You are on page 1of 35

Houlsby, G. T. Gotechnique [http://dx.doi.org/10.1680/jgeot.15.RL.

001]

Interactions in offshore foundation design


G. T. HOULSBY 

This paper presents some examples of design problems for offshore foundations, drawn from the
jack-up industry and the wind turbine industry. The examples are chosen to illustrate some general
points about foundation design, geotechnical engineering and its interaction with other disciplines. The
first example is drawn from the assessment of the safety of installation of jack-up units (large mobile
offshore drilling rigs). It illustrates how more rational approaches can be achieved through a deeper use
of probabilistic methods in both the prediction of performance and the assessment of field observations.
The second example also comes from jack-up practice, but has wider application too: it addresses the
classical problem of the performance of foundations under combined loading, and how this can be
understood in a simple theoretical and practical framework based on plasticity theory. The final
example comes from the renewables sector, where the rapidly expanding offshore wind industry poses
new foundation challenges for geotechnical engineers. Practical and economic foundation solutions are
required if the UK is to meet its ambitious plans to exploit larger turbines in deeper waters. Both
conventional (monopile) and novel solutions (suction caissons, screw piles) to the foundation problem
are discussed. The paper also demonstrates how interactions with other disciplines can enrich
geotechnical engineering, illustrated by specific practical examples from the authors experience.

KEYWORDS: clays; footings/foundations; model tests; offshore engineering; piles & piling; plasticity;
sands; soil/structure interaction; statistical analysis

INTRODUCTION develop detailed numerical models. The models developed


The subject of this paper is the design of foundations also have application to onshore foundations subjected to
for offshore structures. The structures of interest have in the combined loads.
past been mainly for the offshore oil and gas industry, but Part 3 is devoted to the offshore renewables sector, examin-
increasingly during the last decade the focus has shifted to ing the options available for offshore wind turbine foun-
the offshore renewables sector, especially offshore wind. No dations as they move to larger installations in deeper water.
attempt is made here to present a detailed or comprehensive Both conventional (monopile) and less conventional (suction
review of foundations for offshore structures; instead this caisson and screw pile) approaches are discussed.
paper presents a selection of topics that illustrate certain Although geotechnical engineering forms the core of this
important themes, some with application throughout geo- paper, the opportunity is taken to emphasise links with other
technical engineering, not just offshore foundations. disciplines. The interactions in the title deliberately has a
The paper is divided into three main sections. The first two double meaning. First it represents geotechnical problems
use examples from the design and analysis of jack-up units, where the interactions between different forces are important,
which are mobile drilling rigs of enormous importance in the and each component of the problem cannot be treated in
oil and gas industry. Part 1 addresses the prediction of the isolation. Second, there are problems where interactions with
behaviour of a jack-up unit during installation at a particular other disciplines are important for geotechnical engineers,
site an important consideration because each unit may be as it is at these interfaces that the most challenging and
moved to different sites several times in a year. In this section interesting problems arise.
a framework is presented in which the predictions, and the
comparisons of performance with predictions, are made
within a probabilistic framework. This leads to a clear
conclusion that there are opportunities to improve geotech- PART 1: INSTALLATION OF JACK-UP UNITS
nical practice by a more sophisticated application of A jack-up unit, Fig. 1, is a mobile rig mainly used for
probability theory than is current. drilling either exploration or production wells, although it
Part 2 addresses the performance of jack-up units in is also sometimes used for other purposes such as accom-
service, and especially how their safety can be assured during modation. The principal concern here is with the very large
extreme storm conditions. Emphasis is placed on the units used in the oil and gas industry. The large rigs almost
development of models of soil behaviour based on plasticity always have three lattice-work legs, and very occasionally
theory, effectively generalising the concept of bearing four. Typically the legs are up to 180 m long, the spacing
capacity to multi-axial loading. This approach provides a between them is about 60 m, and large rigs can operate in
conceptual basis for jack-up behaviour, and is also used to up to about 120 m of water. A large jack-up has a mass of
around 30 000 t. Much smaller units, of a rather different
design, and with four or more legs, are used for inshore work
in shallower waters.
Manuscript received 22 December 2015; revised manuscript The legs can each be moved up and down independently,
accepted 26 April 2016. usually by a rack-and-pinion system, Fig. 2, with multiple
Discussion on this paper is welcomed by the editor. driving pinions on each leg. On the bottom of the legs are
 Department of Engineering Science, University of Oxford, large, roughly circular footings which are called spudcans.
Parks Road, Oxford, OX1 3PJ, UK. They usually have a shallow conical base, often with a

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
2 HOULSBY

Fig. 1. A typical large mobile jack-up unit


Fig. 2. Rack-and-pinion jacking units (photograph: Keppel FELS)

pointed tip, and are of diameter up to about 20 m. The


precise reason for this design is unclear, but appears to be that it represents a full-scale bearing capacity test in which
based on the concept that the pointed tip provides good the soil is continually brought to failure as the spudcans
location of the rig on initial touchdown of the spudcans penetrate the seabed. Jack-up installation thus provides an
on the seabed, with the flatter section then providing a sub- almost unique opportunity for checking and calibrating
stantial capacity with only a small further penetration. bearing capacity theories at full scale.
Occasionally other designs are used, for instance incorporat- Spudcan installation in both coarse- and fine-grained soils
ing skirts, but these are not discussed here. is important, but the examples pursued here concentrate on
The operation of a jack-up follows the sequence described installation in soft clay, where the penetrations can be sub-
below. stantial, for example up to 30 m. The penetration process in
soft clay occurs as follows, Fig. 4. As the spudcan tip first
(a) The rig is floated to the site on its own hull, with the penetrates the soil the load is low, but rapidly increases as the
legs raised, Fig. 3(a). area of the base of the spudcan in contact with the ground
(b) The legs are lowered to the seabed, and then jacked into increases with penetration. Once the full base of the spudcan
the seabed, Fig. 3(b). is in contact, the rate of increase slows down, and is mainly
(c) The rig is raised on the legs until it is about 2 m out of due to (a) increasing soil strength with depth (as in most
the water, and at that stage the full weight of the rig soft clays the strength increases markedly with depth), and
bears on the three spudcans, Fig. 3(c). This is called (b) geometric effects, which have a smaller but nevertheless
light-ship load. important effect. A certain penetration will be reached at
(d ) Water is then pumped into ballast tanks in the hull, to light-ship load (illustrated as position A to the left of Fig. 4),
increase the weight of the rig, and so push the legs and then as the preload is applied the penetration increases
further into the seabed, Fig. 3(d). This process is called further. When the preload is removed, the spudcan hardly
preloading. Typically at full preload the weight of the moves at all, although there will be a small upward elastic
rig has been increased to twice the light-ship weight. rebound (position B in Fig. 4).
For the smaller rigs with four or more legs the process The vertical load on the spudcan at any penetration is
of diagonal preloading can be used, in which the jacks usually calculated according to standard bearing capacity
are operated in such a way as to load diagonal pairs of theory, equation (1)
legs alternately. V N c su hA 1
(e) The preload is then dumped and the rig returns to the
light-ship weight, Fig. 3(e). with the first term due to the strength of the soil and the
( f ) The hull then climbs further up the legs so that a safe second coming from the overburden pressure once the
air gap, typically of around 20 m, can be established, spudcan is embedded into the soil. An important detail is
Fig. 3(f). The air gap is necessary so that the rig can the definition of the depth h of the spudcan. In offshore
survive extreme storm conditions, Fig. 3(g), without the practice the penetration depth usually refers to the depth of
wave crest level reaching the hull. Even so, the legs may the tip of the spudcan below the seabed. For the purposes of
be subjected to large lateral forces from waves and the bearing capacity calculation, the depth h is usually taken
current, and there will also be wind loads on the as the depth of the shoulder of the spudcan (the lowest point
superstructure. on the spudcan at which the full bearing area is achieved).
Because of the conical base of the spudcan the two values are
The installation and preloading process can be understood as different, and care is necessary in distinguishing between
a bearing capacity problem indeed it is highly unusual in them. Note that because the entire process is submerged in

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 3

(a) (b) (c) (d) (e) (f) (g)

Fig. 3. Stages in the installation and operation of a jack-up unit: (a) float to site; (b) lower legs; (c) light ship load; (d) preload; (e) dump preload;
(f) climb to air gap and operate; (g) storm (adapted from Poulos, 1988)

Light-ship load Load on Undrained


spudcan, V strength, su
A

Preload
B h

Depth, z Depth, z

Fig. 4. Idealised loadpenetration curve for a spudcan footing

150 125 100 075 050 025 025 050 075 100 125 150

025

050

075

100

Fig. 5. Characteristic lines for lower-bound solution for an embedded conical footing (from Houlsby & Martin, 2003)

seawater, the appropriate value for is the submerged key variables


(buoyant) unit weight.  
The most important number in equation (1) is the bearing h D
N c f ; ; ; 2
capacity factor, Nc. This can be calculated by several means; D su0
one quite accurate method is to use the slip line theory of
plasticity theory (see e.g. Houlsby & Martin, 2003). This is In equation (2), Nc is considered as a function of the
essentially the same method used to develop the classical following parameters.
Terzaghi bearing capacity factors, but adapted to account
for the axisymmetric problem (see e.g. Szczepinski, 1974; The angle of the conical base of the spudcan for these
Houlsby & Wroth, 1982; Salenon & Matar, 1982) and a purposes the rather more complicated actual shape is
number of other features. A typical lower-bound mesh of usually replaced by an equivalent cone.
characteristic lines is shown in Fig. 5. This approach allows The roughness of the base of the spudcan, 0   1.
the bearing capacity to be expressed as a function of some This quantity is difficult to determine, and most spudcans

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
4 HOULSBY
Undrained Light-ship Load on
strength load spudcan

Lo
Data

we
rb
ou
nd
Preload

Be
?

st
e
st
im
at
e
W
or
st
ca
se
Depth Depth

Fig. 6. Most probable and worst-case penetration curves for a spudcan

can almost certainly be treated as rough. The maximum diameter and the cone angle (for partial penetration of the
adhesion on the spudcan base is defined as au su. spudcan the diameter in contact with the soil, and therefore
The depth of embedment, expressed as fraction of the the equivalent cone angle, will vary with depth). The strength
diameter, h/D. Importantly, this is a separate effect from at the penetration depth is determined from a design profile,
the weight term h added by the overburden. It arises and averaged over a range of depths if necessary. The approp-
because a more deeply embedded spudcan involves failure riate bearing capacity factor is calculated from equation (2),
of a larger volume of soil. and a single value for the load on the spudcan is determined
The rate of increase of strength with depth, . In most at that depth, using equation (1), possibly modified to ac-
soft clays, where large penetrations occur, the strength count for flow-around conditions, which result in changes in
increases with depth, often approximately linearly. the buoyancy terms as well as in the bearing capacity factor
The rate of increase is expressed in non-dimensional form (Osborne et al., 2011).
as D/su0, following Davies & Booker (1973). This process is repeated at every depth, so that a complete
predicted loadpenetration curve is obtained, extending well
There are two further effects that also need to be taken into below the expected depth of penetration of the spudcan. In
account. First, in more complex soil profiles it is inappropriate practice this process is taken at least one step further: the design
just to use the strength value at the base of the footing, and strength profile is invariably obtained as a fit to some fairly
instead some averaging procedure over a limited depth of soil scattered data, often from multiple measurement devices, for
needs to be used. Avariety of ranges over which this averaging instance a cone penetration test (CPT) profile, an extensive
should be taken have been suggested, see for instance Young number of measurements on samples using perhaps a pocket
et al. (1984). However, the geometry of the plastic failure penetrometer, a torvane and/or a minivane, and a few isolated
mechanism gives a reasonable clue about how far this should UU (unconfined undrained triaxial test) measurements. It is
extend both above and below the footing base. For instance, the usual to derive a best-fit (or most probable) design profile, and
recommendation of the InSafeJIP study (Osborne et al., 2011) also a lower bound (or worst case), Fig. 6. The most probable
was that averaging should take place over a depth extending profile would provide expected penetrations of the spudcans,
from elevation z h  025h to z h 025yc 025D, where and the worst case would give the maximum penetration values
yc is the height of the equivalent cone representing the spudcan. that could reasonably be expected. It is these maximum values
Second, once the spudcan is deeply buried, there is less which will determine whether it is safe to proceed with the
resistance for the soil to flow back into the hole behind the installation operation. If installation does proceed, then the
spudcan, rather than out to the surface. Formally this can be field behaviour would be compared with the predictions (both
addressed through examination of upper-bound failure mech- expected and worst case).
anisms. There is a change of mechanism for deeper pen- The key question is how the lower bound to the strength
etrations, with the flow-around mechanism governing. Hossain profile can be chosen in a rational manner. If the designer
et al. (2005) have suggested an empirical approach for deter- is too cautious, the predicted penetration will be so large
mining the depth at which this transition occurs. Importantly, that, quite incorrectly, it may be concluded that operation
whether or not the hole behind the spudcan is backfilled with at the site is not possible. It is common, for example, to
soil is correctly determined by a criterion based on this adopt characteristic values for strength. These are specified
flow-around mechanism, and not on the basis of collapse for instance in the Eurocode (BSI, 2004) as 5th percentile
of an unsupported cylindrical hole, which would indicate un- values, that is, only one in 20 of the strength measurements
realistically deep penetration before backfilling occurred. is expected to fall below the characteristic value. However,
there is no rational basis for expecting that characteristic
values with this definition lead to an appropriate worst-case
Deterministic calculations estimate of the spudcan penetrations, and confusingly, other
The usual approach is to carry out a deterministic standards adopt other definitions for characteristic values.
calculation for the loadpenetration curve. At every depth This is an important example of a problem where useful
the geometry of the spudcan is determined, for instance its progress can be made by adopting a rather more

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 5
Vertical load, V Vertical load, V

50%
75%
95%

5%

Depth

Depth
25%
Cumulative frequency: %

100 95%
75 75%

50 50%

25 25%
5%

Vertical load, V

Fig. 7. Monte Carlo simulations and their interpretation

sophisticated approach, in which the statistical variations of randomly from the probability distributions for each of the
the parameters governing the problem are addressed. variables. Such Monte Carlo simulations are common
throughout engineering. They are feasible in cases where
the underlying deterministic calculation is fast, so that very
Probabilistic calculations many simulations can be completed. This is the case for
A probabilistic calculation proceeds in a slightly different the simple bearing capacity calculation presented here. The
way. At the heart of the process is exactly the same cal- same approach could not, for instance, be made if the under-
culation as before. First, however, it is now acknowledged lying deterministic calculation relied on time-consuming
that there is not a uniquely known undrained strength at any finite-element analysis.
depth, but that the design profile comes from a probabilistic The Monte Carlo simulation typically involves hundreds
distribution of possible strengths, characterised by a mean of calculations. There are well-established techniques for
value and standard deviation at any depth. The fitting of reducing the number of calculations required to establish
design profiles using rigorous statistical methods, for instance the statistical variation of the results, but a nave approach
those based on Bayesian probability techniques, is an impor- is adopted here, simply drawing values of the parameters
tant topic that is now receiving some attention (e.g. Bienen appropriately from the statistical distributions. Because each
et al., 2010; Cao & Wang, 2013; Houlsby & Houlsby, 2013; calculation chooses slightly different parameter values from
Wang et al., 2014), but is not pursued further here. the distributions, they produce a family of predicted load
Second, it must be acknowledged that the positions of the penetration curves. They all tend to be roughly alike, but each
boundaries between different layers of soil will be uncertain. is slightly different. In particular the curves may cross each
For instance, the borehole used to determine the layering other as, for instance, different strengths may be chosen for
may not be precisely at the spudcan location. The measure- different soil horizons.
ment of depths within the borehole will itself be subject to The large number of curves may be processed in a variety of
some error. Again this uncertainty can be represented as a ways, but one intuitively simple approach is to take a slice
statistical distribution with mean and standard deviation. through them at a particular depth, and rank the predicted
Third, even the geometry of the spudcan itself is not values of the spudcan loads at this depth in ascending order
absolutely certain. In an ideal world every geometric detail in the form of a cumulative distribution curve, see Fig. 7.
would be known, but in the real world there will be some From that curve it is possible to determine, for instance, the
uncertainty. This uncertainty will of course be much less 5th, 25th, 50th, 75th and 95th percentile values at each depth.
than that attached to the soil strength, but there is some All the 5th percentile values at different depths can be joined
uncertainty nevertheless. to form a 5th percentile curve, and the process is repeated for
Finally the bearing capacity theory itself is subject the other percentile values. This procedure was adopted by
to uncertainty. Although accurate bearing capacity factors Houlsby (2010). The meaning of these curves is that, if the
for idealised problems can be determined, slightly different statistical parameters correctly define the true uncertainty, one
approaches with different assumptions lead to different would expect that about 50% of measurements would fall
values. There is of course further uncertainty attached to between the 25th and 75th percentiles, and 90% of measure-
the idealisation of the problem, for instance a spudcan which ments would fall between the 5th and 95th. This approach is
is actually in the form of an irregular hexagonal pyramid may therefore more rational than the worst-case line from the
be idealised as a circular conical footing. The uncertainty in deterministic calculation, which had no rigorous definition.
the applicability of the theory can be captured by assuming Only one in 20 measurements will fall below the 5th percentile,
some statistical distribution for the bearing capacity factor, but note that because all the parameters in the analysis are
Nc, and other factors in the calculation. treated statistically, this does not yield exactly the same result as
Many calculations are then carried out in which, rather simply using characteristic (5th percentile) values of strength in
than always using the mean value, values are chosen a deterministic analysis.

