Buckley Et Al 2020 Full Scale Observations of Dynamic and Static Axial Responses of Offshore Piles Driven in Chalk and

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Buckley, R. M. et al. (2020). Géotechnique 70, No. 8, 657–681 [https://doi.org/10.1680/jgeot.19.TI.

001]

Full-scale observations of dynamic and static axial responses of offshore


piles driven in chalk and tills
ROISIN M. BUCKLEY , RICHARD J. JARDINE†, STAVROULA KONTOE†, PEDRO BARBOSA‡ and
FELIX C. SCHROEDER§

This paper describes and interprets tests on piles driven through glacial tills and chalk at a Baltic Sea
windfarm, covering an advance trial campaign and later production piling. The trials involved six
instrumented 1·37 m dia. steel open-ended tubes driven in water depths up to 42 m. Three piles were
tested statically, with dynamic re-strike tests on paired piles, at 12–15 week ages. Instrumented dynamic
driving and re-strike monitoring followed on up to 3·7 m dia. production piles. During driving, the
shaft resistances developed at fixed depths below the seabed fell markedly during driving, with
particularly sharp reductions occurring in the chalk. Shaft resistances increased markedly after driving
and good agreement was seen between long-term capacities interpreted from parallel static and
dynamic tests. Analyses employing the sites’ geotechnical profiles show long-term shaft resistances in
the chalk that far exceed those indicated by current design recommendations, while newly proposed
procedures offer good predictions. The shaft capacities mobilised in the low-plasticity tills also grew
significantly over time, within the broad ranges reported for sandy soils. The value of offshore field
testing in improving project outcomes and design rules is demonstrated; the approach described may be
applied to other difficult seabed conditions.

KEYWORDS: chalk; full-scale tests; glacial soils; piles & piling

INTRODUCTION located roughly midway between the Rugen and Bornholm


Most offshore oil, gas and wind-turbine structures rely on islands in the German Baltic. Seventy 5 MW wind-turbine
open-ended driven steel piles, whose outside diameters, D, generators (WTGs) and an offshore substation (OSS) were
range between 0·8 and 8 m and penetrations may exceed installed in 2017. Each WTG’s jacket structure is founded on
100 m (Hamre, personal communication, 2018; Jardine, four 2·7 m outside diameter, D, tubular driven steel piles,
2019). Open piles are also driven for bridge, harbour and while the OSS relies on six 3·7 m dia. piles. The piles, whose
other foundations. Field testing is required to verify load- environmental loads invoke axial ‘push–pull’ and lateral pile
carrying capacities under some regulatory regimes, as in the reactions, penetrate through dense/stiff sandy tills over low–
German (BSH, 2015) framework, where Eurocode 7 applies medium density chalk and, in some cases, limestone
in accordance with DIN (2009, 2012). Although static load (Barbosa et al., 2015a, 2015b).
tests are common onshore, they are usually deemed unfea- Iberdrola, the project developers, conducted comprehen-
sible offshore. Checks may be conducted at analogous sive site investigation and ‘pre-construction’ piling cam-
onshore sites (e.g. Al-Shafei et al., 1994; Lahrs & Kallias, paigns at Wikinger. Six instrumented 1·37 m dia. steel
2013) and through dynamic driving monitoring offshore. (grade S355) open-ended piles were driven with dynamic
However, it is often difficult to match field conditions monitoring, in up to 42 m of water, near the WK38, WK43
onshore, or interpret dynamic tests accurately. Onshore and WK70 turbine locations. One pile was tested statically at
tests show axial capacities increasing significantly with time each location after 12–15 weeks of ageing, with novel,
after driving in sands (Jardine et al., 2006), low-plasticity remotely operated seabed testing systems, before conducting
clays (Karlsrud et al., 2014) and in chalk formations, for parallel dynamic re-strikes on adjacent paired piles. Fig. 1
which accurate capacity predictions are particularly hard to sets out the relative positions of the piles, boreholes
make (see Lord et al., 2002; Ciavaglia et al., 2017; or Buckley and piezocone/cone penetration tests (CPTs) at each
et al., 2018a). However, long-term re-strike check tests are location. Additional instrumented dynamic driving and
rarely undertaken offshore. re-strike monitoring followed at five ‘production’ piling
This paper describes and interprets novel static and locations, as summarised in Table 1. A joint project involving
dynamic testing for the Wikinger windfarm, which is Innovate-UK, Iberdrola, Imperial College and Geotechnical
Consulting Group (GCG) supported the offshore research
with parallel onshore pile experiments, laboratory studies
and analytical developments.
Manuscript received 8 July 2018; revised manuscript accepted
13 November 2019. Published online ahead of print 17 December
2019.
Discussion on this paper closes on 1 December 2020, for further BACKGROUND
details see p. ii. Pile shaft capacities, Qs, represent the integration over their
 Formerly Department of Civil & Environmental Engineering,
perimeters πD and shaft lengths, Lp, of the contributions
Imperial College London, UK; now Department of Engineering made by local limiting shaft shear stresses, τf, which usually
Science, Oxford University, UK (Orcid:0000-0001-5152-7759). vary with depth
† Department of Civil & Environmental Engineering, Imperial
College London, UK. ð Lp
‡ Iberdrola Renovables Offshore, Deutschland, Berlin, Germany. Qs ¼ πD τ f dz ð1Þ
§ Geotechnical Consulting Group LLP, London, UK. 0

657

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
658 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER
x: m resistances, qc, choosing between four ‘CPT’ procedures cited
0 10 20 30 in the commentary. The latter include the ICP-05 procedure
0 described by Jardine et al. (2005) in which τf is related to local
WK38-1 (Reaction pile + DLT) shaft radial effective stresses at failure, σrf′ , by the Coulomb
expression given as equation (2)
BH-WK38
2 τ f ¼ σ′rf tan δ′ ð2Þ

WK38-2 (SLT) The radial effective stresses reduce with increasing h/R*
where h is the relative distance from the pile tip and R* is the
pile’s equivalent ‘solid’ radius, R ¼ ðR2  R2i Þ05 where Ri
y: m

4
CPT-WK38 and R are the inner and outer radii. Interface shear angles, δ′,
′ , to ultimate, δult
can vary from peak, δpeak ′ , values when local
WK38-3 (Reaction pile) slip occurs in brittle soils. However, shaft capacities can grow
6
markedly in sand layers over the weeks and months that
follow driving and exceed the medium-term (typically from
ten to 30 day age) estimates provided by approaches such as
Driven pile Borehole CPT
8
the ICP-05 (Jardine et al., 2006).
(a) Offshore designers usually base their shaft capacity
x: m assessments for clay layers on the local undrained shear
0 3 6 9 12 strengths, su, and α coefficients that vary with su/σv0 ′ . Other
0 methods exist to assess shaft capacity in clay – for example
WK43-3 (Reaction pile)
the ‘Fugro-96’ method (Kolk & der Velde, 1996), the
‘NGI = 05’ method (Karlsrud et al., 2005) and the ICP-05
clay method set out by Jardine et al. (2005), which also
10 WK43-2 (SLT)
employs equation (2) and recognises, as with clays, a
dependence of σrf′ on h/R*. Lehane et al. (2013) have also
WK43-1 (Reaction pile + DLT)
proposed direct use of CPT cone resistance, qt, for clays and
report encouraging results for their ‘UWA-13’ methods in
y: m

20 CPT-WK43a comparison with outcomes from 43 static load tests.


Significant effects of pile age after driving on shaft capacity
BH-WK43 CPT-WK43
have also been noted in low-plasticity clays (Karlsrud et al.,
30 CPT-WK43b
2014).
The carefully quality-assured database of Lehane et al.
(2017) included only one site (Cowden in Humberside, UK)
where stiff, low-plasticity glacial tills contributed the main
40 parts of the piles’ shaft capacities. More high-quality tests are
(b) needed to assess how routine design methods apply to such
x: m
strata. Weltman & Healy (1978) presented a review for UK
40
glacial clay tills and concluded that, for the tests they
0 10 20 30
0 analysed, the driven piles’ α coefficients declined, as often
expected, with increasing su to reach 0·4 at the maximum su
( 210 kPa) considered. Well-graded offshore tills with
significant fines contents are sometimes treated as clays,
10 WK70-3 (Reaction pile) WK70-1 (Reaction pile + DLT) especially when they manifest non-hydrostatic trends in
piezocone tests. However, low-plasticity sandy offshore tills,
WK70-2 (SLT) such as those encountered at Wikinger, may not behave in the
same ways as terrestrial tills, marine or alluvial clays.
y: m

20 The most common current industrial practice is to apply


the Construction Industry Research and Information
BH-WK70a+70b
Association’s (Ciria) C574 (Lord et al., 2002) to open piles
CPT-WK70a driven in chalk. A conservative interpretation by Lord et al.
30
(2002) of six static tests led to Ciria C574 recommending
fixed ultimate unit shaft resistances of 20 and 120 kPa for
low–medium and high-density cases, respectively. Noting the
40 paucity of chalk cases, the authors conducted multiple
(c) experiments with driven steel open-ended piles and jacked
highly instrumented Imperial College piles (ICPs) in low–
Fig. 1. Plan showing relative layout of test piles, boreholes and cone medium density chalk at St Nicholas-at-Wade in Kent, UK.
penetration tests at each pre-construction test location: (a) WK38; (b) Buckley et al. (2018a) and Buckley et al. (2018b) drew the
WK43; (c) WK70. Note: each plan based on arbitrary site datum following conclusions.
position

(a) Very high excess pore water pressures develop around


pile tips and low-strength chalk putty annuli form
Offshore pile designers often employ the American around their shafts during driving.
Petroleum Institute (API) RP2 GEO (API, 2014) or ISO (b) Chalk’s relatively high in situ permeability allows
19902:2007 (ISO, 2007) recommendations in sand layers to partial drainage during both CPT penetration and pile
predict τf from the estimated in situ effective overburden installation. Marked water content reductions develop
′ , through β coefficients, or from the local CPT tip
pressure, σv0 as pore pressures dissipate.

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 659
Table 1. Details of the test sites, pile diameters and pile penetrations

Test site Testing stage tw: mm D: m Length of penetration, Lp: m

Total Holocene Glacial till Fluvioglacial till Chalk CPT

WK38 Pre-construction 40 1·37 16·6 1·9 11·7 — 3·0 11·2


WK43 Pre-construction 40 1·37 30·7 — 10·3 — 20·4 11·0
WK70 Pre-construction 40 1·37 31·0 — 6·6* — 24·2 7·4
OSS Production 60 3·67 46·3 2·0 8·3 — 36·0 35·7†
WK08 Production 40·5 3·67 27·9 4·1 9·5 9·5 4·8 49·3
WK11 Production 40·5 2·70 31·1 4·5 12·0 — 14·6 33·3
WK13 Production 40·5 2·70 25·8 3·0 15·9 — 6·9 24·2
WK42 Production 40·5 2·70 19·9 0·1 9·9 9·2 0·7 50·4

*Re-interpretation of SI results indicates glacial till to 7·5 m, underlain by Danian limestone between 7·5 and 14 m, which is in turn underlain
by L-M density chalk.
†Penetration of cone penetration test below seabed level.

(c) Low average shaft resistances, broadly comparable with In situ testing
Ciria C574, apply during and immediately after As indicated in Fig. 1, at least one borehole and CPT were
driving. However, shaft shear stress distributions vary undertaken at each test location. As might be expected,
markedly with depth and show far stronger reductions conditions are highly variable within the low-plasticity
with relative pile tip depth h/R* than apply in clays or glacial and fluvioglacial strata. Augustesen et al. (2015)
sands. and Barbosa et al. (2017) treated the glacial till and
(d ) The shaft radial effective stresses developed during fluvioglacial till layers as clays in their initial design studies.
installation correlate with the CPT cone resistance, Secant shear stiffnesses were measured in the glacial till by
mobilising comparably low σrf′ /qt ratios to crushable high-pressure dilatometer tests (CI, 2013), while shear, Vs,
calcareous sands. and compression wave, Vp, velocities were logged in the chalk
(e) Driven open-ended piles’ shaft capacities can increase through P–S probes suspended in boreholes (CIJV, 2013).
five-fold after driving to give long-term unit shaft Fig. 2 shows profiles of maximum shear modulus, Gvh, in the
resistances far above the Ciria C574 values. Set-up rates chalk, as calculated from Vs measurements at five locations.
are sensitive to site-specific features such as the Consistent with nearby observations by Obst et al. (2017), Vs
discontinuity sets, the installation process and typically ranged from 0·8 to 1·1 km/s, increasing only slightly
physio-chemical processes.
( f ) Equation (2) describes shaft failure accurately in chalks,
with δ′ angles that match laboratory interface tests. As P–S logging Gvh: MPa
in sands, shaft radial effective stresses increase during
0 1000 2000 3000
static loading to failure. 10

The above results derived principally from relatively Test sites


small-scale experiments conducted at an onshore site. This WK70
paper reports full-scale static and dynamic tests that explore WK43
15
the corresponding behaviours of large piles driven for the WK13
Wikinger offshore windfarm project. The Wikinger ground WK11
conditions are considered first, setting out the geotechnical OSS1
parameters required for test analysis. The pile testing
Trend equation (3)
programme and procedures are then outlined before report- 20
ing and interpreting the outcomes. Finally, the offshore data
are employed to assess the applicability of recently proposed
design rules for piles driven in chalk.
Depth: mbsb

25

GROUND CONDITIONS Envelope from


The Wikinger stratigraphy consists of Pleistocene glacial across the site
and fluvoglacial tills over low–medium-density structured 30
Upper to Late Cretaceous chalk, which is incised by narrow
sub-glacial channels infilled with tills; Upper Cretaceous or
Danian limestone is also encountered. The windfarm area is
large and representative pile test locations were chosen that
cover the most adverse conditions (CIJV, 2014a) and 35
encompass the ground profiles and pile penetrations, Lp,
summarised in Table 1. The WK13 and WK38 test piles were
dominated by glacial till, while the WK08 and WK42
production piles penetrated both glacial and fluvioglacial 40
tills. The OSS, WK43 and WK70 piles are dominated by
low–medium-density chalk, while WK11 penetrated almost Fig. 2. Shear modulus, Gvh, in chalk from suspension P–S logging at
equal proportions of glacial till and chalk. the test sites (mbsb = metres below seabed)

