. Material factor for foundations


For cast in place piles without a permanent casing, clause 2.4.2.5 of EC2 Part 1 recommends that the
partial material factor for concrete should be multiplied by an additional factor, k f , for the calculation of
design resistance. The value of k f is given in the UK National Annex as k f =1.1, resulting in a material
factor of γ c = 1.65. This is in conflict with previous UK structural codes (i.e. BS 8110) where the material
factor for concrete was γ c = 1.5. In Germany, Austria, Denmark and Italy, provided the piles conform to
the execution standards of EN 1536, the value of k f = 1.0 is used, resulting in γ c = 1.5 (Fingerloos et al.
2015). Axelsson (2016) reports that in Sweden too, the material factor for concrete is taken as γ c = 1.5.

Bearing in mind that the original intention of Eurocodes was to give designers the opportunity to derive
benefits from recent advances in concrete technology, it would seem appropriate to maintain at least the
same level of reliability with previous UK structural codes by adopting a material factor of γ c = 1.5 in the
UK. Indeed, members of the Federation of Piling Specialists (FPS) in the UK implement a comprehensive
concrete sampling regime and quality control systems, from which the cube strength of as-built piles can
be assessed (see clause B19.8.3 of ICE Specification for piling and embedded retaining walls, 2nd
edition). From a reliability perspective, it would seem more valuable to focus on the quality controls on
site, the site-specific mix design and the method of pile construction, rather than introducing an additional
factor of conservatism, which may not be sufficient to capture the effects of poor quality control on site in
the first place. The design implications of increasing the material factor for concrete from γ c = 1.5 to 1.65
are discussed in section 3.4 of this report.

3.4. Design implications


The reduction in pile diameter for structural calculations (see Section 3.2), in combination with the
increase in the material factor for concrete in piles (see Section 3.3), as specified by the UK NA of EC2,
results in a reduced moment capacity for a given pile compared with previous UK structural codes (i.e.
BS 8110). To illustrate this, Figure 2 compares the ultimate limit state (ULS) moment versus axial load
envelopes by BS8110 and the UK NA to EC2 respectively for a 450mm pile reinforced with 6B25 main
bars and B10 shear links (this reinforcement is the maximum that can be typically installed in a 450mm to
allow proper flow of concrete between bars). For this example, the concrete grade is C28/35 and the cover
to the reinforcement is 75mm. In the EC2 calculation, the diameter reduces to d = 0.95 x d nom = 428mm
and the cover to the reinforcement becomes c = ½ x [428 - (450 - 2 x 75)] = 64mm. For the pile designed
to BS 8110 the material factor for concrete is taken as γ c = 1.5, whereas γ c = 1.65 is used for the pile
designed as per UK NA of EC2 (α cc = 0.85). Figure 1 shows that the reduction in moment capacity varies
from about 10% at zero or low axial stress levels, to about 20% at higher axial loads. This reduction may
seem insignificant, but the example below demonstrates how this reduction in moment capacity could
affect the feasibility of a scheme with 450mm piles. In this example, a 450mm pile is subjected to an
eccentricity moment from an axial load applied at an eccentricity of 100mm at pile cut-off level. The un-
factored pile axial load (SLS) is 1150kN, consisting of 70% permanent actions (G k = 0.7x1150 = 805kN)
and 30% variable actions (Q k = 0.3x1150 = 345kN). The calculations below show how the ultimate limit
state (ULS) structural actions are derived in terms of design axial load (N d ) and design moment (M d ):

BS 8110: N d = 1.40 × DL + 1.60 × LL = 1.40 × 805 + 1.60 × 345 = 1679 kN


M d = N d × ecc = 1679 × 0.100 = 167.9 kNm

EC2 (DA1 ULS-1): N d = 1.35 × Gk + 1.50 × Qk = 1.35 × 805 + 1.50 × 345 = 1604 kN
M d = N d × ecc = 1604 × 0.100 = 160.4 kNm

135
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

The two ULS cases are plotted in Figure 2 with reference to the ULS moment versus axial load (M-N)
envelopes predicted by BS 8110 and the UK NA to EC2 respectively. The BS 8110 ULS moment plots
within the capacity envelope of the BS8110 curve, however the EC2 ULS moment, although reduced due
to the reduced partial factors (1.35 and 1.50 vs. 1.40 and 1.60 respectively) plots outside the capacity
envelope of the UK NA to EC2 curve. This means that the same pile that would be sufficient to resist this
applied moment in a previous code would need to be replaced by a pile with a larger diameter to resist the
same serviceability action. Figure 3 below, shows that there is excellent agreement in pile bending
capacity between BS 8110 and EC2, provided there is no reduction in pile diameter (i.e. d = d nom ) and the
material factor for concrete is taken as γ c = 1.5 (i.e. k f = 1.0).

450mm Pile - C28/35 Concrete - 6B25 main bars with B10 shear links
4000
BS8110_450mm_gamma = 1.50
EC2_428mm_gamma = 1.65
BS Design
3000 EC2 Design

2000
168, 1679
ULS Axial Load (kN)

160, 1604

1000

0
-250 -200 -150 -100 -50 0 50 100 150 200 250

-1000

-2000
ULS Bending Moment (kNm)

Figure 2 : Comparison of pile bending capacity between BS 8110 and EC2 (UK NA)

450mm Pile - C28/35 Concrete - 6B25 main bars with B10 shear links
4000

BS8110_450mm_gamma = 1.50

3000 EC2_450mm_gamma = 1.50

2000
ULS Axial Load (kN)

1000

0
-250 -200 -150 -100 -50 0 50 100 150 200 250

-1000

-2000
ULS Bending Moment (kNm)

Figure 3 : Comparison of pile bending capacity between BS 8110 and EC2 with d =d nom and γ c = 1.5

136
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3.5. Maximum concrete stress in piles


Clause 12.3.1(1) of EN 1992-1-1 specifies that the design stress of plain concrete (unreinforced pile) in
compression (f cc,pl,d ) should be limited as follows:

α cc , pl × f ck
f cc , pl , d = (1)
k f ×γ c

where α cc,pl is a factor taking into account the long-term reduction in strength of plain concrete
f ck is the characteristic compressive cylinder strength of concrete at 28 days
kf is an additional factor for concrete for cast-in-place piles without permanent casing
γc is the partial factor for the strength of concrete

The UK NA recommended values for the above equation are a cc,pl = 0.6, k f = 1.1 and γ c = 1.5. Substituing
the relevant values, gives a maximum stress limit for plain concrete of 0.36f ck = 0.29f cu (f ck = 0.8f cu ) at
Ultimate Limit State (ULS) stress. Assuming that permanent and variable actions are typically 70% and
30% of the total load respectively, the factor (γ ULS ) converting SLS to ULS stress can be approximated as
γ ULS = 1.35 x 0.7 + 1.50 x 0.3 = 1.40. For plain concrete, the SLS stress limit can then be approximated as
0.29fcu/1.40 = 0.21f cu . This is in conflict with previous well established practice in the UK, where the
concrete stress in piles is typically limited to 0.25f cu at safe working load (see clause 7.4.4.3.1 of BS
8004:1986).

If the stress limit of 0.21f cu was to be applied, this would significantly reduce the design compression
capacity of pile sections without reinforcement and would result in more coservative designs. The same
amount of reliability as per previous UK practice can be maintained by applying a factor of a cc,pl = 0.65 in
equation (1) above, in combination with γ c = 1.5 (i.e. k f = 1.0) and no reduction in the nominal pile
diameter, resuling in a stress limit of 0.25f cu at SLS stress. In comparison, the German NA to EC2
recommends α cc,pl = 0.70, γ c = 1.5 (i.e. k f = 1.0) and no reduction in the nominal pile diameter for
structural design calculations. This results in a similar serviceability stress limit of 0.266f cu .

Typically, piles would be reinforced at the head of the pile and over the section where the pile is
reinforced, the coefficient on the compressive strength of concrete would be a cc = 0.85 (see clause
3.1.6(1) of EC2 Part 1 and UK NA). However, most CFA piles are not reinforced full-length and typically
these piles would have a 6m long cage installed from piling platform, which may leave only 3-4m of
reinforcement embedded into the pile below final trim level. If the ground over the top 3-4m of the pile,
does not contribute significantly in reducing the axial load in the pile through shaft friction, the designer
may find that the stress in the un-reinforced section of the pile is not very dissimilar to the stress over the
reinforced section of the pile, in which case the current UK NA stress limit of 0.21f cu would result in a
more conservative designs.

3.6. Minimum reinforcement in piles


Clause 9.8.5(3) of EC2 Part 1 recommends that bored piles should be reinforced with a minimum of 6
longitudinal bars, where piles are subjected to bending or shear. The Standard also recommends that the
minimum bar size should be 16mm and the clear distance between bars should not exceed 200mm
measured along the periphery of the pile. These recommendations are in conflict with the
recommendations of BS EN 1536:2015 and previous UK practice. Clause 7.5.2.3 of BS EN 1536:2015
specifies that for reinforced piles the minimum longitudinal reinforcement shall be four bars of 12 mm
diameter. In addition, clause 7.5.2.5 of BS EN 1536:2015 specifies that the spacing of longitudinal bars
should always be maximized in order to allow proper flow of concrete but should not exceed 400 mm (as
opposed to 200mm in clause 9.8.5(3) of EC2 Part 1). Best industry practice in the UK follows the
recommendations of BS EN 1536:2015. Provided the maximum size of the aggregates does not exceed 20
mm, the clear space between longitudinal bars or bundles of bars is not less than 100mm, but may be
reduced to 80 mm at the lap locations. In the UK, where there is no design shear or bending moment, the
minimum reinforcement provided is in accordance with BS EN 1536:2015. If there is a design shear or
bending moment, EC2 Part 1, clause 9.8.5(3) is used for guidance, but the designer can specifically
design for the minimum resistance offered in any direction by the proposed cage (which may comprise
only 4 or 5 main bars) for an economic design.

137
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3.7. Lap lengths


Equation 8.3 of EC2 Part 1 clause 8.4.3(2) defines the basic required anchorage length (l b,rqd ) as:

σ sd ×ϕ (2)
lb , rqd =
4 × f bd
where: σ sd is the design stress of the bar at the anchorage position (N/mm2)
φ is the bar diameter (mm)
f bd is the ultimate bond stress in the bar (N/mm2) given in equation 8.2 of EC2 Part 1 as:

f bd = 2.25η1 η2 f ctd (3)

where: η 1 is a coefficient related to the quality of bond (1.0 for good bond, 0.7 for poor bond)
η2 is a function of the pile diameter (1.0 for φ ≤ 40mm, as per UK NA to EC2 Part 1)
f ctd is the design value of concrete tensile strength related to the concrete grade (N/mm2)

The design anchorage length is then calculated using equation 8.4 of EC2 Part 1 as follows:

lbd = α 1 α 2 α 3 α 4α 5 lb , rqd (4)

where the coefficients α 1 to α 5 relate to the shape of bars, concrete cover, and confinement (see Table 8.2
of EC2 Part 1). For most bored piles reinforced with straight bars, for which there is no confinement by
traversese welded reinforcement or transverse pressure, the values of coefficients α 1 and α 3 to α 5 are
equal to unity. For convenience and conservatism, it is common to assume that α 2 , the coefficient related
to the design cover to the reinforcement is also equal to unity. However, Figure 4 shows the actual
variation of α 2 , as a fucntion of design cover (c d ) and main bar diameter, as per Table 8.2 of EC2 Part 1.
The design cover, c d , is defined as either the actual cover to the main bar (typically 75mm + link
diameter) or half the clear bar spacing, whichever is the smallest. Typically, the clear bar spacing would
be greater than 100mm, (or 80mm at laps) and this will result in reduced anchorage and lap lengths.

1.00

0.95

0.90

0.85
Bar Diameter
α2 Coefficient

40mm
0.80
32mm
25mm
0.75
20mm

0.70

0.65

0.60
50 60 70 80 90 100 110 120 130 140 150
Design Cover, cd (mm)

Figure 4 : Coeffcient α 2 as a function of design cover and main bar diameter

The design lap length can be calculated from the anchorage length using equation 8.10 of EC2 Part 1 as:

lbd = α 1 α 2 α 3 α 5α 6 lb , rqd (5)

138
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

The coefficient α 6 , relates to the percentage of reinforcement lapped at the same location relative to the
total reinforcement area. The value of α 6 varies from 1.0 to 1.5 and it would typically be equal to 1.5 if
more than 50% of the main bars are lapped at the same section, as per usual piling practice. Substituting
the relevant values for ‘good’ bond conditions and for bars less than 50mm (the UK NA to EC2 Part 1
specifies that large diameter bars are bars with diameter greater than 40mm, clause 8.8(1) of Table NA.1)
gives the following lap lengths for C28/35 piles (Figure 5) and C32/40 piles at different stress levels. The
lap lenghts below are based on a clear bar spacing of 110mm and γ c = 1.65 (as per UK NA to EC2).

80
Lap Lenghts to EC2 for C28/35 Pile

70
Lap Length (number of main bar diameteters)

% Stress in bar
60
60
57
100%
52
50
47 66%
45

40
40 38
35
31
30
30

20
10 15 20 25 30 35 40 45 50
Bar Diameter (mm)

Figure 5: Lap lengths to EC2 for C28/35 pile (clear bar spacing =110mm and γ c = 1.65)

80
Lap Lenghts to EC2 for C32/40 Pile

70
Lap Length (number of main bar diameteters)

% Stress in bar
60
100%
54
51
50
47 75%
43
40 41
40 39
36
32
30
30

20
10 15 20 25 30 35 40 45 50
Bar Diameter (mm)

Figure 6: Lap lengths to EC2 for C32/40 pile (clear bar spacing =110mm and γ c = 1.65)

Figures 5 and 6 show that EC2 lap lenghts under ‘good’ bond conditions may vary between 40 times to
60 times the main bar diameter if the bars are fully stressed at the lalp location. This is in conflict with
previous piling practice in the UK, where typically lap lengths equal to 40 x main bar diameter would be

139
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

sufficient. From a design perspective, laps are usually avoided at locations of high stress (i.e. maximum
bending moment) and provided that the stress in the bar is sufficient low at the lap location, EC2 lap
lengths equal to 40 x main bar diameter may be used. For example, Figures 5 and 6 show that provided
the stress in the bar is less than ⅔ of the ultimate stress in a C28/35 pile, and less than ¾ of the ultimate
stress in a C32/40 pile, respectively, a lap length of 40 x main bar diameter would be sufficient.
Nevertheless, it is becoming increasingly more and more common for consulting engineers to specify lap
lengths well in excess of 40 x main bar diameter assuming full stress conditions and the most onerous
conditions at the lap locations (e.g. α 2 = 1.0).

Excessive lap lengths pose constructability and health and safety risks and in some cases may necessitate
the use of full-tension couplers, as an alternative, which in turn are costly and may require special
measures in place due to manual handling. Therefore, it is important that engineering specifications are
based on the minimum lap lengths that are required based on the specific conditions that apply, rather
than the most conservative estimates that are conceivable within the Code requirements. It is not surpising
that during the transition period between ‘old’ and ‘new’ Codes, engineers are more inclined to yield in
the direction of least doubt, i.e. more conservatism, but it is precicely at this period where engineers
should not doubt the voice of experience.

EC2 gives specific guidance on what constitutes ‘good’ bond and ‘poor’ bond conditions (see Figure 8.2
of EC2 Part 1), from which it is clear that ‘good’ bond conditions exist for most piling applications,
where typically the cover to the main bar is greater than than 83mm. However, there seems to be no
guidance in EC2 on the bond conditions that exist in piles cast under support fluids such as bentonite.
Jones and Holt (2004) concluded that for vertical deformed bars cast into concrete under bentonite,
‘good’ bond conditions exist provided that the cover to the main bar is at least twice the main bar
diameter, which is the case for most piles. This was also acknowledged in a recent Civil Engineering
Standard of a major infrastructure project in the UK, where a 40% increase in the lap length was only
specified for horizontal bars. Due to a widespread confusion in this area and the lack of well-document
case studies, it is not uncommon for consulting engineers to specify by default a 40% increase in the lap
length of all bars in piles cast under bentonite, i.e. considering ‘poor’ bond conditions due to the use of
bentonite. In some extreme cases, engineers have specified an additional factor of 1.4 in combination with
‘poor’ bond conditions resulting in specified lap lengths in excess of 100 x bar diameter, which are
simply impractical to construct. Where full tension couplers have been used, as an alternative, this has
increased considerably construction costs, thus eliminating any economic advantages that may have been
anticipated at earlier stages with the use of Eurocodes. To this end, it would be beneficial, if future
revisions of Eurocodes make specific reference to the assessment of bond characteristics for piles cast
under support fluid (i.e. bentonite), with specific reference to case studies and previous experience.

3.8. Shear reinforcement design for piles


EC2 uses a strut and tie model for the design of reinforced concrete sections for shear. For members that
require shear reinforcement, equation 6.8 of EC2 Part 1 gives the shear capacity of a reinforced concrete
member as follows:
Asw
VRd ,s = z f ywd cot θ (6)
s
where A sw is the cross-sectional area of the shear reinforcement
s is the spacing of the circular shear links
V Rd,s is the design value of the shear resistance provided by the reinforcement alone
z is the level arm between compression and tension cords
f ywd is the design yield strength of the shear reinforcement
θ is the angle between compression strut and the axis perpendicular to the shear force

The EC2 strut and tie model is different to previous UK structural codes (i.e. BS 8110) in that shear
resistance is provided by shear reinforcement alone, ignoring any concrete contrbution, for members that
require shear reinforcement. When the member is in compression, EC2 allows the designer to vary the
angle of the compression chord such that cotθ is within the range 1.0 ≤ cotθ ≤ 2.5 (provided that V Rd,max is
not exceeded, where V Rd,max is the design value of the maximum shear force which can be sustained by
the member, without crushing the compression strut). However, when shear co-exists with tension, the
UK NA to EC2 specifies that cotθ = 1.25, which implies an increase in the amount of shear reinforcement
by a factor of 2.0 irrespective of the amount of tension in the section. In comparison, in the BS8110
model for shear, the angle of the compression chord is effectively fixed, but there is a mechanism to

140
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

capture the effects of axial force through the contributions of the concrete capacity, which is linked to
axial load as per clause 3.4.5.12 of BS 8110.

In order to illustrate the design implications of fixing the value of cotθ to 1.25, irrespective of the amount
of tension, Figure 7 shows comparisons of the shear link spacing required by BS8110 and EC2 for a
600mm pile subjected to various combinations of shear and tension. As the tension increases from -50kN
to -400kN, BS8110 predicts a progressive reduction in the spacing of links to capture the unfavourable
effects of tension, while EC2 predicts the same link spacing irrespective of the amount of tension. This
clearly has cost implications for piles which are subjected to nominal or very small amounts of tension
(see example in section 2.1 of this paper), as these piles would require twice the amount of shear
reinforcement in comparison with previous practice.

Effect of tension on shear reinforcement for 600mm pile resisting ULS shear of 150kN
300
280

280

280

280
BS8110

250
EC2
250

225

200
Spacing of B10 Links (mm)

200

175
150

150

150

150

150

150

150

150
150

100

50

0
-100

-150

-200

-250

-300

-350

-400
-50

Axial Force (kN)

Figure 7 : Comparison between EC2 and BS 8110 shear provisions for a 600mm pile subjected to tension

In the UK, the conventional design practice for shear reinforcement of piles follows the earlier work of
Clarke and Birjandi (1993) and the seminal work of Feltham (2004). More recently, Orr et al. (2010) put
forward a formulation for the shear design of circular sections using the Eurocode 2 truss model.
Traditionally, the effective depth (d eff ) for shear calculations, was taken as the depth to the centroid of the
steel below the centre line of the section (for simplicity the neutral axis was considered to be coincident
with the centroidal axis of the pile). This geometric definition, which was well established due it practical
appeal, is different to the sectional method proposed by Orr et al. (2010), which requires an iterative
analysis for determining the neutral axis along the length of the pile, and as a result does not have the
same practical merit. The paper by Orr et al. (2010) proposes two efficiency factors, λ 1 for circular
sections and λ 2 for helical links, yet the reduction factor λ 2 is not in agreement with the reduction factor
proposed by Feltham (2004) for helical links. Preliminary calculations using the method by Orr et al.
(2010) show that this method gives more conservative predictions for the amount of shear reinforcement
required in piles, compared with previous UK practice (see theoretical predictions compared with
experimental data in Figures 13 and 14 by Orr et al., 2010).

The key limitation, in the current implementation of the EC2 shear calculation model, as per UK NA
guidelines, is that cotθ = 1.25 irrespective of the amount of tension in the section. In comparison, in the
CEB-FIP Model Code 1990, the inclination of the compression cord (θ) is expressed as a function of the
longitudinal strain in the section, giving a more gradual reduction in the value of cotθ as a function of the
tension in the section, as shown in Figure 8. Similarly, Fingerloos et al. (2015) report that in the NA of

141
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

other countries, such as Germany and Greece, the cotθ value is related to the axial stress in the pile, the
applied shear force and the shear capacity of the concrete section and in addition the allowable range of
cotθ is not the same in the NA of each country. Bearing in mind that a reinforced concrete pile with the
same mix design, should not be as variable as the natural ground, it would be beneficial if there was a
harmonisation in the NA recommendations regarding the variation of cotθ with axial strain, such that
countries can benefit from formulations that are safe, economic and well established in other countries.

Compression Tension
cot ϑ UK NA EC2
fib MC2010
2.5

1.0

Axial Strain

Figure 8 : Comparison between EC2 and fib MC2010 variation of cotθ with axial strain

3.9. CRACK WIDTH CALCULATIONS


Guidance on crack control in reinforced concrete structures is given in clause 7.3 of EC2 Part 1, where
recommended values of maximum crack width (w max ) are specified for different exposure classes, under
quasi-permanent load combinations. For most foundations and for the majority of exposure classes, the
maximum limit of crack width under quasi-permanent load combinations is 0.3mm. However, Schiessl
(1975) and subsequentlly Arya and Ofori-Dark (1996) and TR44 (2015) showed that there is no universal
relationship between crack width and the rate of corrosion. Schiessl (1975) carried out exposure tests on
cracked reinforced concrete beams and recorded the depth of corrosion associated with a certain crack
width. Figure 6 (after Schiessl, 1975) shows that the distributions of the measured corrosion depths vary
exponentially and that the distributions are independent of crack width.

For corrosion to occur, carbon dioxide and/or chlorides must be present at the anode, oxygen must be
available at the cathode and the concrete must be sufficiently moist. In addition to these environmental
conditions, the risk of corrosion is a function of various other factors including crack orientation relative
to the reinforcement axis, crack geometry and frequency, thickness of concrete cover, cathode-to-anode
area ratio, quality of concrete and moisture content (TR44, 2015). For example, Arya and Ofori-Dark
(1996), showed that a member with a larger number of narrow cracks will experience more corrosion than
a similar member with fewer wider cracks but the same overall combined surface crack width. This is
because the presence of multiple cracks will shorten the path between anodic and cathodic regions on the
surface of the reinforcing bar. Therefore, it is important that the designer considers the combination of all
these influencing factors rather than simply relying on a crack width value to assess the risk of corrosion.

142
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 9 : Probability of corrosion vs. depth of corrosion for different crack widths (after Schiessl, 1975).

Where crack control is required, the designer may follow the method without direct calculation given in
clause 7.3.3 of EC2 Part 1 by limiting the bar spacing or bar diameter for a given steel stress level.
Alternatively, the designer may use equation 7.8 of EC2 Part 1 (clause 7.3.4) to calculate the crack width
under quasi-permanent actions. Specific problems associated with calculations of crack widths to EC2 for
circular sections have been reported by Kaethner (2011). Apart from the specific modelling difficulties
when dealing with non-rectangular sections, unexpected results have also been reported for rectangular
sections with two rows of bars or large covers. Kaethner (2011) presents examples showing that EC2
predictions of crack width are consistently higher than previous UK structural codes (i.e. BS 8110),
making the case for further guidance and clarity on the EC2 model of crack width calculation. PD 6687-
1:2010 is a welcoming step towards this direction, introducing an alternative calculation model for non-
rectangular sections and irregular bar layouts.

4. CONCLUDING REMARKS
Pile design requires a holistic approach integrating both geotechnical and structural elements in line with
constructability requirements. Significant advances in the area of geotechnical modelling and soil-
structure interaction, may be out-weighed by structural requirements, in the absence of calibration and
verification of partial factors that align new codes with previous practice. This paper gave a brief review
of some of the European practices on the structural design of piles to Eurocodes and focused mainly on
bored piles cast-in place without permanent casing. Projects with partial or full re-use of foundations may
be compromised if the future evolution of structural Eurocodes requires additional reinforcement for piles
which have been succesfully constructed in the past with less steel. From a constructability point of view,
the designer must recognise that excessive amount of reinforcement may put at risk the quality of the end

143
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

product and may have health and safety implications (see Photo 1). This review highlighted several areas
where the current implementation of EC2 is in variance with previous piling practice in the UK and the
following recommendations were put forward:

1. In the UK, the reduction in pile diameter for structural calculations for piles cast-in place without
permanent casing is not necessary provided that the piles are constructed in accordance with the
relevant execution standards.
2. The reduction in pile diameter for piles cast without permanent casing, in combination with the
increase in the material factor for concrete, as specified by the UK NA of EC2, results in a reduced
ultimate moment capacity for a given pile compared with previous UK structural codes.
3. Design lap lengths to EC2 well in excess of 40 x bar diameter are not always required, provided that
good bond conditions exist and the stress in the bar at the lap location is adequately limited.
4. The current implementation of the UN NA EC2 shear model, for piles subjected to shear in
combination with tension, requires that cotθ = 1.25 irrespective of the amount of tension in the
section. This requirement results in additional shear reinforcement where piles are subjected to shear
and nominal tension compared with previous UK structural codes. A more graduation reduction of
cotθ with applied tension would be beneficial, similar to CEB-FIP Model Code 1990 and the German
NA to EC2 .
5. A pragmatic approach is required when assessing crack widths in piles. The amount of corrosion in
reinforced concrete sections is known to be independent of crack width, yet structural codes rely
mostly on crack width to control the risk of corrosion. The amount of corrosion is a function of many
factors including the environmental conditions, the frequency and orientation of cracks and the quality
of concrete (see Section 3.9 for more details). A designer concerned about a crack width of 0.3mm
may find more reasons to be concerned if heavily congested steel, designed to control crack widths,
creates large voids in the concrete due to insufficient flow of concrete between bars.

Photo 1: High density of reinforcement in a pile with heavy-duty couplers requiring manual handling

144
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

REFERENCES
Arya, C., and Ofori-Darko, F. K. (1996). ‘‘Influence of crack frequency on reinforcement corrosion in
concrete.’’ Cement and Concrete Res., 26(3), 345–353.

Axelsson, G. (2016) . Swedish Practice, Design of Piles in Europe How did Eurocode 7 change daily
practice? International Symposium, 28 and 29 April 2016, Leuven, Belgium.

BS 8004 (2015): Code of practice for foundations, BSI.

BS EN 1536 (2015): Execution of special geotechnical works. Bored piles, BSI.

BS EN 12699 (2015): Execution of special geotechnical works. Displacement piles, BSI.

BS EN 14199 (2005): Execution of special geotechnical works. Micropiles, BSI.

BS EN 1992-1-1 (2004): Eurocode 2: Design of concrete structures. Part 1: General Rules and Rules for
Buildings, BSI.

BS EN 1997-1 (2004): Eurocode 7: Geotechnical design. Part 1: General Rules, BSI.

BS 8110-1 (1997): Structural Use of Concrete - Part 1: Code of Practice for design and construction,
BSI.

Burlon, S., Szymkiewicz, F., Le Kouby, A. and Volcke, J. P., (2016) . French Practice, Design of Piles in
Europe How did Eurocode 7 change daily practice? International Symposium, 28 and 29 April 2016,
Leuven, Belgium.

Clarke, J., Birjandi, F. (1993). The behaviour of reinforced concrete columns in shear, The Structural
Engineer, 71 (5), pp 73-81.

Feltham, I. (2004). Shear in reinforced concrete piles and circular columns, The Structural Engineer, 84
(11), pp 27-31.

Fingerloos, F., Finckh, W., Hochreither, H., Ignatiadis, A., Jenisch, F. M., Landgraf, M., Prietz, F.,
Schmitt, W., Schuster, K. and Schwabach, E. (2015). Verbesserung der Praxistauglichkeit der Baunormen
durch pränormative Arbeit Teilantrag 2: Betonbau, Fraunhofer IRB Verla.

Gwizdała, K. and Krasiński, A. (2016). Polish Practice, Design of Piles in Europe How did Eurocode 7
change daily practice? International Symposium, 28 and 29 April 2016, Leuven, Belgium.

Hofmann, R. (2016) . Austrian Practice, Design of Piles in Europe How did Eurocode 7 change daily
practice? International Symposium, 28 and 29 April 2016, Leuven, Belgium.

Huybrechts, N., De Vos, M., Bottiau, M., and Maertens, L. (2016) . Belgian Practice, Design of Piles in
Europe How did Eurocode 7 change daily practice? International Symposium, 28 and 29 April 2016,
Leuven, Belgium.

Igoe, D. J. P. (2016) . Irish Practice, Design of Piles in Europe How did Eurocode 7 change daily
practice? International Symposium, 28 and 29 April 2016, Leuven, Belgium.

Josifovski, J. and Susinov, B. (2016) . Republic of Macedonia Practice, Design of Piles in Europe How
did Eurocode 7 change daily practice? International Symposium, 28 and 29 April 2016, Leuven, Belgium.

Kaethner, S. (2011). Have EC2 cracking rules advanced the mystical art of crack width prediction? The
Structural Engineer 89 (19), pp 14-22.

Kinnunen, J., Riihimäki, T., Tolla, P., Uotinen, V. M. And Länsivaara, T. (2016) . Finish Practice, Design of
Piles in Europe How did Eurocode 7 change daily practice? International Symposium, 28 and 29 April
2016, Leuven, Belgium.

145
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Mandolini, A. (2016) . Italian Practice, Design of Piles in Europe How did Eurocode 7 change daily
practice? International Symposium, 28 and 29 April 2016, Leuven, Belgium.

Møller, O., Kærgaard, J., Augustesen., A. H., Okkels., N. Sørensen , K. K. (2016) . Danish Practice, Design of
Piles in Europe How did Eurocode 7 change daily practice? International Symposium, 28 and 29 April
2016, Leuven, Belgium.

Mets, M. and Leppik, V. (2016) . Estonian Practice, Design of Piles in Europe How did Eurocode 7
change daily practice? International Symposium, 28 and 29 April 2016, Leuven, Belgium.

NA to BS EN 1992-1-1 (2004): Eurocode 2: Design of concrete structures. Part 1: General Rules and
Rules for Buildings, BSI.

NA to BS EN 1997-1 (2004): Eurocode 7: Geotechnical design. Part 1: General Rules, BSI.

Orr, J.K, Darby, A.P, Ibell, T.J., Denton, S.R., and Shave, J. D. (2010). Shear design of circular
concrete sections using the Eurocode 2 truss model, The Structural Engineer, 88 (23/24), pp 26-33.

PD 6687-1(2010). Background paper to the National Annexes to BS EN 1992-1 and BS EN 1992-3, BSI.

Reinders, K., Van Seters, A. and Korff, M. (2016) . Dutch Practice, Design of Piles in Europe How did
Eurocode 7 change daily practice? International Symposium, 28 and 29 April 2016, Leuven, Belgium.