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
6 HOULSBY
Load, V: MN 100
0 50 100 150
0 90
5th percentile
80
25th percentile
Base case 70
5 75th percentile

Actual percentile
95th percentile 60
Bow leg
50
Port leg

10 Starboard leg
40
Depth, z: m

30

20
15
10

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

Fig. 9. Actual and predicted percentile positions of calculated


spudcan loads
25
uncertain parameter in the calculation. As part of the
Fig. 8. Case record A from the InSafeJIP database
InSafeJIP programme, the strength within each soil layer
was fitted by statistical methods, the coefficient of variation
Figure 8 shows an example of a prediction made on (CoV) values for the strength parameters were computed, and
this basis for one case from the InSafeJIP study: this was these were used to guide the values chosen for the predictions.
a joint industry programme addressing the safe installation It was not possible to apply this calibration process for the
of jack-up units (Osborne et al., 2011). The actual installation variability of all the parameters, so for some the uncertainty is
has to remain anonymous for commercial reasons. The con- simply based on a common sense approach.
tinuous lines on Fig. 8 show the various percentile predictions, The overall range of the predictions was checked by the
and the highlighted points show the actual measurements following procedure. There were (by coincidence) precisely
at the site. The difference in response of the three legs is 100 measurements of penetration in clays in the InSafeJIP
fairly typical of these observations, indicating probably slight database. Predictions were made for all these measurements
differences in soil conditions beneath the three legs. Where and the measured percentile position noted for each case.
more extreme differences are encountered, unacceptable tilting These percentiles were assembled in rank order (actual
of the jack-up can occur, although to a certain extent this percentiles). One could, for instance expect about five of the
can be managed by adjusting the distribution of the preload. 100 measurements to fall below the 5th percentile, 20 to be
This site was unusually well monitored: not only are the in the range 5th to 25th, and so on. By plotting the actual
penetrations of all three legs recorded, but these measurements percentile position against the predicted one (Fig. 9), it is
were made throughout the preloading process. Often the data possible to assess the quality of the overall fit. In probability
collected are less comprehensive, for instance the penetration theory this is called a PP plot, which can either be used
may be recorded just at light-ship load and at preload. to compare the distributions of two observations, or to
A geotechnical engineer must make an assessment of the compare a set of observations against a theory. If all the data
measured penetrations, essentially to reassure themselves that points lie on the diagonal line, then the chosen statistical
the observations are in accordance with the understanding variations match the real variability well, but if the dis-
of the site, so that subsequent operations can proceed with tributions are too broad or too narrow, then the data points
some confidence. In this case one could be reassured that would not lie on the diagonal. It can be seen that, with the
the observations do indeed match the predictions well. The values recommended in the InSafeJIP procedures (Osborne
penetrations are all slightly larger than the most probable, et al., 2011), the overall fit is quite good, giving confidence
but almost all of the values are within the 25th to 50th that the true variability has been captured with reasonable
percentile range, and the trend shown by the data is almost precision. Some of the detail is not reproduced, but that is
the same as shown by the predictions. At this particular site, only to be expected.
the underlying model of the ground behaviour appears to The aspect addressed now is how the measurements should
capture the real response in a satisfactory way. be interpreted if they fall near the edge of the distribution,
Such confidence depends, however, on knowing that which on the face of it means that the chosen soil model for
the chosen statistical variation does indeed capture the vari- the site may be incorrect. Fig. 10(a) shows schematically
ability of sites correctly. For instance, if too wide a distribution the case where the measurements are near the edge of the
of results was predicted, then almost every observation would distribution, around the 95th percentile, but the trend is
fall within the 25th to 75th percentile range, and the con- approximately as predicted. An experienced engineer might
clusion would always be that the response was as expected, be fairly comfortable with this situation, as of course about
which would be misleading. It is therefore vital that the one time in ten, one would expect to see values near the edge
statistical variations are properly calibrated, and this can only of the distribution. Fig. 10(b) shows a slightly more
be achieved by collecting the data from a large number of case complicated case. Each of the individual measurements is
records. In practice, however, the variation of the predictions is within the expected range, but the trend is incorrect. Houlsby
dominated by the variability of the strength, as this is the most (2010) suggested that this scenario could be treated in a

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 7
Load, V Load, V penetration calculation is highly non-linear, which results in
the non-Gaussian response. If the measurement is compared
directly with the predicted points, rather than with the fitted
ellipses, this event looks even more unlikely.
The separate probability curves for the two measurements,
Depth, z

Depth, z
as plotted at the top and side of the graph, are called
marginal distributions as they integrate out, or marginalise,
the effect of the other variable. Note that these alone do not
(a) (b) provide enough information for instance if the measure-
ment of penetration at preload had been the same distance
Fig. 10. Schematic diagrams for observations near limits of distri- the other side of the mean, which in the marginal distribution
bution with (a) trend correct and (b) trend incorrect (adapted from is equally likely, then the combination of the two measure-
Houlsby, 2010) ments would be solidly within the 50% contour of the
combined distribution.
Of course it is important to close the loop and determine
heuristic way, setting out a decision table based on where why the initial prediction was apparently inconsistent (in all
the measurements are, and whether or not they follow the probability) with the later observations, and if possible use
expected trend, Table 1. Although this may have some value, the observations to provide a better understanding of the
it is possible to take the statistical approach further to site. Here the problem has been approached in the following
understand the correlations as well as the individual values. way, adopting an approach that has been termed probabil-
This is illustrated by the case in Fig. 11, which shows the istic programming (Gordon et al., 2014; Wood et al., 2014),
predictions and the measured response from another case which implements Bayesian techniques. In essence the
from the InSafeJIP database. The measurements taken at this method proceeds as follows. As before, a large number of
site were for each of the three legs, at light-ship load, where it stochastic simulations of the prediction are made, with values
can be seen that the penetration is somewhat larger than for the various parameters drawn from prior distributions
predicted, and at full preload, where it is rather less than implied by the site investigation data. The predicted loads for
predicted. Neither individual observation is far away from the relevant penetrations are compared with the observed
the prediction, but the observed trend is somewhat different loads at those penetrations from the actual installation
from the predictions. Does this mean that the model for this data. From a large number of simulations those that best fit
site is incorrect? the observations (in a probabilistic manner) can be deter-
Figure 12 shows the results of 500 simulations using the mined within a probabilistic programming Bayesian frame-
statistically varying parameters. The predicted penetration at work. The resulting set of posterior simulations contains
preload is plotted against the penetration at light-ship load. parameters that will, assuming a correct simulator, provide a
The cloud of predicted points forms a roughly elliptical better estimate of the values at the site, and improved,
shape, demonstrating that the two values are closely cor- updated versions of the entire loadpenetration curve based
related: for instance, any individual prediction may use on these posterior simulations. This can be regarded as an
strength from the lower part of the range, and so will predict example of formalising the so-called observational method
larger penetrations at both light-ship load and at preload. (Peck, 1969) within a Bayesian probabilistic setting and
All of the points of course plot above the 1:1 line because extending it to include black-box simulation.
the penetration at preload is always larger than at light-ship Figure 13 shows the calculated loadpenetration curves
load. Also shown in Fig. 12 is the point that would be using the posterior values of the parameters. Of course the
obtained from a single deterministic calculation that would fit to the observed data is much better than in Fig. 11, which
be obtained from the experts fit to the strength data this of used the prior parameters. The fit to the data is still, of
course plots near the centre of the distribution of probabil- course, not exact, as the match is only made in a probabilistic
istic simulations. sense. Note in particular that at depths not coinciding with
The large dot on Fig. 12 shows the measured data, the measured data points there is greater uncertainty about
averaged for the three legs. The distribution curves for the the prediction.
prediction of light ship and preload penetrations are plotted Figure 14 shows the measured strength data at the
at the top and side of Fig. 12. Examining the distribution of site, from four different devices. The black (solid) line
predictions for light-ship penetration, at the top of the graph, shows the design strength profile which was originally fitted
it can be seen that the penetration is higher than expected. by an expert, and was used for the prior estimate of the
Examining the distribution of predictions at preload, at the loadpenetration curve. In the lower stratum the line fits
right-hand side, however, it is seen that the penetration is approximately in the middle of the range of the rather
smaller than expected. scattered data (the level of scatter is absolutely typical of this
The elliptical contours show the curves that would enclose sort of site). The chosen profile within the upper stratum lies
50, 90 and 98% of the predictions they are the two- approximately as an upper bound to the strengths measured
dimensional equivalent of the percentile lines shown in there, and on the face of it looks too high, but it should be
Fig. 8. The measured data point falls almost exactly on the borne in mind that when the expert was fitting this line,
98% curve, indicating that this sort of rare event would be she also had available other secondary information about
expected only one time in 50. That is perhaps sufficient to this site, which led her to believe that the measured strengths
conclude that the model for this site is in fact deficient: in all might be underestimates. However, concentrating just on the
probability some important feature of the soil behaviour has measurements of the minivane (green triangles), the only
not been captured. device that was used in both the upper and lower strata, it can
The ellipses are drawn on the assumption that the process be seen that the chosen design profile does not really capture
is Gaussian the very simplest form of statistical distri- the sudden increase of strength at the interface between the
bution. The fact that the actual cloud of points does not fit two layers.
the ellipses very well is evidence that this process is not The magenta (dashed) line in Fig. 14 is the most probable
Gaussian, which complicates the situation further. Although set of strength parameters as determined by the probabilistic
the variability of each input was treated as Gaussian, the programming approach. It has a slightly lower strength in the

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
8 HOULSBY
Table 1. Decision table based on absolute position and trend of measured data on spudcan penetrations compared to percentile predictions (from
Houlsby, 2010)

Percentile Data follow Example Diagnosis Suggested action


trend of
predictions*
Within 2575% Yes V Good prediction: model closely None required
fits observed behaviour

No V Moderate prediction: model Interrogate assumptions made in model


may not capture mechanism in attempt to identify satisfactory
correctly, in which case fit explanation of discrepancy
may be coincidental

Outside 2575%, Yes V Moderate prediction: model Attempt to identify reason for
but within captures essential trend of systematic error
595% data, but may contain
systematic error

No V Poor prediction: important Need to identify reasons for failure to


features of mechanisms may model important mechanisms and
not be captured establish improved model

Outside 595% Yes V Poor prediction: model may Essential to identify source of
capture trend of data, but systematic error to explain
large systematic error discrepancy

No V Prediction fails to capture Potentially dangerous as the


observed response mechanisms and values assumed in
the predictive model are clearly
inappropriate. Further action
required to understand site
conditions
z

*Note: if only one measured point then no assumption about trend of data can be made.

upper layer (closer to the minivane results), and a very this too is reasonably consistent with the observed soil data
slightly higher rate of increase of strength in the lower layer. (as one would expect it to be), it is less consistent with the
To an experienced engineer the modified profile appears observations than the calculations using the updated strength
every bit as plausible as the original one, and maybe it profile.
suggests that the expert should simply have taken the strength The engineer can, however, be satisfied that the site has
values in the upper layer at face value. now been properly understood. Perhaps more importantly,
Adopting the new profile, and once more carrying out the jack-up installation itself has been used as a gigantic site
a Monte Carlo simulation for the calculated penetration, investigation tool, and has provided a modified strength
Fig. 15 shows the joint probability plot. The measured data profile in which one can have more confidence. The modified
point now fits between the 50 and 90% boundaries, and so profile can be used in any future calculations about the per-
the measured data may reasonably be considered as con- formance of the jack-up at this site, with greater confidence,
sistent with the modified model of the strength profile. In this because the profile combines both the initial site investigation
case the issue has therefore been resolved a small change to data and the observed footing behaviour in a rational way.
the strength profile, still perfectly consistent with the original This is an example of the so-called observational method,
data, is sufficient to explain the observations. Fig. 15 also Peck (1969): large-scale field observations allow improve-
shows the deterministic prediction from the expert: although ment in the level of understanding of ground conditions, and

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 9
therefore allow predictions of future response to be improved, inherently variable properties, which are of necessity only
which is the subject of the next section. sampled in an imperfect manner. The rational treatment of
soils data and measurements requires a statistical and prob-
abilistic approach. And yet the level of knowledge of statistics
Conclusions from part 1 and probability theory of most geotechnical engineers is
Jack-up foundations, because they cause continuous elementary. A deeper understanding of more advanced statis-
failure of the soil during penetration, provide almost tical and probability techniques would be of great value to
unique data from large-scale foundations which can be professionals in geotechnical engineering.
used to verify bearing capacity theories.
Making the link with other disciplines, a rational com-
parison of predictions and measured data during installation PART 2: PERFORMANCE OF JACK-UP UNITS AND
of a jack-up can best be understood in the context of prob- OTHER SHALLOW FOUNDATIONS
ability theory. This conclusion carries over to many other This section addresses the performance of jack-up units
branches of soil mechanics. Soils are materials with under extreme loading conditions. The important interaction
at this stage is with structural engineers, who need to be
Load, V: MN able to analyse the forces in the structure, and for rather
0 20 40 60 80 100 120 flexible structures such as jack-up units the structural forces
0
depend critically on the foundation response. In particular
5th percentile the response of jack-ups depends on the momentrotation
5 25th percentile response of the spudcans, a problem referred to in the indu-
50th percentile stry simply as fixity. The key question asked by the
75th percentile structural engineers is whether the spudcans can be treated
10
95th percentile as pinned, fixed, or somewhere in between as far as the
Bow leg momentrotation response is concerned.
15 Port leg For simplicity, first idealise the rig as a rather stiff hull with
Starboard leg flexible legs, simplifying the problem to planar loading and
Depth, z: m

20 considering the horizontal forces as if they act at the hull


level. The hull is idealised as rigid, and the tops of the legs are
assumed to be fully restrained from rotation by the leg guides.
25 A more sophisticated analysis would recognise the flexibility
of the hull and the fact that that rack-and-pinion leg guides
30 exhibit a complex non-linear response with both rotational
and translational tolerances. Fig. 16(a) shows the reactions at
the footings if the conservative assumption that the legs are
35
pinned to the seabed and free to rotate is made. The reactions
are shown for the case where two of the three legs are to the
40 windward side, and one to the leeward side. The horizontal
load on the rig decreases the combined vertical reactions on
Fig. 11. Case record B from the InSafeJIP database the windward legs by HL/w, and increases the vertical

40
Predictions
Observed
35 50%
90%
98%
Penetration at preload: m

30 Expert's prediction

25

20

15

10
10 15 20 25
Penetration at light ship load: m

Fig. 12. Joint distribution for penetrations at light-ship load and preload for InSafeJIP case record B

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
10 HOULSBY
Load, V: MN Treating the rig as a single-degree-of-freedom system,
0 20 40 60 80 100
the firstqnatural frequency ofq
the rig is doubled, from
0
n 3 EI =mL3 to n 6 EI =mL3 (assuming the
mass is concentrated in the hull). Importantly, for almost
50 all sea states and rig dimensions, this takes the resonance
further away from the main wave excitation frequencies,
100 see Fig. 17, so that dynamic amplification effects are
reduced, and therefore the actual displacements are
150 reduced by fixity by more than a factor of 4.
Depth, z: m