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
660 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER
with depth. The Gmax shear moduli were assessed by CIJV glacial tills as silty sands and sandy silts, while the
(2014b) as varying linearly with depth below the seabed (z, in fluvioglacial tills appear as sandy silts to silty clays.
m) in the low–medium-density chalk Considering the glacial till’s qt range of 6–20 MPa, the
correlation of Baldi et al. (1989) for sandy soils suggests
Gmax ¼ 173z þ 895ðMPaÞ ð3Þ
maximum shear moduli in the 200–1500 MPa range.
The corresponding Vp velocities were all  2 km/s. Most of the Wikinger chalk classifies as structured low–
Figure 3 shows qt profiles assessed at the three static test medium-density with grade A1/A2 (Bowden et al., 2002). It
sites, as well as the WK42 production location where is extremely weak with closely spaced, closed or clean
fluvioglacial till is present. The glacial tills generally fractures. The chalk qt profiles were averaged over 0·3 m
showed 3 , qt , 30 MPa, with peaks up to 50 MPa in penetration intervals, following Smith (2001), generally
isolated, thin, dense sand layers and sleeve frictions, giving 10 to 20 MPa in the structured chalk, but with
100 , fs , 300 kPa. Excess penetration pore pressures, isolated peaks up to 60 MPa. Sleeve frictions generally fell
measured at the u2 position, showed generally negative between 200 and 400 kPa, while u2 values were remarkably
values (100 to 250 kPa) with discrete peaks up to high and increased to reach 10 MPa at 30 m below the
+1 MPa. The relatively deep fluvioglacial till encountered seabed. Field permeability is often sufficiently high in intact
at WK42 showed lower average qt values (generally chalk for it to drain freely under field foundation loading
, 10 MPa), an fs range of 300–600 kPa and positive conditions (Lord et al., 2002). Piezocone dissipation tests
penetration u2 values up to 1 MPa, all suggesting higher performed at St Nicholas-at-Wade in a similar low–medium-
fines contents than in the glacial till. Fig. 4 plots the soil density fractured grade B2/B3 chalk indicated times for 50%
behaviour type index, Ic (Robertson, 1990), in the dissipation of penetration generated excess pore water
glacial/fluvioglacial till layers. This re-analysis classifies the pressures of less than 10 s (Buckley et al., 2018a).

qt: MPa qt: MPa


0 10 20 30 40 50 0 10 20 30 40 50
0 0
Holocene deposits
2
CPT-WK38 Glacial till
4
5
CPT-WK42a
6
Glacial till
Depth: mbsb

Depth: mbsb

8
10
10

Fluvioglacial till
12
Adopted profile Adopted profile
15
14
Mean site qt trend Structured chalk
16
Pile penetration 16·6 m
18 20
Pile penetration 19·9 m
(a) (b)

qt: MPa qt: MPa


0 10 20 30 40 50 0 10 20 30 40 50
0 0
CPT-WK70a
Glacial till
5 Glacial till 5

CPT-WK43a Very high-density chalk


10 10
or Danian limestone
Depth: mbsb

Depth: mbsb

Adopted profile
15 15
Adopted profile
Mean qt trend
20 20
Structured chalk Structured chalk

25 25

Vs correlation Vs correlation

30 30
Pile penetration 30·7 m Pile penetration 31·0 m
(c) (d)

Fig. 3. Site profiles at test sites: (a) WK38; (b) WK42; (c) WK43; (d) WK70

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 661
Soil behaviour type, Ic su: kPa YSR
0 2 4 6 0 500 1000 1500 2000 2500 0 50 100 150 200 250
0 0 0

CPT_WK38
2 CPT-WK42a 2 2

4 4 4

6 6 6

8 8 8 YSR = (su/σ'vo)(1/0·85) ≤ 150

Depth: mbsb
10 10 10

12 12 12

14 14 14
Fluvioglacial till
Fluvioglacial till

16 16 16

18 18 18

20 20 20
(a)

Soil behaviour type, Ic su: kPa YSR


0 2 4 6 0 500 1000 1500 2000 2500 0 50 100 150 200 250
0 0 0

CPT-WK43a
CPT-WK70
2 2 2

4 4 4
Depth: mbsb

6 6 6

8 8 8

YSR = (su/σ'v0)(1/0·85) ≤ 150

10 10 10

Robertson (1990) Soil behaviour type index Ic


1·31 Gravelly sand to sand
12 1·31–2·05 Clean sand to silty sand 12 12
2·05–2·60 Silty sand to sandy silt
(b)
2·60–2·95 Clayey silt to silty clay
2·95–3·60 Clay to silty clay
3·60 Organic material

Fig. 4. Profiles of soil behaviour type index, Ic (Robertson, 1990), undrained shear strength, su, calculated from the CPT cone resistance assuming
clay behaviour and an Nkt factor of 22·5 and yield stress ratio at: (a) WK38 and WK42; (b) WK43 and WK70

Fully continuous chalk CPT profiles were available for applies over the 9·5–63 m depth range considered
only four test sites and some profiles terminated after shallow
penetrations into the chalk (Table 1). As noted earlier, qt ¼ 21Vs11 ð4Þ
correlations have been proposed for various geo-materials
between CPT qt and Gmax or Vs (see e.g. Baldi et al. (1989); Vs is in km/s and qt is in MPa. Equation (4), which may not
Mayne & Rix (1993) or McGann et al. (2015)). A local apply to different chalk sites, grades or densities, led to an
relationship of this type was established for the Wikinger average ratio for the calculated-to-measured qt values of 1·01
chalk. The qt–Vs data pairs presented in Fig. 5, which exclude and standard deviation of 0·16 at Wikinger. Although the qt
peaks associated with flints and the Danian limestone, profiles adopted were averaged following the approach
indicate that the near-linear relationship given by equation (4) described by Smith (2001), some scatter is inevitable due to

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
662 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER
30 driving (Doughty et al., 2018). Table 2 summarises the index
qt = 21Vs1·1 properties of the three main strata; three key additional
R2 = 0·57 n = 75 points are listed below.

(a) Figure 6 identifies the low-plasticity glacial till as


silty/clayey sand, with just over 10% clay on average,
20
confirming the interpretation made using Ic shown in
Fig. 4. The fluvioglacial layers manifest higher fines
fractions and 15–28% clay contents. Five out of eight
oedometer tests on glacial till samples gave
qt: MPa

permeabilities (under in situ stresses)  107 m2/s, while


the remaining three showed  106 m2/s. These
intermediate values fall below the range expected for
10 clean sands and yet significantly above those for
clay-dominated tills. The fluvioglacial till’s permeability
values were generally lower, with five tests indicating less
OSS1 WK18 than 107 m2/s and just one higher value.
WK07 WK19 (b) The chalk classifies as structured low–medium density
WK11 WK35
WK16 WK68 (intact dry density, IDD , 1·5 Mg/m3) grade A1/A2
(Bowden et al., 2002). Its unconfined compressive
0
0 0·2 0·4 0·6 0·8 1·0 1·2 1·4 strength (UCS) ranged from 0·2 to 0·8 MPa, falling
Vs: km/s below the 1·1–5 MPa range proposed for low–
(a) medium-density chalk by Matthews & Clayton (1993).
(c) Either high-density chalk or Danian limestone (with
30 IDD up to  1·9 Mg/m3) was encountered between
Parity
+/–20% approximately 7·5 m and 14 m depth at the WK70
location.

+/–20% The CAU triaxial tests presented in Appendix 1 illustrate


the strongly dilative behaviour manifested by till samples
when sheared undrained towards stable critical states,
20 mobilising large axial strains and high final su values that
qt measured: MPa

would signify high yield stress ratios (YSRs) in clay strata;


YSR often tends to infinity at shallow depths and ratios are
limited here to 150. Location-specific su profiles were
generated for axial capacity calculations (Fig. 4), treating
both tills as clays by applying an Nkt value of 22·5, which was
10 found by site-specific correlation between CPT, CAU and
unconsolidated undrained (UU) triaxial tests. The interpret-
ation led to remarkably high and non-uniform su values. The
resulting su/σv0 ′ ratios were employed as described in
Appendix 2 to generate illustrative YSR profiles, which are
also plotted in Fig. 4. Appendix 2 also presents the interface
ring shear tests performed on glacial (but unfortunately not
0
0 10 20 30
′ , of 26·5–28°. Stiff,
fluvio-glacial) till samples, which gave δult
qt calculated: MPa high-YSR, low-plasticity clay tills are generally insensitive
(b)
(Lehane, 1992; Long & Menkiti, 2007; or Ushev, 2018) and
sensitivity, St, was assumed as unity for the Wikinger tills.
Fig. 5. Correlation between cone resistances and shear wave velocity Buckley (2018) gives further details of the till’s compressi-
in chalk developed from the results at eight WK test locations: (a) cone bility, shear stiffness and critical state parameters.
resistance, qt, plotted against shear wave velocity, Vs (b) measured Turning to the chalk, Appendix 1 reports CAU tests that
plotted against calculated qt showed high stiffness up to the onset (at notably small
strains) of dilation followed by markedly brittle failure,
the significant qt peaks associated with locally denser or more confirming behaviour noted by, for example, Jardine et al.
cemented layers and/or the presence of flints, which may not (1985) and others. The peak shear strengths of low-density
be recognised in the shear wave measurements. Equation (4) samples were consistent with ϕ′ = 36° and c′ = 150 kPa, which
was used to estimate qt when Vs but not CPT data were may reflect cementing. A higher ϕ′ = 36·4° with c′  200 kPa
available. Mean qt trends from locations with similar ground applied to denser samples and these parameters are compa-
profiles were adopted when Vs data were also absent. tible with the lower end of the ranges for c′ of 100 kPa to
.2 MPa and ϕ′ of 36–42° quoted for intact chalk by Lord
et al. (2002). Doughty et al. (2018) showed that intact
Laboratory testing Wikinger chalk, whose natural water contents usually lie
Index, anisotropically consolidated undrained (CAU) close to their liquid limit, degrades readily to putty with
triaxial, oedometer and interface shear laboratory tests su  4 kPa under compaction at constant water content. Fall
were conducted by GEO, Gardline Geosciences Ltd (CIJV, cone tests showed mildly thixotropic behaviour after putti-
2013) and Fugro (2013). Supplementary testing at Imperial fication, with modest increases in su developing through
College examined the interface shear behaviour of the glacial ageing. Far greater increases in su and stiffness could be
till (Buckley, 2018) and the time-dependent behaviour of the obtained by consolidating the putty to higher effective
remoulded ‘putty’ chalk that forms around pile shafts during stresses. Undrained triaxial tests on such consolidated

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 663
Table 2. Index properties from the eight test sites during the pre-construction and production piling stage

Parameter Glacial till Fluvioglacial till Chalk

n Mean Std dev. n Mean Std dev. n Mean Std dev.

Water content, wc: % 173 12·8 3·8 6 13·4 2·1 149 28·4 4·7
Saturated water content, wsat: % — — — — — — 128 28·9 4·2
Liquid limit, wl: % 58 21·5 4·5 5 20·8 3·6 — — —
Plastic limit, wpl: % 58 12·7 1·7 5 12·6 1·3 — — —
Plasticity index, Ip: % 58 8·8 3·4 5 8·1 2·0 — — —
Bulk density, γbulk: Mg/m3 30 2·2 0·2 3 2·2 0·1 143 1·95 0·1
Dry density, γd: Mg/m3 30 1·9 0·2 — — — — — —
Specific gravity, Gs 46 2·69 0·01 4 2·69 0·01 23 2·70 0·02
Intact dry density, IDD: Mg/m3 — — — — — — 143 1·52 0·1
UCS, qu: MPa — — — — — — 55 0·39 0·24
Calcium carbonate: % — — — — — — 30 93·2 9·4

Std dev., standard deviation.

100

80
Percentage passing: %

60

40

Fluvioglacial till (WK42)


Glacial till (WK38, WK42, WK43, WK70)
20

Clay Silt Sand Gravel


0
0·001 0·01 0·1 1 10 100
Particle size: mm

Fig. 6. Particle size distributions in fluvioglacial and glacial till

puttified material indicated c′ = 0 and Mtc = 1·24 (ϕcv′  31°), tested statically, 12–15 weeks after driving, shortly before an
consistent with Clayton (1978) and Razoaki (2000). Interface instrumented re-strike on the adjacent twin pile. The ageing
ring-shear tests demonstrated δult′ angles of 32–34° in the periods were chosen to match the minimum durations
chalk (Fugro, 2013) similar to those reported by Le et al. anticipated between driving and turbine installation.
(2014) and Ziogos et al. (2017) at comparable normal Coupled cylindrical cavity expansion analyses (Randolph &
effective stress levels. Ring shear interface tests by Chan Wroth, 1979) indicate that pore-pressure dissipation rates
et al. (2019) gave comparable δult ′ values for chalk that after driving are governed by non-dimensional time factors
increased only moderately with normal stress, up to the T = tcv/(R*)2. It is difficult to ascribe field cv values
400 kPa maximum investigated. Ziogos et al. (2017) reported accurately for pile equalisation in clays (Lehane et al.,
marked reductions in δ under much higher normal stresses. A 2017). However, piezometers mounted on open-ended,
single value of 33° was adopted in the analyses that follow. 762 mm dia. piles (with R* = 0·17 m) driven in relatively
Doughty (2016) gives further details of the remoulded chalk’s low-plasticity, high-YSR, Lowestoft tills at Tilbrook Grange
compressibility, shear stiffness and critical state parameters. showed around 90% dissipation after 130 days (Clarke et al.,
1993), while 2 m dia. open steel piles (with R* = 0·27 m)
driven for the PISA (Pile–Soil Analysis) programme in
PROGRAMME OF STATIC AND DYNAMIC low-plasticity Cowden till showed more than 90% dissipation
PILE TESTS after 100 days (according to the project’s 2015 field test
Programme factual report, revision B). Noting that the 1·37 m dia.,
The two phases of pile testing at Wikinger are summarised 40 mm thick, Wikinger test piles have R* = 0·23 m and that
in Table 3, which also identifies the individual tests’ codes. the Wikinger tills’ permeabilities (and operational cv values)
The ‘pre-construction’ trial campaign involved six 1·37 m are significantly higher than those at Cowden or Tilbrook
dia. piles, driven in pairs at the WK38, WK43 and WK70 Grange, full dissipation was probably achieved before the
locations with dual sets of accelerometers and strain gauges static trial pile tests. Faster equalisation is expected in the far
attached near the pile heads. One pile from each pair was stiffer and more permeable chalk.