Schiessl, P., "Admissible Crack Width in Reinforced Concrete Structures," Preliminary Report, V. 2 of
IABSE-FIPCEB-RILEM-IASS, Colloquium, Liege, June 1975

Simpson, B. (2007). Approaches to ULS design – The merits of Design Approach 1 in Eurocode 7.
ISGSR2007 First International Symposium on Geotechnical Safety & Risk, Shanghai, China.

Specification for piling and embedded retaining walls (2007), 2nd edition, ICE.

Szepesházi, R. and Scheuring, F. (2016). Hungarian Practice, Design of Piles in Europe How did
Eurocode 7 change daily practice? International Symposium, 28 and 29 April 2016, Leuven, Belgium.

Tadeo C. F. and Estaire, J. (2016) . Spanish Practice, Design of Piles in Europe How did Eurocode 7
change daily practice? International Symposium, 28 and 29 April 2016, Leuven, Belgium.

TR44(2015). The relevance of cracking in concrete to corrosion of reinforcement - 2nd Edition, The
Concrete Centre.

146
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Concrete for Deep Foundations


Karsten Beckhaus, BAUER Spezialtiefbau GmbH, Germany, Karsten.Beckhaus@bauer.de
EFFC/DFI Concrete Task Group, incl. 16 delegates from Specialist Contractors, Consultants and Suppliers [1]

ABSTRACT
Concrete for Deep Foundations can only be specified, designed, produced, supplied and placed
“correctly” – in the meaning of achieving high quality structural members – if all stake holders
collaborate. The client, the designer, the contractor and the concrete supplier shall have a common
understanding of the capability of modern concrete for deep foundations, of the execution tolerances and
of the need for structural limitations in order to let the concrete flow through all openings to embed the
reinforcement and completely fill the excavation. These principals have not changed over the last decades
but the concrete and the design have. A joint Concrete Task Group was set up by EFFC 1 and DFI 2to look
at this issue and carry out a detailed assessment of current best practice and research. The new “Best
Practice Guide to Tremie Concrete for Deep Foundations” deals particularly with aspects of concrete
rheology, its dependence on raw materials and composition and applicable testing methods, and also
provides recommendations on specific design and execution details. It is intended as a practical addition
to existing standards in Europe and the US. The Concrete Task Group is also steering a joint EFFC/DFI
research program, including desk studies, laboratory and on-site testing in Europe and the United States
which will allow definitive concrete acceptance criteria to be presented in a second edition and could be
used for adoption in a new generation of related standards.

1. GENERAL
In addition to strength and durability requirements, both of which have become more stringent within the
last decades, concrete must be suitable for being cast in deep or extremely deep excavations. Today’s
specialist foundation engineering equipment makes it possible to drill or cut down to in excess of 100 m
depth, see figure 1.

Figure 1: Exposed secant pile wall and execution of bored piles and their dimensions, schematically [2]

In deep foundations, the excavation is usually „wet“, i.e. filled partly or fully with water or supporting
slurry, and concrete must be placed with the tremie pipe, see figure 2. The rising concrete must displace
the fluid upwards. If casing is used, in both the dry or wet process, the concrete must still be sufficiently
workable at the time of subsequent extraction of the temporary casings.

1
European Federation of Foundation Contractors – effc.org
2
Deep Foundation Institute – dfi.org

147
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 2: Tremie Concrete (TC) placement; excavation supported by slurry or casing [2]

It is obvious that tremie concrete for such special geotechnical works needs special properties in its fresh
stage (during the casting process) to completely fill the excavation and embed any reinforcement or box-
out, to allow self-levelling and self-consolidation, all without undergoing excessive segregation, bleeding,
or filtration –of water or cement paste to the surrounding soil– until the concrete has been finally placed
and starts to set.

Considering the concrete flow and in particular the resistance to flow, i.e. the reinforcement and the
excavation surface inside and outside of the concrete cover zone, it should also be self-evident, that the
structural design of deep piles or diaphragm walls needs to consider the constraints originating from usual
and inevitable execution tolerances and imperfections, and of the limitations of concrete workability, the
latter which depend on the complex rheology of modern concrete.

2. REGULATIONS IN EUROPEAN NORMS


Eurocodes EC2 and EC7 comprise general rules for geotechnical design and for design on concrete
structures, but no specific regulations for concrete in terms of its placement. Both assume concrete or
reinforced concrete elements of adequate quality and integrity, i.e. with limited imperfections, and EC2
sets requirements only for durability after execution, e.g. minimum concrete cover to protect the
reinforcement. For further regulations EC2 refers to the concrete standard EN 206.

EN 206 comprises general regulations for the composition and quality of fresh concrete, and, in Annex D
of its recent Edition EN 206:2013, particular rules for deep foundation concrete used in bored piles and
diaphragm walls, taken from EN 1536:2010 and EN 1538:2010. Besides the demand of a principal
suitability of fresh concrete, and besides the requested measure for consistence, however, there are
additional important regulations; on the raw materials, and on the composition of concrete, like minimum
cement and minimum fines content, or maximum water-cement ratio.

Specific restrictions for deep foundations for composition and consistence, according to Annex D, depend
on the placement procedure (see Table 1 for bored piles). It has been understood that maximum
consistency has been standardized on the basis of experiences at the time, to reduce the risk of
segregation and associated negative effects. It seems questionable, if such an upper limit disregards the
strong need for higher values corresponding to a sufficient flowability, e.g. in order to flow freely around
and through the reinforcement.

Accordingly, in terms of pile or wall integrity and quality, it is also important to consider requirements for
concrete cover and clearance between reinforcement bars. In this context, the maximum coarse aggregate
size is correctly regulated in the European standards, for deep foundations to a maximum of ¼ of the clear
space between bars, in order to ease the flow through and around reinforcement.

148
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 1: Requirements for Bored Piles, in accordance with Annex D of EN 206:2013


Placement Cement content Water-cement ratio Slump Flow diameter
condition [kg/m³] [-] [mm] [mm]
Dry ≥ 325 ≤ 0.60 150 ± 30 500 ± 30
submerged under and in compliance
with provisions 180 ± 30 560 ± 30
water,
≥ 375 valid for specified
under a stabilizing
exposure classes 200 ± 30 600 ± 30
fluid

In total, with regulations on consistency, concrete composition, cover and clear space, an integral control
system has been established to achieve high quality end products. This has been done empirically, by
experiences which are now old. In other words, for the time being there are fixed values, but these could
be insufficient in order to fulfil the basic principles for concrete like its need to flow properly. Such
conflicts can easily be overseen from parties involved if not experts in deep foundation, and in concrete
technology at the same time.

3. MODERN CONCRETE TECHNOLOGY


Modern design of structures does reflect the possibilities of making economic use of building materials. It
has become very common for the industry to supply or use concrete with low water-cement ratios,
nevertheless highly flowable and set-retarded for several hours if needed. Dense reinforcement in high-
load substructures may already demand special requirements on the passing ability of fresh concrete. It
usually demands well graded aggregates and a stable paste between to avoid blocking. Stable paste is
mainly driven by a low water-cement ratio which is typical also for concrete of higher strength classes or
for specific durability classes, i.e. for a very low porosity which is required for a long service life in
aggressive conditions, e.g. by chlorides.

Today’s concretes consist of comparatively low water contents and low water-cement ratios, often below
0.45. Therefore, modern mix designs clearly deviate from those of “classic concrete” which could also be
named “water type concrete”.

Even if fly ash or other additions can partly advantageously replace cement, such concretes need the
assistance of chemical admixtures, often as a cocktail of superplasticizers and retarders, sometimes even
topped by stabilizers. It is clear that such modern concretes behave differently from the classic types used
15 or even 30 years ago, when today’s standards and regulations were established.

Figure 3: Modern five-component substitutes the old three-component concrete technology

149
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

4. NEED FOR PERFORMANCE BASED QA/QC


As long as the concrete design was rather one-dimensional, by using only the three components of
cement, water and aggregates, and changing its strength class or durability mainly by adapting the water-
cement ratio, the concrete workability was sufficiently described by its consistence as a reliable indicator,
usually measured by slump or flow table test according to EN 12350-2 and -5. However, an individual
practical test for consistence cannot fully describe today’s fresh concrete workability and stability
features, simply because the available tests give only one integral value for the concrete’s rheology but
cannot differentiate between the key rheological parameters [1].

With modern, five-component concrete the options to individually design the concrete for specific
properties are somehow more-dimensional, i.e. the characterisation of fresh concrete has become more
complex and needs at least two rheological parameters to be sufficiently described. One is viscosity and
the other is yield stress. Figures 4 and 5 illustrate the characterisation of fresh concrete by its generic
rheology and should make obvious that the present normative regulations do not and cannot capture the
actual fresh concrete performance.

Figure 4: Plastic behaviour of a Bingham fluid (e.g. concrete) versus a Newtonian fluid (e.g. water), and
comparison of concrete types by rheology [1]

Figure 5: Effect of concrete mix design on rheological behaviour acc. to Wallevik (2003), from [1]

The left graph in figure 4 describes in a simplified way, “that concrete requires a certain amount of energy
to start moving (the yield stress) and, thereafter, it resists this movement (by viscosity)” [1]. However, the
typical behaviour of deep foundation concrete might be easier understood if compared to other concrete
types for different applications, see right graph in figure 4. “Normal concrete, compacted using vibrators,
has both a high yield stress and high viscosity. Self-compacting concrete requires very low yield stress for
self-levelling and compacting by self-weight alone. Tremie concrete needs a low viscosity for a good
filling ability at a relatively high … yield stress … for undisturbed displacing of support fluid and
controlling segregation in deep foundations. As a benefit the large concrete head assists in compaction

150
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

and makes it unnecessary to work at very low yield stress values which might result in sensitive concrete
mixes” [1].

Since the practical possibilities to measure the key rheological parameters directly are limited, because a
special laboratory apparatus (rheometer) is needed, “the ease of flow – as a measure for viscosity – has
been assessed intuitively and qualitatively during placement, for example, through observing and
classifying the difficulty of emptying the tremie pipes or the truck unloading times” [1].

Consequently, considering complex modern concrete technology combined with the demanding design,
and challenging execution of deep foundations, a new approach for the quality assessment, and quality
control system in future seems to necessarily become performance based.

5. EFFC/DFI TREMIE GUIDE


The new EFFC/DFI Tremie Guide is aimed at those involved in the procurement, design, and
construction of bored piles (drilled shafts) and diaphragm walls including Owners/Clients, Designers,
General Contractors and Specialist Contractors. It is intended as a practical addition to existing standards,
not a substitute.

The primary purpose of the Guide is to give guidance on the characteristic performance of fresh concrete
and its method of placement using tremie methods in bored piles and diaphragm walls, allowing
construction of high quality elements. Getting the mix right can only be achieved in a joint approach by
the Specialist Contractor (to achieve the execution requirements), the Designer (to meet the durability and
structural needs), and the Supplier (to produce an economic and practical mix).

In particular, the guide helps to understand concrete’s rheology, as shown above in section 1.4, which is
necessary to reliably understand the main placement characteristics of workability and stability, where
- workability is simply defined as that “property of freshly mixed concrete which determines the ease
with which it can be mixed, placed, compacted, and finished” [1], and
- stability is simply defined as the “resistance of a concrete to segregation, bleeding and filtration” [1].

Figure 6 explains, according to the understanding of the joint EFFC/DFI Concrete Task Group, how the
composition decides about concrete rheology as its generic property which subsequently defines its
practical characteristics, and controls the principal requirements to allow for a high quality placement.

Figure 6: Dependencies between composition, rheology and related characteristics, and overall
requirements [1]

The first edition of the Guide proposes appropriate performance criteria for the concrete together with test
methods and initial recommendations on acceptance values.

The Guide also addresses design considerations including concrete rheology, mix design, reinforcement
detailing and concrete cover as well as best practice rules for placement. A review of methods to test the
as-built elements is presented together with advice on the identification and interpretation of results.

151
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

As the concrete cover zone is potentially critical to concrete flow, and because the structural design and
the execution standards set specific requirements which may seem to be conflicting, the Guide tries to
clarify the requirements for execution as follows: “The designer should specify a nominal cover based on
a minimum cover plus an allowance for construction tolerances. A minimum nominal concrete cover of
75mm is recommended where the concrete is cast directly against the ground. Where casing is used and
the excavation surface can be considered ‘smooth’, the minimum requirement may be reduced to 50mm”,
as shown in Figure 7. “In most cases, the minimum nominal values given above will exceed those derived
from structural and durability requirements. European standard EN 1536:2010 Clause 7.7.3 identifies
particular instances where the minimum nominal cover may need to be increased and these rules should
be followed” [1].

Figure 7: Concrete cover requirements for execution in deep foundations [1]

In an Appendix of the Guide a table summarises various international minimum reinforcement and
spacing for bored piles and diaphragm walls. Additionally to the relative requirement of four times the
maximum aggregate size the European Norm EN1536 also requires an absolute minimum for clear
spacing which is for vertical bars 100 mm, or 80mm for lap length.

6. EFFC/DFI RESEARCH PROGRAM


A second edition of the Guide will be published on completion of additional Research and Development
work, which is aimed to find new acceptance criteria for the characterisation of tremie concrete. The
Research and Development Project is being carried out by the Technical University of Munich together
with the Missouri University of Science and Technology. This project includes desk studies, laboratory
testing, and onsite testing at worksites in Europe and the US. The research work, funded and supported by
many sponsors from all stakeholder parties (named in the Guide [1]), should be completed at the end of
2016.

The first aim of the research project is to develop an advanced test concept for the rheological
characterisation of fresh tremie concrete by means of singular and practical “on-site” tests. Rheological
measurements in a rotational rheometer with vane geometry will be performed in parallel. This procedure
should enable the determination of a correlation between the rheological properties and the workability
characteristics of each concrete. The second aim of the project is the definition of acceptance criteria for
the workability and supplementary stability tests of the new test concepts. To enable this definition, the
values for the rheological, workability and stability testing have to be linked to the quality of the
foundation elements, made with this concrete at the time of testing. The foundation elements produced
with the tested concrete will therefore be rated for their quality after completion either by visual
assessment after exposure (Europe) or by integrity testing (US) [3].

Since, contrary to concrete technology, testing methods have not been investigated according to their
capability to reflect the real workability or stability of concrete, a main goal of the program is to collect
real rheological data from actual construction sites to evaluate and assess the concrete types used, and
their applicability, i.e. the resultant quality of executed piles and diaphragm walls. To accurately measure
the generic rheological parameters viscosity and yield stress, and also thixotropy in special cases [3], a
rotational rheometer with vane geometry is used for the on-site and lab testing.

152
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

However, the obvious and major advantage of simple and practical test methods is the ease of use on
construction sites and the fact that these tests are known to everybody. However, only a limited spectrum
of rheological properties is determinable. While slump, slump flow and spread on the flow table may
permit an indirect characterisation of one parameter, yield stress, other parameters, viscosity and
thixotropy, remain completely unconsidered. And since the form filling properties of fresh concrete are
affected by yield stress as well as viscosity and thixotropy, a reliable characterisation of rheology based
on the normative consistence tests is not possible. To verify the applicability of the new test concept, the
investigation program is planned to be carried out and repeated on seven different deep foundation
construction sites, including additional practical tests rather linked to viscosity using the L-box or
inverted slump cone test.

The investigation program aims to determine the fresh concrete properties and their characteristics over
time, starting with the concrete delivery to site and ceasing upon completion of the tremie placement.
During this period, three stages of the concrete properties are to be determined (see also Figure 8):
- Dynamic properties
Properties of the fresh sheared concrete to account for the concrete behaviour during mixing in the
concrete truck mixer, for the tremie process and for the flow behaviour in the cross-section of the
foundation element directly after pouring
- Thixotropy
Reversible structural build-up of the (undisturbed) concrete within a few minutes at rest to account
for the behaviour in the foundation element after pouring and for interruptions in the placement
process
- Flow retention
Irreversible structural build-up of the concrete (stiffening, setting) within some hours to account for
the extended workability life, required for the placement process for deep foundations

Figure 8: Fresh concrete testing schedule [3]

An accompanying laboratory test program with the same schedule as on site should complete the
information on the influence of specific components used for deep foundation concrete with in the usual
range of composition. The main outputs are the determination of the effect of the type, amount and
interaction of possible constituents, e.g. cement, (mineral) additions and (chemical) additives as well as
the role of the water-binder ratio of the concrete.

Figures 9 and 10 present initial results of the Technical University of Munich on fresh deep foundation
concrete: on the yield stress and viscosity at different times, and on the structural build-up when at rest.
These first findings will be further evaluated and assessed within a wider set of results for more concretes
used on European and US deep foundation sites, and then related to the other observations and findings,
before any conclusions can be drawn on recommendations e.g. for future acceptance criteria.

153
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 9: Initial results on yield stress (upper graph) and viscosity (lower graph) of a deep foundation
concrete, tested on site and in the laboratory, both derived from the rotational rheometer and slump flow
(yield stress) respectively L-Box (viscosity) [3]

400 60

High thixotropy
Static yield stress τ0S [Pa]

Slump flow dS [cm]

τ0S = 0.81tR + 136 58


300
Low thixotropy
56
200
54
Low thixotropy High thixotropy
τ0S = 0,05tR + 72
100
52
Mix 1 Mix 1
Mix 2 Mix 2
0 50
0 60 120 180 240 300 0 60 120 180 240 300

Time at rest [s] Time at rest [s]

Figure 10: Initial results from rheological (left) and workability (right) testing of two different deep
foundation concretes [3]

154
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

7. LOOKOUT
Considering the recent development in concrete technology and its actual state-of-the-art, without having
changed any design restrictions or concrete acceptance criteria at the same time, it is essential to develop
new performance specifications and requirements. In any case, present rules must be checked for
relevance and possibly adapted.

The main goal is the prevention of avoidable defects and anomalies in deep foundation elements. The new
EFFC/DFI Best Practice Guide to Tremie Concrete for Deep Foundations with its specific
recommendations will serve all stakeholders.

The Guide’s Edition 1 already presents supplementary fresh concrete test methods. Further results will be
evaluated from investigations made in today’s deep foundation industry considering current design,
execution methods, and concrete. New criteria shall be presented in the Guide’s Edition 2 which will
better characterise concrete properties which cannot be easily assessed with the knowledge from common
practice. A limited set of criteria characterising different properties quantitatively could substitute present
specifications, or at least lead to a revision of present recommendations which may be inappropriate or
ineffective. Such a new performance based approach to test and assess tremie concrete workability, also
with respect to stability, is overdue. Implementation in the next generation of European standards shall be
considered.

REFERENCES
EFFC/DFI, 2016: Best Practice Guide to Tremie Concrete for Deep Foundations. European Federation
of Foundation Contractors, UK, and Deep Foundation Institute, USA

Beckhaus, K., 2014: Are Specifications for Deep Foundation Concrete up-to-date? DFI/EFFC
Proceedings of the 11th International Conference on Piling and Deep Foundations in Stockholm, Sweden

Kraenkel, T., et al., 2016: Rheology Testing of Deep Foundation Concrete. 25th Conference on Rheology
of Building Materials, University of Technology Regensburg, Germany

MEMBERS of the EFFC/DFI Concrete Task Group [1]:

Karsten Beckhaus (Chairman), Bauer Spezialtiefbau, Contractor


Bartho Admiraal, Volker Staal en Funderingen, Contractor
Björn Böhle, Keller Grundbau, Contractor
Jesper Boilesen, Züblin, Contractor
Michel Boutz, SGS INTRON, Consultant
Dan Brown, Dan Brown & Associates, Consultant
Sabine Darson-Balleur, Soletanche Bachy, Contractor
Thomas Eisenhut, Mapei Betontechnik, Additive Supplier
Peter Faust, Malcolm Drilling, Contractor
Christian Gilbert, Systra, Consultant
Raffaella Granata, TREVI S.p.A., Contractor
Chris Harnan, Ceecom, Consultant
Michael Löffler, CDM Smith, Consultant
Gerardo Marote Ramos, Terratest, Contractor
Duncan Nicholson, ARUP, Consultant
Sarah Williamson, Laing O’Rourke, Contractor

155
156
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Ground Improvement vs. Pile Foundations?


Serge Varaksin, Apageo, France, s.varaksin@apageo.com

Babak Hamidi, GFWA, Australia, b.hamidi@gfwa.com.au

Noël Huybrechts, Belgian Building Research Institute - BBRI and KU Leuven, Belgium, nh@bbri.be

Nicolas Denies, Belgian Building Research Institute - BBRI, Belgium, nde@bbri.be

ABSTRACT
A current trend on the European market is the use of ground improvement concepts as alternative or in
complement with deep foundations realized by piling methods. Several ground improvement techniques
include the use of rigid inclusions. Their construction process, sometimes similar to the typical piling
techniques, makes it difficult to draw a boundary between deep foundation by piles and ground
improvement by rigid inclusions. In practice, this lack of clarity can lead to severe discussions between
the stakeholders of the construction project and sometimes results in a feeling of double standard politics
on the market of foundation contractors with severe specifications for piling contractors and nebulous
requirements for ground improvement contractors. This wrong debate is generally caused by a misuse of
the foundation engineering terminology and by a confusion of the roles and the failure mechanisms
related to the different foundation concepts.
In the present keynote, the concept of rigid inclusions, including the use of a load transfer platform, is
reviewed and illustrated with several case studies. Design methodologies, in line with the Eurocodes, are
discussed and promoted in order to ensure a positive competition or a strong collaboration between both
worlds. The authors give a general overview of the rigid inclusion concept including theoretical, design
and execution aspects. Recent trends on the market (deep mixing, CMC…) are highlighted and research
perspectives proposed to further understand the fundamental principles governing the different
foundation concepts.

1. INTRODUCTION

1.1. Ground Improvement vs. Foundation Piles – the wrong debate


A current trend on the European market, as all over the world, is the use of ground improvement concepts
as alternative to or in complement with deep foundations installed with typical piling methods. The
Belgian market is no exception to the rule. If the design of foundation piles to transmit the structural loads
to the ground is now well established - notably with the development of the Eurocodes and their National
Annexes - the way to design ground improvement concepts still remains unclear. This absence of design
rules is sometimes considered as an advantage for the use of ground improvement methods as lesser rules
equate to lesser constraints and lesser design requirements and control (values). Nevertheless, most of the
time, this lack of design rules leads to severe, long and expensive discussions sometimes resulting in the
elimination of the ground improvement solution only due to the lack of clarity of the design framework of
the concept. This elimination can cost much for all the parties involved in the project. In European
countries - where the design of the engineering solutions is more and more established under the umbrella
of the EN standards - the lack of design rules for ground improvement concepts really represents an
obstacle to the growth of this kind of alternative solution.
A practical example is the replacement of foundation piles by a concept of ground improvement involving
soil mix elements. When we begin to compare both solutions, it is always the beginning of an irrelevant
debate with the following questions: “how to conform the soil mix elements to the severe requirements
imposed to concrete piles on the market”? or “Are the soil mix elements in agreement with the
requirements of Eurocode 7 for the piles”? In this way of thinking, the soil mix element (which can be a
column or a rectangular panel) is directly considered as a pile of lower quality, a cheap pile or a “second-
hand” pile. Indeed, how to compare a well-known concept (reinforced concrete pile or steel pile)
supported by EN design requirements (Eurocode 7 and National Annexes, e.g. Rapport 12 of BBRI in
Belgium) and European (CE) and local (e.g. Benor in Belgium) material markings with a foundation
concept including soil mix elements made of a mix of grout and soil (in the deep mixing process, the
ground is mechanically mixed in place, while a binder, often based on cement, is injected). Finally, if the

157
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

ground improvement concept involving soil mix elements is still selected in place of the pile foundation
solution, there is always the same reaction from the concrete industry and the piling contractors
highlighting the unbalanced design requirements for soil mix elements and concrete products and the
unfair competition between both techniques. “Why do we have to conform to severe (and costly) QA/QC
rules for well-known and recognized technique and material while, when we work with an innovative
technique such as the deep mixing, the requirements are more flexible (or sometimes do not exist)”?
There is thus a feeling of double standard politics on the market of foundation contractors.
Actually, this is a wrong debate. And this latter is caused by the idea that a “soil mix element” can be a
“soil mix pile”. In order to close the debate, it is really important to avoid this terminology. A soil mix
element is not a pile. In the United States of America, for example, the terms used are “soil-cement
columns” or “soil mix panels” but reference is not made to “soil mix piles”. In reality, we are dealing with
different geotechnical concepts. In the classical piling concept, the foundation pile is used to transmit the
structural loads to deeper rock or firm soil layers (= base resistance) at sites presenting
soft/weak/compressible soils at shallow depths and to support loads by shaft resistance (= skin friction).
The resistance of the soft soil located under the concrete slab is generally not considered in the design,
which is not always the case with the ground improvement concepts. In order to fully understand the
issue, it is necessary to return to the roots of foundation engineering.

1.2. Behind the geotechnical concepts – back to the roots


Commonly it is envisaged that shallow footings directly rest on competent ground, and deep foundations
are used to transfer loads to deep firm ground. According to Salgado (2008):
- “Shallow foundations transfer structural loads to relatively small depths into the ground. They range
from isolated foundations, each carrying its own column load, to elements carrying several columns,
walls or even all the loads for a given structure or building”.
- “Piles are long, slender elements made of concrete, steel, timber, or polymer used to support
structural loads. Historically, piles have been used to transfer structural loads to deeper rock or firm
soil layers at sites where soft clays or loose sands exist at shallow depths. In recent decades, the uses
expanded to absorbing tensile and lateral loads, to supporting loads by shaft resistance, and to
reducing the settlement of mat foundations”.
In other references, limitations are introduced in the definitions considering the dimensions of the
different foundation elements. Terzaghi et al. (1996) define a shallow footing as a footing that has a width
equal to or greater than the foundation depth, which is the distance from the level of the ground surface to
the base of the footing, and a pile as a very slender pier that transfers a load through its lower end onto a
firm stratum or else through side friction onto the surrounding soil. Bowles (1996) defines shallow
foundations as bases, footings, spread footings or mats with the ratio of depth of footing to its width being
equal to or less than 1, and defines deep foundations as piles, drilled piers or caisson with ratio of length
to width (or diameter) being equal to or greater than 4. Das (2009) notes that studies show that the ratio of
footing depth to width of shallow footings can be as large as 3 or 4.
While it is very beneficial to have concise definitions for various concepts and behaviors to avoid
confusion and to facilitate the accurate transfer of thoughts and intent, it can be observed that it is not
possible to simply set an integer as the separation point between the two foundation systems and more
insight into the matter may be required. These two terms are only a simplification for explaining the
mechanisms of load transfer to the ground. Bearing of shallow foundations are generally expressed by
shear theories originally developed by Prandtl (1920), Terzaghi (1943), Meyerhof (1951) and Hansen
(1971). Skin resistance (Tomlinson, 1971; Vijayvergiya and Focht, 1972 and Burland, 1973) may become
a major contributor as the ratio of footing height to width begins to increase. At the same time, while a
large based footing may be categorised as a shallow foundation system due to its ratio of depth to width,
the depth of soil within the system may be very deep indeed. Hence, the distinction between both
foundations concepts (shallow or deep) is not always easy to achieve. This topic will later be discussed in
the present keynote.
Moreover, a pile is rarely used alone. The structural loads are often transmitted to the ground with the
help of pile groups involving the consideration of the group effect in the design. In a pile group, the loads
are transmitted to the ground only via the piles, which are linked together by a pile cap. Even if the pile
cap is in practice in contact with the soil, it is not included in the calculation of the bearing capacity. That
is not the case with a pile raft. In a pile raft, the piles and the raft foundation are designed together in
order to reduce potential settlement of the structure. The percentage of the load taken by the slab of the
piled raft solution is normally far more than the load taken by the pile cap of a pile group where only the

158
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

piles are responsible of the load transfer. Pile raft is thus a concept often used in a Serviceability Limit
States (SLS) logic.
To further complicate this simple categorization, it is also possible to convert deep loose or soft soils to
adequately competent soil by ground improvement; thereby safely dissipating the loads without engaging
piles for transferring loads to deeper firm ground.
During the last decades, the importance of the ground improvement market has enormously increased.
New methods, tools and procedures have been developed and applied in practice. A particularity of this
field of geotechnical engineering is that a major part of the advances in ground improvement has to be
credited to the equipment manufacturers and to the specialist contractors. If this constant improvement of
the technologies stretches the boundaries of the practical applications, it sometimes results in a lack of
theoretical concept background to support these applications on the market. Theory and design aspects do
not follow fast enough to support these technological developments and their resulting applications. As a
consequence, geotechnical designers are often exposed to new execution processes without enough
information to support the geotechnical concepts behind the proposed applications.
A first issue is the number of ground improvement methods available on the international market. In order
to answer this question, the ISSMGE Technical Committee TC 211 (former TC17) has adopted a
classification system, shown in Table 1, with the following categories (and methods) for ground
improvement works:
- Category A: Ground Improvement without admixtures in non-cohesive soils or fill materials (dynamic
compaction, vibrocompaction…)
- Category B: Ground Improvement without admixtures in cohesive soils (Replacement, preloading,
vertical drains, vacuum consolidation…)
- Category C. Ground Improvement with admixtures or inclusions (Vibro replacement - stone columns,
sand compaction piles, rigid inclusions…)
- Category D. Ground Improvement with grouting type admixtures (Particulate and chemical grouting,
Deep mixing, jet grouting…)
- Category E. Earth reinforcement (Geosynthetics or MSE, ground anchors, soil nails…)
This classification is based on the broad trend of behaviors of grounds to be improved and whether
admixtures are used or not. Chu et al. (2009) have classified and described the various ground
improvement techniques that are commonly practiced to date. It is not possible to mention all techniques
in this paper; however, separate lists are given on the ISSMGE TC211 website (www.tc211.be).
Some of these techniques are developed to improve the physical and mechanical properties of in-situ soils
without the introduction of imported material, and the outcome will remain as what is classically referred
to as a shallow foundation (e.g. preloading with vertical drains, dynamic consolidation, etc.).
In other ground improvement techniques, higher quality materials are added to the ground as columnar
inclusions. Some inclusions may be very long and can call pile foundations to mind. Nevertheless, in
these geotechnical concepts, foundations are made of classical shallow footings combined with (deep)
soil masses that have been improved by the installation of the inclusions. These foundations cannot be
expressed by the classical shallow foundation approaches as reported in the previous paragraph, and
require further analysis and design of the improved ground as part of the foundation system. The
complication in the categorisation can turn into confusion when columnar inclusions are formed by
installing concrete and grout columns in the ground using piling equipment. That uncertainty particularly
concerns the installation of rigid inclusions (category C5) and the use of soil mix elements (D3) as
alternatives to typical pile foundations.
As defined in Chu et al. (2009), rigid inclusions refer to the use of semi-rigid or rigid integrated columns
or bodies in soft ground to improve the ground performance globally so as to decrease settlement and
increase the bearing capacity of the ground. In the broad sense, the concept of stone columns is a type of
rigid inclusions. However, they are not considered because the materials used for those columns
(generally crushed stones) are disintegrated and the columns formed in the ground are not able to stand
without the lateral support of the soil. Contrarily, the constitutive material of the rigid inclusions is
continuous and presents a high permanent cohesion. Its rigidity is therefore higher than the rigidity of the
surrounding soil. Typical rigid inclusions are concrete columns (possibly installed into the ground with a
classical piling technique), grout columns, soil mix elements (columns, panels, trenches, blocks, etc.),
Controlled Modulus Columns (CMCs), grouted stone columns, etc.