A more sophisticated structural analysis of the rig, account-


200
ing for instance for deck flexibility, a realistic distribution
of the loads, shear deflections of the legs, P effects, and a
250 more realistic mass distribution will alter the above con-
5% clusions slightly, but the broad conclusions are the same.
300 25% The structural design of the rig therefore depends on
50%
how stiff a moment response the spudcan will offer a classic
75%
350 problem of soilstructure interaction. Ideally a structural
95%
Observations
engineer would wish the fixity to be represented as an elastic
spring, which would represent some degree of fixity between
400
fully pinned and fully fixed. This would have the advantage
Fig. 13. Posterior distribution for loadpenetration curves using that linear analysis of the structure could be retained, em-
probabilistic programming approach ploying for example modal analysis techniques. Spudcan
foundations are therefore often modelled by rotational
springs (as shown for the windward leg in Fig. 18), with
Undrained strength, su: kPa horizontal and vertical motion restrained. Sometimes hori-
0 20 40 60 80 100 120 zontal and vertical springs are included too (as shown for the
0 leeward leg in Fig. 18). However, as is well known to geo-
UU
Minivane
technical engineers, the soil response is more complicated,
5
Torvane and cannot just be represented by elastic springs. In the
10
Pocket penetrometer following, the author pursues how the foundation response
Expert's fit can be realistically described in a way that is intelligible and
Probabilistic programming
15 useful to a structural engineer. The method proposed exploits
plasticity theory, which provides the language and math-
Depth, z: m

20 ematics by which geotechnical engineers can readily make


25
knowledge of the response of shallow foundations accessible
to structural engineers.
30 In the following the general relationships between the
load resultants (V, M, H ) on a foundation and the correspond-
35 ing displacements (w, , u) are explored, see Fig. 19. This
approach using force resultants rather than detailed analysis of
40
stresses and strains within the soil is exactly analogous, for
45 instance, to the use of tension, bending moment and shear
force in a beam, rather than the detailed stress distributions.
Fig. 14. Strength data for InSafeJIP case record B

Elastic response
reaction on the leeward one by HL/w. Alternatively, assuming Although the soil is not elastic, it is useful initially to
that the large spudcans on the legs completely restrain the pursue how the elastic response can be represented. Consider
rotation, the reactions are as shown in Fig. 16(b). Note that, first a circular footing on the soil surface. The relationship
compared to the pinned case, the changes in vertical reaction between the loads and displacements of a footing on an
due to the horizontal load are exactly halved to HL/2w. elastic soil can be expressed in the form of a matrix equation.
Based on this simplified analysis of the unit, there are It is convenient to convert this equation to non-dimensional
several benefits of fixity of the foundation. form by dividing the displacements by the footing diameter,
and normalising the forces by the diameter and the soil shear
modulus
If the rotation is fully restrained, as noted above, the 2 3 2 32 3
changes of vertical loads on the spudcans are reduced by V =GD2 K1 0 0 w=D
a factor of 2. This effect is particularly important for the 4 M=GD3 5 4 0 K 2 K 4 54 5 3
leeward footing, which is the most heavily loaded. H=GD2 0 K4 K3 u=D
The maximum bending moment in each leg is halved
from HL/3 to HL/6. However, the moment at the spudcan It is easy to show that four of the nine terms in the matrix
connection is of course increased (it is close to zero in the must be zero, but there are three diagonal terms and two
pinned case and HL/6 in the fixed case), and this has off-diagonal terms in the matrix. The diagonal terms are
implications for fatigue design of the legspudcan easily recognised as representing the vertical, moment and
connection. horizontal stiffnesses. The significance of the off-diagonal
The quasi-static lateral displacement of the deck is terms is less obvious, but they indicate that, for instance, if a
reduced by a factor of 4 by full fixity from HL 3/9EI horizontal load is applied to the footing it does not just
to HL 3/36EI. translate horizontally, but it may rotate too. Vice versa, if a

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 11

40
Predictions
Observed
35 50%
90%
98%
Penetration at preload: m

30 Expert's prediction

25

20

15

10
10 15 20 25
Penetration at light ship load: m

Fig. 15. Joint distribution for penetrations at light-ship load and preload for InSafeJIP case record B: modified strength profile

w H

H H V
V V

HL/3 HL/6
2H/3 H/3 2H/3 H/3

2V HL V + HL 2V HL V + HL
3 w 3 w 3 2w 3 2w 2k k

(a) (b)

Fig. 16. Simplified view of loads on a jack-up unit (two legs to


windward) with (a) pinned spudcans and (b) fixed spudcans
Fig. 18. Spudcan foundations modelled as equivalent springs: wind-
5 ward footing shows rotational spring only, leeward footing shows
Dynamic amplification factor (DAF)

additional vertical and horizontal springs


4

Increased
3 natural V
frequency
M
2 H

u
1 w

0
0 05 10 15 20 25
Excitation frequency/natural frequency
Fig. 19. Conventions for load and displacement of foundation
Fig. 17. Effect of increased natural frequency on dynamic amplifica-
tion factor

are two solutions: K1 2/(1  ) for a smooth footing and


moment is applied, the foundation will translate horizontally 2  [log (3  4)]/(1  2) for the rough case (both of these
as well as rotate. converge to K1 4 for the limiting case of an incompressible
In the standard literature the three diagonal terms material). For the rotational case the solution available is
in equation (3) are reported. For vertical loading there K2 1/[3(1  )], and strictly this is only applicable for a

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
12 HOULSBY
V
M
H

V
H M
V
M
H

(a) (b) (c)

Fig. 20. Possible locations for load reference point: (a) at ground surface; (b) at spudcan widest point; (c) at spudcan tip

Table 2. Stiffness factors for elastic analysis of embedded footings (from Ngo-Tran, 1996)

Embedment Cone angle, : degrees K1 K2 K3 K4

= 02 = 049 = 02 = 049 = 02 = 049 = 02 = 049

zD/R = 00 180 54058 81971 46517 55541 38161 55627 05734 00148
150 55643 82998 49739 59450 40661 57684 07589 00737
120 57790 83429 54976 66326 46563 60640 11966 04862
zD/R = 05 180 59509 88098 61173 72258 45860 65826 08425 04020
150 60893 88102 64086 75634 48308 65728 10197 04504
120 62850 88438 68879 81766 54227 68440 14398 08166
zD/R = 10 180 64950 92537 64992 75899 50156 71627 09360 05691
150 66361 92510 68183 79544 52733 71301 11228 06086
120 68383 92830 73376 86101 58902 73744 15638 09762
zD/R = 20 180 70817 99857 67742 77012 51903 76739 09786 07327
150 72633 100702 71262 81210 55033 78269 11645 07070
120 74950 99811 76962 91741 61621 78945 16233 10750

08 For a spudcan of realistic geometry there are some


06
further important considerations. First, it is vital that the
load reference point (LRP, the point at which loads are
Yield
04 considered to act, and the reference point for displacements)
02
is clearly defined (see Bell, 1991; Dean et al., 1998). Practice
varies, and for instance the LRP may be taken at the mudline
M/su D3

0 (Fig. 20(a)), or at the level of the widest part of the


spudcan (Fig. 20(b)) or at the spudcan tip (Fig. 20(c)).
02
Test H. 100 s. 32 - 05
All have a certain justification and have been used in
04 130 spudcan practice.
V/su A = 21 The selection of the LRP, however, affects the values of the
06
h/d = 05 stiffness factors. If the LRP is moved downwards by some
08 distance z* fD, then some of the stiffness factors do not
01 0 01 02 change, but some do, as noted by Bell (1991).
: rad
2 3 2 32 3
Fig. 21. Momentrotation test on model spudcan (adapted from V *=GR2 K1 0 0 w*=R
de Santa Maria, 1988) 6 7
6M*=GR3 7 6 76 7
5 4 0 K 2 2 fK 4 f K 3 K 4 fK 3 54 * 5
2
4
H*=GR2 0 K 4 fK 3 K3 u*=R

smooth footing, although it is often applied otherwise. For 4


horizontal loading there are two commonly available sol-
utions, K3 4/(2  ) and K3 16(1  )/(7  8) (both of This further raises the possibility of defining the LRP such
these converge to K3 8/3 for the limiting case of an that the off-diagonal term is zero, the elastic metacentre
incompressible material, and the numerical differences are such that K4 fK3 0, or z* DK4/K3. Such an approach
small for other cases); they correspond to subtly different can be convenient in some applications.
boundary conditions on the footing in terms of the vertical Clearly if the correct stiffness factors are to be used, the
displacements. However, in much of the literature the geotechnical engineer must use the same LRP (and sign con-
coupling term K4 has simply been ignored. For instance, no vention) as the structural engineer, otherwise the calculation
mention of this term is made in the commonly used API is meaningless. This seemingly simple matter of communi-
standard for offshore foundations (API, 2010). cation is not always addressed as it should be.
However, there is no rational basis for ignoring the off- Studies of appropriate elastic stiffness factors are reported
diagonal term in the matrix, and its physical significance needs by Bell (1991), Ngo-Tran (1996) and Doherty & Deeks
to be properly assessed. The result is that many conventional (2003), dealing with a variety of geometries of foundation,
calculations that ignore this interaction are incorrect. including parametric studies of the effects of the angle of the

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 13

H H
g g
din Bearing din Bearing
Sli Sli
capacity capacity

Vm V Vm V
(a) (b)

Fig. 22. Hypothetical yield surfaces for vertical and horizontal loading of a flat plate on sand

cone of the spudcan and the depth of embedment, and


provided these are used with care, they provide a sensible 10
modelling of the elastic behaviour. Example values of the stiff- 09
ness factor (taking the load reference point as in Fig. 20(b))
are shown in Table 2 (Ngo-Tran, 1996). Alternatively, with 08
modern finite-element analysis it is relatively straightforward 07
to calculate the various stiffness coefficients directly for a
06
given geometry of spudcan.

V/V0
However, as is widely recognised, soils are not linear 05
elastic except for within a very small range of strains. Fig. 21 04
shows an early test by de Santa Maria (1988) measuring
03
the momentrotation response of a model spudcan on clay
under carefully controlled conditions. The foundation exhi- 02
bits highly non-linear behaviour, with an open hysteresis 01
loop on unloading and reloading. Note that the results are
presented in terms of dimensionless quantities, an important 002 004 006 008 010 012
detail that makes them more readily compared with other M/V0I
results, and applied to field problems.
The importance of addressing the whole of the non-linear Fig. 23. Two-dimensional failure surface (from Roscoe & Schofield,
response is revisited below, but for the time being the curve 1957)
is idealised as if it were elasticplastic with a single yield
point (idealised as the green line in Fig. 21), and the plastic
behaviour is investigated. However, as the foundation is H
subjected to vertical loads and horizontal loads as well as
moments, it is the combination of these loads that causes M
yield. It is necessary therefore to determine the shape of a B
yield surface in (V, M, H ) space. This yield surface limits the
safe loads that can be applied to the foundation, resulting
only in (relatively small) pseudo-elastic deformations.
Consider first just a flat footing with a vertical and
horizontal load. It is tempting to think that failure
will occur either by sliding, at some ratio of horizontal to
vertical load, or by bearing capacity failure at some limiting
vertical load. The shaded area in Fig. 22(a) would show V*
V
the range of allowable loads in this case. (Note that it is
common practice in this application to plot the vertical
load on the horizontal axis, and the horizontal load on the Fig. 24. Three-dimensional failure surface (from Butterfield & Ticof,
vertical axis.) In fact there is a very strong interaction between 1979)
the two failure mechanisms, and any application of horizon-
tal load dramatically reduces the vertical capacity. A more
realistic yield surface is shown in Fig. 21(b). Much research combination of loads that falls inside the surface will be safe,
has been devoted to determining the shape of this surface, whereas large plastic deformations will occur once the
but with the additional complication that a moment is surface is reached: the surface represents the generalisation
applied too. of the idea of bearing capacity for a vertical load, to the more
This is an old problem: Roscoe & Schofield (1957), Fig. 23, complicated case with multiple load components. There
first suggested interaction surfaces for combinations of ver- is an important distinction between a failure surface, which
tical load and moment on shallow foundations; the appli- represents the envelope of all attainable loads, and a yield
cation was to foundations for steel portal frames, but the surface, which encloses just those loads that are accessible by
concept is the same as for offshore foundations. (In their plot solely elastic deformation at a given penetration. If work
the vertical load was in the vertical direction.) Butterfield hardening of the yield surface occurs, accompanied in this
& Ticof (1979) presented the first three-dimensional case by additional vertical penetration, then the failure
yield surface for combined vertical, moment and horizontal surface differs from the yield surface. In the following, it is
loading, based on a series of tests on sand. The surface they entirely the yield surface that is considered.
found has a characteristic rugby-ball shape, with rather This problem has been investigated using the apparatus
pointed ends at high and low loads, Fig. 24. shown in Fig. 25, designed at Oxford by Martin (1994) and
Experiments designed to determine the shape of the further developed by Byrne (2000). The apparatus uses
three-dimensional yield surface are described below. Any stepper motors to drive a model foundation in the vertical

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
14 HOULSBY
direction, in the horizontal direction and to rotate it about a
horizontal axis to apply moment. In this way any planar
movement can be applied, and is measured with a system of
linear variable differential transducers (LVDTs). The corre-
sponding loads are measured using a Cambridge load cell
positioned just above the footing. The apparatus is computer
controlled, so that any combination of loads and displace-
ment can be applied, including some quite complex load
paths. It has been used for many different types of test, for
instance tests in which just a vertical load is applied, or tests
with some constant ratio between vertical and horizontal
displacement, Fig. 26. In the following, however, the focus is
on so-called swipe tests. Tests of this sort were carried out
by Tan (1990) in Cambridge, but have since been adopted
successfully by a number of groups. In a swipe test the footing
is first pushed into the ground to a given vertical load, and
then is moved sideways, keeping the vertical displacement
constant. As the lateral displacement is applied, the hori-
zontal load builds up and the vertical load reduces. It can be
shown theoretically that the load path followed in the test
follows very closely (but not quite exactly) the shape of the
yield surface, see discussion by Martin (1994). A single
test therefore provides the shape of a whole section of the
surface.
By carrying out different types of swipe tests the shape of
the whole yield surface can be explored. For instance,
translating the foundation horizontally principally applies
horizontal load, whereas rotation results mainly in moment
load. Combinations of sliding and rotation are used to
explore the whole yield surface.
Fig. 25. Apparatus for combined loading of model footings Figure 27 shows the results from a swipe test (GG03)
on sand (Gottardi et al., 1999). The model footing is first
loaded vertically to 1600 N, and then translated horizontally.
The test path in (V, H ) space follows a roughly parabolic
track, with the vertical load reducing, and the horizontal
Constant V
load first increasing and then reducing after passing through
a peak. The track of the test follows closely the shape of the
yield surface, which is approximately parabolic in this plane.
Fig. 27 also shows another test GG07, in which the footing
M/2R or H

Constant w
was unloaded to 200 N after the initial loading to 1600 N.
This test is analogous to a test on an overconsolidated
ent material. It confirms the shape of the yield surface at low
lacem loads.
l disp
Radia
Figure 28 shows the results of an equivalent pair of tests in
which the footing is first loaded vertically and then rotated
Vertical loading V rather than translated horizontally, but again at constant
vertical load. During the swipe test the moment increases and
Fig. 26. Load paths attainable with apparatus shown in Fig. 25 (from then decreases, and just as in the horizontal swipe it defines
Gottardi et al., 1999) the shape of the yield surface.