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
664 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER
Table 3. Summary of pile test codes and test histories during test campaigns

Test pile Test code* Pile age Test type

Days Min

WK38-1 WK38-1_DPT_EOD 0 0 Dynamic test end of driving


WK38-1_DPT_BOR 108 278 Dynamic test beginning of re-strike
WK38-2 WK38-2_DPT_EOD 0 0 Dynamic test end of driving
WK38-2_SLT_108 108 — Static test on aged pile
WK38-2_CYC_108 108 — Post static failure cyclic test
WK43-1 WK43-1_DPT_EOD 0 0 Dynamic test end of driving
WK43-1_DPT_BOR 78 335 Dynamic test beginning of re-strike
WK43-2 WK43-2_DPT_EOD 0 0 Dynamic test end of driving
WK43-2_SLT_78 78 — Static test on aged pile
WK70-1 WK70-1_DPT_EOD 0 0 Dynamic test end of driving
WK70-1_DPT_BOR 77 989 Dynamic test beginning of re-strike
WK70-2 WK70-2_DPT_EOD 0 0 Dynamic test end of driving
WK70-2_SLT_77 77 — Static test on aged pile
OSS-C2 OSS-C2_DPT_EOD 0 0 Dynamic test end of driving
OSS-C2_DPT_BOR 0 240 Dynamic test beginning of re-strike
OSS-C4 OSS-C4_DPT_EOD 0 0 Dynamic test end of driving
OSS-C4_DPT_BOR 0 130 Dynamic test beginning of re-strike
WK08-A WK08-A_DPT_EOD 0 0 Dynamic test end of driving
WK08-A_DPT_BOR1 0 14 Dynamic test beginning of re-strike
WK08-A_DPT_BOR2 0 376 Dynamic test beginning of restrike
WK08-A_DPT_BOR3 59 230 Dynamic test beginning of re-strike
WK11-A WK11-A_DPT_EOD 0 0 Dynamic test end of driving
WK11-A_DPT_BOR1 0 15 Dynamic test beginning of re-strike
WK11-A_DPT_BOR2 0 375 Dynamic test beginning of re-strike
WK13-A1 WK13-A_DPT_EOD 0 0 Dynamic test end of driving
WK13-A_DPT_BOR 0 360 Dynamic test beginning of re-strike
WK42-B WK42-B_DPT_EOD 0 0 Dynamic test end of driving
WK42-B _DPT_BOR 67 — Dynamic test beginning of re-strike

*The test code nomenclature gives the pile name (e.g. WK38-1), the type of test (DPT, dynamic pile test; SLT, static pile test; CYC, cyclic pile
test) and the time of the test (end of driving (EOD) or beginning of re-strike (BOR) for dynamic tests and number of days for the static tests).

The second phase of testing included instrumented instrumented pile and pile plug resistance tests to show that
dynamic driving and re-strike monitoring at five production internal shaft resistance reduces dramatically with increasing
locations on six (up to 3·7 m dia.) piles. pile internal diameter and is likely to be both relatively minor
and concentrated towards the bases of large-diameter piles,
offering only a modest contribution to the overall base
Pile and driving details resistance in sands, as has been confirmed by Han et al.
The trial pile driving employed a Menck MHU 800S (2019) through field tests. Analysis of driving records
hydraulic hammer. Paired test piles were set  8 m apart, as indicates that a similar system applied to the Wikinger
shown in Fig. 1, along with a third (un-instrumented) piles, which drove in a fully unplugged coring manner. The
reaction pile. Self-weight penetrations of 2·6 m were noted Impact signal matching code employed by the authors
at WK38, where the soft Holocene cover was relatively thick, involved explicit modelling of the internal shaft resistance
and less than 0·4 m at WK43 and WK70, where the cover (Randolph, 2008). Signal matches in which the ratio of
was thinner. The 2·7 m dia. WTG and 3·7 m dia. OSS internal-to-external shaft resistance was set between 0 and 0·2
production piles were driven with a heavier Menck MHU in both the tills and chalk led to the best fits for the cases
1200S. Penetrations of generally 15–25 mm per blow were considered initially. All shaft resistance was considered as
recorded in the till and 25–50 mm per blow in the chalk applied externally in the final set of analyses reported herein.
during the pre-construction pile installations. The production The pairs of long-term tension and re-strike tests con-
piles showed less variable (10–20 mm per blow) penetrations ducted at WK38, 42 and 70 were planned to allow checking
in both strata. All piles cored fully during installation, with of the static shaft capacities inferred from the dynamic signal
their plugs rising above the seabed. Five to 42 min driving matching analyses. Figs 7(a)–7(c) present profiles of the local
pauses occurred for a variety of operational reasons. (equivalent static) shaft resistance, τs,d, interpreted from EOD
matches on the six test piles. ‘Static’ EOD shaft resistance
appears negligible in the Holocene deposits and varies
Dynamic analysis of driving between 30 and 200 kPa in the glacial till, reducing system-
Back-analysis of the driving signals with Impact atically with h/R*; Fig. 8. The WK38 results (Fig. 7(a))
(Randolph, 2008) allowed assessments of the overall ‘equiv- indicate significant variations in the glacial till’s shaft
alent static’ capacities and shaft load distributions developed resistance profiles between identical piles driven at the
during penetration and at the end of driving (EOD), as well same location, possibly reflecting locally varying ground
as in re-strike tests. conditions, but also demonstrating the degree of variability
Annular piles displace much lower volumes of soil and associated with dynamic test interpretation. Still stronger
develop lower (base and shaft) resistances than closed-ended dependence of τs,d on h/R* was observed in the chalk, as
piles (e.g. Randolph, 2003; Xu et al., 2006; Gavin & Lehane, illustrated in Figs 7 and 8 with τs,d up to 300 kPa on shaft
2007). Chow (1997) and Jardine et al. (2005) employed sections close to the tip and minima around 10 kPa

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 665
τs,d: kPa τs,d: kPa τs,d: kPa
0 200 400 600 0 200 400 600 0 200 400 600
0 0 0
Holocene deposits
Glacial till
5 Glacial till 5

5 Very high-density chalk


10 10 or Danian limestone

Glacial till

15
Depth: mbsb

15

10

20 20
Structured chalk Structured chalk

25 25
15
Structured chalk

30 30

WK38-1_DPT_EOD WK43-1_DPT_EOD WK70-1_DPT_EOD


WK38-2_DPT_EOD WK43-2_DPT_EOD WK70-2_DPT_EOD
20 35 35
WK38-1_DPT_BOR WK43-1_DPT_BOR WK70-1_DPT_BOR
(a) (b) (c)

Fig. 7. Compressive shaft resistance back-analysed using Impact at end of driving and beginning of re-strike: (a) WK38; (b) WK43; (c) WK70

developing higher on the shaft as the tip advanced. The (2015a), could apply a maximum tensile load of 15 MN
chalk’s strong h/R* dependency reflects its markedly sensitive and cycle loads with periods of around one minute.
behaviour. Chalk breaks down readily to putty under Displacements were measured using a Norwegian
laboratory compaction (Doughty et al., 2018) and this Geotechnical Institute (NGI) subsea extensometer system
feature has led to piles ‘running’ or falling under their own connected to an independent reference frame. Pile failure was
self weight to considerable depths in chalk without any defined by either (a) the pile head displacements reaching
hammer blows being applied (Carotenuto et al., 2018). The 137 mm (D/10) or (b) the semi-logarithmic creep rate, kc,
high-density chalk (or Danian limestone) layer is reflected in approaching 4 mm/log cycle of time after 30 min. Four
the WK70 profile (Fig. 7(c)), where EOD τs,d values reached hydraulic actuators were built into a loading beam linked to
80 kPa at 10–15 m depth, far greater than those in the two adjacent reaction piles (Fig. 9). The load steps included
underlying low-to-medium-density chalk. an unload–reload loop, as shown in Fig. 10. The load steps
The EOD shaft capacities assessed from all 12 test piles are were governed by creep rate criteria scaled from the German
summarised in Table 4, along with τavg, the shaft capacities Geotechnical Society recommendations (EA-Pfähle (DGGT,
averaged along each pile’s length. Here too the τavg values 2014)) to reflect pile dimensions and were in keeping with the
reduce with increasing penetration: the lowest average EOD research methodologies applied by Chow (1997) and Buckley
unit shaft resistance (24 kPa) applied to OSS-C2, which had (2018). As discussed later, the frame design was based on
the greatest (36 m) penetration into chalk. Fig. 7 and Table 4 capacity predictions that underestimated the chalk’s long-
indicate the significant variations in EOD shaft resistances term shaft resistances.
between nominally identical piles, which range from ±6%
(for WK38) to ±16% (for WK43).
Glacial till dominated WK38
The re-strike tests indicated shaft capacities increasing
Dynamic behaviour of aged piles markedly with age. Pile WK38-1, driven to 16·6 m through
The dynamic re-strike tests on aged pre-construction piles, primarily glacial till soils, developed a compressive shaft BOR
outlined in Table 3, applied three full-energy blows with a capacity double that measured on the same pile 108 days
Menck MHU 800S hammer with the beginning of re-strike earlier at EOD. The parallel tension static load on WK38-2
(BOR) capacity defined at the first blow. Markedly higher showed the behaviour depicted in Fig. 11, where the static,
shaft capacities were found than at EOD, as listed in Table 4 compressive, shaft capacities interpreted from the dynamic
and in Table 5, especially at the chalk-dominated WK43 and EOD and BOR tests are also marked. The creep displace-
WK70 locations. The shaft shear stress distributions plotted ments observed during maintained load stages followed
in Fig. 7 add further information on how these gains built up semi-logarithmic trends with time, developing gradients, kc
over the pile shafts. (expressed as mm/log cycle of time) that increased system-
atically with average mobilised shaft resistance τavg once a
‘creep yield’ had been exceeded. Fig. 12(a) shows that for
LONG-TERM TESTING WK38, creep yielding took place at around 1/4 of the failure
Static testing procedures load, when τavg = 30 kPa, which is referred to as τcreep-yield. The
The seabed static tests were executed by a maintained load logarithmic plot on the right of Fig. 12(a) indicates that a
procedure with specially developed, remotely controlled power law relationship applied between kc and (τavg–τcreep-yield)
equipment. The system, described by Barbosa et al. for load steps that exceeded the creep yield criterion.

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
666 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER
τs,d: kPa τs,d: kPa
0 100 200 300 400 500 0 100 200 300 400
0 0
Holocene deposits Holocene deposits

Glacial till Glacial till

5 5
Depth: mbsb

Depth: mbsb
Effect of Effect of
increasing h/R* increasing h/R*

Blow 165

10 10

Blow 375 Blow 235

15 15
(a) (b)

τs,d: kPa τs,d: kPa


0 40 80 120 160 0 50 100 150 200 250
0 0
Glacial till Holocene deposits

5
Glacial till
5

10

Very high-density chalk Structured chalk


10
or Danian limestone
Depth: mbsb

15

20
15 Effect of
Effect of increasing h/R*
increasing h/R*
25 Blow 2500
Blow 589
20
Structured chalk 30
Blow 3065

Blow 757
25 35
(c) (d)

Fig. 8. Illustration of dependence of compressive shaft resistance on h/R*: (a) mid driving blows during installation of WK38-1; (b) mid driving
blows during installation of WK38-2; (c) mid driving blows during installation of WK70-2; (d) mid driving blows during installation of OSS-C4

Static tension failure was interpreted at a pile head load of shaft set-up factor of 1·65, assuming tension and compressive
9·33 MN (see Fig. 11) and 20·7 mm (or 1·5%D) displace- shaft resistances are equal. Signal matching analysis indi-
ment, with kc  3·5 mm/log cycle. Any reverse end-bearing cated that 69%, or 6·07 MN, of WK38-2’s shaft capacity was
capacity was considered negligible, as all piles were founded attributable to the glacial till.
into relatively free-draining chalk and the testing rates were Following full unloading, two ten-cycle packets of one-way
slow. The net load found after deducting the submerged axial cyclic loading were applied. The first imposed a
weight of the soil plug and pile is 8·80 MN, giving utilisation ratio (UR = maximum applied cyclic load/static
τavg = 122 kPa. failure load) of 0·62, and the second a UR of 0·84 (Fig. 13).
The static tension capacity proved for WK38 is around The permanent accumulated cyclic displacements, sacc,
10% lower than the compressive static shaft capacity shown on Fig. 14(a), remained below 0·05%D for the first
interpreted from signal matches made for the parallel ten cycles, but increased to 0·07%D/cycle in the second batch
re-strike test. Relating the tensile static capacity to the giving 1·1%D (or 15 mm) of pull-out and cycling halted as
compressive (dynamically measured) EOD for the same cyclic failure appeared imminent. The cyclic loading and
pile (WK38-2) indicates a combined (glacial till and chalk) unloading stiffnesses (kl and kul ) are shown on Fig. 14(b),

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 667
Table 4. Pile shaft resistances at the end of driving and beginning of re-strike for all 12 tests

Test pile End of driving* Beginning of re-strike†

τavg‡: kPa τtill: kPa τchlk: kPa τavg‡: kPa τtill: kPa τchlk: kPa

WK38-1 67 59 88 136 103 284


WK38-2 75 66 98 — — —
WK43-1 26 19 30 139 61 176
WK43-2 36 21 38 — — —
WK70-1 40 16 48 207 39 254
WK70-2 35 14 40 — — —
OSS-C2 24 27 17 78 39 89
OSS-C4 29 45 21 68 48 73
WK08-A 60 36 123 93 42 225
115 59 247
141 76 315
WK11-A 30 22 42 65 22 122
105 19 221
WK13-A 67 48 53 103 79 141
WK42-B 86 73 — 175 152 —

*Final blow at the end of driving.