159
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 1. Classification of ground improvement methods of the ISSMGE TC211 (from Chu et al., 2009)
Category Method Principle
A. Ground A1. Dynamic compaction Densification of granular soil by dropping a heavy weight from
improvement air onto ground.
without A2. Vibrocompaction Densification of granular soil using a vibratory probe inserted
admixtures in into ground.
non-cohesive A3. Explosive compaction Shock waves and vibrations are generated by blasting to cause
soils or fill granular soil ground to settle through liquefaction or
materials compaction.
A4. Electric pulse compaction Densification of granular soil using the shock waves and energy
generated by electric pulse under ultra-high voltage.
A5. Surface compaction (including rapid Compaction of fill or ground at the surface or shallow depth
impact compaction). using a variety of compaction machines.
B. Ground B1. Replacement/displacement (including Remove bad soil by excavation or displacement and replace it
improvement load reduction using lightweight materials) by good soil or rocks. Some lightweight materials may be used
without as backfill to reduce the load or earth pressure.
admixtures in B2. Preloading using fill (including the use Fill is applied and removed to pre-consolidate compressible soil
cohesive soils of vertical drains) so that its compressibility will be much reduced when future
loads are applied.
B3. Preloading using vacuum (including Vacuum pressure of up to 90 kPa is used to pre-consolidate
combined fill and vacuum) compressible soil so that its compressibility will be much
reduced when future loads are applied.
B4. Dynamic consolidation with enhanced Similar to dynamic compaction except vertical or horizontal
drainage (including the use of vacuum) drains (or together with vacuum) are used to dissipate pore
pressures generated in soil during compaction.
B5. Electro-osmosis or electro-kinetic DC current causes water in soil or solutions to flow from anodes
consolidation to cathodes which are installed in soil.
B6. Thermal stabilisation using heating or Change the physical or mechanical properties of soil
freezing permanently or temporarily by heating or freezing the soil.
B7. Hydro-blasting compaction Collapsible soil (loess) is compacted by a combined wetting and
deep explosion action along a borehole.
C. Ground C1. Vibro replacement or stone columns Hole jetted into soft, fine-grained soil and back filled with
improvement densely compacted gravel or sand to form columns.
with admixtures C2. Dynamic replacement Aggregates are driven into soil by high energy dynamic impact
or inclusions to form columns. The backfill can be either sand, gravel, stones
or demolition debris.
C3. Sand compaction piles Sand is fed into ground through a casing pipe and compacted by
either vibration, dynamic impact, or static excitation to form
columns.
C4. Geotextile confined columns Sand is fed into a closed bottom geotextile lined cylindrical hole
to form a column.
C5. Rigid inclusions Use of piles, rigid or semi-rigid bodies or columns which are
either premade or formed in-situ to strengthen soft ground.
C6. Geosynthetic reinforced column or Use of piles, rigid or semi-rigid columns/inclusions and
pile supported embankment geosynthetic girds to enhance the stability and reduce the
settlement of embankments.
C7. Microbial methods Use of microbial materials to modify soil to increase its strength
or reduce its permeability.
C8 Other methods Unconventional methods, such as formation of sand piles using
blasting and the use of bamboo, timber and other natural
products.
D. Ground D1. Particulate grouting Grout granular soil or cavities or fissures in soil or rock by
improvement injecting cement or other particulate grouts to either increase the
with grouting strength or reduce the permeability of soil or ground.
type admixtures D2. Chemical grouting Solutions of two or more chemicals react in soil pores to form a
gel or a solid precipitate to either increase the strength or reduce
the permeability of soil or ground.
D3. Mixing methods (including premixing Treat the weak soil by mixing it with cement, lime, or other
or deep mixing) binders in-situ using a mixing machine or before placement
D4. Jet grouting High speed jets at depth erode the soil and inject grout to form
columns or panels
D5. Compaction grouting Very stiff, mortar-like grout is injected into discrete soil zones
and remains in a homogenous mass so as to densify loose soil or
lift settled ground.
D6. Compensation grouting Medium to high viscosity particulate suspensions is injected into
the ground between a subsurface excavation and a structure in
order to negate or reduce settlement of the structure due to
ongoing excavation.
E. Earth E1. Geosynthetics or mechanically Use of the tensile strength of various steel or geosynthetic
reinforcement stabilised earth (MSE) materials to enhance the shear strength of soil and stability of
roads, foundations, embankments, slopes, or retaining walls.
E2. Ground anchors or soil nails Use of the tensile strength of embedded nails or anchors to
enhance the stability of slopes or retaining walls.
E3. Biological methods using vegetation Use of the roots of vegetation for stability of slopes.

160
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

The concept behind the use of rigid inclusions is not the same as the concept of pile foundations. In the
concept of rigid inclusions, the loads sustained by the soft soil is reduced (usually between 60 and 90%)
in order to reduce the global and differential settlements. Nevertheless, the loads are not directly
transmitted to depth. Here, the soft soil plays a role, and supports part of the load whereas in the pile
foundation concept the soft soil is just bypassed (or used for skin friction consideration).
When we work with rigid inclusions, a Load Transfer Platform (LTP) is often used with a thickness
generally ranging between 40 and 80 cm. In general, the load transfer platform may consist of single or
multiple layers of geosynthetics (often geogrids) placed horizontally in a layer of well compacted
granular material (often crushed stones or gravels). As illustrated in Fig. 1, in comparison with other
commonly utilized foundation concepts, the load transfer platform allows the transfer of the structural
loads to the head of the rigid inclusions by means of an arching effect developing in the granular layer.
This effect is caused by the differential settlement arising between the soft soil and the heads of the rigid
inclusions at the base of the load transfer platform, which also results in the emergence of a negative skin
friction along the rigid inclusions at shallow depth. This negative skin friction is a governing factor of the
load transfer in the concept of rigid inclusions and will be further discussed. Finally, it is to note that the
installation of a load transfer platform also allows the decrease of the bending moments and the shear
stresses in the foundation slab of the structure to be supported.
Previously studied by Combarieu (1988), the concept of rigid inclusions applied with a load transfer
platform has more recently been the subject of an extensive French national research programme called
ASIRI (Améliorations des Sols par Inclusions Rigides, which translates to Ground Improvement by Rigid
Inclusions) (IREX, 2012). The following paragraphs concentrate on the main aspects of this topic.

Figure 1: Type of load transfer in the different usual foundation concepts

2. CONCEPT OF GROUND IMPROVEMENT BY RIGID INCLUSIONS

2.1. Study of the mechanisms at an early stage


Combarieu (1988) has originally studied the behavior of rigid inclusions installed in soft ground. As
shown in Fig. 2, a compressible soil layer of thickness H subjected to an embankment load with intensity
q o will ultimately settle an amount that can be denoted as W s (o). Likewise, the settlement at any depth, z,
can be denoted by W s (z). Considering the addition of an inclusion in the engineering issue, it is interesting
to focus on the settlement of both materials: the soil and the rigid inclusion. The soil conditions at
distances away from the single inclusion are identical to untreated ground after complete stabilisation.
However, the stress and deformations change around the immediate vicinity of the inclusion. The
inclusion settles by an amount equal to W p (z) due to the loading plus a further small amount due to its
own compression. Settlement is higher when the rigid inclusion terminates on soft soil (see Fig. 3)
compared to when it is supported by hard soil (see Fig. 4). In the lower part of the inclusion, where z >
h c , the settlement of the soil is smaller than the inclusion settlement (including its compression); however,
the opposite is true in the upper portion where z < h c . Soil and inclusion settlements are only equal at z =
h c . There is thus the development of a differential settlement between the soil and the rigid inclusion, as
previously stated, with the onset of a negative skin friction along the inclusion when z > h c .
As shown in Fig. 5, Combarieu (1988) has also considered the equilibrium of the forces acting on the
rigid inclusion. The four forces acting on the inclusion at equilibrium are:

161
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

- the driving forces: the vertical load Q acting on the head of the inclusion and the resultant negative
friction, F n , acting along the inclusion segment with a length equal to h c .
- the resisting forces: the positive friction, F p , mobilised in the lower part of the inclusion and along a
segment with length L - h c , and Q p acting at the base of the rigid inclusion. The balance of the forces is
therefore Q + F n = F p + Q p .
According to Combarieu (1988), the computation of the bearing capacity of the rigid inclusion will be
therefore influenced by the development of the negative skin friction along the inclusion.

Figure 2: Ground section without rigid inclusion (Combarieu, 1988)

Figure 3: Ground section with rigid inclusion terminating in soft ground (Combarieu, 1988)

Figure 4: Ground section with rigid inclusion supported by hard ground (Combarieu, 1988)

Figure 5: Forces acting on a rigid inclusion (Combarieu, 1988)

162
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2.2. Study of the mechanisms and design approach – ASIRI project (IREX,
2012)

2.2.1. Load Transfer Platform (LTP) and failure mechanisms


As shown in Fig. 6, it is assumed that rigid inclusions, with diameter D = 2r p , are installed in a square
grid with centre to centre spacing, denoted by s. The thickness of the load transfer platform (LTP) is
denoted by H M , and it is defined by its characteristics (cohesion c’, friction angle φ’ and volumetric
weight γ). The uniformly distributed external load applied to the load transfer platform is indicated by q o .

Figure 6: Cross section illustrating the concept of ground improvement by rigid inclusions, including the
load transfer platform (LTP), for a uniform external loading q 0 (from IREX, 2012)

The ASIRI project (IREX, 2012) has demonstrated that two failure mechanisms are possible for that
foundation concept: the Prandtl’s failure mechanism and the punching shear failure mechanism.

2.2.1.1. Prandtl’s failure mechanism


According to Prandtl (1920), Prandtl’s failure mechanism occurs when the load transfer platform is
covered by a rigid structural element (such as a slab on grade, a raft or footings) or when the embankment
is sufficiently thick to avoid punching failure (which corresponds to the formation of shear cones in the
LTP’s surface).
According to the ASIRI guidelines (IREX, 2012), an embankment is considered thin once:
𝐻𝐻𝑀𝑀 < 𝑂𝑂. 7(𝑠𝑠 − 𝐷𝐷) (1)
As shown in Fig. 7(a), Prandtl’s failure diagram includes a Rankine active limit state domain (I) above the
inclusion head, which is delimited by a logarithmic spiral arc domain (II) and a Rankine passive limit
state domain (III), which is located beyond the inclusion head. q p + is the stress applying on the inclusion
head, and q s + is the stress applying on the in-situ soil.

(a) (b)
Figure 7: (a) Prandtl failure mechanism for slabs on grade, rafts, footings and thick embankments and
(b) punching shear failure mechanism for thin embankments (from IREX, 2012)

163
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

The guidelines of the ASIRI project (IREX, 2012) have been developed in agreement with the philosophy
of the Eurocodes. The maximum load that can be applied to the inclusion head q p + is therefore calculated
at the Ultimate Limit State condition (ULS). This ULS verification is performed by implementing
Eurocode 7 - Design Approach 2 (EN 1997-1, 2004) with the combination of partial factors:
A1 + M1 + R2 (A for Action, M for Material and R for Resistance), which means that load factors on dead
and live loads are respectively 1.35 and 1.50 and the partial material factors are equal to 1.
According to Prandtl’s diagram, q p + can be determined from the stress applied on the supporting soil and
the intrinsic parameters of the load transfer platform:

𝑐𝑐′ 𝛾𝛾 (2)
𝑞𝑞𝑝𝑝 + = 𝑠𝑠𝑞𝑞 𝑁𝑁𝑞𝑞 𝑞𝑞𝑠𝑠 + + 𝑠𝑠𝑐𝑐 𝑁𝑁𝑐𝑐 − 𝑠𝑠𝛾𝛾 𝑁𝑁𝛾𝛾 𝑟𝑟𝑝𝑝
𝛾𝛾𝑐𝑐′ 𝛾𝛾𝛾𝛾

N q , N c and N γ are coefficients that are functions of the friction angle of the constitutive material of the
load transfer platform, and can be calculated from equations (3) to (5):

𝜑𝜑′ (3)
𝜑𝜑′
𝜋𝜋 𝛾𝛾𝜑𝜑′ 𝜋𝜋 tan�
𝛾𝛾𝜑𝜑′

𝑁𝑁𝑞𝑞 = tan2 � + � × 𝑒𝑒
4 2

𝜑𝜑′ (4)
𝑁𝑁𝑐𝑐 = (𝑁𝑁𝑞𝑞 − 1) cot � �
𝛾𝛾𝜑𝜑′

𝜑𝜑′ (5)
𝑁𝑁𝛾𝛾 = 2(𝑁𝑁𝑞𝑞 − 1) tan � �
𝛾𝛾𝜑𝜑′

γ c’ , γ ϕ’ , and γ γ are partial material factors equal to 1 (in the combination A1 + M1 + R2).
The weight of the load transfer platform is typically neglected for a relatively thin platform, and the
superficial (third) term in equation (2) is omitted.
When a purely granular material is used as constitutive material of the load transfer platform, there is no
cohesion, and the related term becomes null. Hence, equation (2) becomes:

𝑞𝑞𝑝𝑝 + = 𝑠𝑠𝑞𝑞 𝑁𝑁𝑞𝑞 𝑞𝑞𝑠𝑠 + (6)

For axisymmetric or plane-strain conditions, s q = 1, and a relationship between q p + and q s +, only function
of φ’, is established:

𝑞𝑞𝑝𝑝 + = 𝑁𝑁𝑞𝑞 𝑞𝑞𝑠𝑠 + (7)

A second equation is still necessary in order to determine the values of q p + and q s +. This is performed
using the load conservation equation:

𝛼𝛼𝛼𝛼𝑝𝑝 + + (1 − 𝛼𝛼)𝑞𝑞𝑠𝑠 + = 𝑞𝑞𝑜𝑜 (8)

where α is the replacement ratio:

𝐴𝐴𝑐𝑐 (9)
𝛼𝛼 =
𝐴𝐴𝑐𝑐 + 𝐴𝐴𝑠𝑠

where A c is the area of inclusion and A s the area of soil.


From equations (7) and (8), q p + and q s + can be expressed as a function of φ’, α and q 0 :

𝑁𝑁𝑞𝑞 (10)
𝑞𝑞𝑝𝑝 + = 𝑞𝑞𝑜𝑜
1 + 𝛼𝛼�𝑁𝑁𝑞𝑞 − 1�

164
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

1 (11)
𝑞𝑞𝑠𝑠 + = 𝑞𝑞𝑜𝑜
1 + 𝛼𝛼�𝑁𝑁𝑞𝑞 − 1�

Within the framework of the ASIRI project (IREX, 2012), further research on Prandtl’s failure
mechanism has been carried out by centrifugal testing with various LTP thicknesses, rigid inclusion
spacing and replacement ratios (Okyay, 2010). In Fig. 8, centrifuge test results are compared with limiting
pressures calculated from Prandtl’s theory, and it can be observed that a very good agreement exists
between measured and theoretical values.

Figure 8: Comparison of measured limiting pressures (from centrifugal testing) with theoretical values
calculated from Prandtl’s theory (from IREX, 2012)

Within the framework of the ASIRI project (IREX, 2012), Prandtl’s approach was also investigated by
performing finite element calculations for various uniformly distributed loads. Figure 9 illustrates the
results of this study. Figure 9 shows the pressures acting on the soil and on the inclusion head
respectively on the abscissa and ordinate. The blue curve is obtained using equation (7). The slanted black
lines correspond to the load conservation equation (8). The pink line is derived from finite element
calculations. The stresses acting on the soil and on the inclusion and the value that can be mobilised at the
head of the inclusion are also showed in Fig. 9. During the investigation, the Young modulus of the
compressive soil was reduced for each uniform loading until the load transfer platform failed. It was
observed that, at the last step prior to the failure, the stress at the inclusion head approached Prandtl’s
limit, but did not intersect it. Prandtl’s failure mechanism can also be visualised by the distribution of the
plastic points that are shown as red dots in Fig. 9.

Figure 9: Comparison of limiting pressures calculated from finite element analyses and according to the
Prandtl’s theory (from IREX, 2012)

165
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2.2.1.2. Punching shear failure mechanism


As illustrated in Fig. 7(b), the second failure mechanism can be modelled by envisaging a vertical shear
cone within the granular layer of the load transfer platform. This failure mechanism only exists for thin
load transfer platforms that are not covered by rigid structural elements, and is associated with the peak
friction angle of the material.
According to Eurocode 7 - Design Approach 2 and from the shear cone geometry, the limit stress at the
inclusion head is determined by using the applied external load, q o , and the properties of the load transfer
platform. Two configurations have to be considered as functions of the geometry of the inclusions and the
load transfer platform, as illustrated in Fig. 10.

(a) (b)
Figure 10: (a) Non-overlapping failure cones and (b) Overlapping failure cones (IREX, 2012)

In the first configuration, as shown in Fig. 10(a), the shear cones, developing above the inclusions, do not
overlap. This condition is encountered when H M < H c , where H c is defined as:

𝑅𝑅 − 𝑟𝑟𝑝𝑝 (12)
𝐻𝐻𝑐𝑐 =
tan 𝜑𝜑′

where R is defined as:

𝑠𝑠 (13)
𝑅𝑅 =
√𝜋𝜋

If the shear cones do not overlap (i.e. when H M < H c ), q p + is the weight of the cone plus the external load
applied on the top circular side of the cone:

𝐻𝐻𝑀𝑀 𝑅𝑅𝑐𝑐 2 𝑅𝑅𝑐𝑐 𝛾𝛾 𝑅𝑅𝑐𝑐 2 1 𝑅𝑅𝑐𝑐 2 𝑐𝑐′ (14)


𝑞𝑞𝑝𝑝 + = � 2 + 1 + � + 2 𝑞𝑞0 + � 2 − 1�
3 𝑟𝑟𝑝𝑝 𝑟𝑟𝑝𝑝 𝛾𝛾𝛾𝛾 𝑟𝑟𝑝𝑝 tan 𝜑𝜑′ 𝑟𝑟𝑝𝑝 𝛾𝛾𝑐𝑐′

where

𝜑𝜑′ (15)
𝑅𝑅𝑐𝑐 = 𝑟𝑟𝑝𝑝 + 𝐻𝐻𝑀𝑀 tan � �
𝛾𝛾𝜑𝜑′

γ c’ , γ ϕ’ and γ γ are the partial material factors equal to 1 (in the combination A1 + M1 + R2).
In the second configuration, as shown in Fig. 10(b), the shear cones, developing above the inclusions,
overlap. This condition is encountered when H M > H c . If the shear cones overlap, q p + is the weight of the
cone, the weight of the soil cylinder above it and the external load multiplied by the unit cell area:

𝐻𝐻𝑐𝑐 𝑅𝑅2 𝑅𝑅 𝑅𝑅2 𝛾𝛾 𝑅𝑅2 1 𝑅𝑅2 𝑐𝑐′ (16)


𝑞𝑞𝑝𝑝 + = � � 2 + 1 + � + (𝐻𝐻𝑀𝑀 − 𝐻𝐻𝑐𝑐 ) 2 � + 2 𝑞𝑞𝑜𝑜 + � � 2 − 1��
3 𝑟𝑟𝑝𝑝 𝑟𝑟𝑝𝑝 𝑟𝑟𝑝𝑝 𝛾𝛾𝛾𝛾 𝑟𝑟𝑝𝑝 tan 𝜑𝜑′ 𝑟𝑟𝑝𝑝 𝛾𝛾𝑐𝑐′

with γ c’ and γ γ are equal to 1.

166
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2.2.2. Ultimate Limit State (ULS) stress domain of the concept


In order to fully understand the concept of rigid inclusions with a load transfer platform, the domain of
admissible stresses at the base of the load transfer platform has to be considered prudently.
Figure 11 graphically presents the ULS stress domain at the base of the load transfer platform.

Figure 11: Domain of admissible stresses (ULS) at the base of the load transfer platform (IREX, 2012)

When failure occurs by Prandtl’s mechanism, regardless of the load level, the stress domain in the load
transfer platform is firstly limited by the Prandtl line according to equation (2) or more generally
according to equation (7) represented by line (1a) in the figure.
The stress on the in-situ soil, q s +, is limited, at ULS, by the allowable stress σ v;d , which is determined on
the basis of the soil study; for example with the appropriate partial factors applied on P LM , the limit
pressure of the Ménard pressuremeter test:

𝑘𝑘𝑝𝑝 𝑃𝑃𝐿𝐿𝐿𝐿 (17)


𝜎𝜎𝑣𝑣;𝑑𝑑 =
𝛾𝛾𝑅𝑅;𝑣𝑣∙ 𝛾𝛾𝑅𝑅,𝑑𝑑

where γ R;v is the partial resistance factor for spread foundations (equal to 1.4 according to Table A.5 of
Eurocode 7), and γ R;d is the appropriate model factor.
This allowable stress, σ v;d , limits the domain with line (2) of Fig. 11.
q p + is limited by the load bearing capacity of the inclusion (according to the philosophy of Eurocode 7)
and by the maximum allowable stress for the constitutive material of the rigid inclusion (according to the
philosophy of Eurocode 2):
𝑅𝑅𝑏𝑏 𝑅𝑅
�𝛾𝛾𝑏𝑏∙𝛾𝛾𝑅𝑅,𝑑𝑑 + 𝑠𝑠�𝛾𝛾𝑠𝑠∙ 𝛾𝛾𝑅𝑅,𝑑𝑑
𝑞𝑞𝑝𝑝+ < 𝑞𝑞𝑝𝑝,𝑚𝑚𝑚𝑚𝑚𝑚 < 𝑚𝑚𝑚𝑚𝑚𝑚 � ; 𝑓𝑓𝑐𝑐,𝑑𝑑 � (18)
𝜋𝜋𝑟𝑟𝑝𝑝 2

where γ b and γ s are the partial factors respectively for the base (R b ) and the shaft (R s ) resistances. The
assessment of the base and shaft resistances and the calculation of the design value f c,d will be the
subjects of the following paragraphs. As observed in Fig. 11, q p,max limits the domain with line (3).
When the load transfer platform is thin and not covered by a rigid structural element (such as a slab on
grade, a raft or footings), the stress domain has to be partially limited. This is illustrated in Fig. 12 where
the load transfer platform is thin, not covered by rigid structural elements and the failure cones do not
overlap. Here, the stress domain is limited by the dashed blue line (1b), which corresponds to
equation (14). If the failure cones overlap, a second curve is added to limit the stress domain, as
illustrated in Fig. 13, with the dashed red curve corresponding to equation (16).
Finally, in order to satisfy the load conservation equation (8), q s + and q p + have to be on the diagonal blue
line that is shown in Fig. 14. For a given load q, the admissible domain will reduce to this segment. The
calculated design value q p,d + is therefore calculated by solving Prandtl’s equation (7) and the load
conservation equation (8) simultaneously. That situation corresponds to the red square in Fig. 14.

167
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 12: Domain of admissible stresses (ULS) at the base of the load transfer platform when the load
transfer platform is thin, not covered by rigid structural element and the failure cones do not overlap
(IREX, 2012)

Figure 13: Domain of admissible stresses (ULS) at the base of the load transfer platform when the load
transfer platform is thin, not covered by rigid structural element and the failure cones overlap (IREX,
2012)

Figure 14: Domain of admissible stresses (ULS) at the base of the load transfer platform with
consideration of the load conservation equation (IREX, 2012)

168
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

As a consequence, q p,d + is a function of


- the load q,
- the diameter of the rigid inclusions,
- the inclusion grid size,
- the thickness of the load transfer platform,
- the parameters of the load transfer platform (c’, ϕ’ and γ).
q p,d + is thus independent of the deformability of the various underlying soil layers.
Nevertheless, while q p,d + is the intersection of Prandtl’s equation (7) and the load conservation equation
(8), the pair (q p +; q s +), which is actually mobilized, can be anywhere on this diagonal segment
(line 4 in Fig. 14), and its actual position will depend on the compressibility of the various soil layers
directly below the load transfer platform. If the soil is very soft, then the mobilized pair will be close to
the limit design value q p,d +, but if the soil is quite dense, then the pair will be away from this limit design
value.
It is important to note that, as shown in Fig. 15, changes of the external load moves the equilibrium in the
plane (q p +; q s +) along a curve that tends towards an asymptote for large loads; i.e. an increase in loading
also increases the efficiency towards its maximum value, but is never able to create internal failure of the
load transfer platform by intersecting Prandtl’s equation (7).

Figure 15: q p,d + and deformability of various soil layers (IREX, 2012)

2.2.3. Geotechnical limit states (GEO) – consideration of the negative skin friction
It must be verified that the forces, mobilized at the base and throughout the rigid inclusion, do not exceed
the limit values computed in agreement with Eurocode 7. The base and the shaft resistances of the rigid
inclusion will be determined using the results of the geotechnical investigation (such as cone
penetrometer tests or Ménard pressuremeter tests). In Belgium, this is realized considering the guidelines
of the Rapport 12 of the Belgian Building Research Institute (BBRI) published in 2009, and which are
currently under revision. In France, the standard NF P 94-262 has to be applied.
If the behavior at the base of the rigid inclusion is commonly regarded to be in accordance with the
principles of Eurocode 7, special attention has to be given to the shaft behavior. Indeed, as previously
mentioned in this keynote, a differential settlement arises between the soft soil and the heads of the rigid
inclusions at the base of the load transfer platform. This differential settlement creates a negative skin
friction along the rigid inclusions at shallow depth. The consideration of this negative skin friction is
therefore relevant for the geotechnical design of the rigid inclusion.
After the verification of the positive fiction arising below depth h c , according to Eurocode 7 and its
various national annexes, the negative skin friction which develops above depth of h c has to be assessed.
As reported in IREX (2012), it must be verified that the friction τ of the soil along the inclusion shaft
(above the depth h c ) does not exceed the following limit value:
𝜏𝜏 < 𝜎𝜎𝑣𝑣′ 𝐾𝐾 tan 𝛿𝛿 (19)
where σ v ’ is the vertical stress calculated along the inclusion. K tanδ is an empirical parameter, which is
given in Combarieu (1985) or more recently in the French standard NF P 94-262.

169
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Within the framework of the ASIRI project (IREX, 2012), an advanced methodology is also provided
allowing the assessment of the negative skin friction and the determination of the critical height, h c .

2.2.4. Structural limit states (STR) of the rigid inclusion – definition of fc,d
In order to design the rigid inclusion, its compressive strength for axial loading (usually denoted by UCS
for Uniaxial Compressive Strength) has to be defined. In the philosophy of the Eurocodes, a design value
of the uniaxial compressive strength of the material, noted f c,d , is then computed.

2.2.4.1. fc,d for a classical concrete pile – according to Eurocode 2


For a classical concrete pile, the designer could refer to the content of Eurocode 2 where f c,d is defined as:
𝛼𝛼𝑐𝑐𝑐𝑐 𝑓𝑓𝑐𝑐,𝑘𝑘
𝑓𝑓𝑐𝑐,𝑑𝑑 = (20)
𝛾𝛾𝑐𝑐

where γ c is the partial safety factor for concrete, α cc is the coefficient taking account of long term effects
on the compressive strength and of unfavourable effects resulting from the way the load is applied and f c,k
is the characteristic (5%) cylinder compressive strength of the concrete material determined in accordance
with European standard EN 206-1. The characteristic strengths for f c,k and the corresponding mechanical
characteristics necessary for design of concrete material, are given in Table 3.1 of Eurocode 2. It is to
note that, in agreement with Eurocode 2, the partial factor for concrete γ c should be multiplied by a factor,
k f , for calculation of design resistance of cast in place piles without permanent casing.

2.2.4.2. fc,d for a rigid inclusion - according to the ASIRI project (IREX, 2012)
Within the framework of the ASIRI project (IREX, 2012), a methodology is proposed for the calculation
of the design and characteristic values of the UCS of the material for the rigid inclusions:

𝑓𝑓𝑐𝑐,𝑘𝑘 𝑓𝑓𝑐𝑐,𝑘𝑘 (𝑡𝑡) 𝐶𝐶𝑚𝑚𝑚𝑚𝑚𝑚
𝑓𝑓𝑐𝑐,𝑑𝑑 = 𝑚𝑚𝑚𝑚𝑚𝑚 �𝛼𝛼𝑐𝑐𝑐𝑐 𝑘𝑘3 ; 𝛼𝛼𝑐𝑐𝑐𝑐 ; 𝛼𝛼𝑐𝑐𝑐𝑐 � (21)
𝛾𝛾𝑐𝑐 𝛾𝛾𝑐𝑐 𝛾𝛾𝑐𝑐

where α cc is a coefficient depending on the presence or absence of steel reinforcement (reinforced = 1,


non-reinforced = 0.8). γ c is a partial coefficient with a value equal to 1.5 at the fundamental ULS and 1.2
at the accidental ULS. f c,k * is the characteristic value of the compressive strength of the concrete, grout or
mortar in the inclusion, as determined in the following formula:
∗ 1
𝑓𝑓𝑐𝑐,𝑘𝑘 = 𝑚𝑚𝑚𝑚𝑚𝑚 �𝑓𝑓𝑐𝑐,𝑘𝑘 (𝑡𝑡); 𝐶𝐶𝑚𝑚𝑚𝑚𝑚𝑚 ; 𝑓𝑓𝑐𝑐,𝑘𝑘 � (22)
𝑘𝑘1 𝑘𝑘2

where f c,k is the characteristic value of the compressive strength measured on cylinders at 28 days of
hardening, f c,k (t) is the characteristic value of the compressive strength measured on cylinders at time t
and C max is the maximum compressive strength value taking into account the required consistency of the
fresh concrete, grout or mortar, depending on the technique used, as shown in Table 2.

Table 2. Assigned values of C max and the coefficient k 1 (from IREX, 2012)
Case Execution mode C max (MPa) k1
1 Drilled inclusions with soil extraction 35 1.3
2 Drilled inclusions using a hollow auger with soil extraction 30 1.4
3 Drilled inclusions using a hollow auger with soil displacement 35 1.3
4 Inclusions either vibratory driven or cast in place 35 1.3
5 Incorporation of a binder with the soil (treated soil-columns, jet (*) (**)
grouting…)
(*) Value to be determined through field testing
(**) Columns of treated soil using a mechanical tool that guarantees the cross-section geometry: k 1 to be determined
on a case-by-case basis with k 1 > 1.3

k 1 is a value depending on both the drilling method and slenderness ratio, as listed in Table 2. According
to IREX (2012), for a soil treated by jet grouting or with a tool that does not guarantee a homogeneous
section geometry, the value of k 1 has to be determined on a case-by-case basis k 1 > 1.5. k 1 may be
decreased by 0.1, only for drilled inclusions when the composition of the ground layers guarantees
stability of the outer borehole walls or when the inclusion is cased and concreted in the dry (a guarantee
of borehole wall stability must be demonstrated using the procedure outlined in the EN 1536 bored pile
execution Standard).

170
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

k2 depends on the slenderness ratio:


k2 = 1.05 for inclusions whose ratio of smallest dimension d to length is less than 1/20;
k2 = 1.3-d/2 for inclusions whose smallest dimension is less than 0.60 m;
k2 = 1.35-d/2 for inclusions combining the two previous conditions.
k 3 depends on the type of control performed on the rigid inclusions (see IREX-2012 for more
information).