250 250

200 200
GG03
GG03
150 150
H: N

100 100
GG07

50 GG07 50

0 0
0 20 40 60 0 200 400 600 800 1000 1200 1400 1600
u: mm V: N
(a) (b)

Fig. 27. Horizontal swipe tests on sand (from Gottardi et al., 1999)

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 15
200 200

GG04
150 150
GG04
GG08
M/2R: N

M/2R: N
GG08
100 100

50 50

0 0
0 20 40 60 80 0 200 400 600 800 1000 1200 1400 1600
2R: mm V: N
(a) (b)

Fig. 28. Moment swipe tests on sand (from Gottardi et al., 1999)

120 hn
015
Best fit
GG03
100 010 GG04
GG05
GG06
005
GG28
080
GG29
0 mn GG07
015 010 005 0 005 010 015 GG08
Best fit
q 060 GG03 GG10
005
GG04 GG12
GG05
040 GG06 010
GG28
GG29
015
GG07
020
GG08 Fig. 30. Representation of section of yield surface perpendicular to
GG10 vertical load axis (from Gottardi et al., 1999)
GG12
0
0 020 040 060 080 100
v how the loads interact in a way that cannot be separated. The
eccentricity of the surface demonstrates that (for instance)
Fig. 29. Unified representation of swipe tests on sand (from Gottardi a clockwise moment combined with a horizontal force to
et al., 1999) the right will have a different effect from an anticlockwise
moment combined with the same horizontal force. This
asymmetry is not recognised by conventional approaches to
For further analysis it is useful to define dimensionless combined loading of foundations using the effective area and
quantities normalising the loads with respect to the inclination factor approach (Brinch Hansen, 1970; Vesic,
maximum load V0 that has been applied to the footing, 1975).
v V/V0, h H/V0 and in addition normalising the Combining the results of several swipe tests, a complete
moments by dividing by the footing diameter m M/DV0. three-dimensional yield surface can be determined that
Normalising the loads in this way allows swipe tests represents the combination of loads that can safely be
with different modes of deformation to be combined applied to the spudcan foundation. Within the yield surface
ontoq
a single plot by defining a generalised load the displacements are small, but once the surface is reached
    there are much larger plastic deformations. A reasonable
q m2 =m20 h2 =h20  2amh=m0 h0 that represents approximation to the surface is given by the expression
the combination of horizontal and moment loads. Note
 2  2  
that this expression includes three parameters m0, h0 and a m h mh
that are used in the fitting of the yield surface dimensions. f 2a  4v1  v2 0 5
m0 h0 m0 h0
The shape of the yield surface is then closely fitted by a
parabola in the (v, q) plane, Fig. 29. which represents a parabolic section in the V, M and V,
Figure 30 shows a section of the yield surface on a plane at H planes and an eccentric ellipse in the M, H plane. More
right angles to the vertical load, eliminating the influence of sophisticated expressions can be used to achieve a closer fit
vertical load (assuming the parabolic shape determined from to the data. The approach allows complete models of
Fig. 29 by introducing the definitions mn m/[4v(1  v)] and foundation behaviour including both the elastic and plastic
hn h/[4v(1  v)]). This plot shows that the shape of the yield behaviour for any load or displacement combination. These
surface for combinations of vertical and horizontal loads is foundation macro-models can then be integrated as part of
roughly elliptical in this plane. Importantly though, the a numerical analysis of a structure. The mathematical details
principal axes of the ellipse are rotated, and not aligned with of some particular models are given, for example, by Martin
the horizontal and moment axes. This is a further example of & Houlsby (2001) and Houlsby & Cassidy (2002).

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
16 HOULSBY
In the analysis of dynamically sensitive structures the and hull are replaced by linear elements. At the bottom of the
importance of the interface between geotechnical engineers legs various assumptions can be made about the way the
and their structural engineering colleagues cannot be over- foundations interact with the ground. Lateral loads from
emphasised. To model the foundation realistically, structural waves can be applied to nodes on the legs of the rig. The
engineers must accept the non-linearity of soil (however model is used to investigate the behaviour as a large wave
inconvenient it may be) and realise that soil cannot just be passes through the jack-up (Cassidy et al., 2001).
represented by linear springs. Geotechnical engineers must Figure 32(a) shows the wave elevations on the upstream
be prepared to cast their soil mechanics models in a way that and downstream legs, plotted against time. The wave itself
is understandable to structural engineers, compatible with is represented by constrained NewWave theory, which
their analyses and not simply reliant on arcane ad hoc realistically captures the shapes of real waves: NewWave
procedures. The appropriate language for this communi- (Tromans et al., 1991) represents well the average shape of an
cation is plasticity theory. extreme wave, but does not capture a realistic pattern of waves
Figure 31 shows a bar stool model of a jack-up rig: a occurring beforehand, which may be important for the
simple two-dimensional structural model in which the legs behaviour of a dynamically sensitive structure. Constrained
NewWave (Taylor et al., 1995) is an improvement in which
the extreme wave is embedded within a realistic background
history of wave elevations. The red (thick) curve in Fig. 32(a)
shows the wave elevation on the upwind legs of the jack-up,
showing an extreme wave that has been focused so that it
reaches its maximum height of 15 m as it passes the legs. The
blue (thin) curve shows it reaching the downwind leg about
3 s later, and with a slightly reduced height.
Figure 32(b) shows the horizontal movement of the deck,
making three different assumptions about the foundation
behaviour. The blue (thin) curve shows the case where the
foundations are pinned, with a maximum displacement of
about 14 m. At the opposite extreme the green (thick) curve
shows the result if the foundations are represented by stiff
elastic springs: the deck displacements are reduced by a
factor of about 35 to just over 04 m. Recalling the expected
factor of 4 if the foundations were fully fixed, it is seen that
the elastic foundations provide a high degree of fixity. Notice
Fig. 31. Jack-up unit and its simple idealisation for numerical too that the natural period of the rig reduces from about 14 s
analysis to about 6 s.

15

Upwave leg Downwave leg


10
Surface elevation: m

10

15
0 20 40 60 80 100 120
Time: s
(a)

16
Pinned
Horizontal hull displacement: m

12 Model C
Linear springs
08

04

04

08

12
0 20 40 60 80 100 120
Time: s
(b)

Fig. 32. (a) Constrained NewWave elevations. (b) Response of jack-up unit to constrained NewWave excitation

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 17

2
H2

D
M2

M3 H3

Fig. 33. General three-dimensional loading on a foundation

In between, the red (thick) curve in Fig. 32(b) shows


the response if the full plasticity model is implemented. The
response is very much like the elastic one, as for most of the
time the foundations do indeed remain elastic. However,
as the largest wave passes there is a permanent movement of
the deck of about 015 m a feature that could not be
captured by the elastic model.
Two important extensions of the above modelling
approach are now addressed. First, so far the loading has
Fig. 34. Apparatus for applying general three-dimensional loading to
been assumed to be planar in two dimensions. In reality a model foundation
three-dimensional analysis is required, and so the behaviour
of foundations subjected to more general loading, Fig. 33,
needs to be understood. Fig. 34 shows a sophisticated testing
The second important extension derives from the well-
rig, developed by Byrne & Houlsby (2005), and unique
known fact that soils exhibit non-linearity at small strain.
worldwide for this application. It uses six actuators to apply,
Fig. 21 shows that the idealisation of elasticity, followed by a
under computer control, any general combination of loads
sudden yield, is an over-simplification of the real footing
or displacements to a model foundation, and uses LVDTs
behaviour. A well-known way of presenting this non-linearity
and a six-axis load cell to measure the corresponding
is the S-shaped curve in a plot of secant stiffness against
displacements and loads.
strain level. Fig. 36 shows the results from some large-scale
As a result, tests are possible with the full range of loads
footing tests (in this case on caisson foundations 3 m in
in three dimensions, Bienen et al. (2006), and these can be
diameter, Houlsby et al. (2006)), showing how the deduced
used to derive the shape of the yield surface under general
secant shear stiffness for the soil gradually changes with the
loading. A wider range of swipe tests has been used to
amplitude of rotation. A more sophisticated model should
identify the yield surface, see Fig. 35. Each of these curves
capture this reduction of stiffness, which is coupled to the
can be imagined as slices through the shape of the yield
gradual development of plastic strain.
surface, which now must be expressed in a six-dimensional
This type of non-linearity can be captured by using a
load space. The general expression for the yield surface
multi-surface model conceptually very similar to the brick
becomes more complicated as it involves six rather than
model described by Simpson (1992), but expressed in terms
three variables, but conceptually it is essentially the same
of force resultants and displacements rather than stresses
as equation (5). Importantly, much of the structure of the
and strains. Rather than a single yield surface, the model
equation can be deduced from the symmetry of the foun-
now employs many yield surfaces, illustrated conceptually
dation about a vertical axis, and only the term involving
in Fig. 37. Each is a different size, but for simplicity they
the torsion about a vertical axis is truly new in the more
are assumed all to be of the same shape. Each surface con-
general equation. In equation (6) below, the more general
tributes a small amount of plastic strain. As the preloading
form of the section in the vertical direction has also been
and unloading is applied, the surfaces are dragged behind
included; this is expressed through power functions, first
the loading point, so that at the end of preloading the
introduced in this context by Nova & Montrasio (1991).
surfaces are distributed as in Fig. 37(a). Later, when a storm
     2  2 loading is applied, the surfaces are again dragged with the
H2 2 H3 2 M2 M3
f load point, and as each surface is encountered the tangent
h0 V 0 h0 V 0 m0 DV 0 m0 DV 0 stiffness gradually decreases. At the point of maximum
   2 loading the load point on a windward leg, and the associated
H 3M 2  H 2M 3 Q
 2a positions of the surfaces, would be as illustrated in Fig. 37(b),
h0 V 0 m0 DV 0 q0 DV 0
showing the reduction of vertical load on the windward leg as
 21  
V V 22 the moment is applied. Complete models of this sort are not
 12 1 0 trivial, but can be expressed in terms of work hardening
V0 V0
plasticity theory (Nguyen-Sy, 2005; Nguyen-Sy & Houlsby,
6 2005).

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
18 HOULSBY
014

012
Horizontal tests

010

H/V0, M/2RV0,Q/2RV0
008

Rotational tests
006

004

002
Twisting tests
0
0 01 02 03 04 05 06 07 08 09 10
V/V0

Fig. 35. Swipe tests under general loading conditions

100
M
90 Jacking
80 SEMV
Hyperbolic curve fit
70
60
G: MPa

50 V
40
30
20 (a)
10
0
M
0000001 000001 00001 0001 001 01
: rad

Fig. 36. Change of rotational stiffness of a foundation as a function of


rotation amplitude (SEMV, structural eccentric mass vibrator; from
Houlsby et al., 2006) V

Figure 38 shows some results testing a simplified version


of the multi-surface model (Byrne et al., 2002a; Byrne &
Houlsby, 2003). Fig. 38(a) shows the results of an experiment (b)
with gradually increasing amplitude of cycles, and Fig. 38(b)
the modelling using the multi-surface approach. This pattern Fig. 37. Yield surfaces in multi-surface model for: (a) preloading; (b)
of loading is particularly challenging to model accurately, storm loading of windward footing
but it can be seen that the multi-surface model is able to cap-
ture the complete non-linear response of the real foundation
with remarkable accuracy. With such sophisticated models behaviour of the jack-up is to be captured then (at the very
the behaviour of spudcans can be modelled quite realistically. least) it is necessary to use quite sophisticated modelling of
Fig. 39 shows the results when such a model is used to both the wave kinematics and of the foundation: both involve
represent the foundations of a jack-up during the application important non-linearities that can only be captured in
of an isolated extreme wave, this time modelled just by numerical analysis.
NewWave, Cassidy et al. (2004). Fig. 39(a) shows the wave Perhaps most importantly, Fig. 41 shows the moment
elevation at the upwind and downwind legs of the jack-up. rotation behaviour of one of the footings. The smaller cycles
Fig. 39(b) shows the corresponding deck displacement: a are almost elastic, whereas in the larger cycles plasticity
maximum displacement of about 023 m occurs, and there is develops. This is extremely important: for low-amplitude
very little permanent displacement after the wave has passed. cycling there is high stiffness, but relatively low damping. As
Fig. 40 shows the load paths followed by the upwind and the cycle amplitude increases the stiffness drops, which is
downwind legs: the upwind leg shows a decrease in vertical disadvantageous as far as the structural response is con-
load as the moment increases, and the downwind leg shows a cerned, but this is offset by the fact that there is energy
corresponding increase in vertical load. At some points absorption through plastic deformation, which is beneficial.
during the wave cycle there are reversals of moments. Note The overall result is that a realistically modelled foundation
also that, because of the non-linearities in the problem, the contributes significant fixity, which plays a vital role in the
actual load paths are more complicated than the lines that performance of the jack-up. Note that only properly formul-
would be predicted by a simplified elastic analysis. If the ated plasticity models offer the realistic combination of high

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 19
100 stiffness at small deformation and energy dissipation at larger
deformations: simpler non-linear elasticity models cannot
capture this behaviour.
Moment load, M/D: N

50 The calibration of such models has principally been


carried out using the results of 1g laboratory tests, and
of course important questions arise regarding the scaling
0
of both strength and stiffness parameters to field scale. Kelly
et al. (2006a) discuss how 1g laboratory test results can be
50
applied to field scale. Centrifuge testing has a role to play in
scaling the results, but it is important to recognise that
centrifuges only allow the scaling of stress level the tests
100 are still conducted at very small scale. The author finds the
10 05 0 05 10 routine reporting of centrifuge tests as if they were at
Rotational displacement, D: mm prototype scale somewhat misleading.
(a) There is, unfortunately, a paucity of case records that
can be used to verify the predicted behaviour at large scale.
100 Fig. 42, shows the measured deck deflection for one case
of a jack-up under extreme wave loading (McCarron &
Broussard, 1992). It can be seen that the character of the
Moment load, M/D: N

50 response is very much as described by the numerical model.


McCarron & Broussard (1992) had some limited success
modelling this case history using a simplified model for
0
the rotational stiffness of the spudcans which employed
vertical springs, with a no-tension cut-off, aligned with
50
the leg chords. Although this modelled some aspects of the
non-linearity, it was unable to capture realistically the
hysteresis in the momentrotation response and permanent
100 rotation of the spudcans.
10 05 0 05 10
Rotational displacement, D: mm
(b)
Onshore application: historic towers
Fig. 38. (a) Experimental measurements of momentrotation There are other applications of these methods. Pisan et al.
response of footing under increasing amplitude of cycling. (2014) used this approach to aid in understanding the
(b) Modelling using a multi-surface model
behaviour of leaning historic towers. Methods based on
yield surfaces for foundation loading do not necessarily

16
Windward leg
12
Leeward leg
8
Wave elevation: m

4
0
4
8
12
16
60 40 20 0 20 40 60
Time: s
(a)

030
Horizontal hull displacements: m

025
020
015
010
005
0
005
010
60 40 20 0 20 40 60
Time: s
(b)

Fig. 39. (a) NewWave surface elevation at windward and leeward legs. (b) Horizontal deck movement of jack-up excited by NewWave loading

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
20 HOULSBY
10 012

8 010
6 A
008
Moment: MN m

M/2RV0
4
006 B
2 C
Windward leg 004
0
Leeward leg
2 002

4 0
0 01 02 03 04 05 06 07 08 09 10
6 V/V0
90 95 100 105
Vertical load: MN
Fig. 43. Estimated load path for the Tower of Pisa remedial works
Fig. 40. Load paths for windward and leeward legs of jack-up
subjected to NewWave loading The initial position for the load point for the Tower of Pisa is
at A. The moment is due to the eccentric loading because of
the inclination of the tower. The ratio of vertical load to
10 moment is about correct, but the value of V0 is not known, so
8 Windward leg it is simply estimated, knowing that the tower is very close
to failure, so the load point is just inside the yield surface.
Moment: MN m

6
When the lead blocks were added, the vertical load
4
increased, but the net moment decreased, moving the load
2 point to B, slightly further away from the yield surface.
0 The extraction of soil from underneath the uphill side was
2 then used to decrease the tilt by about 10%, thus reducing the
4 moment further, while keeping the vertical load constant, the
0005 0 0005 0010 0015 load point moved to C. Finally the lead blocks were removed,
Rotation: deg taking the load point back towards the yield surface at D, but
a little bit further away than initially.
Fig. 41. Moment rotation response of windward foundation of jack-up From this simple analysis it is concluded that the
subjected to NewWave loading
application of the lead weights was indeed a sensible option
as of course it proved to be. This demonstrates that the
removal of the lead blocks, when the stress point moved back
10 By way of guide reaction towards the yield surface, was probably more critical than
By way of accelerations
their initial placement.
08