†First full-energy blow during a re-strike test.
‡Average shaft resistance in the glacial till and chalk.

Table 5. Summary of pile capacities from dynamic and static pre-construction pile tests and proportions in glacial till and chalk

Site Overall shaft capacity, Qs: MN

BOR* SLT (net)† BOR/static % Glacial till % Chalk

WK38 9·72 8·80 1·10 69 31


WK43 18·30 20·90‡ 0·88 14 86
WK70 27·69 22·44‡ 1·23 4 96

*From Impact analysis of re-strike data; BOR, beginning of re-strike.


†SLT (net), static capacity corrected for soil/chalk and steel weight.
‡Based on extrapolation of creep rates to kc = 3 mm/log cycle.

normalised by the values developed at N = 1. While little The power law relationship interpreted between kc and
stiffness change was observed under the first batch of cycles, (τavg–τcreep-yield) in the WK38 static test (which reached
the second ten cycles led to stiffness reductions of 25–35%. full failure) in Fig. 12(a) suggests another approach in
Comparably high levels of one-way repetitive tension loading which failure can be projected as the point at which the
were also able to induce failure with steel piles driven in stiff power law kc trend reaches a specified logarithmic creep rate
Cowden glacial sandy clay till (Ove Arup and Partners limit. This is taken conservatively here as 3 mm/log cycle
(Arup, 1986)). (an increase on the EA-Pfähle (DGGT, 2014) recommen-
dation of 2 mm/log cycle, which reflects the pile dimensions)
and then applied in Figs 12(b) and 12(c). The results from all
Chalk-dominated WK43 and WK70 four extrapolation methods are compared in Table 6. The
The re-strike tests conducted at the chalk-dominated parabolic and hyperbolic methods led to predictions 14
WK43 and WK70 locations showed a more marked ‘static’ to 59% higher than the static result for WK38, where full
shaft capacity set-up. Factors between 5·3 and 5·2 are evident failure was achieved, while the power law kc–(τavg–τcreep-yield)
from Tables 4 and 5 and Fig. 7. Although variations in base extrapolation to 3 mm/log cycle matched the static failure
capacity are more difficult to ascertain, signal matching load to within 5%. Adopting the latter, locally calibrated,
indicated comparatively modest (, 20%) changes in base method indicated net tensile static shaft capacities of
capacity over time. Signal matching for WK43-2 and 20·9 MN (or τavg = 158 kPa) at WK43 and 22·44 MN
WK70-2 indicated that 86% and 96% of the overall shaft (τavg = 168 kPa) at WK70, representing overall set-up
capacities developed within the respective piles’ chalk factors (for glacial till and chalk) of 4·4 and 4·9 compared
sections (see Table 5). to the EOD values. The (compressive) re-strike BOR shaft
Static load tests on the  31 m long WK43-2 and WK70-2 capacities measured at WK43 and WK70 were 13% lower
piles gave the load–displacement outcomes plotted in Figs 15 and 23% higher than the respective extrapolated net tensile
and 16 and creep responses illustrated in Figs 12(b) failure loads. One of the more conservative alternative
and 12(c). Both piles manifested clear creep yielding in extrapolation procedures may have been more applicable to
their maintained load stages. However, neither achieved the WK70 case, although this remains unproven. While the
ultimate failure before the allowable structural limit of dynamic shaft capacities all fall within 23% of the interpreted
the test beam was reached. Methods for extrapolating static values, the estimates for WK43 and WK70 must be
incomplete pile load tests include the hyperbolic and treated with caution, as the good agreement proven at the
parabolic load–displacement curve fitting methods of WK38 location is less certain at the chalk-dominated
Brinch Hansen (1963), Chin (1970) and Decourt (1999). profiles.

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
668 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER

Guide 1800/1720
Yoke ∅1700
A

12 × threaded rods 9·0 m long D = 63·5 mm quality S555/700


As = 3167 mm2; breaking load 2217 kN

Yoke and guide


Load cell

Hydraulic cylinder 4 × 5000 kN/1000 mm

Lifting lug 500 kN 1000


2080

Stroke 1000
1300

Spreader beam two-part


Pressure plate

Positioning tool
1680
680

Inside 1291·6
Locking tool ∅1170 mm
5000 above seabed

Penetration 28·0 m to 32·5 m

1371·6

Pressure pile Test pile


Test-pile D1371·6×40×36·5 m

8330 8330

Fig. 9. Schematic diagram of loading frame and reaction pile arrangement for subsea static load tests at Wikinger; after Barbosa et al. (2015a).
Dimensions in mm, unless stated otherwise

Adding the WK38 result, the BOR ‘static compressive the shaft stresses interpreted from the blows applied
shaft capacities’ are on average 5% higher than the tension immediately prior to, and following after, a range of such
values assessed from the independent static tests at each driving pauses. The combined spread of set-up factors are
location. Given the greater variability and subjectivity plotted against time in Figs 17 and 18 for the till and chalk,
associated with dynamic test interpretation, this degree of respectively.
correspondence is encouraging: even nominally identical Although pore pressure dissipation was likely to have
paired piles can show significantly different EOD capacities, completed in the tills before the static testing was carried out
as shown in Table 4. (15 weeks after driving), shaft capacities may have been
Table 7 presents an overall summary of the interpreted continuing to grow through other ageing processes at that
static test results, which are also annotated on Figs 11, 15 and stage. The set-up ratios, Λ, of 1·6 to 3·2 achieved up to static
16. Signal matching for the WK43-2 and WK70-2 BOR tests testing appeared to follow the approximately semi-
allowed the respective contributions of the static tension logarithmic trend given by equation (5) and plotted on
capacities to be estimated as shown in Table 5 where the Fig. 17
chalk layers are indicated as providing 18·0 MN and  
21·5 MN of the tension shaft capacities interpreted from τ avg ðtÞ t
Λ¼ ¼ 1 þ 04 log ð5Þ
the WK43 and WK70 pile tests. τ avg ðt ¼ tref Þ tref
where τavg(t) is the resistance at time, t, after driving and tref
is an initial reference time which is taken as 0·01 days.
VARIATIONS IN SHAFT RESISTANCE WITH TIME Onshore pile tests in low-plasticity clays and sands by
The shaft resistances interpreted from the BOR blows, as Karlsrud et al. (2014) and Jardine et al. (2006) indicated
listed in Table 4, can be compared to the final EOD blows to ultimately stable Λ values  3 once six to 12 months had
assess changes over time and Tables 5–7 confirm marked elapsed after driving.
set-up over the 12 to 15 week post-driving ageing periods. The equivalent chalk set-up trends presented in Fig. 18
Operational pauses during pile driving provide information appear to vary with the degree by which the pile shafts
on re-strike trends over shorter periods and Table 8 compares penetrated into the chalk Lchalk
p . Set-up was more marked in

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 669
16 kc < 0·5 mm/log cycle kc < 4·0 mm/log cycle
17 Load test limit = 15 MN
16
15
14 Load Holding
13 step period (≥2 min)
1–2 30
12 3 60
12
4 10

Load at pile head, Q: MN


5 30
11 6 10
7–16 30
18 17 60
10
18–20 10
8 21 60
9

4
3 7 19 Rt,c = 2·57 MN
2 4 6
5 20 0·4Rt,c = 1·03 MN
1
21
0
0 200 400 600 800
Elapsed time, t: min

Fig. 10. Static test loading procedure for offshore tests at Wikinger for loading up to 15 MN (where Rt,c is the characteristic design load)

Bore capacity (9·72 MN) achieved in the chalk at BOR are significantly higher
10 000
than those mobilised in the till. Normalising by the
relatively low values mobilised during driving leads to higher
8000 long-term set-up factors in the chalk layers than in the tills.
Scaling up the pore pressure dissipation times discussed
earlier for 43·8 mm dia. piezocones by the ratio (R*/Rcpt)2,
Tension load: kN

6000 where Dcpt = 2Rcpt, indicates times for 50% pore-water


EOD capacity (5·3 MN)
pressure dissipation after installation for 1·37 m to 3·7 m
4000 dia. piles of between 11 and 40 min in chalk. Dissipation
may therefore be important to the initial part of the set-up
trend seen in Fig. 18. However, Buckley et al. (2018a)
2000 hypothesised that an arching mechanism, involving drained
creep relaxing raised circumferential stresses close to the
0
shaft, may also contribute to longer term shaft capacity
0 10 20 30 growth in chalk, as has been postulated previously in sands
Pile head displacement: mm (e.g. Chow et al., 1998; White et al., 2005; Jardine et al.,
2006). Chemical reactions at the interface, internal
Fig. 11. Load–displacement behaviour during static axial tension test re-cementing of the chalk putty and bonding to the interface
WK38-2_SLT_108 may also play a role.
The tests by Ciavaglia et al. (2017) and Buckley et al.
(2018a) on smaller pipe-piles driven in chalk at St
cases where more than 20% of the pile was founded in chalk, Nicholas-at-Wade indicated similarly shaped set-up curves.
reflecting greater scope for capacity growth in the high h/R* However, the smaller onshore piles manifested notably slower
locations where ‘friction fatigue’ losses were most marked set-up rates. The Wikinger fractures are closed or tight,
during driving. Shaft resistances doubled within 10 min and water-filled and spaced at 200–600 mm (indicating grade
re-doubled within 30 min, confirming observations made by A1/A2 chalk), while the St Nicholas-at-Wade fractures are
Dührkop et al. (2017). The set-up tended to stable asymp- open to ,3 mm, air filled and spaced at 60–200 mm. The
totes after around 75 days and equation (6), a hyperbolic arching mechanism may allow more rapid gains in chalk
relationship proposed by Tan et al. (2004), appears appro- masses with tight, widely spaced fractures than when
priate, as shown in Fig. 18 fractures are more open and closely spaced. It is also possible
τ avg ðtÞ that the different groundwater conditions and pile configur-
Λ¼ ations contributed to the onshore piles’ slower set-up trends.
τ avg ðt ¼ tref Þ However, the observation that closed-ended ICPs installed by
  
t=T50 slow cyclic jacking at St Nicholas-at-Wade did not show any
¼ Λult m þ ð1  mÞ ð6Þ shaft capacity increases over extended periods indicates that
1 þ ðt=T50 Þ
set-up is strongly dependent on the installation process; see
Here Λult is the ultimate set-up factor; T50 is the time Buckley et al. (2018b).
required to reach 50% of Λult; and m is a factor applied to
improve the early (t , 1 day) age fitting. The Lchalk
p /Lp . 0·2
case curve corresponds to T50 = 29 min, Λult = 5·6 and SHORT-TERM STATIC SHAFT CAPACITY
m = 0·18, while that for lower Lchalk p /Lp ratios employs PREDICTIONS
Λult = 2·95, T50 = 100 min and m = 0·26. The dashed line The end of driving resistances interpreted along the shaft
shown on Fig. 18 corresponds to an average trend between sections installed in chalk exceed significantly the τf value of
the two curves. Table 4 indicates that the shaft resistances 20 kPa recommended for static design in Ciria C574, giving

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
670 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER

Failure τavg = 122 kPa


5 10

Creep rate, kc: mm/log cycle

Creep rate, kc: mm/log cycle


4 kc = 3 mm/log cycle

kc = 3 mm/log cycle 1

Failure τavg = 122 kPa


3

2
0·1

1 kc = 9·02 × 10–6 τavg


2·58
τcreep-yield = 30 kPa

WK38-2 SLT
0 0·01
0 40 80 120 160 200 1 10 100 1000
τavg: kPa (a) τavg – τcreep-yield: kPa

τavg = 158 kPa


5 10

kc = 3 mm/log cycle
Creep rate, kc: mm/log cycle

Creep rate, kc: mm/log cycle


4

τavg = 158 kPa


kc = 3 mm/log cycle
1
3

2
0·1

1 τcreep-yield = 65 kPa kc = 2·51 × 10–4 τavg


2·07

WK43-2 SLT
0 0·01
0 40 80 120 160 200 1 10 100 1000
τavg: kPa
(b)
τavg – τcreep-yield: kPa

10
5
τavg = 168 kPa

kc = 3 mm/log cycle
Creep rate, kc: mm/log cycle
Creep rate, kc: mm/log cycle

τavg = 168 kPa


4
1
kc = 3 mm/log cycle
3

2
0·1

1 kc = 2·51 × 10–4 τavg


1·927
τcreep-yield = 37 kPa
WK70-2 SLT
0·01
0
0 40 80 120 160 200 1 10 100 1000
τavg: kPa τavg – τcreep-yield: kPa
(c)

Fig. 12. Extrapolation of creep rates to failure: (a) WK38-2; (b) WK43-2; (c) WK70-2. τavg–τcreep-yield is limit below which the creep rates were
negligible

calculated/dynamically measured Qc/Qm capacity ratios of


0·64 ± 0·21. The set-up trends discussed above led to far
Average shaft resistance, τavg: kPa

160 greater divergence in the long-term tests from Ciria C574,


WK38-2 CY1
UR = 0·62
WK38-2 CY1 and lower Qc/Qm ratios. These findings prompted the
120 ≅0·01 Hz
UR = 0·84
≅0·01 Hz development of an effective stress-based ‘Chalk ICP-18’
predictive procedure, which drew on the additional onshore
tests on 139 mm to 762 mm dia. piles, described by Buckley
80
et al. (2018a) and Ciavaglia et al. (2017), respectively.
Reference was also made to offshore observations for
40 monopile driving in chalk, as described by Buckley (2018)
and Jardine et al. (2018).
0
The chalk ICP-18 expressions for driving resistance in
0 400 800 1200 1600 2000 low–medium-density chalk follow the generic ICP-05
Elapsed time, t: s (Jardine et al., 2005) approach and, following Buckley
et al. (2018a): (a) rely on CPT tests to account for variations
Fig. 13. Average shaft resistance mobilised during WK38- in chalk properties; (b) capture the marked tendency of radial
2_CYC_108 plotted against time effective and shear stresses to reduce with h/R*; and (c) match