2.2.4.3. fc,d for soil mix elements - according to the BBRI Soil Mix project (2009-2013)
If the computation of the design value, according to the ASIRI project (IREX, 2012), is not too complex
for a rigid inclusion executed with a classical method (drilled inclusions with or without soil extraction),
the computation of f c,d for a jet grout column or a soil mix element is no easy task according to this
approach. Within the framework of the BBRI soil mix project (2009-2013), a methodology has been
developed to compute the design and characteristic UCS values for soil mix material based on a
simplified approach.
In the BBRI approach, the characteristic UCS value of the soil mix material can be computed with the
help of two different methods depending on the number of test samples defined in agreement with the
quality control requirements of the project (see the SBRCURnet/BBRI soil mix handbook, 2016).
The first approach allows the computation of the characteristic UCS value on the basis of a statistical
analysis. As the theoretical statistical distribution is not always easy to identify (e.g. in the presence of
sub-populations in the histogram of the test results), the UCS characteristic value is computed as the 5%
quantile of the cumulative curve of the results of the UCS tests performed at 28 days of hardening on in-
situ core samples. This approach is only possible if the number of test samples is larger than 20.
Otherwise, a second methodology, which is based on the German standard DIN 4093 (2012), will be
used.
In this second approach, which is valid when the number of test samples is less than 20, the UCS
characteristic value will be computed as the minimum of three values:
- the minimal value of all the test results,
- the arithmetic average value of all the test results multiplied by a constant reduction factor α equal to
0.7.
- a maximum value of 12 MPa.
The constant reduction factor α was determined on the basis of the results of large-scale UCS tests
performed on real-scale soil mix elements and represents the scale effect as discussed in Denies et al.
(2013 and 2014).
In order to consider the heterogeneous character of the soil mix material and the representativity of the
classical test samples (typically 10 cm diameter and 20 cm high), an elimination rule is introduced in the
computational process. For a particular construction site, all samples are tested, then based on the
observations of the test operator, an elimination process is applied according to Ganne et al. (2010) who
have proposed to reject all test samples with unmixed soft soil inclusions that are larger than 1/6 of the
sample diameter, on condition that no more than 15% of the test samples from one particular site would
be rejected. The possibility to reject test samples results from the notion that an unmixed soft soil
inclusion that is 20 mm or smaller does not influence the behavior of a soil mix structure. On the other
hand, an unmixed soft soil inclusion of 20 mm in a test sample of 100 mm diameter significantly
influences the test result. Of course, this condition is only suitable if one assumes that there is no unmixed
soft soil inclusion larger than 1/6 of the width of the in-situ soil mix structure.
Figure 16 summarizes the BBRI methodology to compute the characteristic UCS value of the soil mix
material.
As soon as the UCS characteristic value of the soil mix material at 28 days is determined, the design
value f c,d can be computed:
𝑓𝑓𝑐𝑐,𝑘𝑘
fc,d =αsm β (23)
γSM 𝑘𝑘𝑓𝑓

where:
- α sm is a coefficient that takes the long term effects on the compressive strength and of unfavourable
effects resulting from the way the load is applied (only applied if the lifetime of the soil mix elements
is larger than 2 years) into account. For temporary situations (lifetime < 2 years), α sm is equal to 1 and

171
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

for permanent situations (lifetime > 2 years), α sm is equal to 0.85 according to the SBRCURnet/BBRI
soil mix handbook (2016).

Figure 16: BBRI methodology to compute the characteristic UCS value of the soil mix material

- γ SM is a partial material factor.


- The factor k f , previously mentioned for concrete piles, is defined in the paragraph 2.4.2.5 (2) of
Eurocode 2. k f is considered in order to differentiate the hardening conditions (in laboratory or in-
situ). If the UCS is determined on samples directly cored in the hardened soil mix elements previously
installed in-situ, k f is equal to 1; otherwise k f is equal to 1.1.
- β is a correction factor that takes the age of the soil mix material into account (consult
SBRCURnet/BBRI soil mix handbook 2016 for more information); β is equal to 1 at 28 days of
hardening.
Table 3 summarizes the values used for the different partial factors as used in Belgium and in
The Netherlands according to the guidelines of SBRCURnet/BBRI soil mix handbook (2016).
Table 3: Factors for the computation of the UCS design value, f c,d , of the soil mix material
(SBRCURnet/BBRI soil mix handbook – 2016)
Factors for the computation of the design value f c,d of the soil mix material
(Belgium – The Netherlands)
γ SM α sm kf β
(Persistent-transient/ (temporary/ (UCS based on experience/
accidental ) permanent) UCS tests on in-situ core samples)
ULS DA1-1 Depending on the
(Belgium) age of the material
– see
ULS DA3 SBRCURnet/BBRI
1.50/1.20 1.00/0.85 1.10/1.00
(The soil mix handbook
Netherlands) (2016)
β = 1 at 28
hardening days

2.2.4.4. fc,d for jet grout columns - according to the German standard DIN 4093 (2012)
It can be noted that German Standard DIN 4093 (2012) also proposes a methodology to compute a
characteristic value for the soil mix and the jet grouting material. This methodology is similar to the
simplified approach, as presented in Fig. 16, but with a variable value of α (computed with an iterative
process) and with a maximum value of 10 MPa for the jet grout material. This methodology was notably
discussed in Topolnicki and Pandrea (2012) and in Denies et al. (2013).
It can be noted that all these computational methodologies (ASIRI project, BBRI soil mix project and
DIN 4093) are in line with the philosophy of Eurocodes using partial safety factors.

172
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2.2.5. Edge behavior of the load transfer platform


Stress distribution at the edge of the loading zone, where Prandtl’s failure mechanism does not fully
develop is somewhat different. The horizontal length of development of the Prandtl curve, denoted L max ,
is computed as:
𝜑𝜑′
cos�𝜋𝜋�4− �2� ′ 𝜋𝜋
𝐿𝐿𝑚𝑚𝑚𝑚𝑚𝑚 = 𝜋𝜋 𝜑𝜑′ 𝐷𝐷. 𝑒𝑒 − tan�𝜑𝜑 �∙ �2 (24)
cos� �4+ �2�

As shown in Figure 17(a), Prandtl’s mechanism can fully develop in the load transfer platform if the
overhang length of the footing, denoted L, is greater than L max . The limit pressure at the inclusion head
q p + is then computed as described in the Section 2.2.2 according to the Prandtl mechanism (equation 7).
In the case shown in Figure 17(b), the edge of the inclusion corresponds to the edge of the footing. The
overhang is then zero and the load applied to the footing is nearly fully transmitted on the inclusion head.
The vertical stress on the peripheral soil, which is due to the surrounding ground, is equal to γH.
Considering the decomposition of the engineering issue into an active limit state equilibrium and a
passive limit state equilibrium, as illustrated in Fig. 18, it is possible to obtain the two following equations
describing the role of the stress q in Volume 1 (active earth pressure) and Volume 2 (passive earth
pressure):
𝑞𝑞 = 𝑞𝑞𝑝𝑝+ 𝐾𝐾𝑞𝑞1 (25)
𝑞𝑞 = 𝛾𝛾𝛾𝛾𝐾𝐾𝑞𝑞2 (26)
where K q,1 and K q,2 can be estimated considering the Caquot and Kerisel theory (1966), which is valid for
a weightless medium. Considering equations (25) and (26), it is possible to write q p +/γH = K q,2 /K q,1 =N q *,
which is a bearing factor depending on the friction angles of the material of the load transfer platform
(φ 1 ) and of the surrounding soil (φ 2 ). The N q * values that are computed within the framework of the
ASIRI project (IREX, 2012) considering the Caquot and Kerisel theory (1966) are shown in Table 4.

Table 4: N q * values based on load transfer platform (LTP) and surrounding soil friction angles
Soil φ 2 = 15o Soil φ 2 = 20o Soil φ 2 = 25o Soil φ 2 = 30o
LTP φ 1 N q * (φ 1 )
Nq* Nq* Nq* Nq*
30 18.4 6.98 9.45 13.08 18.43
33 26.1 7.86 10.64 14.71 20.88
35 33.3 8.52 11.53 16.01 22.67
38 48.9 9.68 13.05 18.11 25.80
40 64.2 10.54 14.29 19.71 28.04
As shown in Fig. 17(c), when the footing overhang is between 0 and L max , the limiting pressure at the
inclusion head can be estimated using a linear interpolation between these two extreme values. This is
shown in Fig. 19.
Generally, when more than one inclusion is installed beneath the footing, the edge effect that has been
described is applicable to only a fraction of the inclusion depending on whether the inclusion is at the
footing corner or side. As illustrated in Fig. 20, the edge limit stress, q p +(L), is applicable only to the
exterior portion of the perimeter, the limit stress calculated from Prandtl’s failure mechanism, q p +(P),
applies to the inner portion of the inclusion, and the resulting value must be a weighted average of these
two terms.
By analogy with the distribution of negative friction within a group of piles, IREX (2012) proposes that
limit stress values on the inclusion heads under the footing be determined using the weighting
relationships shown in equations (27) to (31).
For single row inclusions that are shown in Figure 21(a):

1 2 (27)
𝑞𝑞𝑝𝑝,𝑎𝑎 + = 𝑞𝑞𝑝𝑝 + (𝑃𝑃) + 𝑞𝑞𝑝𝑝 + (𝐿𝐿)
3 3

2 1 (28)
𝑞𝑞𝑝𝑝,𝑒𝑒 + = 𝑞𝑞𝑝𝑝 + (𝑃𝑃) + 𝑞𝑞𝑝𝑝 + (𝐿𝐿)
3 3

173
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

(a)

(b)

(c)
Figure 17: Configurations considered for the study of the edge behavior (IREX, 2012)

Figure 18: Decomposition into an active limit state equilibrium and a passive limit state equilibrium
(IREX, 2012)

174
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 19: Principle used to determine the threshold stress on the inclusion head by interpolating
between the extreme values for an overhang greater than L max and a zero overhang (IREX, 2012)

Figure 20: Edge effect combination, modified from ASIRI (IREX, 2012)

For multiple rows of inclusions that are shown in Figure 21(b):

𝑞𝑞𝑝𝑝,𝑖𝑖 + = 𝑞𝑞𝑝𝑝 + (𝑃𝑃) (29)

7 + 5 + (30)
𝑞𝑞𝑝𝑝,𝑎𝑎 + = 𝑞𝑞 (𝑃𝑃) + 𝑞𝑞 (𝐿𝐿)
12 𝑝𝑝 12 𝑝𝑝

5 1 (31)
𝑞𝑞𝑝𝑝,𝑒𝑒 + = 𝑞𝑞𝑝𝑝 + (𝑃𝑃) + 𝑞𝑞𝑝𝑝 + (𝐿𝐿)
6 6

(a) (b)
Figure 21: (a) Edge effect combination for single row of inclusions, and (b) Edge effect combination for
multiple rows of inclusions (IREX, 2012)

2.2.6. Other design aspects


In the ASIRI publication (IREX, 2012), other design aspects, not discussed here, are well considered by
the authors, such as:
- the SLS design approach and the topic of the settlements,
- the design of the foundation slabs installed above the load transfer platform,
- the design of potential geosynthetics installed in the granular load transfer platform,
- the transfer of the lateral loads to the rigid inclusions and the surrounding soil,
- the lateral and flexural behaviors of the rigid inclusions,
- the consideration of seismic loads,
- the soil investigation and testing,
- some execution and quality control (monitoring) aspects,

175
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

- etc.

2.3. Piled embankment – The Dutch approach (CUR-rapport 226)


In the present keynote, the authors mainly present the concept of rigid inclusions combined with a load
transfer platform on the basis of the guidelines published by IREX (2012) as a result of the ASIRI
research project. If there are other guidelines available worldwide focusing on this concept, few
publications present a design methodology in line with the philosophy of the Eurocodes, which is a major
point for European engineers. In order to offer, to the designers, geotechnical alternative solutions
comparable with the classical piling solution, it is quite important to develop design guidelines for the
ground improvement solutions respecting the concepts of the Eurocodes. The engineers have to compare
the two geotechnical solutions (ground improvement and piling) on a similar basis governed by the
principles of the Eurocodes.
The guidelines “Ontwerprichtlijn paalmatrassystemen” of the SBRCURnet (CUR rapport 226, 2010), in
The Netherlands, falls into this category. This publication involves a guideline for the design of piled
embankments with geosynthetic reinforcement generally at the base of the load transfer platform, such as
illustrated in Fig. 22. A large amount of fundamental concepts described in these guidelines can be used
for the design of rigid inclusions combined with a load transfer platform. Subsequent to a survey of the
requirements and the basic principles for the structure as a whole, these guidelines give the design of the
pile foundation, the design of the “mattress” (i.e. the load transfer platform) with the geosynthetic
reinforcements, the execution and the maintenance.
The CUR-rapport 226 (2010) includes many practical design guidelines based on literature studies and on
the results of 2D and 3D finite element calculations. The results of these numerical calculations have been
compared with measurements from practical projects. Considering this comparison, primary directives of
the German EBGEO guidelines, also written in agreement with the Eurocodes, have been adopted.
Indeed, a special interest of the Dutch guidelines is the discussion of different design methodologies
(EBGEO, BS 8006, etc.) and the application and comparison of these methods with practical case studies.
Even if the principles behind piled embankments sometimes differ from the concept of rigid inclusions
combined with a load transfer platform, it can still constitute a source of inspiration for the geotechnical
designers who are specialized in the design of rigid inclusions.
A particular aspect which is deeply investigated in the Dutch approach is the design of the geosynthetic
reinforcements installed at the base of the load transfer platform. This design is performed considering the
particular load transfer distribution that develops in the load transfer platform.

Figure 22: Concept of piled embankment as covered by the Dutch guidelines CUR-rapport 226 (2010)

176
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2.4. Piled embankment - Study of the load transfer distribution


Since many years ago, Suzanne Van Eekelen (Deltares) has studied the nature of load distribution
developing (through the load transfer platform) over the geosynthetic reinforcements and the rigid
inclusions in piled embankments. Van Eekelen et al. (2013) have recently presented a new model, called
concentric arches model, such as illustrated in Fig. 23, to improve the understanding of arching effect in
the load transfer platform and to optimize the design of the concept. In this model, the load is transferred
along the concentric 3D hemispheres towards the geosynthetic reinforcement strips and then via the
concentric 2D arches towards the pile caps.
As reported in Van Eekelen and Bezuijen (2012), the geosynthetic reinforcement strains from the vertical
load are usually calculated in two steps. In the first step, the vertical load is divided into two parts. As
illustrated in Fig. 24, the first part, called “A”, is directly transferred to the piles, and the remainder, is
called “B+C’. Part “A” is relatively large due to arching. EBGEO and CUR-rapport 226 (2010) adopted
Zaeske’s model (Zaeske, 2001) for this calculation step. British standard BS8006 adopted Marston’s
model (1913), and modified it to acquire a 3D model (as described in Van Eekelen et al., 2011 and
Lawson, 2012). In the second step, the vertical load is further divided into load “B” that is transferred
through the geosynthetic reinforcement to the piles and the load “C” that is carried by the subsoil. Loads
“A”, “B” and “C” are all vertical loads.
As described in Van Eekelen and Bezuijen (2012), it is assumed that the strains mainly occur in the
geosynthetic reinforcement strip. Assuming a load distribution on this strip, and the support from the
subsoil (if permissible), the geosynthetic reinforcement strains can be calculated. CUR-rapport 226
(2010) and EBGEO approaches consider a triangular distribution of the load on the geosynthetic
reinforcement strip, while British Standard BS8006 adopts an equally distributed load, and excludes
subsoil support in the calculations. Nevertheless, based on experimental studies, Van Eekelen and
Bezuijen (2012) have highlighted the fact that the distribution of the load on the reinforcement strip
between two piles tends to have the distribution of an inverse triangle (see Fig. 25). To draw this
conclusion, a series of nineteen piled embankment model experiments was carried out in the Deltares
laboratory to understand why the predicted geosynthetic reinforcement strains were always larger than the
measured in-situ geosynthetic reinforcement strains within the framework of actual construction sites.
As shown by the results of this experimental study, there are still research perspectives for the study of
the rigid inclusion concept especially concerning the load transfer distribution developing in the load
transfer platform.

Figure 23: New design model (Concentric Arches model) developed by Van Eekelen et al. (2013)

177
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 24: Load distribution in a piled embankment (Van Eekelen and Bezuijen, 2012)

Figure 25: Measured load distribution on a geosynthetic reinforcement strip in a piled embankment, from
Van Eekelen and Bezuijen (2012) and comparison with the theoretical load transfer distributions
proposed in the different design approaches: British standard BS 8006, ASIRI project (IREX, 2012),
CUR-rapport 226 and German standard EBGEO

178
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3. NUMERICAL MODELING OF RIGID INCLUSIONS


Whilst numerical methods have progressively become the most popular analysis technique, their results
are only as good as the model and the input parameters used for the computation. Moreover blind
implementation of such tools without a deep understanding of the modeling process and suitability of the
characteristic parameters may lead to unforeseen disastrous consequences.
Currently, finite element modeling (FEM) and the finite difference methods (FDM) are globally the most
used numerical methods. Both methods are implemented to solve the hydro-mechanical equations
governing engineering problems (in this case the foundation solution involving a load transfer platform
and rigid inclusions) considered as a continuous medium. The FEM and FDM methods allow the
verification of the bearing capacity and the stability for the modelled geotechnical application.
Nevertheless, the relevance and the accuracy of the numerical results will depend on the relevance of the
constitutive models, on the representativity of the input values adopted for the various material properties
and on the selection of well-suited mesh generation procedures. As underlined in IREX (2012), the
numerical results obtained through these methods yields an approximated solution whose relevance and
accuracy depend on:
- the model constitutive laws of materials and interfaces,
- the mesh generation strategy (discretization of the medium) with the use of finer meshing pattern at
the locations where larger strain field variations are expected,
- the type of elements adopted (number of nodes) and the interpolation laws for each element (e.g.
linear vs. quadratic),
- the use of interfaces between the structural elements and the soil in order to allow the integration of
soil/structure interaction phenomena;
- and the boundary conditions.
It is the authors’ experience that the use of FEM is well-suited for studying the problem when
displacements are small. For large displacements (close to failure or instability), the use of simulations,
including strong “large displacement” assumptions and procedures, becomes necessary (e.g. FDM or the
most-recent developed Material Point Method, called MPM, which is to become soon available in the
commercial software Plaxis 3D).
A great advantage of the numerical methods in ground improvement solutions with rigid inclusions is the
ability to perform parametric studies to optimize the design pattern, especially the height and properties of
the load transfer platform and the quantity and geometrical pattern of the rigid inclusions.
Considering the Dutch and French guidelines (respectively the CUR-rapport 226 and the ASIRI project
report of IREX), the numerical simulations can be conducted with 2D or 3D modelling. According to
IREX (2012), implementation of 2D simulations can be performed with axisymmetric or plane strain
models; both of which have drawbacks. With the axisymmetric configuration, it is only possible to model
a single unit cell (the influence area of an inclusion is considered as a circle) that is located near the
embankment axis. However, the three dimensional inclusion must be transformed into an equivalent 2D
plate in plain strain modelling. In this type of configuration, the rows of inclusion are transformed into
“walls” perpendicular to the sectional plane of the model. Moreover, in 2D models, rigid inclusions may
be represented by either volumetric or beam elements. According to the guidelines of IREX (2012), it is
preferable to model the rigid inclusions using volumetric elements.
Nevertheless, even if 3D models have the advantage of providing a global view of the problem’s
geometry, the high computational time necessary to obtain results for each phase of the project leads the
designers to parsimonious use of 3D simulations in the case of a particular asymmetric geometry, a
reinforced slope, a footing subjected to complex (dynamical) loading or non-uniform loads applied to
slabs on grade. The use of 2D and 3D models is discussed in CUR-rapport 226. Interested readers can
additionally refer to Slaats and van der Stoel (2009) who have conducted a comparative study using
Plaxis 2D and 3D to model a piled embankment with rigid inclusions submitted to lateral forces.
IREX (2012) provides practical guidelines for numerical modelling with consideration of the following
points:
- the definition of the geometric boundaries of the studied domain and the boundary conditions,
- the definition of the constitutive models regarding:
o the identification of the parameters of the problem,
o the hydro-mechanical characteristics of the soils, the load transfer platform, the rigid inclusions
and the potential geosynthetic reinforcements,
o the characteristics of the inclusion/soil interfaces,
o and the characteristics of the slab on grade or raft,

179
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

- and consideration of the construction phases.


Ideally, the numerical model has to be calibrated on the basis of in-situ testing (with soil characterization
results and if possible, the results of a static load test performed on a rigid inclusion).
As underlined in IREX (2012), the realization of numerical calculations has to be performed in large
deformations if geosynthetic reinforcement elements are considered in the problem (second-order effect).
For the purpose of obtaining relevant numerical results, the results must always be considered with
consideration of:
- the basic geotechnical assumptions of the computation model,
- the constitutive models and parameter values used for the different materials (including their drained
or undrained behaviour),
- the incertitude related to the execution mode (probably the most difficult aspect to be assessed in that
kind of problem),
- the different construction phases,
- the long-term behaviour of the construction with
o the use of long-term parameters for the various materials (e.g. the product isochrones curves of
the geosynthetics taking into account the duration of the loading and the lifetime of the structure
to be supported),
o the consideration of the consolidation effects.
A particularity of the numerical methods for the rigid inclusion solutions is the modeling of the
inclusion/soil interface. First, the mesh has to be refined along the inclusion/soil interface. Second, the
constitutive model of the interface has to be considered in detail. As reported in IREX (2012), the
constitutive models typically used for interfaces are of the elastoplastic type. The elastic part allows the
modelling of a gradual mobilization of shear with strain. Two techniques are used for the plastic part:
- either a reduction of φ’ and c’ is applied,
- or a fictitious soil with φ’ = 0 and with a nonzero cohesion for simulating constant friction c’=q s , in
compliance with the limiting values of shaft friction, e.g. as set forth by the French standard for deep
foundations NF P 94-262.
It is to note that this kind of approach should only be conducted by a very experienced geotechnical
engineer, as the use of soil parameters, which do not have a physical specific and special interpretation
can be a very dangerous practice. This particular approach was recently followed by Racinais (2015) who
calibrated soil parameters of a rigid inclusion problem on the Frank and Zhao friction curves (1982) to
obtain better simulations of the rigid inclusion behavior using Plaxis software. Racinais (2015) presented
his numerical model using a case history (plate load test performed on a rigid inclusion in Venette,
France).
Another possibility to design rigid inclusion concepts is the use of homogeneization methods that are only
applicable to vertical loadings. In this method a typical representative unit cell that includes a typical
inclusion, its peripheral soil and load transfer platform volumes is modeled. This unit cell is generally
modeled using an axisymmetric model, and is then compared with a second model that has similar
dimensions but in which the soil and inclusion are replaced with a single homogeneous material. The
characteristics of this material are chosen in such a way to obtain similar results than with the first model.
As highlighted in IREX (2012), this equivalence is typically verified by matching the settlements and
defining an apparent modulus E* for the homogeneous material. The use of this apparent modulus E* is
then extended to model the entire geometry of the problem.
One of the drawbacks of this approach is that it does not consider inclusion/soil interface behavior. The
introduction of a correction coefficient, denoted by β, in the definition of E* can be used in a first
approximation to integrate the interaction at the interface:
𝐸𝐸𝑝𝑝 𝐴𝐴𝑝𝑝
� 𝛽𝛽 +𝐸𝐸𝑠𝑠 𝐴𝐴𝑠𝑠 �
𝐸𝐸 ∗ = (32)
𝐴𝐴

where E p is the deformation modulus of the rigid inclusion, A p is the rigid inclusion area, E s is the
deformation modulus of the soil, A s is the soil area and A is the total considered area. β is thus a
correction coefficient modelling the effect of the inclusion/soil interaction (the sliding behaviour between
both materials at the interface) on the apparent modulus.
A more complex homogeneization method, called the extended biphasic modelling, has been developed
to model the inclusion/soil interface with more accuracy. This method considers the inclusion/soil

180
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

interaction and the modelling of shear and bending forces in the rigid inclusions (Sudret and de Buhan,
2001; Cartiaux et al., 2007; Hassen et al., 2009; Cuira and Simon, 2009; Thaï Son et al., 2009 and 2010).
Regarding the numerical modelling of the load distribution in the load transfer platforms, Van Eekelen et
al. (2013) cite the works of Le Hello and Villard (2009). Le Hello and Villard (2009) focused on load
transfer mechanisms of piled embankments and presented the results of numerical and experimental
studies focusing on the topic of the load distribution. Within this framework, they developed an advanced
numerical model combining 3D Discrete Element Method (DEM) and the Finite Element Method (FEM),
and obtained numerical results for the load distribution in line with the concentric arches model of Van
Eekelen et al. (2013).

4. EXECUTION PROCESSES FOR THE REALIZATION OF RIGID


INCLUSIONS

4.1. Execution methods for the installation of rigid inclusions


As previously mentioned, typical rigid inclusions are concrete columns (possibly installed in the ground
using classical or adapted piling techniques), grout and jet grout columns, soil mix elements (columns,
panels, trenches, blocks…), Controlled Modulus Columns (CMC), grouted stone columns, etc. A list of
rigid inclusions is given in the Table 5 and illustrated in Fig. 26.
Providing a large description of construction processes, the State of the Art Report of Chu et al. (2009)
can be consulted to obtain general information and references concerning all these techniques.
Table 5 does not include typical piling methods – as they are not directly considered as ground
improvement methods – and does not cover deep mixing and jet grouting inclusions – as they are part of
the category D of ground improvement methods dedicated to grouting type admixture techniques.
Nevertheless the realization of rigid inclusions with classical (or adapted) piling methods and with deep
soil mix or jet grout elements certainly represents a growing market in the field of the foundation
engineering.

Table 5. Types of rigid inclusions according to Chu et al. (2009)


Method Description/Mechanisms Advantages Limitations
Controlled modulus columns A borehole is formed by The strength and stiffness of Need special installation
(CMC) pressing and a column of 250 the columns can be controlled. machine
to 450 mm in diameter is The method produces nearly
formed by pressure-grouting. no spoil or vibration.
Multiple stepped pile A borehole is locally enlarged Increase the capacity of grout Used only for soil where
by an opening tool so a or cast in-situ concrete an unsupported borehole
column formed by grout or column without incurring can be formed.
concrete will have enlarged much higher cost.
steps at a given interval.
Grouted gravel or stone A column is formed by Increase the strength of gravel Expensive. Quality control
column forming a gravel or stone or stone columns considerably may be difficult
column and then grouting it by increasing the stiffness of
from the bottom upward using the columns and the interface
a preinstalled grouting tube. friction
Vibro-concrete column Concrete is used to form a Can be used where stone Difficult to control the
column using a method columns are not suitable. uniformity of the column
similar to that for bottom-feed An enlarged bottom can be
dry stone columns. made.
Cast-in-situ, large diameter A large diameter (1 to 1.2 m), More economical and better Need special installation
hollow concrete (PCC) pile hollow concrete pile is cast in- quality control than stone machine
situ using a form of two columns, cement mixing piles
cylindrical casings inserted or concrete piles
into ground.
Y or X shaped pile A grout or concrete pile is Saving cost without Need special installation
formed by compromising bearing machine
inserting a Y or X shaped capacity compared with the
casing as a form circular pile of the same
into ground. diameter

181
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 26: Different types of rigid inclusions, from Chu et al. (2009) and Liu (2007)

For the deep mixing method, the interested reader can refer to the following State of the Art references:
Bruce et al. (1998), Porbaha (1998), Holm (2000), FHWA-RD-99-167 (2001), CDIT (2002),
Eurosoilstab (2002), Terashi (2003), Topolnicki (2004), Larsson (2005), Rutherford et al. (2005),
Essler and Kitazume (2008), ALLU (2010), Denies and Van Lysebetten (2012), Kitazume and
Terashi (2013), FHWA-HRT-13-046 (2013) and GeotechTools (2013), Denies and Huybrechts (2015).
For the design of soil mix elements with a bearing function, the designers can refer to the
SBRCURnet/BBRI soil mix handbook (2016).
For jet grouting columns, the reader can refer to Shibazaki (2003), Essler and Shibazaki (2004), Croce et
al. (2014), Burke (2012) and to the practical guidelines of Maertens et al. (2016).
Within the framework of the present keynote, the authors have still chosen to illustrate the execution of
rigid inclusions within the framework of practical projects involving the CMC and the deep mixing
techniques.

182
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

4.2. Controlled Modulus Columns (CMC) used as rigid inclusions

4.2.1. Execution of Controlled Modulus Columns (CMC)


As shown in Fig. 27 and 28, a Controlled Modulus Column (generally called CMC) is installed in soft
ground using a specially designed auger that is composed of a penetrating helical tip and a cylindrical-like
hollow stem follow-up section. As the auger is thrust and screwed into the soil, the cylindrical extension
displaces the soil laterally, and reduces the amount of spoil that is generated to negligible amounts
compared to cast in-situ piling solutions such as continuous flight auger (CFA) or bored piles. During the
auger extraction, grout is pumped through the hollow stem and auger to form a columnar inclusion with a
diameter that is usually 250 to 450 mm. The strength of the columns can be controlled by varying the
strength of the grout.
The CMC rig should be able to provide a continuous down pull with a high torque in rotation. The torque
is typically in the range of 20 tm, continuous pull down is in the range of 20 t, and rotation speed is in the
order of 15 rpm. Further enhancements to the equipment can include a radio control unit to allow the rig
operator to directly command the concrete pump from his control panel. The control panel displays
torque, speed, depth, down pull force, grout pressure and volume of pumped grout.

4.2.2. Advantages of Controlled Modulus Columns (CMC)


Unlike stone columns whose stability relies on the horizontal containment of the soil (Barksdale and
Bachus, 1983) or deep soil mixing where column strength is dependent on the in-situ soil properties,
CMCs do not rely on external parameters for lateral stability nor are their strengths affected by the
surrounding soil. In fact, column strength can be controlled simply by varying the strength of the grout.
Thus, this method can reduce settlements more efficiently compared to other techniques in which
inclusions are installed in the soil.
As the deformation moduli of CMCs are typically 50 to 3000 times that of the weakest soil stratum
(Masse et al., 2009), it is possible to reduce ground settlements using a lower replacement ratio in
comparison with inclusions that are composed of granular materials (Murayama, 1962 and Aboshi et al.,
1979).
An advantage of the CMC technique, in comparison with stone column type or grouting type techniques,
is its production rate. In order to reach the required size, mixing degree or strength, the ground
improvement techniques involving discontinuous additives (e.g. stone columns) or grouting type
admixtures (e.g. jet grouting or deep mixing) will be time-consuming in comparison with the CMC
technique. For example, introducing and compacting stones for the construction of stone columns must be
carried out in numerous lifts, and time must be allocated to sufficiently enlarge the column and compact
the stones in each lift. Similarly, grout that is produced for jet grouting or deep soil mixing be must
progressively mixed with the soil in place (sometimes with considerable amounts of penetration and
withdrawal phases in the case of the soil mix technique e.g.), which will result in a lower production rate.
CMC grout or concrete is produced in a concrete batching plant, and placement of the material is
performed with the same production rate as obtained for the CFA piling technique.
The amount of vibration that is generated by CMC installation is comparable with CFA piling as the
installation process itself is vibration free. This characteristic can make CMC the preferred choice over
vibratory ground improvement techniques.
The main advantages of CMCs can thus be summarized as:
• strength independency from in-situ soil,
• generally high settlement reduction with lower replacement ratios in comparison to other ground
improvement techniques with additives or inclusions,
• independent from external parameters for lateral stability,
• vibration free installation process,
• installation with lateral displacement of the soil involving negligible volumes of spoil in comparison
with in-situ cast piling systems,
• and high installation rates.