06
Hull dispalcement: m

Conclusions from part 2


04 It is important that interactions are not overlooked when
they may be important, even if they are not obvious or are
02
difficult to calculate. To take the trivial example from the
0 elastic behaviour of footings, in most design codes such as
API (2010) the interaction between horizontal load and
02 moment is ignored. This is not on the rational basis that the
04
effect is small; it is simply that it has not been properly
accounted for but that does not mean that in reality it will
06 not influence the results.
Time 5 s interval More generally, it is vital in problems of soil structure
interaction that there is an effective means of communication
Fig. 42. Lateral motion of jack-up hull under extreme load (from between geotechnical and structural engineers. The struc-
McCarron & Broussard, 1992) tural engineers must accept that soil response cannot be
reduced to the presence of simple linear springs, but geotech-
nical engineers must play their part, expressing soil and
require complex computation, as the example below foundation behaviour in a way that is intelligible to structural
illustrates. engineers and compatible with their numerical analyses.
The stability of the Tower of Pisa is investigated, using a Work hardening plasticity theory, combined with macro-
simple calculation that requires no more than a spreadsheet, models for foundation behaviour, provides an appropriate
and using data entirely available from the literature (Marchi, language for expressing geotechnical knowledge in a way that
2008; Marchi et al., 2011). The question that is addressed is compatible with numerical analysis.
is how risky was the solution that was adopted for saving
the Tower, which involved stacking about 1000 t of lead
weights on the uphill side of the foundation, thus increasing PART 3: FOUNDATIONS FOR OFFSHORE WIND
the vertical load on the foundation. It is initially perhaps TURBINES
counterintuitive to attempt to stabilise a failing foundation Motivation
by increasing the loading on it, but the yield surface approach In this final section the focus moves to another sector the
shows that in this case such an approach was rational. rapidly expanding offshore wind industry, and specifically the
The curve on Fig. 43 shows a best estimate of the yield challenge of foundation engineering for offshore wind
surface for a circular foundation in dimensionless form. turbines. Some particular ideas for more economical

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 21
foundations for offshore turbines are examined, noting that it of increase was less that one part per million per year, and
is estimated that 1922% of the capital total cost of an now it is more than twice that. Over-use of fossil fuels is
offshore installation may be related to the foundation and affecting the atmosphere at an ever-increasing pace. It is well
substructure (Willow & Valpy, 2011). Aspects of the design known that rising carbon dioxide levels are a driver of global
for conventional solutions (monopiles), less conventional warming, with all its consequences. It is essential to find an
solutions (suction caissons) and a novel idea (screw piles) are alternative to fossil fuels.
considered. There will be one general conclusion: that geo- However, there is another quite different reason why
technical engineers need to be fully engaged in, and informed this dependency on hydrocarbons should be reduced, and
about the energy debate. More specifically, cyclic loading is that is that they are a finite resource. Fig. 46 shows the UK
identified as one of the key issues to be addressed for the production of oil and natural gas, mainly from resources in
design of monopiles; for suction caissons there are a number the North Sea. These peaked in 1999 for oil and 2000 for gas,
of important design issues and the focus here is on tension and are now in steady decline as supplies are depleted. This
capacity; finally, screw piles are introduced as a technology same pattern is repeated worldwide, although there is some
that holds excellent promise for turbines in deep water, as debate about whether or not peak oil has been passed yet. In
their combination of good tensile capacity and noise-free broad terms, however, economically accessible worldwide oil
installation meets some key offshore design needs. resources would last in the region of 40 years at present rates
It is useful first to examine why we should look offshore of consumption, and gas perhaps 60 years or so. Recent
for clean renewable power, especially around the UK, and developments in fracking may lengthen this by a few years.
especially exploiting wind power. The primary motivation Alternative sources of energy must be found within the next
comes from the increasing levels of carbon dioxide (CO2) in few decades, and of course the sooner these are developed the
the atmosphere. Fig. 44 shows the famous (or infamous) greater the benefit (see King et al., 2015).
Keeling curve showing the increase of atmospheric carbon There are really only two potential solutions to the
dioxide with time, primarily due to the use of fossil fuels. twin problem of climate change caused by hydrocarbons,
May 2014 was the first month in human history when the and diminishing supplies of those hydrocarbons: nuclear
concentration exceeded 400 parts per million by volume power or renewables (storage and energy efficiency also have
(ppmv) of carbon dioxide, and 2015 was the first year in important roles to play). Nuclear power will undoubtedly
which the annual average exceeded 400 ppmv. The figures play a vital part, but the focus below is on renewables. Wind
are even more alarming when the curve is differentiated to power is one of the most promising of the renewable
obtain the rate of increase of carbon dioxide, Fig. 45. There resources, and offshore wind power has some particular
is some scatter, but the trend is clear: 50 years ago the rate advantages. The primary benefit is that average wind speeds
are typically higher offshore than onshore: in Fig. 47 the
progression from green to yellow shows higher average wind
42000
speeds. Almost as important is the fact that the wind is less
turbulent offshore, making it more suitable for power
Annual average atmospheric

40000
generation. There is also the social advantage that many
carbon dioxide: ppmv

38000
people do not favour wind turbines on aesthetic grounds, but
offshore wind attracts less opposition.
36000 The drawback to offshore wind is of course the additional
cost, associated principally with (a) foundation costs,
34000 (b) transmission distances and (c) costs of operation and
maintenance. Foundation costs can be minimised by seeking
32000 locations with reasonably high wind speed but shallow water.
Fig. 48 shows bathymetry around the UK, and there are
30000 some obvious target areas of shallow water, perhaps most
1950 1960 1970 1980 1990 2000 2010 2020
notably the Dogger Bank. The hatched areas on Fig. 49 show
Year
the regions available for development in round three of the
Crown Estates licensing process. The smaller coloured areas
Fig. 44. The Keeling curve: atmospheric carbon dioxide level as a
function of time (data from Mauna Loa observatory, Hawaii; source: are those that are already developed or under development.
Scripps Institution of Oceanography)

300 6000
Annual increase of atmospheric

250
carbon dioxide: ppmv/year

5000
Power production: PJ

200
4000
150
3000
100
2000
050

1000 Oil production


0
1950 1960 1970 1980 1990 2000 2010 2020 Gas production
Year 0
1970 1975 1980 1985 1990 1995 2000 2005 2010 2015
Fig. 45. Time differential of the Keeling curve: rate of increase of Year
atmospheric carbon dioxide level as a function of time (data from
Mauna Loa observatory, Hawaii; source: Scripps Institution of Fig. 46. Oil and gas production in the UK as a function of time
Oceanography) (source: DECC)

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
22 HOULSBY
60N

Wind speed: m/s


55N > 140
136140
131135
126130
121125
116120
111115
101105
106110
96100

50N
9195
8690
8185
N 7680
7175
<71
10W 5W 0

Fig. 47. Average wind speed around the UK (from DTI, 2004)

The newer developments will be further offshore, and mainly but of course they act much lower on the support structure. It
in somewhat deeper water. is useful to combine the lateral loads into a single force
At the end of 2012 the UK had almost 3 GW of installed
offshore wind power, accounting for almost 60% of the off- H 2 MN from wind 4 MN from wave current
shore wind power worldwide, Table 3. This is in the context 6 MN
that the average electricity demand in the UK is around
45 GW. Of course, because the wind does not blow all the time, which results in a moment
3 GWof installed wind power delivers about 1 GWon average M 2  40 110 MN m 4  40 MN m 460 MN m
(a capacity factor in the region of 30%). The UK therefore
generates around 2% of its electrical power from offshore wind The point of action of the net horizontal load is thus about
(and about twice that amount from onshore wind). There are 80 m above the seabed in this case.
ambitious plans for much larger developments: the UK is one Since the typical weight of the structure and turbine may
of the world leaders in offshore wind developments. be of the order of 10 MN (arising from a mass of 1000 t), the
The design of an offshore wind turbine involves the analysis horizontal load is about 60% of the vertical load. The design
of numerous load cases involving fatigue, serviceability and problem involves very high lateral load with relatively small
ultimate limit states (generating conditions, extreme storm vertical load. Contrast this with the jack-up structures
loading, emergency stop and so on). Much of the design is addressed in Part 2, where the lateral load is rarely more
governed by fatigue life, and this in turn is affected by the than 1015% of the vertical load.
dynamic response of the structure. For instance offshore There are broadly two ways that loads can be transmitted
turbines are usually designed so that their first natural to a foundation. Fig. 51(a) shows a monopile foundation,
frequency falls between the rotational frequency (P) and which must resist directly the vertical and horizontal loads
blade-passing frequency (3P). For the sake of the example, and the overturning moment. Fig. 51(b) shows an alternative
however, all of these cases are reduced to a single set of with multiple foundations: in this case the moment loading
representative loads on a turbine. Fig. 50 shows the order of on each foundation is much less important, because the
magnitude of the maximum forces on a typical modern overall overturning moment is resisted by pushpull action
offshore turbine in around 40 m of water and delivering about of the downwind and upwind foundations. The magnitude of
5 MW. The force of the blades is of the order of 2 MN, and this effect depends critically on the spacing of the foun-
this may act at around 110 m above the waterline. The lateral dations, s. Unless they are very widely spaced, the overturning
forces from waves and current are much larger, about 4 MN, moment would usually be sufficient that the vertical load on

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 23
60N

55N

Depth MSL: m
<5
510
1020
2030
3040
4050
50100
100250
50N
250500
5001000
10002000
20003000
N
30004000
> 4000

10W 5W 0

Fig. 48. Bathymetry around the UK (from DTI, 2004)

Table 3. Status of offshore installed wind capacity at end of 2012 smaller installations in shallow water the monopile is an
(data from IEA, 2013) economical choice. In the future, as larger devices (5 MW
and more) are installed in deeper water, it is unlikely that
Country Total installed Offshore installed large enough monopiles will be feasible, and the foundations
capacity: MW capacity: MW are likely to involve multiple footings of some sort.
China 91 413 428 6%
USA 61 110 0
Germany 34 660 903 14%
Spain 22 959 0
Monopiles
India 20 150 0 In spite of the above observations, the monopile remains
UK 10 861 3653 55% the most important foundation choice for offshore wind tur-
Italy 8554 0 bines at present. Although there is a huge amount of experi-
France 8254 0 ence on piled foundations from the oil and gas industry, it is
Canada 7803 0 important to realise that monopiles for wind turbines present
Denmark 4808 1271 19% a radically different design problem. The most striking
Rest of world 47 533 335 5% difference is their size. In offshore oil and gas, piles are
Total 318 105 6590 100% typically anything from 30 m to about 100 m long (quite
often more), and up to about 2 m in diameter. They are long
and slender, with length-to-diameter ratios of 50 or more.
the upwind footing becomes tensile, so there is the special As a result they behave in a relatively flexible way in bending.
problem of design against tensile loading. In contrast, monopiles are short and fat. Most are around
In general the larger the turbine and the deeper the water, 30 m long, with diameters up to about 6 m. Even larger
the more difficult it is to design a monopile or monopod diameters are under consideration. The length-to-diameter
structure. At present most past installations are in relatively ratio is only about 5, and as a result they are very stiff in
shallow water close to shore, and of devices in the 23 MW bending. This stiffness is of course required as they must
range. Almost all have been founded on monopiles. Fig. 52 resist the large overturning moments directly.
presents the developments in a plot of turbine power This remarkable difference is the driver behind the Carbon
(horizontal axis) and water depth (vertical axis). Blue dots Trust and industry-sponsored PISA project, which is inves-
represent UK developments on monopiles and red dots those tigating how conventional py methods can be adapted for
on multiple foundation ( jacket) structures. In general for short piles with low L/D ratios (Byrne et al., 2015a, 2015b).

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
24 HOULSBY

2 MN

110 m

4 MN

40 m

Seabed

Fig. 50. Dimensions and representative loads on a typical offshore


wind turbine installation
N
0 100 200
km

Fig. 49. UK third-round offshore wind farm sites (Crown Estate)


H
H
The project concentrates on monotonic loading and the first
few cycles of load, and should lead to new design methods.
However, a very important feature of monopiles is the
fact that they are subjected to extremely large numbers V
(.108) of small-amplitude cycles, and the author addresses V

here how the effects of these on wind turbine performance


can be understood. Tests have been carried out on models in V2
a very simple rig designed to apply many thousands of cycles, V1
Fig. 53. A stiff pile is embedded in soil, in this case sand, and M
H
a combination of cyclic horizontal load and moment is
applied by a simple system of weights and pulleys. As a mass H1 H2
rotates, it results in a changing lever arm, which in turn
results in the cyclic load on the pile. The displacements are
monitored by simple dial gauges, a convenient choice that V
avoids the problems of long-term drift of electrical
transducers.
s
The programme of loading was chosen to be representative
of working loads on an offshore turbine. Fig. 54 shows the (a) (b)
estimated position of the failure surface for the pile under
combined lateral and moment loading in a dimensionless Fig. 51. Load transfer mechanisms from offshore wind turbine to
plot with H H=DL2 , M M=DL3 . The failure surface foundation
is almost straight within the relevant range of loads, and
measured failure values correspond closely to the theory. For The study, LeBlanc et al. (2010a), involved one-way load-
design in the field, the ultimate limit state (ULS) can be ing of various amplitudes, dimensioned by b = Mmax/MR,
identified as a point on the yield surface, and then a suitable where MR is a reference (failure) moment, Fig. 55. The effect
safety factor can be applied. It is possible also to identify of the ratio of the minimum to maximum load in the cycle
appropriate points for the serviceability limit state (SLS), c = Mmin/Mmax was also examined, as under field conditions
which would occur many times during the life of the load cycles may not just be in the form of one-way loading.
structure, and the fatigue limit state (FLS) for which the Figure 56 shows the results of a typical test. A relatively
interest may be in millions of cycles. The relevant range of stiff response is observed during each loading cycle, in which
loading is typically between 30 and 50% of the failure values. the soil behaves almost elastically. However, each cycle

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 25
Bearing capacity: test results
60 Bearing capacity: theory
Scaled design loads
10
Cyclic loading ranges: b
50 Jacket
structures?
Beatrice
08
40 Most future
developments?
Water depth: m

Monopiles ULS K = 10
06
30 ULS/135
Walney 2 ~
Gabbard M
Past Walney
developments? Sheringham
20 Thanet London Ormonde 04
Barrow SLS
Lynn
60%
Dowsing
Robin
Lincs
50%
Teesside Rigg FLS
10 North Hoyle Gunfleet Gunfleet 3
02 40%
Blyth Rhvl
30%
Scroby Kentish Burbo
20%

2 3 4 5 6 7 0
0 02 04 06 08 1.0 12
Turbine power: MW
~
H
Fig. 52. Water depth and sizes of wind farm installations
Fig. 54. Estimated failure envelope and loading conditions for cyclic
tests (from LeBlanc et al., 2010a)
Reaction frame

M
10
MR
b = 075 b = 05
05
025
05
0
0
Model c = 0 c =
Mass 05
pile Motor 10

Fig. 55. Variation of amplitude of cyclic loading (left) and form of


Mass loading (right) (from LeBlanc et al., 2010a)
Sand

05
Initial cycles Cycle 103 Cycle 104 Cycle 105
Mass
04
Normalised moment

Fig. 53. Outline of apparatus for cyclic loading of model monopiles


03

involves a small hysteresis loop, implying a degree of frictional 02


damping. Perhaps more importantly, during each one-way
cycle there is a small accumulation of rotation. Although the
01
residual rotation in each cycle is minute, after many cycles a
significant ratcheting rotation is accumulated far more
than the elastic or plastic rotations in any one cycle. The 0
accumulation, however, is at a diminishing rate in the test 0 0001 0002 0003 0004 0005 0006 0007 0008 0009 0010
Normalised rotation
shown in Fig. 56 the first 1000 cycles cause approximately the
same deformation as the next 9000. Although not immedi- Fig. 56. Accumulated displacement during 100 000 one-way cycles
ately visible in Fig. 56, the secant stiffness changes gradually (Abadie, 2015)
with cycling, and the amount of hysteretic damping changes
too. All of these effects (change of stiffness, change of
damping, ratcheting displacement) have implications for the cycling always degrades the performance of a soil. The tests
long-term design of wind turbine foundations. shown were on sand of medium density, and in the case of
The data are presented in terms of quantities defined loose to medium dense sands the stiffness increases with
in Fig. 57. The secant stiffness kN is defined for each cycle. cycling, almost certainly due to slight densification of the
Fig. 58 shows how this stiffness varies with the cycle number. sand close to the pile. However, for wind turbine design either
The four curves are for different amplitudes of cycling. The a decrease or an increase of stiffness can be important, as
smallest amplitude of cycling gives the highest secant stiff- the resonant frequency of the entire foundationstructure
ness, which is consistent with usual observations. However, system should be maintained within a certain range, and the
over 10 000 cycles, the secant stiffness increases in each of the dynamic response of the wind turbine depends partly on the
tests shown. This is contrary to the conventional wisdom that foundation stiffness. LeBlanc et al. (2010a) suggested that

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
26 HOULSBY
10
M k0 kN
M b = 053