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 671
1·2 1·2
WK38-2 CY1 WK38-2 CY2 WK38-2 CY1 WK38-2 CY2
UR = 0·62 UR = 0·84 UR = 0·62 UR = 0·84

0·8 1·0

kl/kl(N=1) or kul/kul(N=1)
sacc/D: %

0·4 0·8

Loading stiffness
Unloading stiffness
0 0·6
0 5 10 15 20 0 5 10 15 20
Cycles, N Cycles, N
(a) (b)

Fig. 14. WK38-2_CYC_108: (a) permanent cyclic displacements normalised by pile diameter; (b) cyclic loading and unloading stiffness

25 000 30 000
BOR capacity (27·69 MN)

Extrapolated static failure (net 20·90 MN)


20 000
BOR capacity (18·30 MN) Extrapolated static failure (net 22·44 MN)

20 000
Tension load: kN
Tension load: kN

15 000

10 000
10 000

EOD capacity (4·78 MN)


5000 EOD capacity (4·61 MN)

0 0
0 10 20 30 0 10 20 30
Pile head displacement: mm Pile head displacement: mm

Fig. 15. Load–displacement behaviour and extrapolated failure load Fig. 16. Load–displacement behaviour and extrapolated failure load
during static axial tension test on WK43-2 during static axial tension test on WK70-2

the interface-shear failure characteristics observed in instru- the pile tip during continuous driving and then later as the
mented ICP field tests. pile tip advances to greater relative depths, h. Lehane et al.
Observations involving a wide range of pile diameters and (2005) addressed this for sand by separating out an assumed
wall thicknesses, tw, indicated that their D/tw ratios affected initial influence of the piles’ effective areas from the
driving resistance, probably because the puttified chalk subsequent ‘friction’ fatigue by employing independent
annuli’s widths depend primarily on tw. The outer shaft parameters for each component, while Lehane et al. (2013)
resistance to driving of open-ended piles (with retained only the h/R* term when dealing with clays. Further
17 , D/tw , 67) can be matched in low–medium-density disaggregation or other modification of the chalk expressions
chalk by substituting the radial effective stresses given by may be possible as additional data become available on
equation (7) into the local Coulomb shaft failure criterion installation resistances. Applying the preliminary expression
given by equation (2) (taking δ′=33° and σrf′ = σri′ ) given by equation (7) to the nine EOD cases (shown in
 0481ðD=tw Þ0145 Table 4) for which CPT data were available in the chalk led to
h ð7Þ an average Qc/Qm ratio of 0·9 and a standard deviation of
σ′ri ¼ 0031qt  For h=R  6 18%. Fig. 19 shows examples of the predicted profiles of shaft
R
stress compared with those interpreted from signal matching.
Equation (7) aims to capture the radial stress reductions The back-analysis of 70 blows reported in Appendix 3
that occur at any given depth as the chalk flows first around proceeded by assuming that the piles’ base resistances

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
672 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER
Table 6. Extrapolation of shaft loads to ultimate conditions using available methods

Site Extrapolated overall net shaft load, Qs: MN Measured or adopted value:
MN
Chin (1970, Decourt Brinch Hansen Creep rate (3 mm/log
1971) (1999) (1963) cycle)

WK38 10·05 14·00 10·71 8·40 8·80*


WK43 21·74 21·98 17·25 20·90 20·90
WK70 17·10 18·14 16·70 22·44 22·44

*Measured value for full failure test.

Table 7. Static tension failure loads and pile capacity predictions for offshore pile tests at Wikinger

WK38-2 WK43-2 WK70-2

Time after driving: days 108 78 77


Tensile failure load, Qt: MN 9·33 21·77* 23·31*
Pile and chalk self-weight: MN 0·53 0·87 0·87
Average tensile shaft resistance, τavg: kPa 122 158* 168*
Compressive EOD shaft load on pile 2: MN 5·34 4·78 4·61
Compressive BOR shaft load on pile 1: MN 9·72 18·30 27·69
Global set-up factor, Λ (static/EOD) 1·65 4·37 4·86

*Based on extrapolation of creep rates to kc = 3 mm/log cycle.

Table 8. Summary of operational pauses in pile driving and interpreted shaft stresses before and following a pause

Test pile Duration: min Beginning of pause End of pause

τtill: kPa τchlk: kPa τtill: kPa τchlk: kPa

OSS-C2 10 26·4 41·8 26 111


OSS-C4 10 28·2 41·1 29 94
WK08-A 7 93·9 — 102 —
WK38-1 42 24·9 — 27 —
29 87·2 — 83 —
WK38-2 43 34·9 — 26 —
11 74·2 — 72 —
32 88·7 — 78 —
WK42 11 24·5 — 24 —
WK43-1 16 13·4 99·5 18 240
6 2·3 51·4 2 106
WK43-2 9 35·3 78·5 41 215
WK70-1 8 26·0 95·3 30 194
WK70-2 9 27·5 54·0 29 118
WK70-2 25 16·4 31·1 20 120

Note: 1, final blow before pause begins; 2, first blow following pause.

developed over their solid tip areas only during driving and applying 100 or more days after driving. As with the
mobilised average annular bearing pressures qba that could be short-term driving case, it is assumed that the Coulomb
related directly to the average local CPT resistance. The latter law applies at the interface (equation (2)) and that the δ′ angle
was characterised as qt, 1·5D, the mean qt averaged 1·5 pile can be predicted from appropriate interface shear tests. The
outside diameters (D) above and below the tip. The resulting static unit shaft shear capacities, τf, which increase signifi-
best-fit qba/qt, 1·5D ratios varied with tip displacement per cantly over time, are calculated from expressions that capture
blow, but indicated a range of 0·16 , qba/qt, 1·5D , 0·8 in both the chalks’ constrained interface dilation, which resembles
tills and chalk, with a mean around 0·50 and standard that seen in sands. While it would be attractive to link the
deviation  0·15. No static compression test data are short-term and long-term stresses through a simple set-up
available to assess whether higher ratios might apply in factor expression, no evidence was found that D/tw affected
monotonic loading tests that are taken to reach failure after the long-term resistances. It appeared that the long-term
displacements of D/10, as has been argued for sands by, for shaft resistances could be captured with a simpler indepen-
example, Byrne et al. (2012). dent expression. Under tension or compression loading,
σrf′ values applied in equation (2) were given as
σ′rf ¼ ðσ′rc þ Δσ′rd Þ ð8Þ
LONG-TERM STATIC SHAFT CAPACITY
PREDICTIONS  052
h
Buckley (2018) and Jardine et al. (2018) also set out σ′rc ¼ 0081qt ð9Þ
effective stress-based ‘Chalk ICP-18’ predictive expressions R
for the ‘long-term’ shaft capacities, which are taken as For h/R*  6

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 673
4 The relative shaft capacity contributions identified from
Re-strike tests
Following driving pauses signal matching for the till and chalk layers to the WK38, 43
and 70 test piles’ shaft capacities allow shaft capacity
3 prediction methods to be assessed for the chalk and glacial
Equation (5) till layers in calculations that are insensitive to the assump-
Set-up factor: Λ

tions made in estimating the minor contributions of the


2
Holocene material, which contributed , 0·5% of the long-
term total capacity at WK38, for example. The soil
parameter profiles presented earlier (and in the
Appendices) were applied and the ICP-05 sand approach
1
Range of was employed to calculate the minor contribution of any
values Holocene cover.
Table 9 summarises the outcomes for glacial till shaft
0 capacity contributions interpreted from the WK38, 43 and 70
1·44 min 14·4 min 2·4 h 1 day 10 days 100 days 2·74 years
tests, as listed in Table 7. Also shown are the predictions
Time since last blow made by applying the ICP-05 sand procedures, which aim to
match medium-term (nominally 10 day age) capacities and
Fig. 17. Set-up trends in glacial till from dynamic re-strike tests and
the API (2014), Fugro-96, NGI-05, ICP-05 clay and
from blows following driving pauses with semi-logarithmic capacity
increase trend without apparent asymptotic value over monitoring UWA-13 clay methods. The ICP sand method under-predicts
period the long-term glacial till resistances by significant margins.
However, the piles were tested around 100 days after driving.
Open-steel piles tested at comparable ages at Dunkirk
7
(Jardine et al., 2006) indicated a set-up factor of  1·9
Following driving pauses between their 10 day and 100 day capacities which, if applied
Re-strike tests
6 Trend for 66% to the ICP-05 sand calculations, raises the predicted-to-
high h/R*
measured shaft capacity ratios to 0·74 , Qc/Qm , 1·37 and
Labels show percentage of pile
penetration in chalk 47% 78%

5 gives an average of 0·96. Adopting any of the cited clay


Set-up factor: Λ

Average trend methods leads to large over-predictions for the resistances


4 available in glacial till units 12–15 weeks after driving,
73% 78% Trend for 18% with 1·5 (for API, 2014) , Qc/Qm , 8·8 (for UWA13).
low h/R*
Surprisingly low average α values ( 0·2 or less) would be
47%
3 27%
17%
71%
59%45% 26%14%
17% needed to obtain matches with the field capacities. Despite
2 17%
the incomplete drainage seen in the piezocone profiling, the
glacial tills appear to have responded to pile driving and
1
testing more like sands than clays, which is consistent with
0
the classification illustrated on Figs 4 and 6.
1·44 min 14·4 min 2·4 h 1 day 10 days 100 days 1000 days No fluvioglacial till was present at the static test locations.
Time since last blow However, larger diameter production piles were driven
through both the glacial and fluvioglacial till sections
Fig. 18. Set-up factors in chalk from dynamic re-strike tests and from considered in Figs 3–6. A long-term re-strike was conducted
blows following driving pauses with interpreted trends for high and low on one WK42 pile 67 days after driving, which provides
h/R* values useful additional evidence. As at the static test locations,
signal matching analysis indicated that the sandy glacial till
section’s contribution exceeded the (nominally 10 day age)
ICP-05 sand method prediction, giving Qc/Qm  0·8. The
It was considered that, as in sands (see Lehane et al., ratio could be expected to fall as ageing continues. The clay
1993; Chow, 1997), the dilatant Δσrd′ component can be methods appeared non-conservative again, giving Qc/Qm
estimated, as between 2·1 (for ICP-05 clay) and 8·6 (UWA-13), in keeping
4GΔr with the glacial tills’ more sand-like behaviour. Applying clay
Δσ′rd ¼ ð10Þ calculation methods to the lower permeability and higher
D fines content fluvioglacial section of WK42 indicated Qc/Qm
where Δr is the radial dilation at the interface required to ratios ranging from 0·5 (for NGI-05 and API (2014)) to 1·2
permit failure and the shear modulus, G, should ideally be (UWA-13) with an average of 0·75; the ICP-05 clay method
measured in the Ghh mode and account for any non-linear could not be applied as no interface shear tests were available
dependence on Δr/D. Buckley et al. (2018b) interpreted a for this unit. However, 10 day age predictions with ICP-05
range of radial dilation of 0·23 to 2·04 μm from instrumented sand gave Qc/Qm  0·1, confirming that the fluvioglacial till’s
ICP tests in chalk and a value of Δr  0·5 μm is rec- behaviour was more clay like than sand dominated. Caution
ommended in the preliminary chalk ICP-18 method, which and field checking are clearly required when designing piles
falls far below the peak-to-trough pile shaft roughness in tills with intermediate sand-to-clay behaviours, such as
measure that is applied in sands, because of the chalk’s far those encountered at Wikinger.
smaller grain sizes. Constrained dilation has a significant The equivalent analyses for the piles’ chalk layers had to
effect with small piles, but is predicted to contribute less account for the 6·5 m thick very high-density chalk, or
(, 5%) to large-diameter offshore piles. Danian limestone, layer encountered at WK70. The Ciria
While the scalar coefficient (0·081) included in equation C574 calculation adopted a ‘high-density’ 120 kPa shaft
(9) is notably higher than that expected in sands, the ‘friction resistance, while qt = 50 MPa was assumed for the chalk
fatigue’ exponent of 0·52 in equation (9) suggests a similar ICP-18 assessment of WK70’s ‘limestone’ section, matching
rate of local stress degradation, and therefore shape of shear the operational maximum to which the deployed field CPT
stress profile, to that anticipated for silica sands by the equipment operated. Higher qt values may have been
ICP-05 and UWA-13 methods. observed if higher capacity cones had been available.

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
674 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER
τs,d or τf: kPa τs,d or τf: kPa
0 50 100 150 0 40 80 120 160 200
10
18
OSS-C2 WK11-A
OSS-C4 Prediction equation (7)
20 Prediction equation (7)
Depth: m 21

Depth: m
30
24

40 27

50 (a) 30 (b)

τs,d or τf: kPa τs,d or τf: kPa


0 40 80 120 160 200 0 40 80 120 160 200
10 10

WK43-1
WK43-2 High-density
15 Prediction equation (7) 15 chalk

WK70-1
Depth: m

Depth: m

WK70-2
20 20 Prediction equation (7)

25 25

30 30

(c) (d)

Fig. 19. Examples of shear stress profiles at EOD interpreted from signal matching; profiles predicted using equations (2) and (7): (a) OSS,
(b) WK11-A, (c) WK43, (d) WK70

Table 9. Summary of static pile capacity predictions in the glacial till

Site Measured* Shaft load Qs in glacial till layer: MN

ICP-05 sand† API (2014)† Fugro-96† NGI-05† UWA-13†,‡ ICP-05 clay†


§ § § §
WK38 6·07 2·29 (0·38) 9·21 (1·52) 11·62 (1·91) 9·52 (1·57) 19·53 (3·22) 12·39 (2·04)§
WK43 2·93 1·19 (0·41) 7·48 (2·56) 6·37 (2·18) 7·52 (2·57) 13·95 (4·77) 6·04 (2·07)
WK70 0·90 0·64 (0·72) 4·49 (5·00) 3·41 (3·80) 4·45 (4·94) 7·89 (8·79) 3·85 (4·29)

*Proportion based on static values given in Table 5.