183
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 27: Photography of a CMC auger

Figure 28: CMC auger used for CMC installation, from Chu et al. (2009)

184
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

4.2.3. Case histories relating to the use of Controlled Modulus Columns (CMC)

4.2.3.1. Previous world records for the installation depth of CMCs


In 2006, the CMC method was used for the foundation system of WeeHawken residential resort opposite
Manhattan along the Hudson River in New Jersey, USA. This site had very poor soil conditions with
miscellaneous heterogeneous fills over a thick layer of soft, highly compressible organic clay. The
displacement auger that was used in this project is shown in Fig. 28 and 29. A very large piling rig was
used to install 2100 CMCs with depths varying from 21 m to 30 m, which was the deepest installation
depth for CMC in the world at that time.

Figure 29: CMC installation on the construction site of the WeeHawken project in New Jersey, with the
courtesy of Menard Company

This world record was broken a few years later with the installation of CMCs in Louisiana, USA.
Buschmeier et al. (2012) have reported the installation of 34 m long CMCs to support oil tanks.
As part of the development of an oil terminal located on the banks of the Mississippi River in New
Orleans, five steel tanks, each 12.8 m high have been constructed. Three tanks had diameters of 39.6 m
and the two other tanks’ diameters were of 45.7 m. The maximum pressure that the tanks exert to the
ground is 130 kPa, and an additional pressure of 16 kPa can be added due to the installation of the load
transfer platform.
The superficial fill layer of the load transfer platform that is approximately 0.15 to 1.2 m thick is
underlain by soft to medium stiff silty clay (presenting some traces of organic matter and localised sand
pockets), which extends to depths varying between 4 and 6 m. This layer is followed by very soft clay
(with a silt and sand content) reaching depths ranging between 20 and 24 m. A thin sand layer was also
identified at approximately 21 m depth. Medium stiff to stiff clay (with fine sand pockets and shell
fragments) was observed up to 32 m depth followed by stiff to very stiff silty to sandy clays over a very
dense layer of silty sands at a depth of about 34 m. The groundwater level was located at less than 1 m
below the ground level.
Project specification stipulated tanks’ maximum and central settlements as given in Table 6.

185
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Table 6: Tank Performance Criteria (Monitoring Period = 3 years after hydrotest)


Tank bottom Uniform
Tank Center Deflection
settlement Settlement
50% of API 653
Steel bottom 100 mm 200 mm
Standard
CMCs were installed with a higher replacement ratio down to the depth of about 21 m where a sand layer
with reduced compressibility was identified, and with a lower displacement ratio down to the maximum
treatment depth at about 34 m. Figure 30 illustrates the installation of the CMCs on the construction site.
Due to the variations of the soil profile, it was necessary to design each tank individually. As shown in
Fig. 31, analyses included three dimensional digital modeling of a quarter of a tank, three dimensional
modeling of a thin slice of the tank and analytical calculations of rafts on floating piles. Details and
results of the numerical simulations are given in Bushmeier et al. (2012).
Figure 32 presents a schematic view of the 3D slice model concept finally used for the design of all the
tanks. As a result of these numerical analyses, CMCs were installed to depths of 21 and 34 m with
diameters that were respectively 318 and 470 mm. Installation depth was variable for each tank.

Figure 30: Installation of the CMCs on a construction site (oil terminal) located on the banks of the
Mississippi River in New Orleans, USA

186
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 31: Results of 3D FEM analysis for quarter of tank model (left) and thin slice model (right), from
Buschmeier et al. (2012)

Figure 32: 3D thin slice model concept used for the design of all the tanks, from Bushmeier et al. (2012)

4.2.3.2. Current world record for the installation depth of CMCs


Hamidi et al. (2016) have recently reported the installation of 42 m long CMCs that, to the knowledge of
the authors, are the world’s deepest CMCs ever executed. This project is also located near New Orleans,
and includes four oil tanks, a water tank, two shop and maintenance buildings and ancillary structures.
The diameter and height of the oil tanks are respectively 43.3 m and 11 m, and each tank will be filled
with a product that will apply a design pressure of 120 kPa to the bottom of the tank.
Prior to construction, the site was relatively level and approximately at elevation ±0 m RL (reduced
level). Initially, the uppermost 0.3 m of the ground was treated and modified to cement-stabilised clay.
The site was then elevated with sand to +1.2 m RL. The tanks will be built on a pad that has been further
raised by 0.3 m.
The upper 1.5 to 3 m thick layer of ground consisted of a crust of desiccated over consolidated clay with
an over consolidation ratio (OCR) of 4. Below this layer was a very soft clay layer that extended to depths
of approximately 33.5 m. The OCR for this layer was assumed to decrease with depth from 3 in the upper
part to 1.2 in the lower layers. A sand layer was present at depths of approximately 33 m to 36 m in some
areas, but in other areas this layer was replaced by a stiff to medium stiff layer of clay roughly up to
depths of 51 to 57 m. This lower clay was understood to have an OCR of 1.1. Cone penetration test (CPT)
profiles at the location of the four tanks showed that the cone resistance was almost consistently very low
and negligible in the soft clay. The CPT and soil profile of one of the tanks in shown in Fig. 33.
Initial calculations indicated that the tanks were susceptible of undergoing settlements in the magnitude of
1.5 m to 1.8 m without implementation of an improved foundation system.

187
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Since expected settlements exceeded the tank’s design criteria, limiting long term settlements under the
tanks to 300 mm and differential settlements to the values specified in API 653 (2001), specific
geotechnical measures had to be implemented. The specialist contractor who was awarded the project
proposed that three oil tanks be supported at the edge by a concrete slab and a geotextile-reinforced gravel
ring wall while the fourth tank was designed to be supported by a load transfer platform and a geotextile-
reinforced gravel ring wall. All structures were to be supported by square grids of CMCs with typical
centre to centre spacing of 1.7 to 2.5 m and diameters of 396 mm.
The design included an iterative approach and finite element analyses using a combination of three
different modelling techniques; i.e. axi-symmetrical modeling, three dimensional strip modeling, three
dimensional global modeling. The calculation process was iterative as parameters needed to be adjusted
in such a way that the various types of models yielded similar results. The purposes of these various
computational models are described in Hamidi et al. (2016). This approach led to approximately 280 mm
of long term settlement under the centre of the tank, of which, 75% occurred below the toes of the CMCs
rather uniformly.
As a result of the numerical analyses, more than 2600 CMCs were installed under the structures to an
average depth of 36.5 m during a period of approximately 4 months.
Installing CMCs, with a specially adapted equipment presenting a drilling capacity up to 45m depth,
down to the depth of approximately 42 m in this project has increased the world record described in
Bushmeier et al. (2012) by 8 m.

Figure 33: CPT and soil profile of one of the tanks (note: units in empirical system)

188
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

5. HYBRID CONCEPT OF FOUNDATIONS WITH RIGID INCLUSIONS


INSTALLED WITHOUT LOAD TRANSFER PLATFORM

5.1. Case history – construction of an eleven-story building in Leuven (Belgium)

5.1.1. Introduction to the project: the REGA-Instituut (KU Leuven)


The present case history illustrates the use of soil mix panels as an alternative to pile groups for the
foundation of an eleven-story building in Leuven (Belgium).
This building is the REGA-Instituut of KU Leuven. A few years ago, the university has actually decided
on the construction of a new laboratory for microbiology and medicinal chemistry. The project consisted
of a L-shaped building, such as illustrated in Fig. 34. The project includes a lot of laboratory facilities, an
animal house, several offices, an underground parking, a lunch room, meeting rooms and a 180-person
amphitheater. The construction of this L-shaped building is complicated by the realization of an
underground tunnel (for the supply of the laboratories) connecting the REGA-Instituut to an adjacent
research building of KU Leuven.

Figure 34: Project presentation for the REGA-Instituut of KU Leuven - with the courtesy of SVR-
Architects

189
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

5.1.2. CSM-panels used as bearing elements as alternative to pile groups


Originally, the foundation of the building consisted of a number of pile groups supporting the columns or
pillars of the building frame. Figure 35 presents a schematic view of a part of the initial foundation plan.
The pile groups supporting the pillars of the building frame are represented on the foundation plan as well
as the “supply” tunnel. At the beginning of the project, this concept was discussed and an alternative was
proposed with the installation of soil mix panels (realized with the CSM technique) as a replacement for
the pile groups. Indeed, this solution was cost-effective and presented the advantage to decrease the
constraints related to the consideration of the (potential seismic) lateral loads imposed in the design
requirements and job specifications for the building.

Figure 35: Schematic view of a part of the original foundation plan of the REGA-Instituut with the
different pile groups supporting the pillars of the building frame and the connection tunnel

190
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 36 presents the new concept of foundations made of CSM-panels. As illustrated in Fig. 36, unique,
double and triple adjacent CSM-panels were executed in place of the pile groups. In the following
paragraphs, the terminology “CSM-bearing elements” is used to describe the three foundation patterns
(unique, double and triple CSM-panels). Figure 37 presents an aerial view of the construction site during
the foundation works. Figure 38 presents two views of the CSM-panel caps after execution during
summer 2013 by the Belgian company Soetaert nv.

Figure 36: Schematic view of a part of the new foundation plan of the REGA-Instituut with the different
soil mix panels (CSM-panels) supporting the pillars of the building frame and the connection tunnel

191
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 37: Aerial view of the construction site during the execution of the foundation works (Google view
in July 2013)

192
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 38: Different views of the CSM-panel caps after execution during summer 2013

193
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

5.1.3. Design of the CSM-panels as bearing elements


The design of the CSM-bearing elements will be performed as follows. First, the CSM-bearing elements
were not directly connected to the slabs of the building pillars allowing lateral movements on the top of
the soil mix panels in case of seismic lateral loading. No rigid connections with steel reinforcement is
thus foreseen between the tops of the CSM-bearing elements and the slabs supporting the building pillars.
The CSM-panels were therefore mainly designed to support vertical compressive loads. For that reasons,
no steel reinforcement was placed in the different CSM-bearing elements during execution.
Considering the structural design of the CSM-bearing elements, it was imposed that the UCS design value
of the soil mix material was always larger than the structural vertical stress (mainly due to the weight of
the building): f c,d > q. The design value of the UCS of the soil mix material was verified with the help of
lab tests performed on cored samples (the boreholes are visible in Fig. 38 at the top of several CSM-
bearing elements). A vertical coring on the full depth of, at least, one CSM-bearing element was required
in order to control the UCS of the soil mix material until the base of the bearing element. The design
methodology of BBRI, presented in Section 2.2.4.3 of the present keynote, was followed for the
computation of f c,d .
The geotechnical design of the CSM-bearing elements was verified only considering the base resistance
of the element. This base resistance (= toe resistance) was computed on the basis of CPT results
according to the De Beer procedure, as required in the Belgian National Annex of Eurocode 7 for piles
(see Rapport 12 of BBRI). No shaft resistance was considered in the calculation of the bearing capacity of
the different CSM-bearing elements. According to this design approach, the different CSM-bearing
elements were considered as a pile foundation with only the base resistance acting in the computation of
the bearing capacity. That assumption will be later discussed considering the dimensions of the different
CSM-bearing elements.
Figure 39 presents the different design configurations for the installation of the unique, double and triple
CSM-panels. The three diagrams present the variation of the cone resistance, q c (MPa), and the base
resistance, q b (MPa), computed according to the De Beer approach, as a function of the depth respectively
for the three patterns (unique, double and triple). The q c -curve is therefore the same on the three
diagrams; only the q b -curve varies as a result of the different dimensions of the CSM-bearing elements.
According to the rapport 12 of the BBRI (2009), the base resistance (R b ) was computed for each CSM-
bearing element as follows:
𝑅𝑅𝑏𝑏 = 𝜖𝜖𝑏𝑏 𝑞𝑞𝑏𝑏 𝐴𝐴𝑏𝑏 𝛼𝛼𝑏𝑏 𝛽𝛽 (33)
where ε b (-) is different from 1 in tertiary clay (in the present case, the CSM-panels are installed in
tertiary sand: ε b = 1), A b (m²) is the pile base, q b (MPa) is the base resistance according to the De Beer
approach, α b (-) is an empirical factor depending on the execution process (in this case α b = 0.5 as for the
bored piles) and β is a shape-factor taking into account the shape of a non-circular pile base:
𝐵𝐵
1+0.3 2
𝐵𝐵1
𝛽𝛽 = (34)
1.3

where B 1 and B 2 are respectively the longer and the shorter side of the pile base.
A design value for the bearing capacity, R c,d , (only based on the base resistance) was then computed
according to the principle of Eurocode 7 (with the application of the different partial factors) and this
value was compared, for each CSM-bearing element, with the vertical compressive stress due to the
structural loads:
𝑅𝑅𝑏𝑏
𝑅𝑅𝑐𝑐,𝑑𝑑 = < 𝑞𝑞 (35)
𝛾𝛾𝑏𝑏 𝛾𝛾𝑟𝑟𝑟𝑟 𝜁𝜁

where γ b is the partial factor on base resistance (in this case: γ b = 1.35), γ rb is the model factor (in this
case: γ rb = 1.15) and ζ is the correlation factors taking into account the density of the soil tests on the
construction site (in this case: ζ = 1.2).
For example, considering a triple CSM-bearing element (as represented in Fig. 39), it is possible to
compute the value of R c,d considering the base area of the triple element (A b = 4.32 m²), the value of q b at
the foundation base (q b = 6.67 MPa), the value of β according to equation (34) (β = 0.94). The design
value of the base resistance of a triple CSM-bearing element is thus equal to 7287 kN (~730 tons per
element).

194
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 39: Different design configurations for the installation of the unique, double and triple CSM-
panels, as installed on the construction site of the REGA-Instituut

195
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Especially due to the importance of their base areas, the different CSM-bearing elements have brought a
cost-effective solution to the design of the building foundations. Considering this case history, it is still
possible to highlight the following discussion topics.
In this kind of design concept, the slabs under the building pillars are just installed on the top of the CSM-
bearing elements. There is no load transfer platform foreseen to spread the stresses. That involves a direct
transmission of the vertical (and potential lateral) loads on the top of the CSM-bearing elements.
The compressive stresses will be therefore maximum at the top of the CSM-bearing elements. As a
consequence, considering this kind of concept, it will be therefore necessary to provide a CSM-bearing
element presenting a high quality material on the first meter. And this is not an easy task. Indeed,
Ganne et al. (2010) have observed on different construction sites that the strength of the soil mix material
over the first meter is strongly influenced by the execution process (e.g. infiltration of rinsing water and
falls of lumps of earth in the fresh executed soil mix panel). With this kind of concept, a special attention
has to be brought during the execution process to ensure a full-quality product on the first meter on the
soil mix elements.
Another point of discussion concerns the geotechnical design of such kind of bearing elements. In the
present case history, the CSM-bearing elements, formed of several CSM panels, have been computed as
pile foundations only considering a base resistance and taking into account an installation factor (α b =
0.5) used for the bored piles according to the Rapport 12 of BBRI (Belgian National Annex of Eurocode 7
for piles).
One can still question that approach considering the dimensions of each CSM-bearing elements.
Theoretically, considering its dimensions, a bearing element will respectively be computed as a shallow
foundation, a pile foundation or a pier foundation (sometimes called “caissons”). Each type of foundation
actually presents his own design approach corresponding to his own failure mechanism. And the value of
the bearing capacity obtained with a design approach or another will sometimes lead to severe
differences.
In Belgium, for example, according to the revised Rapport 12 of BBRI and the guidelines for shallow
foundations (both in the pipeline for publication), a foundation element will be considered as
- a pile foundation as soon as the ratio pile length/diameter is larger than 5;
- a shallow foundation as soon as the ratio footing depth/width is smaller than 2.5;
- a pier foundation (generally called “puit” or “put” in Belgium) once the ratio footing depth/width is
ranging between 2.5 and 5.
According to the Belgian National Annexes of Eurocode 7:
- The foundation pile will be computed according to the De Beer procedure based on the CPT results
and particularly on the cone resistance.
- The design value of the bearing capacity of a shallow foundation will be computed considering the
formulae (36) and (40) respectively in drained and undrained soil conditions.
In drained conditions:
𝑅𝑅𝑑𝑑 1
= 𝑐𝑐 ′ 𝑁𝑁𝑐𝑐 𝑏𝑏𝑐𝑐 𝑠𝑠𝑐𝑐 𝑖𝑖𝑐𝑐 𝑔𝑔𝑐𝑐 𝑑𝑑𝑐𝑐 + 𝑞𝑞 ′ 𝑁𝑁𝑞𝑞 𝑏𝑏𝑞𝑞 𝑠𝑠𝑞𝑞 𝑖𝑖𝑞𝑞 𝑔𝑔𝑞𝑞 𝑑𝑑𝑞𝑞 + 𝛾𝛾′𝐵𝐵′𝑁𝑁𝛾𝛾 𝑏𝑏𝛾𝛾 𝑠𝑠𝛾𝛾 𝑖𝑖𝛾𝛾 𝑔𝑔𝛾𝛾 𝑑𝑑𝛾𝛾 (36)
𝐴𝐴′ 2

where A’ (m²) is the effective design surface of the foundation element and B’ (m) is the effective
width of the foundation element. c’ (kPa) is the effective cohesion in drained conditions, q’ (kPa) is
the effective vertical design stress near the foundation at the foundation base and γ’ (kN/m³) is the
effective volumetric weight of the ground under the foundation. N c , N q and N γ are the bearing
capacity factors.
According to Chen (1975), N q and N c can be defined as follows:
𝜋𝜋 𝜑𝜑
𝑁𝑁𝑞𝑞 = 𝑒𝑒 𝜋𝜋𝜋𝜋𝜋𝜋𝜋𝜋𝜋𝜋 𝑡𝑡𝑡𝑡𝑡𝑡² � + � (37)
4 2

and
𝑁𝑁𝑐𝑐 = �𝑁𝑁𝑞𝑞 − 1�𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 (38)
and according to Eurocode 7, N γ will be defined as:
𝑁𝑁𝛾𝛾 = 2�𝑁𝑁𝑞𝑞 − 1�𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 (39)

196
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

The factors b c , b q and b g depend on the gradient of the foundation element. s c , s q and s g are the form
factors. i c , i q and i g depend on the inclination of the load and the factors g q , g c and g γ are given for an
inclined ground surface. The d-factors are the footing depth factors of Meyerhof.
In undrained conditions:
𝑅𝑅𝑑𝑑
= (𝜋𝜋 + 2)𝑐𝑐𝑢𝑢 𝑏𝑏𝑐𝑐 𝑠𝑠𝑐𝑐 𝑖𝑖𝑐𝑐 + 𝑞𝑞 (40)
𝐴𝐴′

where c u (kPa) is the undrained cohesion.

- The bearing capacity of a pier foundation will be computed as the average of the two following values:
o the bearing capacity of a shallow foundation using Meyerhof depth factors calculated for a ratio
footing depth/width equal to 2.5;
o the bearing capacity of a bored pile without shaft resistance.

If this is a clear computational procedure which allows to determine a bearing capacity for the three kinds
of foundation generally considered in practice, the design procedure proposed for the pier foundations in
the Belgian National Annex of Eurocode 7 mainly concerns the realization of foundations of intermediate
dimensions ranging between the typical dimensions of a shallow foundation and a pile. The design of pier
foundations of larger dimensions, as executed on the construction site of the REGA-Insituut, for example,
could be regarded with a further attention. As illustrated in Fig. 39, on this construction site, the different
CSM-bearing elements can be considered as pure shallow foundation (e.g. in the case of elements with L
= 1.2 m and B 2 = 0.55 m; ratio = 2.1), as typical intermediate pier foundation (e.g. elements with L = 4.85
m and B 2 = 1.2 m; ratio = 4) and as pile foundation (e.g. elements with L = 8 m and B 2 = 0.55 m; ratio =
14.5).
Considering the CSM-bearing elements installed in triple configuration, the larger element presents the
following dimensions: L = 8 m and B 2 = 1.8 m. This element will be considered as a pier foundation and
its bearing capacity will be computed without taking into account the friction resistance arising along its
shaft according to the Belgian National Annex of Eurocode 7. Nevertheless, considering the theory
behind the drilled shaft foundations (presenting similar dimensions and installed with comparable method
than the bored piles), the shaft resistance really plays an important role in the bearing capacity of such
large bearing elements (Kulhawy, 1991). In the Foundation Engineering Handbook (1991), Kulhawy
gives a methodology to design this kind of foundation. The particularities of this approach, well-adapted
to deep foundations presenting a large diameter or large horizontal dimensions, are:
- the consideration of the weight of the foundation element in the design,
- the computation of the base resistance considering the bearing capacity equations (36) and (40) with
potentially adapted bearing factors,
- the consideration of the side resistance.
In the case of the REGA-Instituut, a safety design approach was followed with success. But it will be
really interesting in the future to monitor this kind of bearing element for the purpose of analysing the
global behaviour of the foundation with regard to the development of a shaft friction or not along the soil
mix elements. Moreover, in this kind of hybrid foundation concept, with rigid inclusions but without a
load transfer platform, the role of the surrounding soil in the design should be investigated. The way the
load is transferred from the slab on grade or from the structure to the bearing elements and possibly to the
surrounding soil has to be clarified.
To conclude this paragraph, it can be noted that the design of alternative bearing elements by the help of a
particular design method should always be performed regarding the assumptions behind this design
method; in particular considering the dimensions of the elements. A bearing element will often be
designed as a shallow or a pile foundation only regarding the value of its dimensions. But a special
attention should be brought to the use of a design method when these dimensions or the ratios between
these dimensions are close to the boundaries of the field of application of the method. To complicate this
issue, the boundaries between a pile, a pier or a shallow foundation (in terms of definitions) are often
different between countries certainly involving a lack of clarity for the geotechnical designers playing
with new alternative techniques such as the use of soil mix panels or columns as bearing elements.

197
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

6. CONCLUSION
A current trend on the European market is the use of ground improvement concepts as alternative to
typical piling methods (e.g. the use of soil mix panels or columns as bearing elements in place of piles).
The lack of design requirements concerning the ground improvement techniques was, for a long time, an
obstacle to the development of such techniques on the European market - subjected to a strict control by
means of the Eurocodes and the European standards - and sometimes leads to severe discussion between
the GI contractors and the pile contractors denouncing a double standard politics on the foundation
market. Nevertheless, this last decade, several design methods have been established for ground
improvement techniques in line with the Eurocodes.
The present keynote concentrates on the presentation of the rigid inclusion concept involving the use of a
load transfer platform and it particularly highlights the content of the ASIRI guidelines (IREX, 2012).
The authors explain the failure mechanisms associated to this concept and the role of the load transfer
platform considering the stress domain, the geotechnical and the structural limit states associated to the
rigid inclusions. It is to note that the design of rigid inclusion concepts according to this approach falls
within the philosophy of the Eurocodes.
The authors also refer to the Dutch guidelines (CUR rapport 226, 2010) involving the design of piled
embankments and the installation of a geosynthetic reinforcement generally at the base of the load
transfer platform. The consideration of these guidelines certainly provides an additional information to
the geotechnical designer interested in the design of rigid inclusion concepts. Particular aspects which are
deeply investigated in the Dutch approach are the design of the geosynthetic reinforcements installed at
the base of this load transfer platform and the study of the load transfer distribution developing in this
load transfer platform. The recent concentric arches model of Van Eekelen et al. (2013) is highlighted and
the distribution of the line load on the reinforcement strip between two piles is discussed.
In the present keynote, the author focus on the use of rigid inclusions as an alternative to vertically loaded
pile foundations. It can be noted that the design principles, presented in Section 2, can be extended to the
application of lateral loads on the slabs on grade of the structure. Indeed, the load transfer platform will
have a positive effect on the transmission of the lateral loads to the rigid elements.
Numerical modelling of the rigid inclusion concept is treated considering the guidelines of IREX (2012)
underlining the relevant parameters and assumptions to be used for the modelling. Advantages and
drawbacks of the 2D- and 3D-modeling are discussed regarding the complexity of the design.
A particularity of the numerical methods modelling the rigid inclusion solutions is the consideration of
the inclusion/soil interface. As reported in IREX (2012), the constitutive models typically used for
interfaces are elastoplastic. The elastic part allows the modelling of a progressive mobilization of the
shear with strain. As for the plastic part, the authors report the recent work of Racinais (2015) who
models a fictitious soil with φ’ = 0 and with a nonzero cohesion for simulating constant friction c’=q s , in
compliance with the limiting values of shaft friction set forth by the French standard for deep foundations
NF P 94-262.
Racinais (2015) calibrates the soil parameters of a rigid inclusion problem on the Frank and Zhao friction
curves (1982) in order to obtain better simulations of the rigid inclusion behaviour with the help of the
Plaxis software for a practical case.
After design considerations, the authors review several execution techniques allowing the installation of
rigid inclusions and they focus on impressive case histories illustrating the realization of Controlled
Modulus Columns (CMCs) at large depths in USA.
In the last part of the keynote, the authors finally report a case history in Leuven (Belgium), where CSM-
panels were used as soil mix bearing elements as an alternative to pile groups to support the pillars of an
eleven story-building. The design of these CSM-bearing elements is discussed regarding the nature of the
bearing elements and their dimensions. Structural and geotechnical designs are both considered in line
with the principles of Eurocode 7. A particular attention is given to the suitability of a design method with
regard to the dimensions of the bearing elements.
If the use of rigid inclusions in combination with a load transfer platform is now well-established, the
hybrid foundation concept, such as presented in the last part of the keynote, with the use of CSM-panels
in place of pile groups has to be more deeply investigated. In this concept, no load transfer platform is
used to transfer the structural loads to the foundation elements. The soil has probably (?) no more any role
in the bearing capacity and in the mitigation of the settlement contrarily to the concept of the rigid
inclusions with a load transfer platform. Further research has to be conducted to investigate the real
behaviour of such bearing elements in the ground (in particular with regard to the shaft resistance).

198
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

In the case of the REGA-Instituut, the CSM-panels were not reinforced with the help of steel beams; only
vertical loads applying on the top of the CSM-bearing elements. Nevertheless, the reader can refer to the
SBRCURnet/BBRI soil mix handbook (2016) to obtain more information on the design requirements of
soil mix elements used with a bearing function. In this handbook, the design of the bearing capacity of the
soil mix element is detailed according to a procedure in line with Eurocode 7 and requirements are given
for the corrosion protection of the steel beams in agreement with the content of the EN 1993-5 for the
piles.
The respect of guidelines (IREX, 2012; CUR-rapport 226, 2010; SBRCURnet/BBRI soil mix handbook,
2016…) in line with the Eurocodes should lead to a growing and sustainable use of alternative ground
improvement techniques on the market with a positive competition between the techniques and the
contractors.
In order to conclude, if ground improvement and piling techniques are often used in competition,
geotechnical designers begin to use them in combination. Indeed, we move towards geotechnical designs
more and more based on the functions and lifetime of each geotechnical element combining them to reach
an optimized design. For example, in 2013, Yamashita et al. report measurements performed underneath a
piled raft completed with grid-form deep cement mixing walls installed to reduce the risks of structural
damages potentially caused by liquefaction in case of earthquake. The structure was a twelve-storey
office building. Figure 40 presents a schematic view of the building and foundation. Figure 40 illustrates
the layout of piles and grid-form deep cement mixing walls. The load distribution between the piles, the
soil mix walls and the surrounding soil has been monitored during a period of three years. As a result of
the monitoring measurements, 70 % of the load was taken by the piles, 14 % by the soil mix walls and
15% by the soil. A settlement of 20 mm was observed after three years of monitoring. The measurements
also learned that the magnitude 9.0 Tohoku earthquake (on March 11, 2011) had almost no influence on
the settlements and on the load distribution. As for this case, it is certain that in a near future, the
combination of ground improvement with typical piling techniques will be more and more regarded in
view of optimizing design and costs of construction project.

199
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 40: Schematic view of the building and foundation (Yamashita et al., 2013)

Figure 41: Layout of piles and grid-form deep cement mixing walls (Yamashita et al., 2013)

200
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

REFERENCES
Aboshi, H., Ichimoto, E. & Harada, K. 1979. The Compozer: A Method to Improve Characteristics of Soft
Clays by Inclusion of Large Diameter Sand Columns. International Conference on Soil Reinforcement;
Reinforced Earth and Other Techniques, 2, Paris, 20-22 March, 211–216

ALLU. 2010. Mass Stabilisation Manual, ALLU Finland Oy

American Petroleum Institute. 2001. API Standard 653: Tank Inspection, Repair, Alteration, and
Reconstruction, 3rd Edition, 112

Barksdale, R. D., and Bachus, R. C. 1983. "Design and Construction of Stone Columns, Volume 1,
FHWA/RD-83/026." 194

BBRI CSTC WTCB. 2009. Rapport 12. Directives pour l’application de l’Eurocode 7 en Belgique.
Partie 1 : dimensionnement géotechnique à l’état limite ultime de pieux sous charge axiale de
compression. Richtlijnen voor de toepassing van de Eurocode 7 in België. Deel 1: het grondmechanische
ontwerp in de uiterste grenstoestand van axiaal op druk belaste funderingspalen

BBRI Soil Mix project. 2009-2013. IWT 080736 soil mix project. SOIL MIX in constructieve en
permanente toepassingen – Karakterisatie van het materiaal en ontwikkeling van nieuwe mechanische
wetmatigheden. See Denies et al. (2012) for more information

Bruce, D. A., Bruce, M. E. C. and DiMillio, A. F. 1998. Deep mixing method: a global perspective. ASCE,
Geotechnical special publication, n°81, pp. 1-26

Bowles, J. E. 1996. Foundation Analysis and Design, 5th Ed., New York, McGraw Hill, 1175

BS 8006-1. 2010. Code of practice for strengthened/reinforced soils and other fills

Burke G.K. 2012. The State of Practice of Jet Grouting. ASCE Geotechnical Special Publication No. 228.
Proceedings of the 4th Interantional Conference on Grouting and Deep Mixing. Pp. 74-88

Buschmeier, B., Masse, F., Swift, S. & Walker, M. 2012. Full Scale Instrumented Load Test for Support of
Oil Tanks on Deep Soft Clay Deposits in Louisiana Using Controlled Modulus Columns. International
Symposium on Ground Improvement (IS-GI) Brussels 2012, 3, Brussels, 31 May - 1 June, 359-372

Burland, J. B. 1973. Shaft Friction Piles in Clay - a Simple Fundamental Approach. Ground Engineering,
6, 3, 30-42

Caquot A., Kerisel J. 1966. Traité de mécanique des sols, Gautier-Villars

Cartiaux F.-B., Gellee A., Buhan (de) P., Hassen G. «Modélisation multiphasique appliquée au calcul
d’ouvrages en sol renforcé par inclusions rigides ». Revue française de géotechnique, n° 118, 2007, p.
43-52.

CDIT. Coastal Development Institute of Technology. 2002. The Deep Mixing Method – Principle, Design
and Construction. Edited by CDIT, Japan. A. A. Balkema Publishers/Lisse/Abingdon/Exton (PA)/Tokyo

Chen, W. F. 1975. Limite analysis and soil plasticity, Elsevier, Amsterdam.

Chu, J., Varaksin, S., Klotz, U. & Mengé, P. 2009. State of the Art Report: Construction Processes. 17th
International Conference on Soil Mechanics & Geotechnical Engineering: TC17 meeting ground
improvement, Alexandria, Egypt, 7 October 2009, 130

Combarieu, O. 1988. Amélioration des sols par inclusions rigides verticales. Application à l'édification
de remblais sur sols médiocres. Revue Française de Géotechnique, 44, 57-79

Croce, P., Flora, A. & Modoni, G. (2014). Jet Grouting Technology, Design and Control, CRC Press,
Taylor & Francis Group.