Normalised accumulated rotation


(N) b = 040
s
Mmax b = 027
1 b = 020

01

Mmin = k N 031
0 N static

Fig. 57. Definitions of static and accumulated rotation (from LeBlanc


001
et al., 2010a) 1 10 100 1000 10 000 100 000
Number of cycles

500
One-way cycling
Fig. 59. Variation of accumulated rotation with load amplitude and
450 b = 020
cycle number
400
Normalised stiffness

b = 027 Figure 60(b) is more interesting, as it shows the depen-


350 dence on the form of cycling. Symmetric cycling (c =  1) of
300 b = 040 course produces no accumulated deformation. One-way
cycling (c = 0), however, does cause accumulated rotation.
b = 053
250 The value c = 1 represents a constant applied load with no
200 cyclic component, and again produces no accumulation.
What is most striking, however, is that the maximum
150
accumulated deformation is caused by cycles some way
100 between one-way and two-way cycling. In other words, the
1 10 100 1000 10 000 100 000 largest accumulations occur when the pile is first pushed in
Cycle number one direction, then partially pushed backwards, then pushed
again in the first direction. Cyclic loading test programmes
Fig. 58. Variation of secant stiffness with load amplitude and cycle often concentrate on one-way and two-way loading, and may
number miss this more damaging intermediate condition.
Cyclic testing is often only carried out for a few tens
the dependence of stiffness on cycle number from these of cycles, and these data are unusual in extending to over
laboratory tests could be quite well fitted (at least within the 20 000 cycles. Although the slopes in Fig. 59 are almost
range of cycle number tested) by the logarithmic expression constant over three decades of cycle number (approximately
k N k 0 Ak ln N 7 10 to 104), the field problem involves of the order of 108
cycles in the lifetime of the turbine, and there are serious
with the initial stiffness in turn a function of the cycle size questions about extrapolation of these data to very large
and form, k N K b b K c c . Note that the results of these numbers of cycles. Furthermore, the real loading involves a
laboratory tests at 1g should be applied to field cases with complex combination of cycles of different magnitudes and
caution. Field monitoring of natural frequency can shed in different directions. LeBlanc et al. (2010b) suggested a
some light on changes in foundation stiffness, and Kallehave methodology for how the effects of these could be combined.
et al. (2015) detected no significant change of natural There is much further important work currently underway
frequency with time over a period of 25 years for a monopile on the performance of monopiles, which will continue to find
in dense sand. wide, and even dominant, use in the offshore wind industry,
The accumulated displacement during many cycles, nor- but this topic is not pursued further here.
malised by dividing by the displacement in the first cycle, is
presented in Fig. 59. In each case shown the accumulated
rotation, after normalising by the rotation in the first cycle,
increases roughly as a power law with number of cycles, and Suction caissons
can be quite accurately fitted (within the range of cycles Consideration will now be given to suction caissons, which
tested) by the expression are widely regarded as attractive alternatives to piling for
offshore wind turbines. Often informally termed suction

T b  T c  N 031 8 buckets they are large, thin-walled cylindrical structures,
static sealed by a stiffened plate on the top and open at the base.
First note the observation that the total accumulated defor- They are installed into the ground first by penetration under
mation increases as a power function with number of cycles. their own weight, and then by pumping water out of the
The exponent of about one third seems quite well established caisson to create a suction that forces the caisson into the
from a number of different tests. As a result, the incremental ground, Fig. 61. In clay the mechanism of penetration is
accumulation in any one cycle decreases with cycle entirely driven by the differential pressure across the lid of the
number: (d/dN ) / N 0311 N 069. caisson (Houlsby & Byrne, 2005a), but in sand there is an
The absolute magnitude of the accumulated deformation additional beneficial effect that a seepage pattern is set up,
is expressed through two empirical factors Tb(b) and Tc(c), which reduces the effective stresses inside the caisson, further
multiplying the power function of the number of cycles. assisting penetration (Houlsby & Byrne, 2005b).
Fig. 60(a) shows the dependence Tb(b) on the magnitude of The caissons themselves may be no less expensive
the cycle. Unsurprisingly, larger cycles produce more accu- than piles of equivalent capacity, but they have two great
mulated rotation, and Tb is also found to be a function of the advantages. First, they do not require expensive equipment
relative density. for installation essentially all they require is a pump and

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 27
025 5
Rd = 38%
Rd = 4%

020 4 Rd = 38%

015 Rd = 4% 3
Tb Tc

010 2

005 1

0 0
0 02 04 06 08 10 05 0 05 10
b c
(a) (b)

Fig. 60. Effects of (a) cyclic load amplitude and (b) cyclic load form on accumulated deformation (from LeBlanc et al., 2010a)

Flow 1400
1200
1000
Vertical stress: kPa

800
600
Pressure W 400
differential
200
0
200
400
200 210 220 230 240 250 260 270
Displacement: mm
Flow
Fig. 62. Cyclic loading of a model suction caisson in sand

(b) The tensile capacity of caissons, under both


serviceability and ultimate conditions, when they are
Fig. 61. Installation of a suction caisson used in a multiple footing configuration to support a
jacket.
(c) Performance under cyclic loading.
control system and second, the installation is almost silent.
There is a major concern that the vibration from pile driving Of the above issues, the one that is most frequently raised
operations is sufficiently detrimental to marine mammals by designers is the question of how much tension can be
dolphins and seals that strict limits are being placed on the reasonably allowed in design, and so just the tensile loading
noise from offshore operations, especially driven piling problem is addressed here. On average a caisson supporting a
(Bailey et al., 2010). wind turbine will be in compression, but it will be subjected
There are a number of issues around the design of suction to small-amplitude cyclic vertical loading during operating
caissons for offshore wind turbines, and these were the subject conditions, and occasional much larger cycles during extre-
of a major design study undertaken for the Department of me conditions. Fig. 62 shows the results of a model test of a
Trade and Industry (DTI), the Engineering and Physical caisson in saturated sand with packets of gradually increasing
Sciences Research Council (EPSRC) and a consortium of load cycles. In this and the following figures, the vertical load
companies (Byrne et al., 2002b; Doherty et al., 2005; Houlsby is converted to an equivalent stress by dividing by the area of
& Byrne, 2005a, 2005b; Kelly et al., 2006a, 2006b; Villalobos the caisson. Examining the first four packets, each of ten
et al., 2009, 2010). The most pressing issues are as follows. cycles, not surprisingly the hysteresis in the cycles increases
with amplitude, as well as the accumulated deformation in
(a) Assessment of whether the caisson can be installed at a each cycle. In the fifth packet, of five cycles, the amplitude is
given site. There seems little doubt that they can be sufficient that tension is just reached at the limit of the cycles:
installed in relatively homogeneous sand, even very again the hysteresis and accumulated deformation increase.
dense sands, and in most clays. However, very stiff or In the final packet of five cycles the amplitude of cycling
fissured clays may be a problem, as might very was only slightly larger, but involved a more significant
coarse-grained soils, in which it may not be possible to tension. It can be seen that there is a dramatic change in
form a satisfactory seal to apply the suction. Isolated the nature of the response, with each application of tension
larger stones may also cause a problem. Installation in involving significant upward movement. On removing the
some layered or inhomogeneous soils may also not be tension and reapplying compression there is equally large
possible. However, the range of soils where installation downward movement, so that the hysteresis loop takes on a
can be achieved is quite wide. characteristic banana shape. Note that these cycles result in

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
28 HOULSBY
the important phenomenon that in each cycle there is a large Displacement: mm
residual downward movement, even though this is triggered 150 160 170 180 190 200 210
0
by the application of a tensile (upward) load in part of the A slow
cycle. The implication is that if a turbine on multiple caissons Direction of
were to be subjected to very large overturning moments from 100
movement 100 kPa

Vertical stress: kPa


wind loading, sufficient to cause substantial tension, then the
accumulated displacement would cause the turbine to tilt B fast
towards the prevailing wind direction a somewhat para- 200
doxical result. The explanation surely lies in the fact that the
average load on each foundation is downwards, and the large 200 kPa
C fast + 200 kPa pressure
cyclic displacements caused by tensile loading result in 300
damage to the soil fabric, resulting in an overall weakening
of the foundation.
At the end of this test a tensile pullout test was applied, and 400
the tensile stress, even at quite large displacements (approxi-
mately 250 kPa), was much smaller than the compressive Fig. 64. Tensile loading of suction caissons in sand
capacity, which was not even approached in the largest cycles,
and must be significantly greater than 1300 kPa. rapid wave loading that a turbine would experience during a
The results in Fig. 62 were from small-scale tests, and the storm. The capacity is limited by cavitation of the pore fluid
question of course arises whether they are also applicable in underneath the caisson. In this case the test is carried out at
the field. It is very costly to carry out cyclic tension tests on atmospheric pressure, so cavitation occurs at approximately
full-size caissons, but there are data from quite substantial 100 kPa gauge pressure and, comparing to curve A, the
field trials on a 15 m dia. caisson, Houlsby et al. (2006). capacity is increased by approximately this amount (in fact
Fig. 63 shows the results of cyclic tests from four different rather more, presumably because the negative pore pressures
tests at different sizes and with different modes of installa- enhance the effective stresses acting on the skirts, and hence
tion. In each of the four tests, three selected hysteresis loops enhance the frictional component of the capacity).
at different amplitudes have been extracted from a longer Curve C shows the result of a fast test on the same
sequence of cycling. After normalising the results using the caisson, but carried out at an elevated pressure in a pressure
procedures in Kelly et al. (2006a), the hysteresis loops from chamber. This is used to simulate the water pressure conditions
the different tests are remarkably similar. The scaling involves on the seabed. In this case the pressure is elevated by 200 kPa,
normalising the applied vertical load as V/D 3 and displace- and so cavitation now occurs after reduction of gauge pressure
ment as (w/D)  ( pa/D)05. The second part of the displace- by a correspondingly larger pressure change. The ultimate
ment normalisation accounts approximately for the change capacity is seen to increase by a little more than 200 kPa,
of soil shear modulus with stress level. The results in Fig. 63 confirming that the capacity is determined by the onset of
give confidence that, provided appropriate normalisation cavitation, which depends on the water pressure at the seabed.
techniques are used, laboratory experimental results can be The result is that in the field the ultimate tensile capacity
scaled to predict field conditions. will be very high. However, concentrating on just the early
It is generally accepted though that the ultimate capacity of stages of all three of the tests, it is seen that they each show an
a suction caisson in tension will be very large. Fig. 64 shows initially very soft response, followed in the rapid loading
the results for tensile loading of a model caisson in sand. cases by an increase in load. When scaled up to field
First consider curve A, which is from a test involving a very dimensions using the procedures described above, the
slow application of the tensile load. The net vertical stress on displacements in this initial phase would be unacceptable
the caisson is small, and equates approximately to the friction on serviceability grounds. Note, however, that tensile capacity
that is developed on the outside and inside of the skirts, as approximately equal to the frictional resistance is developed
during the slow extraction excess (negative) pore pressures during the rapid tests at very small incremental displacement.
beneath the caisson can dissipate. The conclusion is that, on serviceability grounds, tension on
Curve B shows the result of an equivalent test in which caissons should be limited to the frictional capacity, and no
the same caisson is pulled very fast: the ultimate capacity is account should be taken of the end bearing which is limited
much larger, as negative excess pore pressures are developed by cavitation, except under truly exceptional load conditions.
underneath the caisson. This case is more relevant to the If deformations are to be limited to acceptable values,
significant tension must be avoided.
It is worth noting that the tensile capacity limited by
15 m field, suction installed
015 m lab, suction installed cavitation is one of the exceptionally rare cases in soil
5 02 m lab, pushed mechanics where the soil behaviour is not determined by
015 m lab, pushed effective stresses, but is limited by a total stress value. Such
Dimensionless vertical load

4
cases only arise when the absolute pressure in the pore fluid
affects the overall soil response, which is of course the case
3
when cavitation is relevant.
2

1 Screw piles
Screw piles, or helical piles, have not yet been adopted
0 as foundations for offshore wind turbines, but several of
their characteristics suggest that they are potentially attractive
1
0 001 002 003 004 005 and economical for this application. A screw pile consists
Dimensionless vertical displacement of a shaft to which are fixed large-diameter helical plates,
Fig. 65. Screw piles are installed by twisting them into the
Fig. 63. Application of normalisation techniques to load cycles of a ground by a hydraulic torque motor. Some additional vertical
suction caisson in sand load (crowd) can be applied, which assists installation. They

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 29
The design in both clay and sand can be simplified by
normalising the capacity in an appropriate way. The relevant
dimensionless groups are V=su D2p and V =D3p , suggesting
that for clay the capacity increases with the square of the helix
diameter, and for geometrically similar piles in sand it
increases with the cube of the diameter. However, in sand it is
more likely that the length may be chosen independently of
p
diameter, and for this case the capacity varies approximately
with the square of the diameter, suggesting that it would be
more appropriate to use V =LD2p.
L
Dp The installation torque can also be expressed in terms of
S dimensionless groups, and here it can be seen, for instance,
that the torque in clay scales with the cube of the diameter as
the relevant group is T=su D3p. For geometrically similar piles
in sand the relevant group is T=D4p, but for the reasons
discussed above it may be more appropriate to use T=LD3p.
Some key ratios are useful in design. The first is the
capacity, multiplied by the plate diameter and divided by the
installation torque, VDp/T. This number would be expected
to be roughly constant for piles of similar geometry and
D
in similar soil types (although the ratio may be expected
to differ for sands and clays, and to vary with, for instance,
Fig. 65. Schematic diagram of screw pile angle of friction or overconsolidation ratio). Note that
onshore the current practice is to use the capacity to torque
ratio, V/T, but this has the odd dimension of 1/length, and
are robust and simple, can be installed in many different soil thus depends on the scale of the pile.
types, and have good tensile capacity. Onshore, screw piles The second useful ratio is simply the ratio of tensile
are now routinely used for installations that involve light capacity Vt to compressive capacity V.
vertical loads, but also require tensile capacity, most usually Perko (2009) gathered data from various sources (onshore
due to overturning moments. Examples are the gantries for practice and model tests), from piles typically only 300 mm
signs over motorways. Onshore the piles are typically small, in diameter, and up to 900 mm; the loads were correspond-
with the helical flights up to about 600 mm in diameter. ingly small. Data from Perko (2009), together with other
The loading of an offshore wind turbine results in similar sources, are collated in Table 4. Although there are some
load ratios: relatively low vertical load combined with a outliers, the geometric mean of the ratio VDp/T is 823 for
tension caused by overturning moment, suggesting the use of compressive loading, and for VtDp/T it is 666 for tensile
screw piles offshore (Byrne & Houlsby, 2015). There are key loading. The geometric mean value of Vt/V is 085. (Note
additional advantages. that there is a slight inconsistency in these values because
gaps in the data sets mean that slightly different data are used
(a) The installation is potentially almost silent. This is a to determine the different factors.) There were insufficient
significant advantage as it is now recognised that the data to be able to determine these ratios with greater accuracy
noise from driven piling for large numbers of wind for particular soils, although there is some indication that the
turbines is damaging to marine mammal populations. lower VDp/T and VtDp/T values are associated with coarser
Driven piles are often now ruled out on environmental grained materials (sands). These figures, however, provide the
grounds. empirical background against which any design methods for
(b) As the screw pile is installed the torque can be offshore piles can be calibrated.
monitored, and the torque can be closely correlated with There are essentially two ways to calculate the capacity for
the load capacity (see below), so that the installation compressive loading. In the first, Fig. 66(a), it is assumed that
provides a confirmation of the future performance. each helix acts as an independent plate, and the capacity is
calculated as a function of plate bearing and shaft friction
However, there are also challenges in moving this technology components. Alternatively, Fig. 66(b), it can be assumed that
offshore. First, the loads offshore are vastly greater than the soil fails on a cylindrical envelope around all the plates.
those encountered onshore, so screw piles need to be scaled End bearing is only developed on the bottom plate, but the
up by a factor of 3 or 4, to the order of 2 m dia. Second, side friction is over a much larger area. Of course the design
installation equipment must be developed that can cope capacity will be the lower of these two values, and an efficient
with the required torque, and be operated in the offshore design may involve adjusting the plate spacing so that both
environment. There are also non-geotechnical challenges, occur almost simultaneously.
but such problems are not insuperable. For instance the The tensile capacity can be calculated in a very similar way,
shaftplate connection is vulnerable to fatigue, suggesting again treating the plates as independent or as a single unit,
that cast construction (successfully used in other offshore Figs 67(a) and 67(b), although the detailed shape of the
applications (Broughton et al., 1997)) could be employed. breakout surface is uncertain, and is still a subject for
In developing design methods for much larger piles, it is research. The problem in clay is reasonably well understood
important to identify the key dimensionless groups that (Merifield, 2011), but there remains considerable uncertainty
define the pile behaviour. First, there are groups that rep- about tensile capacities in sand.
resent the geometry of the pile for instance the ratio of shaft Calculations of the installation torque can similarly be
to helix diameter, D/Dp (see Fig. 65), length to diameter ratio, made, at least in clay, assuming either that each plate cuts a
L/Dp, and the number, spacing ratio, s/Dp, and pitch ratio, path through the soil, or that an entire block of soil rotates
p/Dp, of the helix plates. There is no particular reason to with the pile. The latter would be unsatisfactory, and
depart from well-established onshore practice in choosing for realistic combinations of design parameters it does not
these ratios, at least for initial design. occur.