†Values in parentheses are ratios of calculated to measured capacity.
‡Uses UWA-13 method with inclusion of angle of interface friction (equation (8)).
§
Adopts ICP-05 calculation in top 2 m of Holocene sand.

The outcomes summarised in Table 10 show Ciria C574 to regarded as a preliminary proposal that may well require
be markedly over-conservative at all three test sites, with updating as new findings emerge.
Qc/Qm outcomes between 5 and 10. In contrast, the chalk
ICP-18 expressions matched the test capacities to within
+/20%. Further independent checking is required as the SUMMARY AND CONCLUSIONS
latter approach was developed to fit the field behaviour High costs and logistical difficulties have led to offshore
observed at Wikinger and St Nicholas-at-Wade. It is static pile load testing and/or long-term dynamic re-strike
encouraging that Buckley (2018) and Jardine et al. (2018) testing being extremely rare. However, the Wikinger case
found fair capacity matches for other sites where the history shows that field testing can be highly cost-effective,
necessary CPT profiles and test records are available, provided it is conducted in advance of final design. Barbosa
including closed-ended piles installed by impact driving. et al. (2017) describe how major project risks were eliminated
Jardine et al. (2019) describe research that is underway to and total pile lengths reduced by 3 km for Wikinger, saving
develop the method further. However, chalk ICP-18 must be 8000 t of steel and 16 000 t of carbon dioxide emissions.

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 675
Table 10. Summary of static pile capacity predictions in chalk for the test piles’ ages. However, they fell far below
predictions made with five offshore design methods
Long-term (.100 day) shaft load Qs in chalk layer: MN for clays. In contrast, re-strike test analysis involving
the higher clay content and lower permeability
Site Measured* Ciria C574† Chalk ICP-18†,‡ fluvioglacial layers indicated both better
correspondence with clay method predictions and
WK38 2·73 0·27 (0·10) 3·31 (1·21) resistances far higher than expected by the ICP-05
WK43 17·97 1·76 (0·10) 14·23 (0·79)
WK70 21·54 4·83§ (0·22) 21·67∥ (1·01)
sand method.
(g) The Ciria C574 shaft capacity method significantly
*Proportion based on static values given in Table 5. under-predicted the driving resistance experienced in
†Calculation assumes Δr = 0·5 μm with G calculated from equation the chalk and greatly underestimated, by factors of 5 to
(3). 10, the 100 day age shaft capacities.
‡Values in parentheses are ratios of calculated-to-measured capacity. (h) The chalk ICP-18 expressions for driving resistance led
§
Adopts 120 kPa in high-density chalk/Danian limestone layer. to good representations of overall field capacities at the

Adopts qt = 50 MPa in the high-density chalk/Danian limestone end of installation and captured the chalk’s marked
layer. h/R* trends. The long-term shaft capacity expressions
reflected equally well the field capacities observed at
ages exceeding 100 days.
Large sums were recouped from supply, fabrication and (i) The results have important economic consequences for
installation costs. Highly significant benefits may be taken projects such as large offshore windfarms. Full-scale
from considering ageing trends in both tills and chalk, static testing was shown to be feasible offshore and
provided the piles can be driven well before they have to carry highly cost-effective to the Wikinger project. Highly
their design loads. Still greater structural savings could have significant benefits may be obtained by recognising the
been achieved at Wikinger if the testing had been conducted favourable effects of pile ageing, updating design
at an even earlier stage. procedures and supporting engineering assessments
Parallel industrial–academic research, including pile through careful field checking.
experiments at an onshore site, laboratory testing and
analysis aided the interpretation of the novel Wikinger field
tests and helped to frame ten main conclusions regarding MD: 36·4°
piles driven in dense/stiff low-plasticity tills and low– 1500
LD: 36·0°
medium-density chalk.

(a) Static testing leads to more reliable measurements of


pile capacity than the more complex and less objective
process of analysing instrumented hammer blows 1000
dynamically. However, careful signal matching of pile
monitoring data and analysis of static tension tests
q: kPa

conducted at Wikinger led to broadly compatible field


measurements.
(b) The piles’ shaft resistances apply principally on their 500
outside areas during driving and re-striking. The local
shaft shear stresses reduce markedly with increasing c' = 200 kPa
c' = 150 kPa Site Depth Material Kc ef
relative tip penetration (h/R*) in tills and, still more WK11 22·3 m Structured LD 0·4 0·98
WK13 19·5 m Structureless 0·4 0·76
sharply, in chalk. WK13 45·6 m Structured MD 0·4 0·89
(c) Shaft capacity set-up progressed gradually in the tills, WK38 38·9 m Structured LD 0·4 0·94
0 WK55 46·3 m Structured MD 0·4 1·65
leading after 15 weeks to resistances around 2·4
(+/0·8) times those available at the end of driving. 0 500 1000 1500
(d ) More significant shaft increases applied in chalk, with p': kPa
average ultimate set-up factors greater than 5·5 that (a)
followed a hyperbolic trend with time and depended on
h/R*. Smaller factors applied to piles with relatively low 1600
penetrations into the chalk.
(e) Independent dynamic and full static tests to failure
conducted on one pair of offshore piles showed good 1200
agreement regarding overall shaft capacity. Signal
matching analyses also allowed robust estimates to be
q: kPa

800
made of how the piles’ long-term (108 day age) shaft
capacities were distributed between the till and chalk
layers. Although subject to greater uncertainty, the 400
capacities interpreted from dynamic re-strikes added
value at two other chalk-dominated sites where the
static tests did not reach full failure. 0
( f ) The integrated programme of static and dynamic 0 5 10 15 20 25
testing indicated that the shaft capacities developed Axial strain, εax: %
in the tills were highly sensitive to local clay content (b)
and permeability. The low-plasticity glacial till’s
interpreted long-term shaft capacity contributions were Fig. 20. Triaxial tests on samples of intact structured and structure-
broadly compatible with estimates made with the less chalk: (a) effective stress paths; (b) deviator stress plotted against
ICP-05 sand approach, when allowance was made axial strain (tests conducted by GEO)

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
676 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER
( j) Independent checking of the preliminary chalk ICP-18 structureless, structured low- and medium-density chalk. Note that
approach at other sites has given encouraging results. the externally measured strains are likely to be substantially higher
New research is underway to help further refine the than the local values due to compliance effects (see Jardine et al.,
approach. 1985). The overall behaviour is consistent with that reported in the
literature for intact chalk (e.g. Jardine et al., 1985; Addis & Jones,
1990; Leddra et al., 1993). The low-density samples are interpreted
ACKNOWLEDGEMENTS as showing peak values of 36° with c′ of 150 kPa. Cohesion appears
The authors gratefully acknowledge the major funding to increase with IDD, consistent with these chalks possessing a more
from Innovate-UK (grant no. 101968) to undertake the cemented fabric; peak ϕ′ in medium-density samples was similar at
research described, as well as Bilfinger Construction GmbH 36·4°; however, cohesion was closer to 200 kPa.
(principal contractor) and Allnamics Geotechnical B.V. who
carried out the pile testing. The authors are also grateful to
Professor Barry Lehane for his useful correspondence and A1.2 Glacial till
Professor Mark Randolph for the use of Impact. They also The undrained effective stress paths and stress–strain plots from
acknowledge the work of Gardline Geosciences Ltd, GEO typical CAU tests, conducted by GEO, on glacial till samples from
Copenhagen and Fugro Geoconsulting Ltd in their site the test sites are shown in Fig. 21. The maximum su values found
investigations and laboratory testing. from CAU tests range from  200 to 1000 kPa. The specimens
′ = 35·5°) with their
appear to reach critical state with Mtc = 1·44 (ϕcv
end points falling on a unique line in specific volume–mean effective
stress space, with slope λ = 0·07 (Cc = 0·161) and Γ at p′ = 1 kPa of
APPENDIX 1. TRIAXIAL BEHAVIOUR OF 1·68. The latter value of Cc is consistent with values reported by
WIKINGER CHALK AND TILL Gens (1982) on samples of Cromer till, by Jardine (1985) for
Appendix 1 provides illustrations of the behaviour seen in triaxial Magnus till and by Ushev (2018) for Cowden till.
tests on high-quality samples of chalk and till taken during the
Wikinger windfarm site investigations.
APPENDIX 2. ICP-05 CLAY METHOD PARAMETERS
The ICP-05 clay design procedure relies critically on establishing
A1.1 Chalk the profiles of YSR, defined as
The effective stress paths and stress–strain behaviour observed in
CAU triaxial tests is shown in Fig. 20 for intact samples of σ′vy
YSR ¼ ð11Þ
σ′v0
2000 ′ is the effective overburden pressure and σvy
here σv0 ′ is the effective
vertical yield stress which can sometimes be obtained from

1500 40
Interface friction angle, δ': degrees

δ'ult= 26·5 – 28°


30
q: kPa

1000

20

500
Site Depth: m Kc ef
WK31 8·1 m σ 'rc= 310 kPa
WK08 7·7 0·42 0·28 10
WK31 10·1 m σ 'rc= 290 kPa
WK38 11·3 0·51 0·23 Glacial till WK37 5·8 m σ 'rc= 210 kPa
OSS 3·1 0·40 0·28 Ring shear tests
0
0
0 400 800 1200 1600 2000 0 5 10 20 25
p': kPa Shearing displacement: mm
(a) (a)
2·5
0·1
Change in sample height, Δh: mm

Dilation

2·0
Mtc = 1·44
φ'cs = 35·5° 0
1·5
q/p'

1·0
–0·1

0·5
Contraction
–0·2
0 0 5 10 20 25
0 5 10 15 20 25 Shearing displacement: mm
Axial strain, εax: % (b)
(b)
Fig. 22. Ring shear tests on samples of glacial till conducted at
Fig. 21. CAU triaxial tests on samples of glacial till: (a) effective Imperial College: (a) interface friction angle plotted against displace-
stress paths; (b) q/p′ plotted against axial strain (tests conducted by ment; (b) change in sample thickness plotted against displacement
GEO) (Buckley, 2018)

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 677
oedometer tests, but may be difficult to resolve in tests on stiff glacial of materials. The above relationship is only valid for low-plasticity
tills (Lehane, 1992). Estimates can also be made from the ratio clays that fail in a ductile manner such as at Wikinger and a value of
′ , which is often related to YSR using relationships such as the
su/σv0 0·3 is adopted.
one given by Jardine et al. (2005) Site-specific drained interface shear angles are also critical to any
  ICP-05 analyses undertaken in clay layers. The large strain interface
su su shear behaviour was investigated by Buckley (2018) through ring
¼ YSR085 ð12Þ
σ′v0 σ′v0 nc shear interface tests in the Bishop ring shear apparatus at Imperial
College (Bishop et al., 1971) using interfaces with similar rough-
′ )nc is the value from CAU compression tests on K0
where (su/σv0 nesses to that of a driven pile (Ra = 10–15 μm). Three tests were
consolidated samples, which lies between 0·25 and 0·35 for a range carried out in which the glacial till was first sieved to remove

Table 11. Summary of equations and adopted parameters using models in Impact

No. Holocene and glacial till Chalk

Shaft resistance τ ¼ ks w þ cs v  τ stat G = G1 Gmax = : 895 + 17·3z: MPa


′ : MPa
G1 = 200σv0 G = G1 = 0·2 Gmax: MPa
G ρs = 2·2 G/cm3 ρs = 2·2 G/cm3
Spring ks ¼
πD αs = 1·15 αs = 1·1
pffiffiffiffiffiffiffiffi βs = 0·2 βs = 0·2
G
Radiation dashpot Cs ¼ ¼ Gρs ν = 0·5 ν = 0·5
Vs
"  βs #
Δv
Viscous effects τ inter ¼ τ stat 1 þ αs
v0
Base resistance Q ¼ Kb w þ Cb v  qb;stat
4GR
Spring Kb ¼
ð1  νÞ
4R2 pffiffiffiffiffiffiffiffi
Radiation dashpot Cb ¼ 08 Gρs
ð1  νÞ
4R3 ρs
Subsidiary mass m0 ¼ 016
ð1  νÞ

40 000
Measured F
Measured Zv
30 000 Calculated Zv
Displacement: m

20 000

10 000

0
F used as input
20 40 60 80
–10 000 Time: ms
(a)

10 000 –0·01
Measured Measured
Calculated Calculated
5000
0
Displacement: m

20 40 60 80
Displacement: m

0 Time: ms
20 40 60 80
0·01
Time: ms
–5000

0·02
–10 000

–15 000 0·03


(b) (c)

Fig. 23. Example signal match from blow WK38-2_DPT_EOD: measured and calculated (a) force (F ) and velocity times impedance (Zv); (b)
upward travelling force; (c) displacement

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
678 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER
particles . 425 μm ( 10–14%). Sieving is necessary as coarse ch coefficient of horizontal consolidation
particles can affect the results due to the limited specimen thickness, D diameter of pile
but could lead to results that are not fully representative of in situ F force in the pile
behaviour. The remaining material was then remoulded and placed Fup upward travelling wave
in two layers, as close to natural moisture content as practicable. The fs CPT sleeve friction
test procedure followed the recommendations of Jardine et al. G shear modulus
(2005). G1 secant shear modulus

The results, shown in Fig. 22, indicate that in all three cases δpeak Ghh shear modulus (propagating horizontally and polarised
and δult′ were reached at displacements of , 2 mm and , 10 mm, horizontally)
respectively. When examined upon completion of the test, all three Gmax maximum shear modulus
samples had adhered to the interface. There was no evidence of Gs specific gravity
polishing or formation of a shear surface at the interface. The values Gvh shear modulus (propagating vertically and polarised
of δult ′ values. While Lehane & Jardine
′ are , 0·5° lower than the δpeak horizontally)
(1994) report lower ultimate angles (22–24°) for a glacial clay at h distance from the pile tip
Cowden, where the fines content was significantly higher (30% clay Ir rigidity index = G/su
and 50% silt), Ushev (2018) found higher angles for Cowden till and K0 coefficient of earth pressure at rest (in situ)
Jardine (1985) reported higher values for Magnus till. Iverson et al. Kb base spring constant (dynamic soil resistance models)
(1998) demonstrated the tendency for ultimate friction angles in Kc coefficient of radial effective stress (shaft) after full
glacial till to reduce with increasing clay content. equalisation = σrc′ /σv0