201
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Cuira F., Simon B. 2009. Deux outils simples pour traîter des interactions complexes d’un massif
renforcé par inclusions rigides. Proc. 17th ICSMGE, Alexandrie, M. Hamza et al. (Eds.), IOS Press,
2009, p. 1163-1166.

CUR-rapport 226. 2010. Ontwerprichtlijn paalmatrassystemen.

Das, B. M. 2009. Shallow Foundations Bearing Capacity and Settlement, 2nd Ed, Boca Raton, FL, USA,
Taylor & Francis Group.

Denies, N., Huybrechts, N., De Cock, F., Lameire, B., Vervoort, A., Van Lysebetten, G. and Maertens, J.
2012. Soil Mix walls as retaining structures – mechanical characterization. International symposium of
ISSMGE - TC211. Recent research, advances & execution aspects of ground improvement works. 31
May-1 June 2012, Brussels, Belgium

Denies, N. and Van Lysebetten, G. 2012. General report Session 4 – Soil Mixing 2 – Deep Mixing.
Proceedings of the International symposium of ISSMGE - TC211. Recent research, advances & execution
aspects of ground improvement works. N. Denies and N. Huybrechts (eds.). 31 May-1 June 2012,
Brussels, Belgium, Vol. I, pp. 87-124

Denies, N., Van Lysebetten, G., Huybrechts, N., De Cock, F., Lameire, B., Maertens, J. and Vervoort, A.
2013. Design of Deep Soil Mix Structures: considerations on the UCS characteristic value. Proceedings
of the 18th International Conference on Soil Mechanics and Geotechnical Engineering, Paris, France,
2013, Vol. 3, pp. 2465-2468

Denies, N., Van Lysebetten, G., Huybrechts, N., De Cock, F., Lameire, B., Maertens, J. and Vervoort, A.
2014. Real-Scale Tests on Soil Mix Elements. Proceedings of the International Conference on Piling &
Deep Foundations, Stockholm, Sweden, 21-23 May 2014, Eds. DFI & EFFC, pp. 647-656

Denies, N. and Huybrechts, N. 2015. Deep Mixing Method – Equipment and Field of Applications,
Chapter 11 of the book: Ground Improvement Case Histories. Elsevier

DIN 4093. 2012. Design of ground improvement – Jet grouting, deep mixing or grouting. August 2012

EBGEO. 2011. Recommendations for Design and Analysis of Earth Structures using Geosynthetic
Reinforcements – EBGEO. Ernst & Sohn – A Wiley company. DGGT - Deutsche Gesellschaft für
Geotechniek e. V. - German Geotechnical Society

EN 206. 2013. Concrete – Part 1: Specification, performance, production and conformity

EN 1536. 2015. Execution of special geotechnical work - Bored piles

EN 1992-1-1. 2004. Eurocode 2: Design of Concrete Structures – Part 1-1: General – Common Rules
for Building and Civil Engineering Structures

EN 1993-5. 2007. Eurocode 3: Design of steel structures - Part 5: Piling

EN 1997-1. 2004. Eurocode 7: Geotechnical Design - Part 1: General Rules 1

Essler and Shibazaki. 2004. Jet grouting. In M. P. Moseley & K. Kirsch (Eds.), Ground improvement, 2nd
ed., Spon Press

Essler, R. and Kitazume, M. 2008. Application of Ground Improvement: Deep Mixing. TC17
website: www.tc211.be

Eurocode 2 – see reference EN 1992-1-1

Eurocode 7 – see reference EN 1997-1

Eurosoilstab. 2002. Development of design and construction methods to stabilise soft organic soils.
Design Guide Soft Soil Stabilisation. EC project BE 96-3177

202
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

FHWA-RD-99-167. 2001. An introduction to the Deep Soil Mixing Methods as used in geotechnical
applications: verification and properties of treated soil. Prepared by Geosystems (D.A. Bruce) for US
Department of Transportation, FHWA, p. 434

FHWA-HRT-13-046. 2013. Federal Highway Administration Design Manual: Deep Mixing for
Embankment and Foundation Support. US Department of Transportation, Federal Highway
Administration, October 2013, p. 244

Frank, R. & Zhao, S. R. 1982. Estimation Par Les Parametres Pressiometriques de 1'enfoncementt Sous
Charge Axiale de Pieux Fores dans des Sols Fins. Bulletin Laison Laboratoire Central des Ponts et
Chaussees, 119, 17-24.

Ganne, P., Huybrechts, N., De Cock, F., Lameire, B. & Maertens, J. 2010. Soil mix walls as retaining
structures – critical analysis of the material design parameters, Geotechnical challenges in megacities.
International Geotechnical Conference, Moscow, Vol. 3, pp. 991-998.

Geotech Tools. 2013. Geo-construction Information & Technology Selection Guidance for Geotechnical,
Structural, & Pavement Engineers. 2013. http://geotechtools.org/ Copyright © 2010–2013 Iowa State
University

Hamidi, B., Masse, F., Racinais, J. & Varaksin, S. (in print) The Boundary between Deep Foundations
and Ground Improvement. Geotechnical Engineering.

Hamidi, B., Nikraz, H., and Varaksin, S. 2009. "Arching in Ground Improvement." Australian
Geomechanics Journal, 44(4 (December)), 99-108.

Hansen, J. B. 1970. A Revised and Extended Formula for Bearing Capacity. Danish Geotechnical
Institute Bulletin, 28 (successor to Bulletin No 11), 21.

Hassen G., Dias D., Buhan (de) P. 2009. Multiphase constitutive model for the design of
piledembankments: comparison with three-dimensional numerical simulations. International Journal of
Geomechanics, vol. 9, n° 6, 2009, p. 258-266.

Holm, G. 2000. Deep Mixing. ASCE, Geotechnical special publication, N°112, pp. 105-122

IREX. 2012. Recommandations pour la conception, le dimensionnement, l’exécution et le contrôle de


l’amélioration des sols de fondation par inclusions rigides (for the French version). Recommendations
for the design, construction and control of rigid inclusion ground improvements (for the English version).
Projet National ASIRI. Presses des Ponts, ISBN 978-2-85978-462-1 (available in French and English).

Kitazume, M. and Terashi, M. 2013. The Deep Mixing Method. CRC Press/Balkema. Taylor and Francis
Group, London, UK

Kulhawy, F. H. 1991. Drilled shaft foundations. Chapter 14 of the Foundation Engineering Handbook,
second edition. Van Nostrand Reinhold.

Larsson, S.M. 2005. State of practice report - Execution, monitoring and quality control, International
Conference on Deep Mixing, pp. 732-785

Lawson, C.R. 2012. Role of Modelling in the Development of Design Methods for Basal Reinforced Piled
Embankments, Proceedings of EuroFuge 2012, Delft, the Netherlands.

Le Hello, B., Villard, P. 2009. Embankments reinforced by piles and geosynthetics - numerical and
experimental studies with the transfer of load on the soil embankment. Engineering Geology 106, 78-91.

Liu, H.L. 2007. New piling techniques for soil improvement in China, Proc 13 Asian Regional Conf on
Soil mech and Geot Eng. Kolkata.

Maertens, J., De Vleeschauwer, P. and Langhorst, O. 2016. Jet grouting: State of the Art.

203
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Masse, F., Pearlman, S. L. & Taube, M. G. 2009. Controlled Modulus Columns for Support of above
Ground Storage Tanks. 40th Ohio River Valley Soil Seminar (ORVSS), Lexington, Kentucky, November
13.

Meyerhof, G. G. 1951. The Ultimate Bearing Capacity of Foundations. Geotechnique, 2, 4, 301-333.

Murayama, S. 1962. Vibro-Compozer Method for Clayey Ground (in Japanese). Mechanization of
Construction Work, 150, 10-15.

NF P 94-262. 2012. Justification des ouvrages géotechniques. Norme d’application nationale de


l’Eurocode 7 - Fondations profondes.

Okyay, U. S. 2010. Etude expérimentale et numérique des transferts de charge dans un massif renforcé
par inclusions rigides. Application à des cas de chargements statiques et dynamiques. Civil and
Environmental Engineering. Lyon, Institut National des Sciences Appliquées de Lyon, 402.

Porbaha, A. 1998. State of the art in deep mixing technology: part I. Basic concepts and overview.
Ground Improvement, Vol. 2, pp. 81-92

Prandtl, L. 1920. Uber Die Härte Plastischer Körper Nachrichten Von Der Königlichen Gesellschaft Der
Wissenschaften, Gottingen, Math.- Phys. Klasse, pp. 74-85.

Racinais, J. 2015. Calibration of Rigid Inclusion Parameters Based on Pressuremeter Test Results. 16th
European Conference on Soil Mechanics and Geotechnical Engineering (XVI ESCMGE), Edinburgh, 13-
17 September, Presentation.

Rutherford, C., Biscontin, G. and Briaud, J.-L. 2005. Design manual for excavation support using deep
mixing technology. Texas A&M University. March 31, 2005

Salgado, R. 2008. The Engineering of Foundations. McGraw-Hill International Edition.

SBRCURnet/BBRI. 2016. Soilmix-handboek [in Dutch].

Schweiger, H. F. 2002. Results from Numerical Benchmark Exercises in Geotechnics. 5th European
Conference Numerical Methods in Geotechnical Engineering (NUMGE 2002), Paris, 4-6 September,
305-314.

Shibazaki, M. 2003. State of Practice of Jet Grouting. ASCE Geotechnical Special Publication No. 120.
Proceedings of the Third International Conference on Grouting and Ground Treatment, pp. 198-217.

Slaats, H. and van der Stoel, A.E.C. 2009. Validation of numerical model components of LTP by means of
experimental data. Proceedings of the 17th International Conference on Soil Mechanics and
Geotechnical Engineering (ICSMGE 2009).

Sudret B., Buhan (de) P. 2001. Multiphase model for inclusion-reinforced geostructures. Applications to
rock-bolted tunnels and piled raft foundations. International Journal for Numerical and Analytical
Methods in Geomechanics, vol. 25, 2001, p. 155-182.

Thai Son Q., Hassen G., Buhan (de) P. 2009. Dimensionnement sous sollicitation sismique de sols de
fondations renforcés par inclusions rigides. Proc. 17th ICSMGE, Alexandrie, M. Hamza et al. (Eds.), IOS
Press, 2009, p. 606-609.

Thai Son Q., Hassen G., Buhan (de) P. 2010. Seismic stability analysis of piled embankments: a
multiphase approach. International Journal for Analytical and Numerical Methods in Geomechanics, vol.
34, 2010, p. 91-110.

Terashi, M. 2003. The State of Practice in Deep Mixing Methods. Grouting and Ground Treatment (GSP
120), 3rd International Specialty Conference on Grouting and Ground Treatment New Orleans,
Louisiana, USA, pp. 25-49

Terzaghi, S. 1943. Theoretical Soil Mechanics, New York, John Wiley & Sons, 510.

204
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Terzaghi, K., Peck, R. B. & Mesri, G. 1996. Soil Mechanics in Engineering Practice, 3rd Edition, New
York, John Wiley and Sons, 512.

Tomlinson, M. J. 1971. Some Effects of Pile Driving on Skin Friction. Conference on Behaviour of Piles,
London, 107-114.

Topolnicki, M. 2004. In situ soil mixing. In M. P. Moseley & K. Kirsch (Eds.), Ground improvement, 2nd
ed., Spon Press

Topolnicki, M. and Pandrea, P. 2012. Design of in-situ soil mixing. International symposium of ISSMGE
- TC211. Recent research, advances & execution aspects of ground improvement works. 31 May-1 June
2012, Brussels, Belgium

Van Eekelen, S.J.M.; Bezuijen, A. and Van Tol, A.F., 2011. Analysis and modification of the British
Standard BS8006 for the design of piled embankments. Geotextiles and Geomembranes 29 (2011)
pp.345-359.

Van Eekelen, S.J.M. and Bezuijen, A. 2012. Basal reinforced piled embankments in the Netherlands, Field
studies and laboratory tests. International symposium of ISSMGE - TC211. Recent research, advances &
execution aspects of ground improvement works. 31 May-1 June 2012, Brussels, Belgium

Van Eekelen, S.J.M., Bezuijen, A., van Tol, A.F., 2013. An analytical model for arching in piled
embankments. Geotext. Geomemb. 39, 78-102.

Vijayvergiya, V. N. & Focht, J. A. 1972. A New Way to Predict Capacity of Piles in Clay. 4th Offshore
Technology Conference, Houston, paper 1718.

Yamashita, K., Wakai, S. & Hamada, J. 2013. Large-scale piled raft with grid-form deep mixing walls on
soft ground. 18th International Conference on Soil Mechanics and Geotechnical Engineering, 2-6
September, Paris, Vol. 3, pp. 2637-2640.

Zaeske, D., 2001. Zur Wirkungsweise von unbewehrten und bewehrten mineralischen Tragschichten über
pfahlartigen Gründungselementen. Schriftenreihe Geotechnik, Uni Kassel, Heft 10, February 2001 (in
German).

205
206
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

General report
J. Powell, Geolabs & Chairman CEN TC341 “Geotechnical investigation and testing” & Chairman ISO TC182
“Geotechnics”

207
208
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Advances and innovations in measurement techniques


and quality control tools
Noël Huybrechts, Belgian Building Research Institute - BBRI & KU Leuven, Belgium, nh@bbri.be

Monika De Vos, Belgian Building Research Institute - BBRI, Belgium, mdv@bbri.be

Gust Van Lysebetten, Belgian Building Research Institute - BBRI, Belgium, gvl@bbri.be

ABSTRACT
This contribution gives an overview of some recent advances with regard to measurement techniques
applied in geotechnical engineering practice. It focusses mainly on the use of fibre optic technology and
especially on the experience of the Geotechnical Division of the BBRI with the multi-point FBG optical
fibre technique. Since 2007 BBRI applied this measurement technique in many research programs and on
several real job sites. The examples in this paper show that fibre optic sensors or flexible, performant and
accurate measurement devices that can be applied for and integrated in almost all kind of geotechnical
applications where deformation and/or temperature measurements need to be carried out.

1. INTRODUCTION
A large range of measurement techniques and monitoring devices are applied in the field of geotechnical
engineering and piling in particular. In general, one can make a subdivision in measurement techniques
and monitoring that are applied before, during and after the pile installation works.

Before the start of the construction works these techniques are mostly applied in the framework of soil
investigation and soil characterization or in the framework of preliminary testing of piles to evaluate and
study the driveability or buildability, the environmental impact like noise or vibrations of the selected
installation procedure and to determine the pile bearing capacity (e.g. Figure 1).

Figure 1: Full-scale driveability test in Antwerp to evaluate and optimize different installation methods
and to assess environmental impact: settlements, vibrations and noise (de Nijs et. al , 2015).

209
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Measurement techniques that are applied during the work have mostly the aim to control the quality of the
pile installation by monitoring and recording an extended set of execution parameters and to mitigate
environmental effects (noise, vibrations, settlements,…). The last decade it is becoming common practice
to register the relevant execution parameters by means of sensors that are integrated in the piling
equipment and/or drilling tools. Real-time visualization of the measured parameters to the operator of the
piling rig is a very helpful tool to assure the quality of the pile installation. Two examples are given in
Figure 2.
The minimum requirements for quality control measurement are specified in the pile execution standards
EN 1536 and EN 12699. For requirements with regard to admitted vibration or noise levels several
national/local standards or directives are available or in force; e.g. DIN 4150-1 to 3 and the SBR
directives A to C (2006).

Figure 2: Examples of integrated sensors in CFA drilling rig – courtesy of Jean Lutz (left) and DAT
instruments (right).

After pile installation, measurement techniques and monitoring can be applied for several reasons:
− to verify the quality of the concrete and the integrity of the pile, e.g. by means of the sonic echo
method, the mechanical impedance method, cross-hole sonic logging, gamma-gamma logging and
more recently by means of temperature measurements;
− to verify the quality of pile-soil interaction by means of comparative soil investigations (e.g. CPT,
DMT, …) at short distance from the pile shaft and under the pile base;
− to determine the bearing capacity of piles and in particular the normal load distribution in the pile in
order to deduce the shaft resistance and the pile base resistance in the case of axially loaded piles or
the bending moments in the case of horizontally loaded piles;
− to monitor the long term behaviour of the pile foundation.

Several reference documents with regard to the more classical type of measurement devices that can be
used in geotechnical engineering and piling in particular can be found in the literature, amongst others in
Dunicliff (1993) and Hertlein, B. et al. (2006).
The major advances that took place in the ITC sector have found the last decade also their way to the
geotechnical sector, leading to significant ameliorations of the classical measurement devices.
Miniaturization, integrated electronics, wireless data transmission, real time and online visualization, etc.
have been introduced and applied on regular base, - see e.g. (Soga, 2010) and (Van Alboom, 2012) -, but
the most important advancement in the last decade originate, according to the authors’ opinion, from the
fibre-optics technology. Optical fibres are very small and fragile materials and the use of them seems on
first sight not coherent with the generally required robustness of sensors applied in geotechnical
engineering. However, they show some important advantages as illustrated further. This contribution will
therefore focus mainly on recent practical experiences with different types of instrumentation based on
the fibre-optics technology that have been applied in deep foundation elements. The basic principles of
different types of optical fibre technologies are given in the next paragraph. Examples of practical
applications are illustrated in paragraph 3.

210
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2. OPTICAL FIBRE TECHNOLOGY: AN OVERVIEW


Many different fibre optic technologies are available and can be applied as optical sensing devices.
With regard to classical sensing devices these technologies have some specific advantages:
- the optical fibre itself is the sensor and the information carrier for the light waves; generally spoken a
fibre optic sensor works by modulating one or more properties of the propagating/reflected light wave
(intensity, phase, polarization, frequency, …) in response to the parameter (e.g. deformation,
temperature) that is being measured;
- the core of the optical fibres consists of a thin strand of glass (about 10 μm diameter) in which light is
transmitted; glass is an inert material so very suitable in harsh and aggressive environment;
- all the sensing and transmitting components of a fibre optic sensor are non-electrical, so disturbance of
the measurements due to electro-magnetic interference, stray currents, corrosion or short circuits due
to e.g. water infiltration are not an issue; they can also be applied in an explosive environment;
- the measured reflections of the light waves are stable on the long-term (no zero drift);
- in the case of multiplexed or distributed optical sensors a multitude of sensor points are available on
one single optical fibre. Above that, the dimensions of optical fibres are so small that they can easily
be integrated in all kinds of geotechnical bearing elements. With some basic knowledge of the
technology it is possible to find and prototype a measurement solution for almost all types of deep
foundation elements at a reasonable cost.
This last point is according to the authors’ opinion one of the main advantages when fibre optic sensors
are applied in geotechnical engineering.
Multiplexed or multi-point sensing, which means that along the length of an optical fibre several sensing
areas are present at well-defined positions, can be carried out with the Fibre Bragg Gratings technique
(FBG).
Distributed sensing, which means that the entire fibre length is used as sensor, can be carried out by
means of backscattering technologies such as Brillouin scattering, Raman scattering and Rayleigh
scattering.

The working principle of the Fibre Bragg gratings technology (FBG) is illustrated in in Figure 3. One
single continuous optical fibre is at several well-defined positions provided with a Bragg grating that
serves as a sensor. At the position of such a Bragg grating an incident spectrum from a light source is
reflected at a specific (predetermined) wavelength. By providing Bragg gratings at several positions on
the optical fibre and at the condition that each Bragg grating reflects the incident spectrum at a different
wavelength, a multi-point sensor can be obtained. For the moment it is possible to provide up to 30
sensors on one single continuous fibre, which is a huge reduction of the number of transmitting cables in
comparison with classical electrical sensors.

Figure 3: Illustration of the working principle of the FBG/DTG optical fibre technology. In the black box
at the upper right corner: an example of the reflected spectrum of an optical fibre installed in a soil nail
with 11 Fibre Bragg Gratings (FBGs). All 11 FBG sensors reflect the light at a certain wavelength,
dependent on the manufacturing properties and the deformation and temperature of the FBG.

211
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

With this technology, variations of parameters such as temperature, strain and pressure at the location of
the Bragg grating shift in a linear way the reflected wavelength, reason why this technology can be used
as sensor device. The main advantage of this technology with regard to the distributed sensing systems is
the high measurement precision that can be obtained, as well as the fact that it is also possible to sample
measurements at high frequencies (up to 4 kHz or even more). Furthermore, the interpretation of the
optical measurements in terms of reflected wavelengths at the position of the Braggs is very straight-
forward and does not require specific/detailed knowledge of fibre-optic technology.

The main advantage of distributed sensing with the scattering techniques is the fact that an optical fibre
can serve as a sensor along its entire length. The basis of this technology is the use of optical time-
domain Reflectometers (ODTR’s), which are normally used in telecom to evaluate the quality of optical
fibres and connections through the analysis of the measured backscatter when a narrow pulse of light is
send into an optical fibre. The principles of distributed sensing are illustrated in Figure 4. Generally, the
sampling rate of distributed sensing systems is limited. This technique is thus only applicable for the
measurement of slow-changing parameters. A main advantage of distributed sensing is the fact that the
sensor cost is considerably low.

Figure 4: Distributed sensing principle – backscattered light (after Soga 2011).

With the Brillouin scattering technique, which is based on acoustic vibrations stimulated by incident light,
temperature and strain variations can be deduced along the entire optical fibre length and this over very
long distances of several kilometres. Spatial resolutions equal about 0.5 to 1 m (or even less when the
total sensor length is reduced), which means that an average temperature or strain measurement per 0.5 to
1 m is obtained along the fibre. In comparison with the FBG technology, the precision of temperature
and deformation measurements are in general lower (see Table 1). The last decade, this technology has
often been applied in geotechnical engineering and several cases are reported in literature, e.g. Klar et al.
(2006), Mohamad et al. (2011), Iten et al (2012), Van Alboom et al. (2012), Bell et al. (2013) and De
Vos et al. (2013).

The Rayleigh scattering technique, for which the scattering is caused by density and composition
fluctuations that are created in the material during the manufacturing process, has the advantage that the
precision of strain and temperature that can be obtained are higher and even comparable with FBG
technology, and that the spatial resolution is very high (5 mm to 10 mm). However, the total sensor length
is, compared to the other distributed sensing technologies, limited to typically 80 m. Nevertheless, recent
developments in read-out units make it possible to apply this technology on optical fibres with a length of
up to 2 km, but in that case with a considerably decreased sample frequency. The Rayleigh scattering
technique offers a high precision in combination with distributed sensing and high spatial resolution, and
seems therefore very promising for future applications in geotechnical engineering.

In all the technologies mentioned before (FBG, Brillouin scattering and Rayleigh scattering), the optical
sensors react both on strain and temperature variations, which means that interference between both
physical phenomena need to be avoided or corrections need to be carried out. When measurements of
temperature are envisaged, e.g. in a structural or foundation elements, one have to package the optical
sensor lines in such a way that no deformations are transmitted to the sensor line.
When measurement of strains are principally envisaged in an environment where temperature variations
are to be expected during the monitoring period, it is necessary to perform temperature measurements in
order to correct the strain measurements. Temperature measurements can then be performed by means of
the same optical fibre technique, for which an appropriate packaging is provided, by classical

212
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

measurement techniques (e.g. thermocouples), or by means of the fibre-optic distributed Raman scattering
technology.
The distributed sensing technique based on the Raman Scattering, for which the scattering is caused by
the molecular vibrations of the glass fibre stimulated by incident light, is namely insensitive for
deformations and therefore very suitable for temperature measurements. With this technology,
temperature measurements with a precision of 0.1 °C can be obtained along fibre lengths up to several
kilometres with a spatial resolution of 0.5 m to 1 m. The Raman scattering technique is regularly used in
geothermal applications (vertical borehole heat exchangers or in energy piles) in order to evaluate the
thermal conductivity of the subsoil with depth by means of distributed thermal response testing or
enhanced thermal response testing. In the latter one, a hybrid cable (optical fibre + electrical conductors)
is integrated in the borehole or the energy pile and while heat is injected via the electrical conductors the
temperature evolution along the optical fibre, which depends on the local heat dissipation to the
surrounding soil, is measured and evaluated.
Interesting (future) applications for distributed temperature sensing in geotechnical engineering are
thermal pile integrity testing, leakage detection in retaining walls, evaluation of the concrete cover in piles
and micropiles and the dimensions of foundation elements (e.g. jet grout columns).

For more details with regard to optical sensing techniques, reference is made to the NI tutorial (NI, 2011)
and the specialised literature referenced in this tutorial.

Table 1: Comparison of different multi-point and distributed fibre optic technologies


Strain Temperature Readout Sensor
OF Techn. Type of measurement
[accuracy] [accuracy] cost cost
FBG Yes Yes Multi-point Up 30 sensors on 1 sensor line Low- High
[1 µstrain] [0.1 °C] Single ended Lengths up to 1 – 10* km* Moderate
Brillouin Yes Yes Distributed ε av or T av for 0.5-1.0m Moderate- Low
scattering [10-30 µstrain] [1.0 °C] Double-ended Length up to 10-25* km High
Raman No Yes Distributed ε av or T av for 0.5-1.0m Moderate Low
scattering [0.1 °C] Both Length up to 10-15* km
Rayleigh Yes Yes Distributed ε av or T av for 5 mm -10 mm High Low
scattering [2-5 µstrain] [0.2-0.4 °C] Single-ended Length up to 80m and recently
1-2 km
*higher lengths can be found in literature

3. EXPERIENCES OF BBRI WITH OPTICAL FIBRE TECHNOLOGY

3.1. Introduction
The Belgian Building Research Institute (BBRI) started to explore the application of fibre-optic sensing
in 2007. The reason for that was a particular question of the design service of the railway company to
elaborate a long term monitoring solution for the load distribution in jet grout nails, which were installed
to stabilize the existing railway embankments in Ottignies, Belgium. The stabilization of the
embankments was necessary to extend the existing railway tracks from two to four tracks (see Figure 5).
The jet grout nails were executed with 3 m long lost hollow drill rods, which were during installation
coupled with coupling sleeves. Their total length was about 15 m and their inner diameter 24 mm,
locally even 18 mm in the coupling sleeves. It was very difficult to find an instrumentation solution with
classical measurement techniques. In collaboration with Tuc Rail, FOS&S (actually FBGS), and Jan
Maertens bvba, the BBRI elaborated a monitoring solution based on the FBG/DTG technology that could
be integrated in the soil nail after its installation in order to monitor the load distribution in the jet grout
nail.

213
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 5: Stabilization of railway embankments in Ottignies, Belgium, and extension of the number of
railway tracks from two to four. The jet grout nails E and F were instrumented with a lost FBG/DTG base
extensometer system.

This solution has extensively been described by Huybrechts et al (2009), but can be summarized as
follows:
− an optical fibre provided with Bragg gratings (FBG/DTG) each meter was transformed into an
extensometer system by fixing physical anchors to the optical fibre each meter and by making the
zone between the anchors, containing a Bragg grating, free from its environment by a double tubing
(see principle in Figure 6); the total diameter of this system was 6 mm;
− this system was attached to a hollow steel tube with an outer diameter of 8 mm; this steel tube allowed
to integrate the FBG/DTG extensometer system into the hollow jet grout nail that was rinsed after its
installation and to pre-tension the FBG array slightly;
− finally the FBG/DTG extensometer system was fixed to the jet grout nail by means of a grout that was
injected through the hollow 8 mm-steel tube;
− after hardening of the grout the FBG/DTG extensometer system will follow quasi perfectly the
deformation of the jet-grout nail, an accuracy of better than 5 µstrain can be obtained by this system.

Figure 6: Principle of the FGB/DTG extensometer system.

In order to validate this solution, BBRI carried out an extensive preliminary test program consisting of:
− a study of the effect of the grout composition and the hydration shrinkage of the fixation grout on the
measurements and the attenuation of the signal;
− comparative measurements with classical and proven measurement techniques (see Figure 7)
− in situ validation tests with prototypes of the FBG/DTG extensometer system installed in self-drilling
hollow rod anchors that were load tested beyond the elastic limit of the steel (Figure 8). These tests
were carried out in the framework of a real scale research test campaign of the BBRI on ground
anchors and are published in BBRI (2008).

214
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 7: Validation tests of the FGB/DTG extensometer system by laboratory testing.

6000

5000 Series1
Series2
Series3
Series4
Deformation (µstrain)

4000
Series5
Series6
Series7
3000
Series8
Series9
Series10
2000
Series11
Series12
Series13
1000
Series14

0
0 500 1000 1500 2000 2500 3000 3500 4000 4500
n° measurement

Figure 8: Validation tests of the FGB/DTG extensometer system by in situ load testing.

As the results of the deformation measurements with the FBG/DTG extensometer fixed with a grout into
the hollow steel rods were very satisfying, the principles of this measurement system have been applied
on many other and different deep foundations. By attaching steel reservation tubes to the geotechnical
element, this FBG/DTG extensometer system can be installed after the foundation installation procedure,
having the advantage of nearly zero-risk regarding damage to the instrumentation.

215
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Over the years the system has been optimised: the system has been ruggedized, grout compositions have
been adapted taken into account the expected maximum deformations of different foundation types and to
deal with tension and compression, the handling and grout injection tools have been improved, …
An overview of some practical cases is given in the next paragraph.

3.2. Applications of the lost FBG/DTG extensometer system fixed with grout
connection
In this paragraph an overview is given of some practical applications of the lost FBG/DTG extensometer
system (real job sites and real-scale research tests) of which the principles have been highlighted in the
previous paragraph. As explained before, the principle consists of connecting a steel reservation tube to
the foundation element. After installation of the deep foundation element, the optical fibre line is
integrated in the reservation tube, pre-tensioned and fixed to the tube by means of a cement-based grout.
As illustrated in the cases presented in Figure 9 to 16, the system can be used for all kind of foundation
elements (soilmix, jet-grout, concrete piles, diaphragm elements, steel piles and profiles) and can even be
used as extensometer to measure soil heave or settlement. Depending on the expected maximum
deformation and the type of deformation (tension or compression), the grout composition needs to be
adapted. Also the dimensions of the system (in particular the anchors) can be adapted in function of the
application and the available space.

Figure 9: Application of the lost FBG/DTG extensometer during load testing of soil mix columns and
comparative measurements with a classical retrievable extensometer system – Test field BBRI,
Limelette, Belgium 2009.

216
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 10: Application of the lost FBG/DTG extensometer in 24 m and 36 m long bored piles, Andermatt,
Switzerland 2012 : Instrumented preliminary static load tests (compression). The FBG/DTG
extensometer system has been installed in the reservation tubes that have been used for the ultrasonic
integrity tests. To be published by Chitas et al. (2017) - (Courtesy of Besix).

217
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Axial force (kN)


-600 -500 -400 -300 -200 -100 0 100
0
100kN
-1
200kN
-2 300kN
-3 400kN
-4 T=30°C

Depth (m)
-5 T=40°C

-6 T=-4°C

-7

-8

-9

-10

-11

-12

Figure 11 : Application of the lost FBG/DTG extensometer system in 4 displacement screw piles and 1
cfa pile executed with temporary casing – Test field “Energy piles – Oostende, Belgium” execution of
thermo-mechanical loading tests in the framework of the Smart Geotherm project (BBRI, 2011 – 2017)
and (Allani et al., 2016).

218
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 12: Application of the lost FBG/DTG extensometer in 2 cast in situ driven piles and 2 cfa piles
with large hollow stem – Bypass Mechelen, Belgium 2015 : Instrumented preliminary static load tests
(compression).