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
30 HOULSBY
Table 4. Onshore and model data for helical piles: geometric means of measured dimensionless parameters

Source Test type Soil No. of tests VDp/T VtDp/T Vt/V

Ghaly et al. (1991) Laboratory Sand 40 433


Rao et al. (1991) Laboratory Soft clay 17 071
Tsuha et al. (2010) Centrifuge Sand 18 809
Livneh & El Naggar (2008) Field Silt 8 1233 065*
11 796
Sakr (2011) Field Sand 7 661 083*
4 550
Sakr (2009) Field Oil sand 2 958 054*
3 520
Cerato & Victor (2009) Field Layered 8 1115
Perko (2009) Various Various 117 809 092*
86 741
Weighted geometric mean 134 823 085
170 666
*Implied.

V V

Wpile
Wpile
Wsoil

Envelope
friction

Shaft friction

Plate bearing
Plate bearing
Shaft friction

End bearing
End bearing

(a) (b)

Fig. 66. Compressive loading of a screw pile: (a) individual plate failure; (b) envelope failure

Figure 68 shows the results of a calculation for an offshore the necessary equipment. There is a golden opportunity to
screw pile with realistic dimensions for a four-legged jacket achieve a step change in the technology employed offshore.
structure supporting a large wind turbine at a deep-water site Although not used in recent practice offshore, there is
in the North Sea (the details are confidential). The soil is clay. ample historical evidence of their successful use offshore.
The steps in the capacity come as each new helix enters The earliest known offshore structure supported on screw
the ground, and in this case the independent capacity of the piles was the Maplin Sands lighthouse. The foundation was
plates is always a little less than the envelope capacity. Fig. 69 designed by an Irish engineer, Alexander Mitchell, remark-
shows the dimensionless torque ratio VtDp/T for the same able for his achievements not least because he was totally
site, giving values a little over 10 slightly higher than the blind from the age of 21. The lighthouse, see Fig. 71, was
typical values expected from the empirical database, but well supported on nine screw piles, eight forming an octagon
within the variability observed. The calculations also give a and one in the centre. They were driven into the sand by a
ratio of tensile to compressive capacity, Vt/V, in the range man-powered capstan system. Each pile had a single helix of
from about 07 to 095, Fig. 70, which is also consistent with diameter 12 m (4 feet), and was driven 7 m (22 feet) below
the empirical database. These comparisons provide consider- the mudline. The lighthouse was operated for over 90 years
able confidence in the design method. until 1931, but scouring around the structure had gradually
It would be desirable to compare these predictions with undermined the foundation, and the lighthouse finally
measurements of performance of screw piles in the North Sea, collapsed in 1932.
but alas that is yet to happen. However, screw piles could offer Mitchell had a clear motivation for using screw piles. In
an economical and environmentally acceptable solution to the his paper to the Institution of Civil Engineers in 1842
foundation problem for deep-water wind turbines. What is (reprinted Mitchell, 1848) he made it clear that the tensile
needed at this stage is for an installation contractor to develop capacity of screw piles was one of their attractions. He also

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 31

Vt

Wpile
Wpile
Wsoil

Envelope
friction

Shaft friction

Plate bearing

Shaft friction

(a) (b)

Fig. 67. Tensile loading of a screw pile: (a) individual plate failure; (b) envelope failure

Total bearing load, kN


0 5000 10 000 15 000 20 000 25 000 30 000
0

Minimum compression
50
Independent compression
Interacting compression
100 Tension
Pile tip depth: m

150

200

250

300

350

Fig. 68. Predictions of ultimate loading on a screw pile for a realistic deep-water North Sea site

correctly identifies depth, helix area and soil type as the key Conclusions from part 3
elements affecting the capacity. Offshore wind power can and should be an important
Whether this broad spiral flange, or Ground Screw, as contributor to the UKs energy supply in the future. As
it may be termed, be applied to the foot of the pile to turbines are installed offshore and in deeper water, and as the
support a superincumbent weight, or be employed as a turbines themselves become larger (5 MW or more), the
mooring to resist an upward strain, its holding power trend will be away from monopile foundations towards jacket
entirely depends upon the area of its disc, the nature of the structures on multiple footings.
ground into which it is inserted, and the depth to which it In order to drive costs down, novel solutions will be required.
is forced beneath the surface. One such option is the use of suction caissons, which have
already been the subject of much study and are now beginning
The proper area of the screw should, in every case, be to be deployed in this sector. A more radical solution would be
determined by the nature of the ground in which it is to be the use of helical piling, already in use over 150 years ago, but
placed, and which must be ascertained by previous not currently deployed offshore. As the performance charac-
experiment. (Mitchell, 1848: p. 110) teristics of helical piles are well matched to the loading con-
Note that Mitchell recommends testing to prove the ditions offshore (principally because of their tensile capacity),
capacity of his innovative design. Over 160 years later, they could offer an economically viable and environmentally
Mitchells wise advice still holds true. acceptable solution for large turbines on multiple piles.

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
32 HOULSBY
Torque ratio, Vt Dp /T
0 2 4 6 8 10 12
0

50

100
Pile tip depth: m

150

200

250

300

350

Fig. 69. Normalised torque ratio for a screw pile

Tension/compression capacity ratio, Vt /V


0 01 02 03 04 05 06 07 08 09 10
0

50

100
Pile tip depth: m

150

200

250

300

350

Fig. 70. Tensile to compressive capacity ratio for a screw pile

OVERALL CONCLUSION
This paper is intended to encourage geotechnical engineers
to be outward looking towards other disciplines.
Throughout soil mechanics, geotechnical engineers deal
Fig. 71. Maplin Sands lighthouse (from Mitchell, 1848)
with a material which is intrinsically variable: as illustrated
in Part 1, it is vital that geotechnical engineers embrace more
sophisticated statistical and probabilistic techniques to stands for the offshore environment; and the earth is the
address this variability in a rational way. subject of Gotechnique. All the four elements come together
In Part 2 the behaviour of a jack-up unit was used to in the area of offshore structures, and it is the way they
illustrate how rigorous and well-defined mathematical tech- interact that creates the fascination of the topic.
niques such as plasticity theory can be used to define models
for soil mechanics problems, especially for foundations. These ACKNOWLEDGEMENTS
models allow the behaviour of footings to be framed in a way The author is very grateful to the British Geotechnical
that is compatible with advanced numerical codes, thus Association for giving him the opportunity to present a
allowing geotechnical expertise to be encoded and commu- Rankine Lecture and this accompanying paper. This paper
nicated to structural engineers and to others. For their part, draws on work carried out in collaboration with many
structural engineers must accept that foundation response colleagues, too many for all to be named individually here,
cannot adequately be represented by elasticity theory. but importantly including Dr Christelle Abadie, Ross Bell,
Geotechnical engineers have an important role to play Associate Professor Britta Bienen, Professor Roy Butterfield,
in contributing to the future energy supply, one of the most Professor Mark Cassidy, Professor Guido Gottardi, Dr Neil
important challenges of the twenty-first century. They can Houlsby, Mike Hoyle, John Huxtable, Dr Richard Kelly,
contribute to the solutions, and should engage knowledge- Dr Christian LeBlanc, Professor Chris Martin, Dr Luan
ably with the wider energy debate, drawing on past experi- Ngo-Tran, Dr Nguyen-Sy Lam, Julian Osborne, Brooks
ence to seek novel solutions to challenges such as those Paige, Yura Perov, Professor Mark Randolph, Dr Paulo de
offered by the design of offshore wind turbine foundations, as Santa Maria, Dr Teh Kar-Lu, Dr Richard Thompson,
illustrated in Part 3. Tor-Inge Tjelta, Dr Felipe Villalobos, Professor Frank
Wood and Professor Zhu Bin. The author is grateful to
Professor Alan Lutenegger for generously providing infor-
Postscript mation about helical piling. Finally, the author is particularly
The four elements recognised in classical times were fire, grateful to Professor Byron Byrne, who led much of the
air, water and earth. The fire represents the oil and gas that experimental work presented here. The author also wishes to
has driven so much of offshore engineering; the air represents express his deep sense of gratitude for the wise guidance from
wind power that should be exploited economically; the water his mentor, the late Professor Peter Wroth.

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 33
NOTATION foundations on sand. Gotechnique 56, No. 6, 367379, http://
A area of foundation dx.doi.org/10.1680/geot.2006.56.6.367.
a dimensionless factor defining eccentricity of elliptical yield Bienen, B., Cassidy, M. J., Randolph, M. F. & Teh, K. L. (2010).
surface section Characterisation of undrained shear strength using statistical
B footing width methods. Proceedings of 2nd international symposium on
D diameter of circular footing or pile frontiers in offshore geotechnics, ISFOG 2010, Perth, Australia,
Dp helix diameter of screw pile pp. 661666.
EI bending stiffness of leg of jack-up Brinch Hansen, J. (1970). A revised and extended formula for bearing
G soil (small strain) shear modulus capacity, Bulletin No. 98, pp. 511. Copenhagen, Denmark:
H horizontal load on jack-up unit or foundation Danish Geotechnical Institute.
h depth of embedment of footing Broughton, P., Hayes, R., Wood, A. & Komaromy, S. (1997). Cast
h0 dimensionless factor defining maximum horizontal capacity nodes for the Ekofisk 2/4J jacket. Proceedings of the Institution of
K1 dimensionless vertical stiffness factor Civil Engineers Structures and Buildings 122, No. 3, 266280.
K2 dimensionless rotational stiffness factor BSI (2004). BS EN 1997-1: 2004 Eurocode 7: geotechnical design
K3 dimensionless horizontal stiffness factor part 1: general rules. London, UK: BSI.
K4 dimensionless stiffness factor coupling horizontal and Butterfield, R. & Ticof, J. (1979). The use of physical models in
rotational response design. Discussion. In The measurement, selection and use
k spudcan rotational stiffness of design parameters in geotechnical engineering: proceedings of
kN pile rotational stiffness on unloading cycle N the 7th European conference on soil mechanics and foundation
k0 pile rotational stiffness on first unloading engineering, vol. 4, pp. 259261. London, UK: British
L leg length of jack-up; length of pile Geotechnical Association.
M moment on footing Byrne, B. W. (2000). Investigations of suction caissons in dense sand.
MR reference (failure) moment DPhil thesis, University of Oxford, Oxford, UK.
m mass of jack-up Byrne, B. W. & Houlsby, G. T. (2003). Foundations for offshore
m0 dimensionless factor defining maximum moment capacity wind turbines. Phil. Trans. R. Soc. London, Ser. A 361, No. 1813,
Nc bearing capacity factor 29092930.
Q torsional load on foundation Byrne, B. W. & Houlsby, G. T. (2005). Investigating 6 degree-
q generalised horizontal/moment load on foundation of-freedom loading on shallow foundations. Proceedings of
R radius of footing international symposium on frontiers in offshore geotechnics,
s spacing of piles; spacing of helix plates on screw pile Perth, Australia, pp. 477482. London, UK: Taylor and Francis.
su undrained strength Byrne, B. W. & Houlsby, G. T. (2015). Helical piles: an innovative
su0 undrained strength at soil surface foundation design option for offshore wind turbines. Phil.
T installation torque for screw pile Trans. R. Soc. 373, No. 2035, 20140081, http://dx.doi.org/
Tb factor in rotation accumulation equation (8) accounting 10.098/rtsa.2014.0081.
for load magnitude Byrne, B. W., Houlsby, G. T. & Martin, C. M. (2002a). Cyclic
Tc factor in rotation accumulation equation (8) accounting loading of shallow offshore foundations on sand. Proceedings
for load type of international conference on physical modelling in geotechnics,
u horizontal displacement of foundation St Johns, Newfoundland, Canada, pp. 277282.
V vertical (compressive) load on jack-up unit or foundation Byrne, B. W., Houlsby, G. T., Martin, C. M. & Fish, P. (2002b).
Vt vertical tensile load on screw pile Suction caisson foundations for offshore wind turbines. Wind
V0 reference vertical load, bearing capacity under pure vertical Engng 26, No. 3, 145155.
load Byrne, B. W., McAdam, R., Burd, H. J., Houlsby, G. T., Zdravkovic,
v normalised vertical load, V/V0 L., Taborda, D. M. G., Potts, D. M., Jardine, R. J., Sideri, M.,
w vertical (downward) displacement of foundation; spacing Schroeder, F. C., Martin, C. M., Gavin, K., Doherty, P., Igoe, D.
between legs of jack-up unit Muir Wood, A., Kallehave, D. & Skov Gretlund, J. (2015a).
wpile weight of pile New design methods for large diameter piles under lateral
wsoil weight of soil surrounding screw pile loading for offshore wind applications. In Frontiers in offshore
yc height of equivalent cone for spudcan idealisation geotechnics III (ed. V. Meyer), pp. 705710. Leiden,
z depth below mudline the Netherlands: CRC Press.
roughness factor on conical footing Byrne, B. W., McAdam, R. A., Burd, H. J., Houlsby, G. T., Martin,
angle of base of conical footing C. M., Gavin, K., Doherty, P., Igoe, D., Zdravkovic, L.,
unit weight of soil (submerged unit weight for offshore Taborda, D. M. G., Potts, D. M., Jardine, R. J., Sideri, M.,
calculations) Schroeder, F. C., Muir Wood, A., Kallehave, D. & Skov
b dimensionless factor defining cyclic magnitude Mmax/MR Gretlund, J. (2015b). Field testing of large diameter piles under
c dimensionless factor defining cyclic load type Mmin/Mmax lateral loading for offshore wind applications. In Geotechnical
rotation of foundation engineering for infrastructure and development: XVI European
rate of increase of undrained strength with depth conference on soil mechanics and geotechnical engineering (eds
n first natural frequency of jack-up M. G. Winter, D. M. Smith, P. J. L. Eldred and D. G. Toll), vol. 3,
pp. 12551260. London, UK: ICE Publishing.
Cao, Z. & Wang, Y. (2013). Bayesian approach for probabilistic site
characterization using cone penetration tests. Proc. ASCE,
REFERENCES J. Geotech. Geoenviron. Engng 139, No. 2, 267276.
Abadie, C. N. (2015). Cyclic lateral loading of monopile foundations Cassidy, M. J., Eatock Taylor, R. & Houlsby, G. T. (2001). Analysis
in conhesionless soils. DPhil thesis, University of Oxford, of jack-up units using a constrained NewWave methodology.
Oxford, UK. Appl. Ocean Res. 23, No. 4, 221234.
API (2010). RP 2A-WSD recommended practice for planning, Cassidy, M. J., Martin, C. M. & Houlsby, G. T. (2004). Development
designing and constructing fixed offshore platforms. Washington, and application of force resultant models describing jack-up
DC, USA: American Petroleum Institute. foundation behaviour. Mar. Structs (special issue on jack-up
Bailey, H., Senior, B., Simmons, D., Rusin, J., Picken, G. & platforms) 17, No. 34, 165193.
Thompson, P. M. (2010). Assessing underwater noise levels dur- Cerato, A. B. & Victor, R. (2009). Effects of long term dynamic
ing pile-driving at an offshore windfarm and its potential effects on loading and fluctuating water table on helical anchor perform-
marine mammals. Marine Pollution Bull. 60, No. 6, 888897. ance for small wind tower foundations. Proc. ASCE,
Bell, R. W. (1991). The analysis of offshore foundations subjected to J. Performance of Constructed Facilities 23, No. 4, 251261.
combined loading. MSc thesis, University of Oxford, Oxford, UK. Davies, E. H. & Booker, J. R. (1973). The effect of increasing strength
Bienen, B., Byrne, B. W., Houlsby, G. T. & Cassidy, M. J. (2006). with depth on the bearing capacity of clays. Gotechnique 23,
Investigating six-degree-of-freedom loading of shallow No. 4, 551563, http://dx.doi.org/10.1680/geot.1973.23.4.551.