Ks shaft spring constant (dynamic soil resistance models)
kc displacement creep rate
kl cyclic loading stiffness
kul cyclic unloading stiffness
APPENDIX 3. SUMMARY OF DYNAMIC ANALYSES Lp length of pile penetration
The analyses of the Wikinger dataset employed the Impact Lchalk
p length of pile penetration in chalk
software (Randolph, 2008) adopting the Randolph & Simons M empirical parameter used to assess set-up factor in Skov
(1986) soil model for the shaft and the Deeks & Randolph (1995) & Denver (1988) equation
model at the toe, as summarised in Table 11 and by Buckley et al. Mtc stress ratio at critical state in triaxial stress space
(2017). In both models, the primary input parameters of shaft and m empirical parameter used to assess set-up factor in Tan
base resistance, soil density and shear modulus are linked to et al. (2004) equation
measurable soil properties. The values of G were secant values, G1 m0 supplementary lumped mass connected through pile
degraded from the small strain, Gmax values to account indirectly for base node
soil non-linearity, following Alves et al. (2009) and Salgado et al. N number of axial cycles applied
(2015). In the Holocene and glacial till, the best matches were p′ mean effective stress
obtained taking G1 close to 200σv0 ′ , following the recommendations pa atmospheric pressure
of Lee et al. (1988) for sandy soils, which resulted in G1/Gmax ratios Qb pile base axial load resistance (capacity)
of , 0·3. In the chalk, the trend with depth obtained from the P–S Qs pile shaft axial load resistance (capacity)
logging Vs measurements was used to estimate Gmax with depth Qtot pile total axial load resistance (capacity)
(equation (3)), which was then reduced to G1 = 0·2Gmax. The same qb,0·1D pile end-bearing unit resistance at a displacement of 10%
values of shear modulus were adopted for both EOD and BOR of pile diameter
analyses. The viscosity parameters, αs and βs, along the shaft were qba pile end-bearing resistance under annulus
calculated from the correlations given by Loukidis et al. (2008). For qb,stat limit base stress plastic slider (dynamic soil resistance
clays, βs = βb = 0·2 and the value of αs at the shaft is given by model)
  qt total cone resistance (= qc + (1  a)u2)
su
αs ¼ 165  075 ð13Þ qt,1·5D average net CPT tip resistance ±1·5D around pile base
pa qu unconfined compressive strength
R pile radius
where pa is atmospheric pressure. For both the Holocene/till and the
R* equivalent radius for open-ended piles
chalk, βs was taken as 0·2, consistent with the recommendation of
Ra average centre-line roughness
Randolph (2008). The adopted value of αs in the chalk was 1·1,
Ri internal pile radius
taking su from remoulded samples to reflect the soft behaviour
Rt,c designer’s characteristic tensile capacity
expected in the annulus of chalk putty close to the shaft. However,
St sensitivity
substitution of the intact strength for glacial till into equation (13)
sacc accumulated permanent cyclic displacement
gives a negative value of αs, as the correlations were not developed
su undrained shear strength
for such high-strength insensitive materials. While Brown & Hyde
T dimensionless time factor
(2008) reported minimal rate dependence in Statnamic loading tests
T50 time to achieve 50% of ultimate pile set-up
conducted in glacial tills where the in situ moisture content was close
t time
to the plastic limit, the authors are not aware of any similar findings
t50 time for 50% dissipation of excess pore water pressures in
applying to such tills during fully dynamic driving. Lehane &
a CPT dissipation test
Jardine (1994) showed that rate effects had a significant impact on
tref reference time to assess pile ageing
axial installation resistance in Cowden glacial till, for which loading
tw pile wall thickness
rate has also been found to have a considerable influence on lateral
u2 CPT excess pore water pressures measured at the
pile capacity (McAdam et al., 2019). The αs value of 1·15 which gave
shoulder position
the best quality signal matches for the Wikinger glacial till is close to
Vp elastic compression wave velocity
the value of 1·0 adopted by Randolph (1993) for similar analyses on
Vs elastic shear wave velocity
0·762 mm open piles driven at Tilbrook Grange in stiff, low-
v velocity
plasticity, glacial Lowestoft till.
v0 reference velocity (= 1 m/s)
An example signal match is shown in Fig. 23.
w displacement
wc water content
wl liquid limit
NOTATION wpl plastic limit
a cone area ratio wsat saturation moisture content
Cb base dashpot constant (dynamic soil resistance models) Z pile impedance
Cc compression index z depth
Cs shaft dashpot constant (dynamic soil resistance models) α α-type driven pile design method factor
c′ effective cohesion intercept αs shaft viscosity parameter (soil resistance model)

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 679
βs shaft viscosity parameter (soil resistance model) Barbosa, P., Geduhn, M., Jardine, R. J., Schroeder, F. C. & Horn, M.
Γ specific volume on the critical state line at p′ = 1 kPa (2015a). Offshore pile load tests in chalk. In Geotechnical
γbulk bulk density engineering for infrastructure and development: proceedings of
γd dry density the XVI European conference on soil mechanics and geotechnical
Δh change in sample height during shearing engineering (eds M. G. Winter, D. M. Smith, P. J. L. Eldred
Δr average movement of the soil grains due to dilation and D. G. Toll), vol. 6, pp. 2885–2890. London, UK: ICE
Δv relative velocity between the pile and soil Publishing.

Δσrd change of radial effective stress during shearing due to Barbosa, P., Geduhn, M., Jardine, R. J. & Schroeder, F. C. (2015b).
dilation Full scale offshore verification of axial pile design in chalk. In
δ′ interface angle of shearing resistance Frontiers in offshore geotechnics III (ed. V. Meyer), pp. 516–520.
δf′ failure interface angle of shearing resistance Leiden, the Netherlands: CRC Press/Balkema.

δpeak peak interface angle of shearing resistance Barbosa, P. M., Geduhn, M., Jardine, R. J. & Schroeder, F. C.

δult ultimate interface angle of shearing resistance (2017). Large scale offshore static pile tests –practicality and
εax axial strain benefits. In Proceedings of 8th international conference on
Λ set-up factor on shaft capacity offshore site investigation and geotechnics, London, UK,
Λult ultimate set-up factor on shaft capacity pp. 644–651. London, UK: Society for Underwater Technology.
λ slope of the isotropic compression line Bishop, A. W., Green, G., Garga, V. K., Andresen, A. & Brown, J.
ν Poisson ratio (1971). A new ring shear apparatus and its application to the
ρs soil or chalk mass density measurement of residual strength. Géotechnique 21, No. 4,
σrc′ radial effective stress after equalisation 273–328, https://doi.org/10.1680/geot.1971.21.4.273.
σri′ short-term radial effective stresses Bowden, A. J., Spink, T. W. & Mortimore, R. N. (2002). The
σrf′ radial effective stress at failure engineering description of chalk: its strength, hardness and

σv0 vertical effective stress density. Q. J. Engng Geol. Hydrogeol., 35, No. 4, 355–361.

σvy vertical yield stress Brinch Hansen, J. (1963). Discussion: hyperbolic stress strain
τavg average shaft shear stress at failure from static or response: cohesive soils. J. Soil Mech. Found. Div., 89, No. 4,
dynamic test 241–242.
τavg (t) average shaft shear stress at failure from static or Brown, M. J. & Hyde, A. F. L. (2008). Rate effects from pile
dynamic test at time t shaft resistance measurements. Can. Geotech. J. 45, No. 3,
τcreep-yield average shaft shear stress below which static test creep 425–431.
rates are negligible BSH (Bundesamt für Seeschifffahrt und Hydrographie) (2015).
τinter limit shaft shear stress at low strain rates (dynamic soil Minimum requirements concerning the constructive design of
resistance model) offshore structures within the exclusive economic zone (EEZ).
τf calculated long-term local shear stress at static failure Hamburg, Germany: Bundesamt für Seeschifffahrt und
from design methods Hydrographie.
τs,d local shear stress at failure interpreted from dynamic Buckley, R. M. (2018). The axial behaviour of displacement piles in
tests chalk. PhD thesis, Imperial College London, London, UK.
τstat limit shear stress plastic slider (dynamic soil resistance Buckley, R. M., Kontoe, S., Jardine, R. J., Maron, M.,
model) Schroeder, F. C. & Barbosa, P. (2017). Common pitfalls of pile
ϕ′ angle of shearing resistance driving resistance analysis – a case study of the Wikinger

ϕcv angle of shearing resistance at critical state offshore windfarm. In Proceedings of the 8th international
conference on offshore site investigation and geotechnics,
London, UK, pp. 1246–1253. London, UK: Society for
Underwater Technology.
Buckley, R. M., Jardine, R. J., Kontoe, S., Parker, D. &
REFERENCES Schroeder, F. C. (2018a). Ageing and cyclic behaviour of
Addis, M. & Jones, M. (1990). Mechanical behaviour and strain rate axially loaded piles driven in chalk. Géotechnique 68, No. 2,
dependance of high porosity chalk. In Chalk: proceedings of the 146–161. https://doi.org/10.1680/jgeot.17.P.012.
international chalk symposium held at Brighton Polytechnic Buckley, R. M., Jardine, R. J., Kontoe, S. & Lehane, B. M. (2018b).
on 4–7 September 1989, pp. 239–244. London, UK: Effective stress regime around a jacked steel pile during
Thomas Telford. installation ageing and load testing in chalk. Can. Geotech. J.,
Al-Shafei, K., Cox, W. & Helfrich, S. (1994). Pile load tests in dense 55, No. 11, 1577–1591, https://doi.org/10.1139/cgj-2017-0145.
sand: analysis of static test results. Proceedings of the offshore Byrne, T., Doherty, P., Gavin, K. & Overy, R. (2012). Comparison of
technology conference, Houston, TX, USA, paper OTC-7381- pile driveability methods in North Sea sand. In Proceedings of
MS. the 7th international conference on offshore site investigations and
Alves, A. M., Lopes, F. R., Randolph, M. F. & Danziger, B. R. geotechnics, London, UK, pp. 481–488. London, UK: Society
(2009). Investigations on the dynamic behavior of a small- for Underwater Technology.
diameter pile driven in soft clay. Can. Geotech. J., 46, No. 12, Carotenuto, P., Meyer, V., Strøm, P. J., Cabarkapa, Z., St. John, H. &
1418–1430. Jardine, R. J. (2018). Installation and axial capacity of the
API (American Petroleum Institute) (2014). API RP 2A-WSD: Sheringham Shoal offshore wind farm monopiles – a case
recommended practice for planning, designing and constructing history. In Engineering in chalk: proceedings of the chalk 2018
fixed offshore platforms – working stress design, 22nd edn. conference (eds J. A. Lawrence, M. Preene, U. L. Lawrence and
Washington, DC, USA: American Petroleum Institute. R. Buckley), pp. 117–122. London, UK: ICE Publishing.
Arup (Ove-Arup & Partners) (1986). Research on the behavior of Chan, L. D., Buckley, R. M., Liu, T. & Jardine, R. J. (2019).
piles as anchors for buoyant structures – summary report, offshore Laboratory investigation of interface shearing in chalk. In 7th
technology report, OTH 86 215. London, UK: Department of International symposium on deformation characteristics of geo-
Energy. materials (IS-Glasgow 2019) (eds A. Tarantino and E. Ibraim),
Augustesen, A. H., Leth, C. T., Ostergaard, M. U., Moller, M., E3S Web of Conferences vol. 92, article 13009. London, UK:
Dührkop, J. & Barbosa, P. (2015). Design methodology for EDP Sciences.
cyclically and axially loaded piles in chalk for Wikinger OWF. In Chin, F. K. (1970). Estimation of the ultimate load of piles not
Frontiers in offshore geotechnics III (ed. V. Meyer), pp. 509–514. carried to failure. Proceedings of the 2nd Southeast Asian
Leiden, the Netherlands: CRC Press/Balkema. conference on soil engineering, Singapore, pp. 81–92.
Baldi, G., Belloti, R., Ghionna, V. N., Jamiolkowski, M. & Lo Chin, F. K. (1971). Discussion: pile tests – Arkansas river project.
Presti, D. C. F. (1989). Modulus of sands from CPTs and DMTs. J. Soil Mech. Found. Div., 97, No. 6, 930–932.
In Proceedings of the 12th international conference on soil Chow, F. C. (1997). Investigations into displacement pile behaviour for
mechanics and foundation engineering, Rio de Janeiro, Brazil, offshore foundations. PhD thesis, Imperial College London,
pp. 165–170. Rotterdam, the Netherlands: Balkema. London, UK.