219
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 13: Application of the lost FBG/DTG extensometer in a cast in situ driven pile – Bypass
Mechelen, Belgium 2015 : Instrumented preliminary static load tests (Horizontal). Reservation tubes
have been attached to the reinforcement cage at opposite sides from the neutral line and in line with the
direction of the horizontal load (Verstraelen et al., 2015).

220
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 14: Application of the lost FBG/DTG extensometer on 30-m long sheetpiles and in a clay layer -
BAM-Oosterweel link test pit – Site Noordkasteel, Antwerp, Belgium 2013 - 2015 – Determination of
normal loads and bending moments in the sheet piles upon a 21-m deep excavation and the swelling of
the stiff tertiary Boom clay layer situated under the excavation layer of the pit (Couck et al., 2015).

221
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

1.03_B Inclino
1.03_A

1.03_B 1.03_A Inclino

Kant 1.03_B
uitgraving

1.03_A

Inclino

Kant
spoorweg

Verticale rek (microstrain) Trek = positief, druk = negatief Verticale rek (microstrain) Trek = positief, druk = negatief
-200 -150 -100 -50 0 50 100 150 200 -200 -150 -100 -50 0 50 100 150 200
17 17
16 16
15 15
14 14
13 13
12 12
11 11
10 10
9 9
8 8
7 7
Peil (mTAW)

Peil (mTAW)

6 6
5 5
4 4
3 3
2 2
1 1
03/10/2014 03/10/2014
0 0
17/11/2014 -1 17/11/2014 -1
08/12/2014 -2 -2
08/12/2014
-3 -3
23/01/2015 -4 23/01/2015 -4
11/02/2015 -5 11/02/2015 -5
-6 -6
13/04/2015 -7 13/04/2015 -7
25/06/2015 -8 25/06/2015 -8
-9 -9
26/08/2015 26/08/2015
-10 -10
04/11/2015 -11 04/11/2015 -11
-12 -12
23/11/2015 23/11/2015
-13 -13
-14 -14
3.03 kant Colomalaan 3.03 kant uitgraving
-15 -15
-16 Rek t.o.v. nulmeting Rek t.o.v. nulmeting
-16
-17 van 23/09/2014 -17 van 23/09/2014

Figure 15: Application of the lost FBG/DTG extensometer on 30-m long diaphragm walls – Bypass
Mechelen, Belgium 2013-2015 – Determination of bending moments in the diaphragm walls upon
excavation (cut & cover tunnel). The FBG/DTG extensometer system has been installed at both sides of
the neutral line of the diaphragm wall in reservation tubes that were provided for the ultrasonic integrity
tests.

222
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 16: Application of the lost FBG/DTG extensometer system in soil nails Antwerp left river bank
(2015-2016) – LT monitoring of soil nail reinforced slope.

The cases in Figures 9 to 16 show the possibilities of the FBG optical fibre technology. Most of these
cases were very successful, however in a few cases some damage occurred after a certain time, probably
due to progressive glass fissuring at the interface between the anchor and the optical fibre. This on his
turn is probably due to handling of the very fragile FBG during its preparation in the laboratory.
In order to avoid to a maximum this kind of problems, very strict procedures have been elaborated in
order to avoid damage to the coating of the fibre during its preparation/packaging in the laboratory and a
double ended backup connection is systematically foreseen.
Furthermore, the fibre optic sensors industry evolves very fast and some products, which are very
interesting for the construction industry, appeared the last years on the market, in particular the reinforced
FBG optical fibres, e.g. the GFRP (glass fibre reinforced polymers).

Figure 17 : Lost FBG/DTG extensometer system prepared and packaged by BBRI (left); ∅ 1mm GFRP
optical fibre with FBGs (middle); ∅ 2mm GFRP optical fibre with FBG/DTGs (right).

223
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Very recently, the BBRI had the opportunity to perform comparative testing with 2 types of GFRP FBG
arrays during a static pile load test. The following systems (see Figure 17) were integrated in a reservation
tube that was foreseen in the pile foundation and fixed into it by means of a cement grout:
− the lost FBG/DTG extensometer system as described in the previous paragraphs
− a GFRP FBG optical fibre with a total diameter of 1 mm
− a GFRP FBG optical fibre with a total diameter of 2 mm

The pile was submitted to compressive loading and Figure 18 shows the results of the measurements
during one of the last load steps. From these results one can conclude the following:
− The deformations measured with the GFRP fibres coincide relatively well with the FGB/DTG
extensometer systems.
− The strains measured with the GFRP FBG fibres are sometimes smaller, sometimes higher and
sometimes the same; this proves in the authors’ opinion that the GFRP FBG fibre is rather a local
measurement (comparable to a strain gauge or a strain gauge transducer) than an extensometer based
measurement that gives the average deformation between the anchor points.
− The ∅ 2 mm GFRB seems to stick closer to the extensometer based system than the ∅ 1 mm GFRP
fibre, from which one can conclude that deformation measured with the ∅ 2 mm is less local than the
∅ 1 mm fibre.

Anyway, these results are promising for the future. However, these conclusions are only valid for
compressive deformations up to 500 µstrain. For larger compressive deformations and deformations
under tension, supplementary comparative testing is necessary.

Figure 18: Comparative testing of the lost FBG/DTG extensometer system a ∅ 1mm GFRP optical fibre
and a ∅ 2mm GFRP optical fibre.

224
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

3.3. Applications of the lost FBG/DTG extensometer system fixed to the steel
reinforcement
Another way of applying FBG optical fibres to foundation elements is by attaching them to the
reinforcement steel. The Bragg gratings, that have in general a length of a few centimetres, can then be
glued directly to the steel according to a procedure comparable to that of strain gauges. This results in a
local strain measurement. Many of these systems are also commercially available and packaged in all
kinds of forms and can be fixed to the structure/reinforcement by gluing, spot-welding, bolting, etc.
However, BBRI is confronted regularly with applications, where commercially packaged sensors give no
satisfying solution. For that reason, sensor prototypes are mostly fabricated by BBRI starting form raw
FBG/DTG optical fibre arrays.
Contrary to the lost FBG/DTG extensometer systems highlighted in paragraph 3.2, integrating the optical
fibre arrays into the reinforcement elements means that the sensor lines are present during the installation
of the deep foundation element and that a real risk exists to damage the sensors during foundation
installation. Protection of the sensor lines is thus an important issue.

BBRI often prefers, for reasons of interpretation, average deformations over a certain length of the
foundation (typical 0.5 to 2 m), rather than local measurements. For that reason, BBRI connects the FBG
sensors to the reinforcement elements mostly according to the extensometer principle. In most cases and
in order to protect the optical fibre, a continuous incision of 1.5 to 2 mm² is firstly made along the
reinforcement beam or the rebar. Afterwards, the fibre is integrated and glued in this sleeve at several
positions that serve as anchor points. The Bragg gratings are situated and pre-tensioned between those
“anchor” points and can move freely in the sleeve due to a protective 0.9 mm tubing.
Figure 19 illustrates the application of the lost FBG/DTG extensometer system integrated in the steel
reinforcement of 17 large-scale bending tests on real scale soil mix elements (CSM-panels and columns).
These tests were executed in the framework of the BBRI research project on soil mix (BBRI, 2009-2013)
in order to determine the bending stiffness and the bending resistance of soil mix walls. The results were
reported by Denies et al.( 2015).

This instrumentation principle has also been applied for the monitoring of the bending moments in jet
grout columns that were reinforced with steel beams. This case was extensively reported by Van Alboom
et al. (2012), De Vos et al. (2013) and Verstraelen et al. (2012). For this case comparative measurements
with other optical fibre technologies (a.o. distributed sensing with the Brillouin scattering method) were
performed as well.

225
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 19 - Application of the lost FBG/DTG extensometer system on steel reinforcement – 17 large-scale
bending test on real scale soil mix elements (CSM panels and columns) in order to determine bending
stiffness and bending resistance of soil mix walls in the framework of the BBRI research project on soil
mix (BBRI, 2009-2013) and (Denies et al., 2015).

226
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Another important case where the lost FGB/DTG extensometer system was connected directly to steel
elements is the test pit for the Oosterweel link in Antwerp (see also Figure 14 for the grout connected
sensors). Figure 20 shows the instrumentation of the struts and some of the measurements as the
excavation of the test pit advances. In situ comparison with classical strain gauges showed very well
agreement with the FBG measurement on the struts. Furthermore, the 30-m long sheet piles were
instrumented with optical fibres that were glued on the sheet piles (see Figure 21). BBRI installed a
FBG/DTG sensor array, according to an extensometer principle, in a small sleeve of 2 mm² and protected
it with a 0.2 mm thick steel plate that was spot welded to the sheet pile. The Geotechnical Division of the
Flemish Ministry of Public works installed as well an optical fibre in a somewhat wider sleeve. This
optical fibre was from the distributed sensing type (Brillouin scattering) and because it was provided with
a ruggedized coating, it was installed without any mechanical protection but the glue and a steel shoe at
the base.
The main objective of the instrumentation integrated in the sheet pile was to evaluate if such
measurement systems could survive hard driving and hammering work, which seemed to be the case. In
Figure 21 some of the measurement results of the distributed sensing fibre during the different phases of
the excavation of the test pit are shown.

Figure 20: Application of the lost FBG/DTG extensometer on 10 steel struts (5 levels) - BAM-Oosterweel
link test pit – Site Noordkasteel, Antwerp, Belgium 2013 - 2015 (Couck et al., 2015).

227
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 21: Application of the lost FBG/DTG extensometer on 30-m long sheetpiles - BAM-Oosterweel
link test pit – Site Noordkasteel, Antwerp, Belgium 2013 - 2015 – Evaluation of robustness of optical fibre
measurements systems subjected to hard driving work. (Couck et al., 2015).

228
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Finally, another case were the lost FBG/DTG extensometer system was applied is shown in Figure 22. In
the framework of a research program of Parkwind (2013-2016), which is a company that develops,
finances, builds and operates wind farms in the North Sea. For this project BBRI prototyped strain
sensors consisting of reinforcement bars with integrated FBG/DTG sensors. During the construction of
the windmill, these reinforcement bars were integrated in the zone between the monopile and the
transition piece (in yellow) on which the windmill is constructed. It was aimed to monitor on the long
term the strains in the grout connection between the monopile and the transition piece. The measurements
are still on-going.

Figure 22: Application of the lost FBG/DTG extensometer for monitoring of the strains in the grout
connection between monopole and transition piece 2013-2016 – Parkwind (2013-2016).

3.4. Retrievable fibre optic extensometer system


In the previous paragraphs monitoring solutions with optical fibres were presented. In all the cases it
concerns optical fibres that are connected to the foundations by means of a grout or fibres that are glued
to steel reinforcement, which mean that they are lost. For applications where the instrumentation serves
only for a limited time, e.g. in the case of pile load tests, the investment is rather high. For that reason the
BBRI has developed a retrievable extensometer system with FBG optical sensors.
As illustrated in Figure 23 this system consists of anchors that are pneumatically fixed in a reservation
tube that is integrated in the reinforcement cage and the pile. Between each couple of pneumatic anchors
an OF-extensometer, which measures the displacement of the pile segment between the anchors, is
present.

229
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

The advantage of the use of optical sensors is that the number of extensometers can be increased with
regard to classical retrievable extensometer systems. The reason for this is the fact that, because of the
multi-point type of sensors, the number of cables that have to pass through the pneumatic anchors, is
limited. With this retrievable extensometer system up to 28 modules with lengths varying between 1 to 2
m each can be connected in series. This system was very recently tested in situ and compared with other
measurement systems, leading to very satisfying results.

Figure 23: Principles of a retrievable extensometer system (left); multi-point retrievable extensometer
system with optical fibre sensors (middle); example of the deformation measurements with depth of a 20-
point retrievable optical fibre extensometer system (right).

4. CONCLUSIONS AND PERSPECTIVES


In this contribution an overview has been given of some recent advances with regard to measurement
techniques in geotechnical engineering. It focussed especially on the experience of the Geotechnical
Division of the BBRI with the multi-point FBG/DTG optical fibre technique that, with some basic
knowledge of optical fibre technology, can be applied as a very flexible, performant and accurate strain
measurement device in almost all kind of geotechnical applications where deformation measurements
need to be carried out.
For most of the applications that have been illustrated in the previous paragraphs, the distributed optical
fibre sensing technology could be applied as well, but the measurement precision of this system is up to
now somewhat less. However, this technology evolves very fast, which is very promising for the future as
the sensor cost of distributed optical sensors is low and the amount of information that can be obtained is
very high. This evolution is for our domain very positive as it might lead to an important increase of the
monitoring of geotechnical structures in the near future. This will undoubtedly lead to a better
understanding and new insights with regard to the behaviour of geotechnical structures and deep
foundations and this (r)evolution should in the author’s opinion also be taken into account in the 2nd
generation of Eurocodes, which should include incentives to promote an increased use of (long term)
monitoring in (geotechnical) structures. This is also in line with the desirable evolution to a more
reliability based design approach.

Another important evolution with regard to measurement techniques in geotechnical engineering that has
not been assessed in detail in this contribution but which is worth to be mentioned, is the use of
temperature measurements in foundations. The aim of these methods is to measure the temperature profile
along the foundation depth during the exothermic hydration reaction of concrete and grout as it sets and
hardens. The temperature evolution with time at a certain point depends on the thermal conductivity of
materials surrounding the temperature sensor. In case of an important deviation in the characteristics of
the concrete (e.g. due to segregation or the presence of a bentonite/soil inclusion), this impacts the
thermal characteristics of the zone around the temperature sensor and thus the temperature evolution with
time. These measurements are thus very suitable to detect imperfections, to evaluate the concrete cover,
the dimensions of deep foundation elements, and become more and more used on the market. The
principles of this method and some cases are reported a.o. by Mullins (2010), Likins et Al. (2011),

230
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Piscsalko et al. (2011) and Piscsalko (2014). In 2014 an American standard dealing with thermal
integrity testing was published as well.
One wishes also to refer to research on the use of this thermal method to detect anomalies in the joints of
diaphragm walls, e.g.Niederleithinger et al. (2014) and Spruit (2015).

With the view on the application of temperature profiling to assess the integrity, the quality and the
dimensions of deep foundations, distributed optical fibre sensors can be a very suitable technique to be
applied, in particular the Raman scattering technique, which is not sensible to deformations and can
deliver very accurate temperature measurements.
Above that, an interesting application of the distributed Raman scattering technique is the use of “hybrid”
cables (see Figure 24). A hybrid cable consists of the combination of one or more optical temperature
sensing fibres (Raman principle) and a number of copper heating cables. This type of cables is actually
used for Enhanced Thermal Response Tests on vertical geothermal boreholes, which allow the
determination of soil thermal conductivity along the borehole length. During such a test a user-defined
amount of energy is dissipated into the soil by heating the copper cable electrically, while the temperature
evolution along the optical fibre is monitored. An example of the results of such a test is given in Figure
25.

Optical Hybrid cable


fibres

Diam. 12 mm

4 copper cables

Figure 24: Hybrid cable containing optical temperature sensing fibres and copper heating cables.

Thermal conductivity (W/mK)


0 1 2 3 4 5 6 7 8 9 10
0
10 Thermal
conductivity by
20 Enhanced TRT
30
Undisturbed
40
temperature T_0h
T_0h T_1h
Depth (m)

50 (24/07/2015) → T_3h
60 → T_96h
70
80
90
100
110
120
5 6 7 8 9 10 11 12 13 14 15 10 15 20 25 30 35 40 45 50
Soil temperature (°C) Soil temperature (°C)

Figure 25. Result of an Enhanced Thermal Response Test (ETRT): thermal conductivity along the 120 m
deep borehole and undisturbed temperature profile (left); Temperature evolution during heating of the
copper cable (right).

According to the authors, this “hybrid” measurement and testing methodology also has a huge potential to
be used in many geotechnical applications for quality control (integrity, concrete cover, dimensions of

231
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

deep foundations) and for the detection of leakage in retaining structures and dikes, and real applications
are expected to be reported soon.

REFERENCES
Allani, M., Van Lysebetten, G. & Huybrechts, N. 2016. Experimental and numerical study of the thermo-
mechanical behaviour of energy piles for Belgian practice. Proceedings of the 1st International
Conference on Energy Geotechnics - ICEG2016 organized by ISSMGE TC308 on Energy Geotechnics -
Kiel, Germany, August 29-31, 2016 (to be published).

ASTM D7949. 2014. Standard Test Methods for Thermal Integrity Profiling of Concrete Deep
Foundations.

BBRI. 2008. Proceedings of the international symposium “Ground anchors – Limelette test field results”,
Brussels, 14.05.2008 (proceedings volume 1, 2 and 3, available on www.tis-sft.wtcb.be).

BBRI. 2011-2017. Smart Geotherm project funded by the Flemish Agency for Innovation and
Technology. www.smartgeoherm.be.

Bell, A., Soga, K., Ouyang, Y., Yan, J. & Wang, F. 2013. The role of fibre optic instrumentation in the re-
use of deep foundations, Proceedings of the 18th International Conference on Soil Mechanics and
Geotechnical Engineering, Paris.

CEN. 2010. EN 1536:2010. Execution of special geotechnical works – Bored piles.

CEN. 2015. EN 12699:2015. Execution of special geotechnical works – Displacement piles.

Chitas, P., Bauduin, C., Raucroix, X., Goffinet, M. 2017 Design of a piled raft foundation for a building
at an alpine valley in Switzerland. Soil-structure interaction analysis and comparison with pile test
results. Proceedings of the 18th ICSMGE, Seoul (to be published).

Couck, J., Van Royen, K., anssens, B., de Nijs, R., Van Lysebetten, G. 2015. Proefcampagne voor de
Oosterweelverbinding in Antwerpen. Geotechiek, jaargang 19, nr. 5, december 2015 - Speciale uitgave
n.a.l.v. de Geotechniekdag ‘Proef op de Som”, 03.11.2015, Breda.

De Vos, L. Van Alboom, G. & Haelterman, K. 2013. Comparison of monitoring techniques for measuring
deformations in an excavation. Proceedings of the 18th International Conference on Soil Mechanics and
Geotechnical Engineering, Paris.

de Nijs, R.E.P., Kaalberg, F.J., Osselaer, G. Couck, J. and Van Royen, K. 2015. Full scale field test
(sheet)pile driveabilitity in Antwerp (Begium). Proceedings of the European International Conference on
Soil Mechanics and Geotechnical Engineering, Edinburgh.

DIN 4150-1. 2001. Erschütterungen im Bauwesen – Teil 1 : Vorermittlung von Scwhingungsgrossen.

DIN 4150-2. 1999. Erschütterungen im Bauwesen – Teil 2 : Einwirkungen auf Menschen in Gebaüden.

DIN 4150-3. 2015. Erschütterungen im Bauwesen – Teil 3 : Einwirkungen auf bauliche Anlagen

Dunnicliff, J. 1993. Geotechnical instrumentation for monitoring field performance. John Wiley & Sons,
Inc.

Hertlein, B. and Davis, A. 2006. Nondestructive testing of deep foundations. John Wiley & Sons, UK.

Huybrechts, N. , Ganne, P., Maekelberg, W., Maertens, J. & Vlekken, J. 2009. Toepassing van optische
vezel-technologie voor de lange termijn monitoring van vernagelde taluds te Ottignies (GEN).
Geotechiek, juli 2009.

232
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Iten, M., Hauswirth, F., Fischli, F., Puzrin, A.M. 2012. Distributed fiber-optic sensors in geotechnical
engineering monitoring – Technical Article. ISSMGE Bulletin, Volume 6, Issue 5, available
on www.isssmge.org.

Klar, A., Bennet, K., Soga, K., Mair, R.J., Tester, P., Fernie, R., St John, H.D. & Torp-Peterson, G. 2006.
Distributed strain measurements for pile foundations. Geotechnical Engineering 159 Issue GE3.

Likins, G. & Mullins, G. 2011. Structural Integrity of drilled shaft foundations by thermal measurements.
Structural Engineer, November 2011.

Mohamad, H., Soga, K. and Pellow, A. 2011. Performance Monitoring of a secant piled wall using
distributed fibre optic strain sensing. Journal of Geotechnical and Geoenvironmental Engineering, Vol.
137, No. 12, pp. 1236-1243.

Mullins, G. 2010. Thermal Integrity profiling of drilled shafts. DFI Journal, Vol.4, N° 2 December 2012.

NI, 2011. Overview of Fiber Optic Sensing Technologies Publish Date. http://www.ni.com/white-
paper/12953/en/.

Niederleithinger, E., Amir, J.M., Schneider, N.S. 2014. New Possibilities for quality assurance of
diaphragm wall joints? Global Perspective on Sustainable Execution of Deep Foundation Works.
Proceedings of the DFI-EFFC International Conference on Piling & Deep Foundations, Stockholm.

Piscsalko, G., Cotton, D., October 2011. Non-Destructive Testing Methods for Drilled Shaft and ACIP
Piles. Proceedings from Deep Foundations Institute 36th Annual Conference on Deep Foundations:
Boston, MA; 252-532.

Piscsalko, G., May 2014. Non-Destructive Testing Of Drilled Shafts and CFA Piles - Current Practice and
New Method. Proceedings of the DFI/ EFFC 2014 International Conference on Piling and Deep
Foundations: Stockholm, Sweden; 533-546.

SBRCURnet. 2006. SBR Meet- en beoordelingsrichtlijn: Trillingen – Deel A - Schade aan gebouwen.

SBRCURnet. 2006. SBR Meet- en beoordelingsrichtlijn: Trillingen – Deel B – Hinder voor personen.

SBRCURnet. 2006. SBR Meet- en beoordelingsrichtlijn: Trillingen – Deel C – Storing aan apparatuur.

Soga, K. 2010 Innovation in monitoring technologies for tunnels, Jornada Técnica: Instrumentacion de
tuneles y excavaciones, Barcelona, 21.10.2010.

Spruit, R. 2015. To detect anomalies in diaphragm walls – phd. TU-Delft. http://repository.tudelft.nl/.

Van Alboom, G., De Vos, L., Haelterman, K. & Maekelberg, W. 2012. Innovative monitoring tools for
online monitoring of building excavations.Proceedings of the ISSMGE-TC211 International symposium
on Ground Improvement, Brussels, Vol. IV pp. 327-338.

Verstraelen, J., Maekelberg, W. Lejeune, C., De Clercq, E. & De Vos, L. 2013. Design and performance
of a jet grout retaining wall in a railway embankment on soft soil, Proceedings of the 18th International
Conference on Soil Mechanics and Geotechnical Engineering, Paris.

Verstraelen, J. & Maekelberg, W. 2015. Ontwerpen op basis van resultaten van laterale
paalbelastingsproeven. Geotechiek, jaargang 19, nr. 5, december 2015 - Speciale uitgave n.a.l.v. de
Geotechniekdag ‘Proef op de Som”, 03.11.2015, Breda.

ACKNOWLEDGEMENTS

The authors wish to thank their colleagues R. Bonsangue, B. Andre and T. Leduc for their contributions
to the development and prototyping of optical sensor solutions and their hard work on all the different job
sites.

233
234
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Research and Development Activities on Pile Foundations in Europe


Kenneth Gavin, TU Delft, Netherl°ands, k.g.gavin@tudelft.nl

David Igoe, Gavin and Doherty Geosolutions, Dublin, Ireland, digoe@gdgeo.com

Kenny Kataoka Sorensen, Department of Engineering, Aarhus University, Denmark, kks@eng.au.dk

Noël Huybrechts, Belgian Building Research Institute-BBRI & KU Leuven, nh@bbri.be

ABSTRACT
The paper describes some recent field test programmes performed across Europe. Given the increased
need to develop alternative, clean energy sources the paper presents details of R&D programmes with a
focus on reducing costs offshore developments and increases safety. Work on the development of design
codes for large diameter piles is presented. New areas of research are highlighted in particular ongoing
research is needed to look at effects such as cyclic loading and dynamic interaction. In the offshore field
the use of piles that reduce environmental impact is an area of ongoing research that is addressed.
Interesting full-scale field tests looking at the problem of negative skin friction are also considered.

1. INTRODUCTION
Whilst clean, renewable energy solutions are being developed there is a continued need for offshore oil
and gas platforms to provide an important part of our energy mix. Jacket structures for oil and gas
platforms are commonly founded on (predominantly axially loaded) open-ended steel tubular piles. The
challenge to companies in this sector are to reduce costs and at same time increase safety. New
exploitation opportunities are rare and are typically in deep water areas with challenging geology.
Reassessment of and life-time extension of old platforms is also becoming more important. One major
benefit for developers and operators are that the new CPT based design approaches for axially loaded
piles now included in the offshore design standards typically allow for increased axial pile capacities to
be adopted, particularly in dense sand deposits. The CPT based design approaches that account for
friction fatigue, stress level, interface shearing characteristics and plugging have been shown to be more
reliable than traditional design approaches in database assessments by Chow (1997) and Schneider (200)
amongst others. However, topics such as the effect of pile ageing and cyclic loading on axial pile capacity
remain uncertain. These issues are addressed in Section 2 of the paper.

Many countries around Europe have developed ambitious targets for energy generation from renewable
sources. Offshore wind farms are seen as key to achieving this aim. Siting turbines offshore provide
several benefits including: (i) availability of high unrestricted wind speeds, (ii) the ability to use larger
turbines and (iii) the ability to develop combined wind and wave/solar energy installations as alternative
renewable energy solutions which become economically viable. The majority of turbines constructed to
date have been founded on monopile foundations. Because of increased water depths and larger turbines,
monopile diameters are increasing, with planned developments having piles of 10 m diameter. The cost
of foundations can represent over 30% of the development cost for an offshore wind farm and advances in
pile design that allow for cost savings are urgently required. In addition critical issues related to the
impact of cyclic loading, soil-structure interaction for dynamic analyses and environmental concerns
around the impact of installing large-diameter piles have led to the development of large industry-
academia R&D projects. These issues are addressed in Section 3 of the paper. In the final section issues
related to onshore piling such as the use of low-displacement piles in addressing environmental concerns
and field tests investigating negative skin friction on piles are discussed.

2. FIELD TESTS INVESTIGATING PILE AGEING AND CYCLIC


LOADING OF PILES
The positive gain in capacity with time (i.e. set up) of driven piles in clay is well understood due to
general acknowledgement of the effective stress changes associated with the dissipation of excess pore
pressures induced by driving. Set-up for piles in sand is a less well known phenomenon. A number of
field studies of pile ageing in sand have been conducted in recent years. Four major field studies have
recently being conducted by researchers at Imperial College London (IC), the Norwegian Geotechnical

235
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Institute (NGI) and University College Dublin (UCD). The work described in detail by Gavin et al.
(2015) is summarised herein.

2.1. Imperial College tests at Dunkirk


The Imperial College Dunkirk pile test site at Dunkirk, Northern France is comprised of a 3 m thick layer
of hydraulic sand fill over Flandrian marine sand. The ground water table is approximately 4 m below
ground level (bgl). The sand is a fine to medium, predominantly silica deposit. The geotechnical
characteristics are described in detail by Chow (1997). In summary the deposit is dense to very dense
with relative density of ≈ 75% .The sand has a peak friction angle φ′ p in the range 35-40°, constant
volume friction angle φ′ cv of 32° and interface friction angle δ of ≈ 27°. A typical CPT q c trace showed
end resistance, q c values in the range 10 to 30 MPa over the depth of interest for piling at the site and the
presence of a weak organic layer at 8 m bgl.

An investigation of ageing of 325 mm, OD 11 m to 22 m long steel pipe piles (Chow et al. 1998) showed
that the tension shaft capacity of a pile rose by 85% between tests performed over a five year period.
Jardine et al. (2006) undertook further pile driving and testing at the same site that was designed to
specifically investigate pile ageing. They drove six steel tubular test piles (designated R1 to R6) had a
diameter, D of 457 mm, a wall thickness which varied between 13 and 20 mm. They were driven to
penetrations of between 18.9 and 19.4 m. Monitoring of the soil core during pile installation revealed that
partial plugging occurred with plug length ratios (PLR = soil core length/pile penetration length) of
approximately 60%. Slow, load-controlled, tension load tests were performed on the piles at various time
periods after installation. One of the most striking features of the response was the strong gain in capacity
observed during first time load tests on fresh piles which were tested between 9 days and 235 days after
installation with the resistance increasing from 1450 kN to 2400 kN, See Figure 1.

Figure 1: Effect of time on load-displacement response of fresh piles (after Jardine et al 2006)

The data from Figure 1 was added to dynamic End of Driving (EoD) capacity assessments made for the
instrumented piles at the same site reported by Brucy et al (1991) and the tension capacity from a pile test
on a five year old, 22 m long pile reported by Chow (1997), to propose an intact ageing characteristic line
(IAC), see Figure 2. In this plot the shaft capacity, Q s at a given time, is normalised by the shaft capacity
predicted using the ICP pile design approach, Q sICP . The data suggests that the semi-logarithmic rates of
capacity increase are modest initially before finally stabilising.

236
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 2: Effect Intact ageing curve for the Dunkirk pile tests

2.2. University College Dublin Pile Tests in Blessington Sand


This study on pile ageing involved load testing of four, 7 m long open-ended steel piles. The test piles
were driven into dense sand at the UCD Blessington test site in Ireland. The glacially deposited fine sand
has a relative density close to 100% and the ground water table was approximately 14 m below the piling
platform level. At each pile test location two CPT tests were performed prior to pile installation. The CPT
q c resistances measured were relatively consistent across the test location, See Figure 3. Additional in-
formation on the ground conditions at the site and the sand properties have been reported in Gavin et al.
(2009), Doherty et al. (2012) and Kirwan (2015).The piles had identical geometries with an external
diameter of 340 mm and a wall thickness, t of 14 mm. One of the piles (designated Pile S5) was
instrumented with Kyowa miniature radial stress sensors located at distances from pile toe, h, of 1.5, 5.5,
and 10.5 and 17.5 pile diameters. The piles were subjected to maintained tension load tests at time
intervals of 1 day, 11 days, 30 days and 220 days after driving respectively. Details of the pile
construction, instrumentation and calibration can be found in Kirwan (2015).

Figure 3: CPT profiles conducted beside piles S2 – S5

237
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Pile driving was performed using a Junntan PM-16 drop hammer with a drop-height of 0.3m. The pile
driving records and soil core development (in terms of Incremental Filling Ratio, IFR = incremental
change of soil plug length/change in pile penetration) for three of the piles (S2, S3 and S5) were similar.
Pile S4 had a similar response in terms of plug development and blow-counts until a penetration depth of
4m. Thereafter, the degree of coring increased and the number of blow-counts for this pile was
significantly lower than the other test piles.

The load-displacement response measure during first-time loading on fresh piles is shown in Figure 4.
There was a generally a strong trend for the tension capacity of the piles to increase with time. A notable
exception is Pile S4. It is possible that Pile S4 had a significantly smaller capacity than piles S2, S3 and
S5 at the end of driving due to the much higher IFR and lower installation resistance noted over the last 1
m of penetration and the notably lower blow counts recorded during installation.