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
34 HOULSBY
Dean, E. T. R., James, R. G., Schofield, A. N. & Tsukamoto, Y. Kelly, R. B., Houlsby, G. T. & Byrne, B. W. (2006b). Transient
(1998). Drum centrifuge study of three-leg jackup models in clay. vertical loading of model suction caissons in a pressure chamber.
Gotechnique 48, No. 6, 761786, http://dx.doi.org/10.1680/ Gotechnique 56, No. 10, 665675, http://dx.doi.org/10.1680/
geot.1998.48.6.761. geot.2006.56.10.665.
de Santa Maria, P. E. L. (1988). Behaviour of footings for offshore King, D., Browne, J., Layard, R., ODonnell, G., Rees, M., Stern, N.
structures under combined loads. DPhil thesis, University of & Turner, A. (2015). Global Apollo programme to combat climate
Oxford, Oxford, UK. change. London, UK: London School of Economics. See
Doherty, J. P. & Deeks, A. J. (2003). Elastic response of http://www.globalapolloprogram.org/ (accessed 29/07/2016).
circular footings embedded in a non-homogeneous half-space. LeBlanc, C., Houlsby, G. T. & Byrne, B. W. (2010a). Response
Gotechnique 53, No. 8, 703712, http://dx.doi.org/10.1680/ of stiff piles in sand to long-term cyclic lateral loading.
geot.2003.53.8.703. Gotechnique 60, No. 2, 7990, http://dx.doi.org/10.1680/geot.
Doherty, J. P., Houlsby, G. T. & Deeks, A. J. (2005). Stiffness of 7.00196.
flexible caisson foundations embedded in nonhomogeneous LeBlanc, C., Byrne, B. W. & Houlsby, G. T. (2010b). Response of
elastic soil. Proc. ASCE, J. Geotech. Geoenviron. Engng 131, stiff piles to random two-way lateral loading. Gotechnique 60,
No. 12, 14981508. No. 9, 715721, http://dx.doi.org/10.1680/geot.09.T.011.
DTI (2004). Atlas of UK marine renewable energy resources, Livneh, B. & El Naggar, M. H. (2008). Axial testing and numerical
Technical report no. R.1106. London, UK: Department of modelling of square shaft helical piles under compressive and
Trade and Industry. tensile loading. Can. Geotech. J. 45, No. 8, 11421155.
Ghaly, A. M., Hanna, A. & Hanna, M. (1991). Installation torque Marchi, M. (2008). Stability and strength analysis of leaning towers.
of screw anchors in dry sand. Soils Found. 31, No. 2, 7792. PhD thesis, Universit di Parma, Parma, Italy.
Gordon, A. D., Henzinger, T. A., Nori, A. V. & Rajamani, S. K. Marchi, M., Butterfield, R., Gottardi, G. & Lancellotta, R. (2011).
(2014). Probabilistic programming. In Proceedings of the future Stability and strength analysis of leaning towers. Gotechnique
of software engineering, pp. 167181. New York, NY, USA: 61, No. 12, 10691080, http://dx.doi.org/10.1680/geot.9.P.054.
Association for Computing Machinery (ACM). Martin, C. M. (1994). Physical and numerical modelling of offshore
Gottardi, G., Houlsby, G. T. & Butterfield, R. (1999). The plastic foundations under combined loads. DPhil thesis, University of
response of circular footings on sand under general planar Oxford, Oxford, UK.
loading. Gotechnique 49, No. 4, 453470, http://dx.doi.org/ Martin, C. M. & Houlsby, G. T. (2001). Combined loading
10.1680/geot.1999.49.4.453. of spudcan foundations on clay: numerical modelling.
Hossain, M. S. Hu, Y., Randolph, M. F. & White, D. J. (2005). Gotechnique 51, No. 8, 687700, http://dx.doi.org/10.1680/
Limiting cavity depth for spudcan foundations penetrating clay. geot.51.8.687.40470.
Gotechnique 55, No. 9, 679690, http://dx.doi.org/10.1680/ McCarron, W. O. & Broussard, M. D. (1992). Measured jack-up
geot.2005.55.9.679. response and spudcan-seafloor interaction for an extreme storm
Houlsby, G. T. (2010). A probabilistic approach to the prediction of event. In Proceedings of the 6th international conference on
spudcan penetration of jack-up units. Proceedings of the 2nd behaviour of offshore structures, BOSS 92 (eds M. H. Patel
international symposium on frontiers in offshore geotechnics, and R. Gibbins), pp. 349361. London, UK: BPP Technical
ISFOG 2010, Perth, Australia, pp. 673678. Services.
Houlsby, G. T. & Byrne, B. W. (2005a). Design procedures for Merifield, R. S. (2011). Ultimate uplift capacity of multiple helical
installation of suction caissons in clay and other materials. type anchors in clay. Proc. ASCE, J. Geotech. Geoenviron. Engng
Proceedings of the Institution of Civil Engineers Geotechnical 137, No. 7, 704716.
Engineering 158, No. 2, 7582, http://dx.doi.org/10.1680/geng. Mitchell, A. (1848). On submarine foundations; particularly the
158.2.75.61630. screw pile and moorings. Minutes Proc. ICE 7, 108146.
Houlsby, G. T. & Byrne, B. W. (2005b). Design procedures for Ngo-Tran, C. L. (1996). The analysis of offshore foundations
installation of suction caissons in sand. Proceedings of the subjected to combined loading. DPhil thesis, University of
Institution of Civil Engineers Geotechnical Engineering 158, Oxford, Oxford, UK.
No. GE3, 135144. See http://dx.doi.org/10.1680/geng.158.3. Nguyen-Sy, L. (2005). The theoretical modelling of circular shallow
135.66297. foundation for offshore wind turbines. DPhil thesis, University of
Houlsby, G. T. & Cassidy, M. J. (2002). A plasticity model for the Oxford, Oxford, UK.
behaviour of footings on sand under combined loading. Nguyen-Sy, L. & Houlsby, G. T. (2005). The theoretical modelling of
Gotechnique 52, No. 2, 117129, http://dx.doi.org/10.1680/ a suction caisson foundation using hyperplasticity theory. In
geot.52.2.117.40922. Frontiers in offshore geotechnics: proceedings of the international
Houlsby, G. T. & Martin, C. M. (2003). Undrained bearing symposium on frontiers in offshore geotechnics, IS-FOG 2005
capacity factors for conical footings on clay. Gotechnique (eds M. Cassidy and S. Gourvenec), pp. 417422. London, UK:
53, No. 5, 513520, http://dx.doi.org/10.1680/geot.2003.53.5. Taylor and Francis.
513. Nova, R. & Montrasio, L. (1991). Settlements of shallow foun-
Houlsby, G. T. & Wroth, C. P. (1982). Direct solution of plasticity dations on sand. Gotechnique 41, No. 2, 243256, http://dx.doi.
problems in soils by the method of characteristics. Invited org/10.1680/geot.1991.41.2.243.
keynote paper. Proceedings of the 4th international conference on Osborne, J. J., Teh, K. L., Houlsby, G. T., Cassidy, M. J., Bienen, B.
numerical methods in geomechanics, Edmonton, Alberta, & Leung, C. F. (2011). Improved guidelines for the prediction
Canada, vol. 3, pp. 10591071. of geotechnical performance of spudcan foundations during
Houlsby, G. T., Kelly, R. B., Huxtable, J. & Byrne, B. W. (2006). installation and removal of jack-up units, Report of InSafeJIP
Field trials of suction caissons in sand for offshore wind turbine Joint Industry Study (multiple sponsors). London, UK: RPS
foundations. Gotechnique 56, No. 1, 310, http://dx.doi.org/ Energy.
10.1680/geot.2006.56.1.3. Peck, R. B. (1969). Advantages and limitations of the observational
Houlsby, N. M. T. & Houlsby, G. T. (2013). Statistical fitting of method applied in soil mechanics. Gotechnique 19, No. 2,
undrained strength data. Gotechnique 63, No. 14, 12531263, 171187, http://dx.doi.org/10.1680/geot.1969.19.2.171.
http://dx.doi.org/10.1680/geot.13.P.007. Perko, H. A. (2009). Helical piles: a practical guide to design
IEA (2013). IEA wind: 2012 annual report. Boulder, CO, USA: and installation. Hoboken, NJ, USA: Wiley.
International Energy Agency, PWT Communications. Pisan, F., di Prisco, C. G. & Lancellotta, R. (2014). Soil
Kallehave, D., Thilsted, C. L. & Troya, A. (2015). Observed foundation modelling in laterally loaded historical towers.
variations of monopile foundation stiffness. In Frontiers in Gotechnique 64, No. 1, 115, http://dx.doi.org/10.1680/geot.
offshore geotechnics III (ed. V. Meyer), pp. 717722. Leiden, 12.P.141.
the Netherlands: CRC Press. Poulos, H. G. (1988). Marine geotechnics. London, UK: Unwin
Kelly, R. B., Houlsby, G. T. & Byrne, B. W. (2006a). A comparison Hyman.
of field and laboratory tests of caisson foundations in sand and Rao, S. N., Prasad, T. V. S. N. & Shetty, M. D. (1991). The behaviour
clay. Gotechnique 56, No. 9, 617626, http://dx.doi.org/10.1680/ of model screw piles in cohesive soils. Soils Found. 31, No. 2,
geot.2006.56.9.617. 3550.

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.
INTERACTIONS IN OFFSHORE FOUNDATION DESIGN 35
Roscoe, K. H. & Schofield, A. N. (1957). The stability of short pier thermomechanical principles, to more application-oriented
foundations in sand. Discussion. Br. Weld. J., January, 1219. plasticity solutions to problems in soilstructure interaction.
Sakr, M. (2009). Performance of helical piles in oil sand. Can. Tonight we are probably grateful that Guy has not chosen to
Geotech. J. 46, No. 9, 10461061. deliver a treatise on the principles of hyperplasticity a topic
Sakr, M. (2011). Installation and performance characteristics of high
that he virtually invented but the scientific quality of his
capacity helical piles in cohesionless soils. Deep Found. Inst. J. 5,
No. 1, 3957. contributions is equally evident in the topics he has covered,
Salenon, J. & Matar, M. (1982). Bearing capacity of axially drawn from the world of foundations for offshore structures.
symmetrical shallow foundations. J. de Mcanique Thorique Many moons ago we shared the same PhD supervisor,
et Applique 1, No. 20, 237267. the late Peter Wroth, and this early training is evident through-
Simpson, B. (1992). Retaining structures: displacement and design. out tonights lecture in Guys emphasis on dimensional
Gotechnique 42, No. 4, 541576, http://dx.doi.org/10.1680/ analysis, addressing any new problem by first establishing
geot.1992.42.4.541. appropriate dimensionless groups of the relevant parameters.
Szczepinski, W. (1974). Stany graniczne i kinematyka osrodkw Those groups then guide equally the setting up of model tests,
sypkich. Warsaw, Poland: Pan stwowe Wydawnictwo Nakowe ensuring they represent the real problem, and performing
(in Polish).
Tan, F. S. C. (1990). Centrifuge and theoretical modelling of conical
analytical parametric studies that lead to a design framework.
footings on sand. PhD thesis, Cambridge University, Cambridge, In the first part of his lecture, Guy referred to the
UK. parametric solutions he developed for the bearing resistance
Taylor, P. H., Jonathan, P. & Harland, L. A. (1995). Time domain of spudcan foundations, but his main focus was not so much
simulation of jack-up dynamics with the extremes of a Gaussian the deterministic solutions but on statistically sound methods
process. In Proceedings of the 14th international conference on to quantify confidence limits in predicting the penetration-
offshore mechanics and arctic engineering (OMAE), vol. 1-A, resistance profiles for such foundations. This is an essential
pp. 313319. New York, NY, USA: American Society of part of all geotechnical design, but perhaps particularly in
Mechanical Engineers. the offshore world where soil data are expensive to acquire,
Tromans, P. S., Anaturk, A. R. & Hagemeijer, P. (1991). A new
and as a result sparse. His message was clear, that we need to
model for the kinematics of large ocean waves applications
as a design wave. Proceedings of the 1st international offshore grasp the nettle of non-deterministic design calculations,
and polar engineering conference, Edinburgh, UK, vol. 3, quantifying the reliability of our predictions.
6471. Interactions are an essential part of all aspects of our work,
Tsuha, C. H. C., Aoki, N., Rault, G., Thorel, L. & Garnier, J. (2010). and those among you who have interacted with Guy will
Physical modelling of helical screw piles in sand. In Physical know that he does not suffer fools gladly, being somewhat
modelling in geotechnics: proceedings of the 7th international dismissive of fuzzy thinking. He has made extensive
conference on physical modelling in geotechnics, ICPMG 2010 contributions to the so-called fixity problem that impacts
(eds S. Springman, J. Saue and L. Seward), vol. 2, 841846. the design of the truss legs for mobile drilling rigs, building
Leiden, the Netherlands: CRC Press. gradually from simple elastic foundation response to yield
Vesic, A. S. (1975). Bearing capacity of shallow foundations. In
Foundation engineering handbook (eds H. F. Winterkorn and
envelopes and elegant kinematic hardening models that
H. Y. Fang), pp. 121147. New York, NY, USA: Van Nostrand describe the full non-linear rotational response of foun-
Reinhold. dations. Quite apart from the scientific advances, this has
Villalobos, F., Byrne, B. W. & Houlsby, G. T. (2009). An involved gradual education of those responsible for rig
experimental study of the drained capacity of suction caisson design, convincing them of the multiplicity of rotational
foundations under monotonic loading for offshore applications. responses of the foundations, depending on the load path.
Soils Found. 49, No. 3, 477488. The final part of his lecture has summarised succinctly
Villalobos, F. Byrne, B. W. & Houlsby, G. T. (2010). Model testing the pressing need for alternative energy sources and the role
of suction caissons in clay subjected to vertical loading. Appl. within that of offshore renewable energy from wind and
Ocean Res. 32, No. 4, 414424.
water. The offshore wind industry, where the UK has shown
Wang, Y., Huang, K. & Cao, Z. (2014). Bayesian identification of
soil strata in London Clay. Gotechnique 64, No. 3, 239246, leadership in respect of current and planned developments,
http://dx.doi.org/10.1680/geot.13.T.018. is at a transition point in terms of the increases in power
Willow, C. & Valpy, B. (2011). Offshore wind: forecasts of future costs output of each installation and water depth. New types of
and benefits, Report to RenewableUK. Swindon, UK: BVG structure and foundation approaches are needed, and Guy
Associates. has outlined alternatives from suction caissons to screw piles.
Wood, F., van de Meent, J. W. & Mansinghka, V. (2014). A new He has emphasised the damaging role of cyclic loading,
approach to probabilistic programming inference. Proceedings of ranging from fatigue and the gradual accumulation of defor-
the 17th international conference on artificial intelligence and mations over tens of thousands of cycles, to the more abrupt
statistics, Reykjavik, Iceland, pp. 10241032. loss of strength that can occur as a result of shear stress
Young, A. G., Remmes, B. D. & Meyer, B. J. (1984). Foundation
performance of offshore jack-up drilling rigs. J. Geotech. Engng,
reversals in sand. This is clearly an area where geotechnical
ASCE 110, No. 7, 841859. engineers must work closely with other disciplines in order to
develop cost-effective but robust foundation options, because
these influence the structural design.
As I commented earlier, we have been spared intense
VOTE OF THANKS analysis, although its presence was evident in support of the
PROFESSOR MARK RANDOLPH, Fugro Chair in material presented and I look forward to the written version
Geotechnics, Centre for Offshore Foundation Systems, The of the lecture where no doubt more such details will be
University of Western Australia. provided.
Ladies and gentlemen, I am privileged to have been a We have also been given our homework, to address our
friend and colleague of Guys for over 30 years, and was own deficiencies in the areas of statistics, mathematical
delighted when I was asked by the British Geotechnical rigour and true engagement with the energy debate. Above
Association to deliver this vote of thanks. all, however, we have enjoyed an outstanding 54th Rankine
The hallmark of Guys work is scientific rigour, and he Lecture, reflecting some of Guys contributions over his
has made numerous fundamental contributions to soil career, but forward-looking and provocative in articulating
mechanics and geotechnical engineering, ranging from new the challenges to be faced. I ask you all to join me now in
approaches to constitutive modelling of soils, based on showing our thanks and warm appreciation to Guy.

Downloaded by [ WESTERN UNIVERSITY] on [13/09/16]. Copyright ICE Publishing, all rights reserved.

You might also like