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
680 BUCKLEY, JARDINE, KONTOE, BARBOSA AND SCHROEDER
Chow, F. C., Jardine, R. J., Brucy, F. & Nauroy, J. F. (1998). Effects PhD thesis, Imperial College London (University of London),
of time on capacity of pipe piles in dense marine sand. London, UK.
J. Geotech. Geoenviron. Engng, 124, No. 3, 254–264. Jardine, R. J. (2019). Geotechnics, energy and climate change: the
CI (Cambridge-Insitu) (2013). Wikinger offshore wind farm ground 56th Rankine lecture. Géotechnique, https://doi.org/10.1680/
investigation 2013 – results of pressuremeter tests, CIR1286/13. jgeot.18.RL.001.
Cambridge, UK: Cambridge Insitu Ltd. Jardine, R. J., Brooks, N. J. & Smith, P. R. (1985). The use of
Ciavaglia, F., Carey, J. & Diambra, A. (2017). Time-dependent uplift electrolevel transducers for strain measurements in triaxial tests
capacity of driven piles in low to medium density chalk. on weak rock. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr.,
Géotechnique Lett. 7, No. 1, 90–96, https://doi.org/10. 22, No. 5, 331–337.
1680/jgele.16.00162. Jardine, R. J., Chow, F. C., Overy, R. & Standing, J. R. (2005). ICP
CIJV (COWI/IMS Joint Venture) (2013). Wikinger offshore wind- design methods for driven piles in sands and clays. London, UK:
farm: geotechnical site survey report based on main geotechnical Thomas Telford.
campaign, A031412-005-RP. Kongens Lyngby, Denmark: Jardine, R. J., Standing, J. R. & Chow, F. C. (2006). Some
COWI/IMS Joint Venture. observations of the effects of time on the capacity of piles
CIJV (2014a). Pile load test concept based on geotechnical main driven in sand. Géotechnique 56, No. 4, 227–244, https://doi.
campaign, WIK-CIV-IB-0816. Kongens Lyngby, Denmark: org/10.1680/geot.2006.56.4.227.
COWI/IMS Joint Venture. Jardine, R. J., Buckley, R. M., Kontoe, S., Barbosa, P. &
CIJV (2014b). Wikinger offshore windfarm: soil and foundation Schroeder, F. C. (2018). Behaviour of piles driven in chalk. In
expertise report based on main geotechnical campaign, WIK- Engineering in chalk: proceedings of the chalk 2018 conference
CIJV-IB-0488. Kongens Lyngby, Denmark: COWI/IMS Joint (eds J. A. Lawrence, M. Preene, U. L. Lawrence and
Venture. R. Buckley), pp. 33–51. London, UK: ICE Publishing.
Clarke, J., Long, M. & Hamilton, J. (1993). The axial tension test of Jardine, R. J., Buckley, R. M., Byrne, B., Kontoe, S., Macadam, R.
an instrumented pile in overconsolidated clay at Tilbrook Grange. & Vinck, K. (2019). The ALPACA research project to improve
Large-scale pile tests in clay. London, UK: Thomas Telford driven pile design in chalk. In Proceedings of the XVII European
Publishing. conference on soil mechanics and geotechnical engineering,
Clayton, C. R. I. (1978). Chalk as fill. PhD thesis, University of Reykjavík, Iceland: geotechnical engineering, foundation of the
Surrey, Guildford, UK. future (eds H. Sigursteinsson, S. Erlingsson and B. Bessason),
Decourt, L. (1999). Behavior of foundations under working load paper 0071. Reykjavík, Iceland: Icelandic Geotechnical Society.
conditions. Proceedings of the 11th Pan-American conference on Karlsrud, K., Clausen, C. & Aas, P. (2005). Bearing capacity of
soil mechanics and geotechnical engineering, Foz do Iguassu, driven piles in clay, the NGI approach. In Frontiers in offshore
Brazil, pp. 453–488. geotechnics (eds S. Gourvenec and M. Cassidy), pp. 775–782.
Deeks, A. J. & Randolph, M. F. (1995). A simple model for inelastic Boca Raton, FL, USA: CRC Press.
footing response to transient loading. Int. J. Numer. Methods Karlsrud, K., Jensen, T. G., Lied, E. K. W., Nowacki, F. &
Geotech. Engng, 19, No. 5, 307–329. Simonsen, A. S. (2014). Significant ageing effects for axially
DGGT (German Geotechnical Society) (ed.) (2014). Recommen- loaded piles in sand and clay verified by new field load tests.
dations on Piling (EA-Pfähle). Berlin, Germany: Ernst Proceedings of the offshore technology conference, Houston, TX,
& Sohn. USA, paper OTC-25197-MS.
DIN (Deutshces Institut für Normung) (2009). DIN-EN-1997-1: Kolk, H. & der Velde, E. (1996). A reliable method to determine
Eurocode 7: geotechnical design – part 1: General rules; English friction capacity of piles driven into clays. Proceedings of
version. Berlin, Germany: DIN. the offshore technology conference, Houston, TX, USA,
DIN (2012). DIN-1054 Subsoil: verification of the safety of paper OTC-7993-MS.
earthworks and foundations, supplementary rules to DIN EN Lahrs, T. & Kallias, A. (2013). Probebelastungen von
1997-1. Berlin, Germany: DIN. Stahlrohren in Kreide für den Offshore-Windpark Baltic 2.
Doughty, L. (2016). Laboratory testing of chalk. MSc thesis, Proceedings of the Pfahl symposium, Braunschweig, Germany,
Imperial College London, London, UK. pp. 451–466 (in German).
Doughty, L. J., Buckley, R. M. & Jardine, R. J. (2018). Investigating Le, T. M. H., Eiksund, G. R. & Strøm, P. J. (2014). Characterisation
the effect of ageing on the behaviour of chalk putty. of residual shear strength at the Sheringham Shoal offshore wind
In Engineering in chalk: proceedings of the chalk 2018 conference farm. Proceedings of the 33rd international conference on ocean,
(eds J. A. Lawrence, M. Preene, U. L. Lawrence and offshore and arctic engineering, San Francisco, CA, USA.
R. Buckley), pp. 695–701. London, UK: ICE Publishing. Leddra, M. J., Jones, M. E. & Goldsmith, A. S. (1993). Compaction
Dührkop, J., Maretzki, S. & Rieser, J. (2017). Re-evaluation of pile and shear deformation of a weakly-cemented, high porosity
driveability in chalk. In Proceedings of 8th international sedimentary rock. In Proceedings of conference on engineering
conference on offshore site investigation and geotechnics, geology of weak rock, Leeds, UK (eds J. C. Cripps,
London, UK, pp. 666–673. London, UK: Society for J. M. Coulthard, M. G. Culshaw, A. Forster, S. R. Hencher
Underwater Technology. and C. F. Moon), pp. 45–54. Rotterdam, the Netherlands:
Fugro (2013). Wikinger offshore windfarm: advanced laboratory and Balkema.
drivability assessment, J22026-1. Wallingford, UK: Fugro Lee, S. L., Chow, Y. K., Karunaratne, G. P. & Wong, K. Y. (1988).
Geoconsulting Ltd. Rational wave equation model for pile-driving analysis.
Gavin, K. & Lehane, B. M. (2007). Base load–displacement J. Geotech. Engng, 114, No. 3, 306–325.
response of piles in sand. Can. Geotech. J., 44, No. 9, 1053–1063. Lehane, B. M. (1992). Experimental investigations of pile behaviour
Gens, A. (1982). Stress–strain and strength characteristics of a low using instrumented field piles. PhD thesis, Imperial College
plasticity clay. PhD thesis, Imperial College London (University London, London, UK.
of London), London, UK. Lehane, B. M. & Jardine, R. J. (1994). Displacement pile behaviour
Han, F., Ganju, E., Prezzi, M., Salgado, R. & Zaheer, M. (2019). in glacial clay. Can. Geotech. J., 31, No. 1, 79–90.
Axial resistance of open-ended pipe pile driven in gravelly Lehane, B. M., Jardine, R. J., Bond, A. J. & Frank, R. (1993).
sand. Géotechnique, https://doi.org/10.1680/jgeot.18.P.117. Mechanisms of shaft friction in sand from instrumented pile
ISO (International Standards Organisation) (2007). ISO 19902: tests. J. Geotech. Engng, 119, No. 1, 19–35.
2008-07: Petroleum and natural gas industries – fixed steel Lehane, B. M., Schneider, J. A. & Xu, X. (2005). The UWA-05
offshore structures, 1st edn. Geneva, Switzerland: International method for prediction of axial capacity of driven piles in sand. In
Standards Organisation. Frontiers in offshore geotechnics (eds S. Gourvenec and
Iverson, N. R., Hooyer, T. S. & Baker, R. W. (1998). Ring-shear M. Cassidy), pp. 683–689. Boca Raton, FL, USA: CRC Press.
studies of till deformation: Coulomb-plastic behavior and Lehane, B. M., Li, Y. & Williams, R. (2013). Shaft capacity of
distributed strain in glacier beds. J. Glaciol. 44, No. 148, displacement piles in clay using the cone penetration test.
634–642. J. Geotech. Geoenviron. Engng 139, No. 2, 253–266.
Jardine, R. J. (1985). Investigations of pile–soil behaviour, with Lehane, B. M., Lim, J. K., Carotenuto, P., Nadim, F., Lacasse, S.,
special reference to the foundations of offshore structures. Jardine, R. J. & van Dijk, B. (2017). Characteristics of unified

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.
FULL-SCALE OBSERVATIONS OF THE AXIAL RESPONSES OF OFFSHORE PILES 681
databases for driven piles. In Proceedings of 8th international Randolph, M. F. (2008). IMPACT – dynamic analysis of pile
conference on offshore site investigation and geotechnics, London, driving, manual. Crawley, Australia: University of Western
UK, pp. 162–194. London, UK: Society for Underwater Australia.
Technology. Randolph, M. F. & Simons, H. A. (1986). An improved soil model
Long, M. & Menkiti, C. O. (2007). Geotechnical properties of for one-dimensional pile driving analysis. In Proceedings of
Dublin boulder clay. Géotechnique 57, No. 7, 595–611, the 3rd international conference on numerical methods in
https://doi.org/10.1680/geot.2007.57.7.595. offshore piling, Nantes, France, pp. 3–17. Paris, France:
Lord, J. A., Clayton, C. R. I. & Mortimore, R. N. (2002). Editions Technip.
Engineering in chalk, Ciria C574. London, UK: Construction Randolph, M. F. & Wroth, C. P. (1979). An analytical solution for
Industry Research and Information Association (Ciria). the consolidation around a driven pile. Int. J. Numer. Analyt.
Loukidis, D., Salgado, R. & Abou-Jaoude, G. (2008). Assessment of Methods Geomech., 3, No. 3, 217–229.
axially-loaded pile dynamic design methods and review of INDOT Razoaki, R. N. (2000). Effect of ageing on mechanics of chalk
axially-loaded design procedure, Publication FHWA/IN/ slurries. PhD thesis, University of Portsmouth, Portsmouth, UK.
JTRP-2008/6. West Lafayette, IN, USA: Joint Transportation Robertson, P. K. (1990). Soil classification using the cone pen-
Research Program, Indiana Department of Transportation and etration test. Can. Geotech. J. 27, No. 1, 151–158.
Purdue University. Salgado, R., Loukidis, D., Abou-Jaoude, G. & Zhang, Y. (2015).
Matthews, M. C. & Clayton, C. R. I. (1993). Influence of intact The role of soil stiffness non-linearity in 1D pile driving
porosity on the engineering properties of a weak rock. In simulations. Géotechnique 65, No. 3, 169–187, https://doi.
Proceedings of international symposium on geotechnical engin- org/10.1680/geot.13.P.124.
eering of hard soils and soft rocks, Athens, Greece (eds Skov, R. & Denver, H. (1988). Time-dependence of bearing capacity
A. Anagnostopoulos, R. Frank, N. Kalteziotis and of piles. In Proceedings of the 3rd international conference on the
F. Schlosser), pp. 693–702. Rotterdam, the Netherlands: application of stress-wave theory to piles, Ottowa, Canada (ed. B.
Balkema. Fellenius), pp. 879–888. Richmond, BC, Canada: BiTech.
Mayne, P. W. & Rix, G. J. (1993). Gmax–qc relationships for clays. Smith, A. K. C. (2001). Interpretation of cone penetration tests in
Geotech. Test. J. 16, No. 1, 54–60. chalk. Ground Engng 34, No. 9, 30–35.
McAdam, R. A., Byrne, B. W., Houlsby, G. T., Beuckelaers, W. J. A. Tan, S. L., Cuthbertson, J. & Kimmerling, R. E. (2004). Prediction
P., Burd, H. J., Gavin, K. G., Igoe, D. J. P., Jardine, R. J., of pile set-up in non-cohesive soils. In Current practices and
Martin, C. M., Muir Wood, A., Potts, D. M., Skov Gretlund, J., future trends in deep foundations (eds J. A. DiMaggioand
Taborda, D. M. G. & Zdravković, L. (2019). Monotonic lateral M. H. Hussein), GSP 125, pp. 50–65. Reston, VA, USA:
loaded pile testing in a dense marine sand at Dunkirk. American Society of Civil Engineers.
Géotechnique, https://doi.org/10.1680/jgeot.18.PISA.004. Ushev, E. (2018). Laboratory investigation of the mechanical
McGann, C. R., Bradley, B. A., Taylor, M. L., Wotherspoon, L. M. properties of Cowden till under static and cyclic conditions. PhD
& Cubrinovski, M. (2015). Applicability of existing empirical thesis, Imperial College, London, UK.
shear wave velocity correlations to seismic cone penetration test Weltman, A. & Healy, P. (1978). Piling in ‘Boulder Clay’ and other
data in Christchurch New Zealand. Soil Dyn. Earthq. Engng 75, glacial tills, DoE/Ciria report PG 5. London, UK: Ciria.
76–86. White, D. J., Schneider, J. A. & Lehane, B. M. (2005). The influence
Obst, K., Nachtweide, C. & Müller, U. (2017). Late Saalian and of effective area ratio on shaft friction of displacement piles
Weichselian glaciations in the German Baltic sea documented in sand. In Frontiers in offshore geotechnics (eds S. Gourvenec
by Pleistocene successions at the southeastern margin of the and M. Cassidy), pp. 741–748. Boca Raton, FL, USA:
Arkona Basin. Boreas 46, No. 1, 18–33. CRC Press.
Randolph, M. F. (1993). Analysis of stress-wave data from pile Xu, X., Liu, H. & Lehane, B. M. (2006). Pipe pile installation effects
tests at Pentre and Tilbrook. In Large-scale pile tests in clay in soft clay. Proc. Instn Civ. Engrs – Geotech. Engng 159, No. 4,
(ed. J. Clarke). London, UK: Thomas Telford. 285–296.
Randolph, M. F. (2003). Science and empiricism in pile foundation Ziogos, A., Brown, M., Ivanovic, A. & Morgan, N. (2017). Chalk–
design. Géotechnique 53, No. 10, 847–875, https://doi.org/ steel interface testing for marine energy foundations. Proc. Instn
10.1680/geot.2003.53.10.847. Civ. Engrs – Geotech. Engng 170, No. 3, 285–298.

Downloaded by [ University Of Western Australia] on [03/09/24]. Copyright © ICE Publishing, all rights reserved.

You might also like