Figure 4: Load-displacement response measured during first time load tests

One of commonly assumed mechanisms for pile ageing is that breakdown of hoop stresses set-up during
installation would allow stationary radial stress measured on the pile to increase with time. Such increases
have been recorded on closed-ended concrete driven piles by Briaud et al. (1989) and Axellson (2000).
The radial stress sensors on pile S5 at Blessington were logged periodically until the load test was
performed 217 days after driving. The variation of radial stress over the setup period is shown in Figure 5
together with the stationary radial stress profiles predicted using the IC-05 and UWA-05 design methods.
The following points are noteworthy:

1. At the end of installation the radial stresses measured at all sensor locations were much higher than
predictions made using either of the design approaches.
2. Rather than increasing with time, relaxation of the radial stress occurred during the set-up period. This
relaxation was evident at all sensor levels.
3. At the end of the set-up period, the average stationary radial stresses had reduced significantly. Whilst
the values measured near the pile tip were still significantly higher than the values predicted by both
the ICP and UWA design methods, it should be noted that both design approaches were developed
based on pile test data with a minimum h/D value of 4. At locations further from the pile tip, the radial
stresses are much closer to the predictions made using the CPT based design approaches.
4. During static loading, as the test load increased the radial stress sensors at all locations showed a trend
for large increases in radial stress due to dilation. Near the pile tip, the radial stress doubled, reaching
a value just below the end of installation radial effective stress. Large stress increases were noted at
the other two sensor levels.

238
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 5: Variation in radial stress measured with time

The study found in general that the axial capacity of the piles increased by 185% over 220 days.
However, one pile which experienced less plugging during installation mobilised a lower resistance than
the other three test piles. During a 220 day equalisation period, the radial stress measured near the pile toe
reduced significantly. During loading, large increases in radial effective stress occurred at all levels of the
pile as a result of dilation.

2.3. Norwegian Geotechnical Institute Field Tests at Larvik and Ryggkollen


A major joint industry project (JiP) on pile ageing that included load tests at two sand sites in Norway
were reported by NGI (2014) and Karlsrud et al. (2014). Axial load tests were performed on piles at each
site where six open-ended piles with diameters of 400-500 mm and driven lengths ranging from 15 to
20.1 m were installed and load tests were performed over a two year period. The sand deposit at Larvik
site is a recent, alluvial deposit of loose, fine silty sand with some interbedded layers of clayey silt,
constituting 9 to16% of the test bed. The q c values in the range 1.5 to 10 MPa, with a relative density
interpreted to be in the range 20 to 40%. The water table at the site was close to ground level. The sand at
the Ryggkollen test site is a gravelly medium sand formed by an end moraine. The site was located in a
sand quarry. CPT tests gave q c values in the range 20 to 30 MPa, which suggested relative density is in
the range 50 to 80%. The water table was at 10 m bgl. Detailed soil properties can be found in NGI
(2014).

The test piles were driven using a hydraulic hammer. Casings were used to isolate the pile response from
overlying soil layers; the casing depth was 1.4 m at Larvik and 5 m at Ryggkollen. The piles installed at
Larvik were 508 mm diameter, with a wall thickness of 6.3 mm and were driven to an embedded length
of 20.1 m. At the end of installation PLR values were between 83 and 87%. The piles installed at
Ryggkollen had an outer diameter of 406.4 mm, with a wall thickness of 12.5 mm, and were driven to an
embedded length of 15 m. At the end of driving, PLR values were generally between 63 and 84%.
However, one pile (R5) developed a relatively short soil plug (PLR= 32%) after hitting a stone during
installation. The load test set-up is shown in Figure 6.

239
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 6: Typical load testing arrangement, NGI tests (after NGI, 2014)

The axial capacity measured during first-time load tests on piles is shown in Figure 7, where the capacity
Q 0 is the reference (10 day capacity of the piles). The pile capacities at both test sites show a trend for
substantial increase with time. However, comparing these site with pile ageing tests at Dunkirk and in
Blessington sand (See Gavin et al. 2013) show the growth rate is quite variable. Also the nine month pile
capacity at Larvik was higher than that measured in the 12 and 24 month tests. Karlsrud et al. (2014)
attribute this trend to the local variability of soil conditions at the site, which caused variation in both the
driving resistance and degree of plugging experienced. This normalisation approach suggests that both the
rate effect (i.e. the time when the largest gain in capacity is developed) and the maximum set-up vary
from site to site. The piles at Ryggkollen were slowest to show any ageing effects. The interpreted
increase in ultimate shaft resistance after 12 months varies from a factor of 1.75 to 2.55, with Larvik
giving the lowest and Blessington the highest.

Figure 7: Proposed curve-fitted capacity build-up with time (after Karlsrud et al, 2014)

Repeated load tests on the same pile at both Larvik and Ryggkollen showed a significantly lower capacity
at the same time after pile installation than the corresponding fresh tests. Karlsrud et al (2014) therefore
warned against using repeated load testing to failure as basis for addressing real time effects. The piles
subjected to sustained loading over 12 months before being loaded to failure also showed less gain in
capacity than the fresh piles. The gain after 24 months dropped by a factor of respectively 1.42 (Larvik)
and 1.16 (Ryggkollen). Karlsrud et al (2014) recommend that this aspect must be considered when
applying time effects in design practice. Further investigations are needed to assess how this negative
impact of sustained loading depends on the level of sustained loading, and if it is any different for piles
loaded in compression as compared to the tension loading applied in NGI’s tests.

240
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

2.4. The Effect of Cyclic Loading


The cyclic response of piles in sand has been shown by Chan and Hanna (1980) to be affected by the
number (N) of load cycles, the mean shaft load (Q mean ) and the shaft cyclic amplitude (Q cyc ) among other
factors, as shown in Figure 8. While a number of model laboratory scale cyclic pile studies have been
conducted, the dearth of field cyclic tests on driven piles in sand has been highlighted by Tsuha et al.
(2013). Limited suites of field cyclic load tests on instrumented piles have been conducted by Imperial
College at Labenne and Dunkirk in France (Lehane 1992 and Chow 1997) and by University College
Dublin (UCD) at the Blessington test site (Gavin and O’Kelly, 2006, Igoe et al. 2011).

Q Cyclic Period T

Qmax

Qcyc

Qmean

Time
Qmin

Figure 8: Definition of cyclic loading parameters (after Tsuha et al. 2012)

A range of static and dynamic load tests were undertaken by Jardine et al. (2008) in order to evaluate the
effects of prior cyclic or static failure on the ageing process. The effects were striking with the authors
concluding that any test to static or cyclic failure degrades the pile capacity and can lead to a brittle load-
displacement response. While the pile capacities recovered with time, the growth rates were lower than
with fresh piles. Data from retests performed on pile R1 shown (as the lower line) in Figure 2 reveal that;

(i) The capacity at any time for a reloaded pile is lower than the capacity of a fresh pile.
(ii) Performing a subsequent reload test on the same day verified the reduction in the pile capacity
caused by the preceding failure, explaining the zigzag pattern in Figure 2.
(iii) The capacity of the statically pre-tested piles increased with time. However, the rate of set-up was
lower than for fresh piles.
(iv) The trend for capacity variation for time for the reloaded pile was similar to the average trend line
proposed by Chow et al. (1998) from their database study. Most of the entries in the latter databases
had involved multiple re-tests on the same piles.

Comprehensive testing on multiple piles by these author showed that cycling above certain combinations
of mean and cyclic loads led to changes in capacity;

- Low level stable cycling, where the pile head displacements accumulate slowly over hundreds of
cycles, was shown to accelerate the ageing process resulting in increased capacities. High level
unstable cycling, where head displacements develop rapidly leading to failure within 100 cycles, was
shown to degrade the capacity of aged piles. Metastable load cycling, where head displacements
accumulated at a moderate rate leading to failure between 100 – 1000 cycles, also resulted in
significant capacity reductions.
- Cyclic interaction diagrams were proposed to express how cyclic loading parameters N, Qmean and
Q cyc affect the piles cyclic response relative to the max static shaft capacity, Q s,max as seen in Fig. 9.
Boundaries for stable, metastable and unstable behaviour were proposed and it was noted that the
boundaries did not appear to be sensitive to whether the piles had experienced prior cyclic or static
failure.

241
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

1 No cyclic failure

0,9 Previously Failed

First Failure
0,8

0,7 Black Symbol = Unstable


Grey Symbol = Metastable
0,6 White Symbol = Stable
Qcyc/Qs,max

Unstable
0,5

0,4 Metastable
0,3

0,2
Stable
0,1

0
-0,1 0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1
Qmean/Qs,max

Figure 9: Cyclic Interaction Diagram based on Dunkirk cyclic tests (after Jardine and Standing 2012)

3. MONOPILES FOR THE OFFSHORE WIND INDUSTRY

3.1. Background
The limitation of the current design codes (that were developed for long-slender piles) to the design of
laterally loaded monopiles used in the offshore wind sector are well known (See Doherty and Gavin
2011).

3.2. PISA Project

3.2.1. Background
The PISA (PIle Soil Analysis) project is a JiP established to develop new design methods for the very
large diameter, relatively rigid monopiles with diameters up to 10 m used to support turbines. The project
involves: (i) numerical modelling using 3D finite element analysis; (ii) the development of a new design
method, and (iii) field testing at two separate sites, dense sand at Dunkirk and stiff clay at Cowden. The
work is being directed by an Academic Work Group (AWG), led by the University of Oxford and
comprising researchers from Imperial College and UCD in collaboration with the Geotechnical
Consulting Group and supported by the larger PISA industrial consortium. The comprehensive 3D finite
element analyses that were completed for the development of the design method and to inform the field
test programme are described by Zdravkovic et al. (2015).

3.2.2. Design Models


A new design approach that draws on the existing p-y method and is based on 1D finite element model of
a monopile has been developed. In addition to the standard distributed load (p) curve See Figure 10, the
model includes a distributed moment curve (associated with the vertical shear stresses on the pile shaft), a
base shear curve, that defines the relationship between the shear force, S and lateral displacement, and a
base moment curve, M b considering the rotation at the pile toe. Two approaches are considered for
determining the soil reaction curves; (i) a rules based method with recommended parameters determined
from basic strength and stiffness parameters and (ii) a more detailed approach using bespoke finite
element analyses to derive site specific soil response functions.

242
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 10: Soil reaction components applied to a monopile (Byrne et al 2015a)

The field test programme had a vital role in the development of the new design methods, by validating
both the numerical models and the parameterisation of the soil response.

3.2.3. Field Tests


The major field test programme undertaken at the two sites performed between October 2014 and July
2015, summarised below is described in detail in Byrne et al (2015a). The field tests were performed in
an onshore setting to allow for multiple tests to be performed that investigate a range of the parameters
that govern pile performance and to investigate the effects of scale. The test piles were loaded at a height
h, above ground level of between 5 and 10m in order to consider normalised Moment, M, horizontal load,
H and pile diameter, D ratios representative of those arising from wind and wave loads applied to
offshore wind turbines. The piles were instrumented with fibre optic strain gauges, inclinometers and
extensometers in the buried section (piezometers were placed in the soil at Cowden). In addition above
ground instruments included inclinometers, LVDT’s and load cells, See Figure 11.

RAMSTROKE

LOAD CELL - LCL

STICK UP
INCLINOMETERS - SUI

DISPLACEMENT
TRANSDUCERS - LVDT M/H

AH PH

AL PL

EXTENSOMETERS - IPE z
INCLINOMETERS - IPI
L
FIBRE OPTICS - FOS

PIEZOMETERS - PZT

D
Figure 11: Instrumentation used in the PISA test programme (McAdam pers com.)

In total there were 14 test piles at each site. Three pile diameters were tested, 0.273m, 0.762m and 2m.
The majority of tests were performed on the mid-size 0.762m diameter piles. The focus of the PISA
project was on monotonic loading and tests were carried out as a series of maintain load steps, which

243
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

targeted specific displacements and included a main unload-reload loop after the 3rd load step. A limited
number of additional load tests were performed to explore rate effects, cyclic loading and damping.

3.3. REDWIN Project


The REDWIN (REDucing cost in offshore WINd by integrating structural and geotechnical design) aims
to bring together the structural and geotechnical analyses performed for offshore wind turbines. The
project will develop integrated soil-structure interaction models, See Figure 12 that account for stiffness,
damping, drainage, degradation and long-term behaviour. The foundation models being developed based
on the concept of macro-elements, and soil interface models are based on the concept of distributed
foundation springs such as generalized p-y models.

Figure 12: Illustration of soil interaction models: a) foundation model, and b) soil interface model (from
www.ngi.no/en/Project-pages/REDWIN)

The project is led by the NGI and partners including research and industrial organisations; NTNU, Statoil,
Statkraft, IFE and Olav Olsen AS. Technical work packages will include a work package aimed at
compiling measurement data such as the natural frequency response of turbines in operation and data
from model and full-scale pile tests from the literature. A work package will be dedicated to development
of a library of 3D soil models capable of converting raw soil date into input data for advanced soil-
structure interaction models. Work package four concentrates on implementing the 3D Soil models into a
time domain aero-elastic code.

3.4. Vibro Project


Environmental concerns regarding pile driving is driving research into alternative installation methods as
the noise associated with driving large diameter piles has been shown to affect sea mammals (Igoe et al
2013). Whilst vibratory installation is relatively common in the offshore setting, uncertainties regarding
the potential effect of installation method on the capacity of monopiles were the driver for the Vibro
project. The project was led by RWE and the consortium included major utilities; engineering
consultants, pile manufactures, authorities, certifying body’s universities and research institutes. The
project involved a major field trial where six 4.3 m diameter, 18 m long, instrumented monopiles were
installed in medium-dense sand at Cuxhaven, See Figure 13. Three of the piles were driven and three
were vibrated in place. The piles were statically load tested in pairs (i.e. a driven pile and vibrated pile
were loaded of each other) to assess the effect of installation on the lateral capacity and stiffness of the
piles. An interesting feature of the project was that CPT tests were performed before and after pile
installation.

244
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Figure 13: Artist impression of test site in Cuxhaven. ©VIBRO-project by RWE

4. ONSHORE PILING PROJECTS

4.1. Introduction
In addition to new markets in offshore engineering a number of interesting research topics are ongoing in
the onshore field. These relate for example to different pile installation methods with screw piles
becoming increasingly used and problems such as the development of negative skin friction. Another
topic that is currently being studied extensively are thermally activated pile foundations (so-called energy
piles) for heating and cooling purposes. Finally, the mechanical behaviour of micropiles, a foundation
system with a large potential in a wide range of applications, is also subject of research. Following
sections discuss briefly the ongoing research on these topics.

4.2. Screw piles

4.2.1. Background
Screw piles are finding increased market share due to several advantages including; (i) the low-
displacement nature of their installation, (ii) low noise and (iii) their installation speed. A major research
project was performed by the Belgian Building Research Institute (BBRI) organized, with the financial
support of the Belgian Federal Ministry of Economic Affairs (BBRI, 1998-2000 & 2000-2002) to
investigate the behaviour of screw piles and to compare this to pre-cast piles. The research was performed
at two sites. At the first site located in Sint-Katelijne-Waver, five types of screw piles and driven precast
piles were installed in over-consolidated tertiary Boom clay. In total thirty pile load tests (including static,
dynamic and Statnamic loading) were performed. A similar test campaign was performed at the second
site in Limelette where the subsoil consists of quaternary silty layers (loam) and tertiary Ledian-
Bruxellian sand. This summary of the work concerns results from the first test site.

4.2.2. Site description and pile details


A major field investigation programme include in-situ CPT, pressuremeter, SPT, DMT and geophysical
testing was performed. This was augmented with laboratory tests including triaxial strength tests with
bender elements. Huybrechts (2001) gives details of the installation procedure for the six different types
of piles were installed and tested, five cast-in-place screw piles and one pre-cast driven pile;

- Atlas pile, installed by the company Franki Geotechnics B


- De Waal pile, installed by the company De Waal Palen
- Fundex pile, installed by the company Fundex
- Olivier pile, installed by the company Olivier
- Omega pile, installed by the company Socofonda

4.2.3. Load Tests


For each test site a series of static, dynamic and kinetic pile load tests were performed that included:

- Instrumented static load tests (two tests per pile type at each site);
- Dynamic pile load tests (two tests per pile type on each site and a second series of twelve pile load tests
after one year on the Sint-Katelijne-Waver test site, in order to study the time effect);
- Statnamic load tests (one test per pile type on each site)

245
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

The results of the static pile load tests are summarized in figures 14, where the measured loads are
normalised by estimates of the pile capacity made using CPT based design methods and the vertical
displacement is normalised by the pile diameter. The load test programme proved that the screw piles had
comparable capacities (considered at normalised displacement of 10%) to the driven pre-cast piles. Given
the environmental and time savings achieved by screw piles this was a very positive outcome.

Figure 14: Normalised load settlement curves static pile load tests Sint-Katelijne-Waver

4.3. Development of negative skin

4.3.1. Background
In 2013 a research and development project was initiated by the Engineering College, Aarhus University
in collaboration with Per Aarsleff A/S with the aim to further document the influence of bitumen coating
and the development of negative skin friction on precast concrete piles installed in soft soils. As part of
the project a novel full-scale pile test setup, involving instrumented piles and a ground monitoring
program was established at Randers Harbour, Denmark.

4.3.2. Site conditions and test setup


The test setup consists of four test piles (T) and five reaction piles (R) installed in a row with 3 m centre-
centre spacing, as illustrated in Figure 15. Figure 15b shows details of the pinning of the test piles to the
HE 280 M crossbeam, which enables measurement of the total down drag force on the pile using a
standard load cell. The toes of the test piles are located in the soft organic post glacial marine Gytja, while
the reaction piles are embedded into the underlying layers of marine sand. All piles are precast reinforced
concrete piles with a square cross-section (0.25 m width). Bitumen coating was applied to test piles T1
and T3, while Piles T2 and T4 are uncoated.

Monitoring equipment:
Piezometers (4)
Magnetic
Extensometer (4)
Settlement plates (3)
Pile strain gauges (16)

All levels are in meters above


mean sea level (DVR90).
T4 T3 (B) T2 T1 (B)
Distances are in meters.
R5 R4 R3 R2 R1 Artesian water table in sand at Figure 15b. Photo showing
~16 m roughly +1 m..
c/ c 3 m pinning of test pile to cross-
beam.
Figure 15a. Longitudinal section view and situation plan of the
pile test setup before terrain loading (after Ventzel and Jensen,
246
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Multipoint vibrating wire piezometer (4 piezometers), settlement plates (3) and a magnet extensometer (4
probes) were used to monitor pore water pressures and settlements, while a total of 16 embedded
vibrating wire strain gauges were cast into the test piles to measure internal pile strains. In addition to
CPT and field vane testing carried out as part of the site investigation, laboratory testing (oedometer tests,
interface shear tests) have been carried out to obtain soil and interface parameters, while the pile stiffness
has been estimated from in-situ loading of the test piles in addition to tension tests on representative pile
sections.
After pile installation and a period of rest (~ 200 days), 0.8 m of fill was placed around the piles to initiate
settlement in the soft soil and hence to generate negative skin friction on the piles.

4.3.3. Preliminary key findings


Figures 16a shows the observed development in the total down drag force on the test piles until 315 days
after ground loading (fill placement), while Figure 16b shows the relationship between measured total
down drag forces and changes in the average excess pore water pressure in the top layers. Based on the
preliminary observations the following key findings have been made.

 Significant reduction seen in total down drag force due to bitumen coating. A reduction greater than
80% is supported by the findings, cf. Figure 16b.
 The study has indicated that the negative skin friction may be significantly overestimated if
calculation is carried based on an undrained analysis (α-method). The results suggest that a drained
analysis may be more appropriate.
 The mobilisation of down drag forces on both coated and uncoated piles is seen to be proportional to
changes in excess pore water pressure / effective normal stress on the pile.
 The on-going study has furthermore highlighted a high degree of uncertainty in the estimation of
stress distribution on the test piles. The uncertainty stems from the use of a limited number of single
point measurements of strains in addition to a significant influence of concrete shrinkage on the strain
measurements and uncertainties about the concrete stiffness.

Time [days] after loading Avr. excess pwp in top layers [kPa]
-200 0 200 400 -5 0 5 10 15
-10 0
Total down drag force [kN]
Total down drag force [kN]

0 10
10 Coated
T1 20
20
30 T3 30
40 T2 40
T1
50 Uncoated T4
50 T3
60 T2
60
70 T4
80 70

Figure 16a. Development of total down drag Figure 16b. Changes in total down drag
forces on test piles after loading forces on test piles vs. changes in the average
excess pore water pressure

Monitoring of the field test is still on-going, and the results are yet to be published. Preliminary and
selected results have been presented by Sorensen (2015) and in this paper. To verify if the full down drag
forces have been mobilised on the test piles further ground loading is planned.

4.4. Thermo-mechanical behaviour of energy piles

4.4.1. Background
Although the integration of heat exchangers in foundation piles to thermally activate them as shallow
geothermal source has become common practice, the thermal and thermo-mechanical behaviour remains a
topic of discussion. Former and current research work focuses on the short and long term energetic
performance and thermo-mechanical behaviour in terms of bearing capacity and settlements.

4.4.2. Energy pile testing projects


Based on three studies in Austria (Brandl, 1998), Switzerland (Laloui et al., 2006) and the UK (Bourne-
Webb et al., 2009) including in-situ tests, a simplified descriptive framework regarding the thermo-

247
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

mechanical performance of energy pile foundations has been developed (Bourne-Webb et al., 2013). This
framework takes into account the effect of heating and cooling, pile shaft – soil interaction, mechanical
loading and end-restraints. An example is shown in Figure 17, illustrating the influence of heating and
cooling of a mechanically loaded, at both ends restrained foundation pile on the vertical load distribution
and shaft friction.

Figure 17: Impact of end-restraint on thermal load effects (in dashed line the effect of pure mechanical
loading). (Left) Heating and both end restraints; (Right) Cooling and both ends restraints (from Bourne-
Webb et al., 2013).

Meanwhile, several other energy pile test programs have been and are being carried out, for example in
the USA, Australia, UK, France, Spain and Belgium. Figure 18 shows some energy piles before and after
installation, as well as the test setup for mechanical loading in the framework of a research program in
Ostend (Belgium). Small-scale and centrifuge testing and numerical modelling (Finite Element, Finite
Difference, …) are performed as well aiming to improve the fundamental understanding of energy pile
thermo-mechanical behaviour. An extensive overview of on-going research is given by Olgun and
Loveridge (2015).

.
Figure 18: Energy piles installed for a BBRI test program in Ostend (Belgium). The heat exchangers are
attached to the reinforcement cage prior to installation. The energy piles are fully monitored with strain
and temperature sensors.

In all research projects the influence of several parameters is studied more detailed such as pile type, pile
diameter, soil type and end-restraints. In future work, these results will be related to the existing
framework, for example within the scope of the European COST Action TU 1405 ‘European network for
shallow Geothermal energy Applications in Buildings and Infrastructures (GABI)’. This project also aims
to propose design rules for energy piles, based on a benchmark about energy pile modelling and design

248
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

throughout Europe and the world. The workgroup meetings are open to everyone who is interested. In
2012, the Ground Source Heat Pump Association (GSHPA) already published a Thermal Pile Standard
(GSHPA, 2012). This standard was based on the GSHPA borehole loop guide, but expanded to energy
pile applications and already contains some design considerations.

4.5. Micropiles
Micropiles are widely used in areas that are difficult accessible for large drilling rigs. Typical applications
are underpinning works, extension or reinforcement of existing infrastructure, stabilisation of slopes,
tension piles for the lateral support of retaining structures, etc. Many installation methods and materials
are available on the market and the bearing capacity of these elements depends strongly on the installation
method, including grout injection parameters.
In order to tackle this problem of the bearing capacity of isolated micropiles a load testing program is
going on in Belgium BBRI (2014-2016). The main objective of this research program is to link the
execution and grout injection method to the bearing capacity of the micropile systems, to categorise these
foundation elements and to translate the results in the framework of the Belgian semi-empirical design
method according to Eurocode 7. Some first results were published in the Volume 3 of BBRI (2008).
Besides the behaviour of isolated micropiles, the group effect and the effect of cyclic loading is assessed
in the research program by an experimental test set up as well as by 3D FEM numerical modelling (see
figure 19). The publication of the results of this research are expected by the end of 2016 – begin 2017.

Figure 19: Group effect on micropiles - numerical 3D FEM simulations..

5. CONCLUSIONS
Despite the relatively advanced state of the European piling industry many critical issues remain and
these are being addressed through research and development projects. Issues such as new applications of
existing technology, e.g. offshore wind developments, climate change and environmental concerns are the
drivers for this research need. This paper provides a brief overview of projects been undertaken by
members of ETC3. The positive collaboration between industry and academia in many JiP projects is an
indicator of the importance of this ongoing work and many challenges remain.

249
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

REFERENCES
Axelsson (2000a) Long term set-up of driven piles in sand, PhD Thesis, Dept. of Civil and Environ-
mental Engineering. Stockholm: Royal Institute of Technology. Set-Up of Driven Piles in Sand

BBRI (1998-2000 & 2000-2002). Soil displacement screw piles – calibration of calculation methods and
automatisation of the static load test procedure : stage 1 - friction piles & stage 2 – end-bearing piles.
Research programs subsidised by the Belgian Federal Ministry of Economical Affairs, convention
numbers CC-CIF-562 and CC-CI-756.

BBRI (2014-2016), Micropieux – développement d’une méthode de dimensionnement belge intégrée.


Recherche subsidiée par le NBN et le SPF-Economie, convention nr. CCN/NBN/PN14A14 & PN14B14

BBRI. 2008. Proceedings of the international symposium “Ground anchors – Limelette test field results”,
Brussels, 14.05.2008 (proceedings volume 1, 2 and 3, available on www.tis-sft.wtcb.be)

Briaud, J.L., Tucker, L.M., and Ng, E., (1989). “Axially loaded 5 pile group and single pile in sand”,
Proceedings of the International Conference of Soil Mechanics and Foundation Engineering, Rio de
Janeiro, August 1989.

BW Byrne, RA McAdam, HJ Burd, GT Houlsby, CM Martin, K Gavin, P Doherty, D Igoe, L Zdravković,


DMG Taborda, DM Potts, RJ Jardine, M Sideri, FC Schroeder, A Muir Wood, D Kallehave, J Skov
Gretlund, (2015a) Field testing of large diameter piles under lateral loading for offshore wind
applications Proceedings of the 16th European Conference on Soil Mechanics and Geotechnical
Engineering, Edinburgh, UK

BW Byrne, R McAdam, HJ Burd, GT Houlsby, CM Martin, L Zdravković, DMG Taborda, DM Potts, RJ


Jardine, M Sideri, FC Schroeder, K Gavin, P Doherty, D Igoe, A Muir Wood, D Kallehave, J Skov
Gretlund (2015b) New design methods for large diameter piles under lateral loading foroffshore wind
applications International Symposium on Frontiers in Offshore Geotechnics, Olso, June

Brucy, F., Meunier, K. and Nauroy, J. F. (1991) "Behaviour of pile plug in sandy soils during and after
driving," Proc., Offshore Technology Conference, pp 145-154.

Bourne-Webb P, Amatya B, Soga K et al. (2009) Energy pile test at Lambeth College, London:
Geotechnical and thermodynamic aspects of pile response to heat cycles. Géotechnique 59(3): 237–248.

Bourne-Webb, P. J., Amatya, B., Soga, K., 2013. A framework for understanding energy pile behaviour.
Proc. of the Inst. Of Civ. Eng., 166. pp. 170-177.

Brandl H (1998) Energy piles and diaphragm walls for heat transfer from and into the ground.
Proceedings of the 3rd International Geotechnical Seminar, Deep Foundations on Bored and Auger Piles
(BAP III), University of Ghent, Ghent, Belgium, vol. 1, pp. 37–60.

L Zdravković, DMG Taborda, DM Potts, RJ Jardine, M Sideri, FC Schroeder, BW Byrne, R McAdam, HJ


Burd, GT Houlsby, CM Martin, K Gavin, P Doherty, D Igoe, A Muir Wood, D Kallehave, J Skov
Gretlund, (2015) Numerical modelling of large diameter piles under lateral loading for offshore wind
applications International Symposium on Frontiers in Offshore Geotechnics, Olso, June

Chow, F. C. (1997). Investigations into behaviour of displacement piles for offshore structures. PhD
Thesis, Civil Engineering, University of London (Imperial College).

Chow. F.C., Jardine, R.J., Brucy, F and Nauroy, J.F. (1997). “Time related increases in the shaft ca-
pacities of driven piles in sand”, Geotechnique, Vol 47, No.2, pp 353-361.

Chow F.C., Jardine R.J., Naroy J.F. and Brucy F. (1998) ”Effects of time on the capacity of pipe piles in
dense marine sand”, J. Geotech. Eng., ASCE vol. 124(3), pp 254-264

Doherty. P, Kirwan. L, Gavin, K, Igoe, D, Tyrrell. S, Ward, D and O’Kelly, B (2012a), Soil Proper-ties at
the UCD Geotechnical Research Site at Blessington, Paper accepted to BCRI Confer-ence, Dublin,
September 2012.

250
ISSMGE - ETC 3 International Symposium on Design of Piles in Europe. Leuven, Belgium, 28 & 29 April 2016

Igoe, D. Gavin, K and O'Kelly, B (2013) 'An investigation into the use of push-in pile foundations by the
offshore wind sector'. International Journal of Environmental Studies, 70 (5):777-791.

Schneider, J.A. (2009), Uncertainty and bias in evaluation of LRFD ultimate limit state for axially loaded
driven piles, Deep Foundations Institute Journal, Vol.3, No.2, November.

Gavin, Kenneth, Igoe, David, Doherty, Paul : Piles for offshore wind turbines: A state of the art review.
Proceedings of the ICE - Geotechnical Engineering, 164 (4) 2011-08-01, pp.245-256.

Gavin K.G. and O’Kelly B.C. (2007). “Effect of friction fatigue on pile capacity in dense sand”, Journal
of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 133, No. 1, pp 63-71.

Huybrechts, N. (2001). Test campaign at Sint-Katelijne-Waver and installation techniques of screw piles,
Proceedings of the symposium Screw piles – installation and design in stiff clay, March 2001, Brussels

Jardine, R. J., Chow, F. C., Overy, R. F. & Standing, J. (2005). ICP Design Methods for Driven Piles in
Sands and Clays. Thomas Telford, University of London (Imperial College).

Karlsrud, K., Jensen, T.G, Wensaas Lied, E.K, Nowacki, F, and Simonsen, A.S. (2014), Significant ageing
effects for axially loaded piles in sand and clay verified by new field load tests. Proceedings of the
Offshore Technology Conference, Houston, May 2014, Paper No. OTC-25197-MS.

Kirwan, L. (2015) Investigation into Ageing Mechanisms for Axially Loaded Piles in Sand, PhD under
review, University College Dublin

Laloui L, Nuth M and Vulliet L (2006) Experimental and numerical investigations of the behaviour of a
heat exchanger pile. International Journal of Numerical and Analytical Methods in Geomechanics.

Lehane, B. M. (1992). Experimental investigations of pile behaviour using instrumented field piles. Civil
Engineering. London, University of London (Imperial College).

Lehane, B. M., Schneider, J. A. & Xu, X. (2005). The UWA-05 method for prediction of axial capacity of
driven piles in sand. Frontiers in Offshore Geotechnics: ISFOG. Perth, University of Western Australia

Norwegian Geotechnical Institute (2014) Time effects on pile capacity. Summary and evaluation of test
results, Report 20061251-00-279, Rev.1 March 2014.
Olgun, G., Loveridge, F. (2015). Energy geo‐structures and storage of thermal energy in the ground: task
force 3 position paper. International Symposium on Energy Geotechnics (SEG). "SEG 2015". Universitat
Politècnica de Catalunya ed. Barcelona, 2015.

Sørensen, K. K. (2015), Fuldskala pæleforsøg på Randers havn/Full scale pile testing at Randers
Habour, presentation, www.danskgeotekniskforening.dk, DGF meeting no. 4/2015

251
252
GOLD Sponsors

Organising Secretariat

You might also like