You are on page 1of 23

OTC 23521

Foundation Modeling and Assessment in the New ISO Standard 19905-1


P.C. Wong, ExxonMobil Development Company; J.S. Templeton III, SAGE USA, Inc.; O.A. Purwana, Keppel
Offshore & Marine; H. Hofstede, GustoMSC Engineers; M.J. Cassidy & M.S. Hossain, U. of Western Australia;
and C.M. Martin, U. of Oxford

Copyright 2012, Offshore Technology Conference

This paper was prepared for presentation at the Offshore Technology Conference held in Houston, Texas, USA, 30 April–3 May 2012.

This paper was selected for presentation by an OTC program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Offshore Technology Conference and are subject to correction by the author(s). The material does not necessarily reflect any position of the Offshore Technology Conference, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Offshore Technology Conference is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of OTC copyright.

Abstract
This paper presents the new foundation assessment provisions in ISO 19905-1. The paper discusses the improvements to
spudcan penetration analysis achieved by the refinement of bearing capacity formulations and by adoption of a fundamentally
new approach to backflow prediction which accounts for the changes in soil flow regime and the evolving pattern of soil
deformation in the vicinity of the spudcan. It also addresses the upgrades made to the SNAME 5-5A approach to foundation
capacity and foundation stiffness reduction in sand and clay as well as the means used to better account for the effects of deep
spudcan penetrations in clay. The change to, and basis for, the use of gross foundation capacity in some of the calculations
are also discussed. Finally, the paper presents the new foundation acceptance checks framework consequent to the change
from available capacity/reaction concept in SNAME 5-5A to gross capacity/reaction approach in ISO 19905-1.

At the time of this paper submission, it was expected that ISO 19905-1 would be isssued prior to the presentation of the paper
at the 2012 Offshore Technology Conference. Upon the issue of this standard the new foundation assessment provisions
described in this paper will be required for jack-up site-specific assessments worldwide.

Introduction and Background


The jack-up geotechnical assessment approach in ISO 19905-1 is the result of the concerted effort of Panel 4 under the
auspices of ISO TC67/SC7/WG7. The development of the latest assessment approach by Panel 4, made up of experienced
geotechnical practitioners in the industry and well-respected researchers in academia, spanned almost two decades. Using
SNAME TR5-5A Revision 2 as the starting point, the geotechnical assessment approach was revised and upgraded based on
findings from industry studies and academic research projects in understanding spudcan foundation behavior over the years.
The prior and existing recommended practices for assessment of jack-up foundations and their performance under storm
loading conditions stemmed from industry studies in the 1980s and early 1990s. Those recommendations were substantially
conservative by intention, and they were limited by the research of the time. As a result they were reliable but, in many cases,
overly restrictive for jack-up location approvals. Not only are the new foundation assessment requirements of ISO 19905-1
better grounded in current research, they are less restrictive and more realistic while remaining reliably conservative. This
paper provides an overview of five main geotechnical areas in the geotechnical assessment approach, namely:
• Site investigation requirements
• Spudcan penetration and foundation bearing capacity
• Spudcan foundation response under combined load during storm
• Special spudcan foundation considerations (fatigue, earthquake)
• Spudcan foundation acceptance criteria

Site Investigation Requirements


While it has long been generally recognized that site-specific assessment for a jack-up rig requires geotechnical data, there
has never been general agreement in the industry as to the optimal geotechnical site investigation field program and the type
and amount of specific geotechnical data required. Historically the sources of geotechnical data considered adequate for jack-
up location approvals have covered a very broad range, including:
• Detailed site-specific investigation specifically for the rig and site (though this is a rarity)
2 OTC 23521

• Detailed investigation at same site-specific platform location


• Detailed site investigation at a nearby platform location (usual best case for storm assessment)
• Limited site investigation at site prior to going on location (usual best case for installation)
• Limited site investigation from the jack-up after going on location
• No site investigation but prior experience with a similar rig at same location
• No site investigation but prior relevant experience nearby
• Geotechnical inferences from regional catalogues or atlases
ISO 19905-1 provides much more detailed and specific guidance in this area than has been available previously. It
provides much information regarding various available site investigation techniques and the circumstances in which each is
appropriate. ISO 19905-1 has made a significant advance in this area of site investigation requirements, as can be seen from a
direct comparison of the ISO and SNAME provisions.
The basic guidance in SNAME 5-5A (2.4.1) is:
“Site-specific geotechnical information must be obtained. The type and amount of geotechnical data required will depend on
the particular circumstances such as the type of jack-up and previous experience of the site, or nearby sites, for which the
assessment is being performed. Such information may include shallow seismic survey, coring data, cone-penetrometer tests,
side-scan sonar, magnetometer survey and diver's survey.”
The basic provision in ISO 19905-1 is quite similar:
“Site-specific geotechnical information applicable to the anticipated range of penetrations shall be obtained from or on behalf
of the operator. The type and amount of geotechnical data required depends on the particular circumstances, such as the type
of jack-up and previous experience at the site or nearby sites. Such information can include geophysical survey (sub-bottom
profiler; side-scan sonar; bathymetry; magnetometer) data; boring/coring data; in-situ and laboratory test data; and diver's
survey data.”
This is followed by non-prescriptive guidance on specific information including:
• Bathymetric survey via acoustic reflection systems
• Sea floor survey: side scan sonar or high-resolution multibeam echo sounder, ROV, divers, shallow seismic survey, high
resolution acoustic reflection techniques
• Geotechnical investigation: undisturbed soil sampling, in-situ testing and laboratory testing, additional soil testing to
provide shear moduli for cyclic or dynamic behavior
• Integration of geotechnical results with geophysical survey results
Specific guidance on the number and positioning of boreholes and piezocone penetration test holes at a given jack-up site
(open water site or work-over site) with a given geological setting (simple, complex, or very complex) is provided in Annex
D of ISO 19905-1.
Finally, one of the most notable provisions in ISO 19905-1 is the clarification that the site-specific geotechnical
information is to be “obtained from or on behalf of the operator.”

Bearing Capacity in Clay


In ISO 19905-1, predictions of leg penetration in clay are made with the spudcan idealized as a flat circular foundation of
diameter B embedded at depth D below the mudline. This is the same as the SNAME 5-5A approach, but there is additional
illustrated guidance explaining how to calculate the equivalent dimensions B and D for both partially- and fully-penetrated
spudcans. The gross vertical bearing capacity is denoted QV (rather than FV in SNAME 5-5A) and is calculated as
Q V = (s u N c s c d c + p′o ) πB 2 4 (1)
where
su = design undrained shear strength (discussed below)
Ncsc = bearing capacity factor for circular footing = 6.0
dc = depth factor = 1 + 0.2D/B ≤ 1.5
p′o = effective overburden pressure at depth D
The corresponding equation in SNAME 5-5A is
FV = (c u N c s c d c + p′o ) πB 2 4 (2)
where
cu = undrained shear strength at depth D + B/4
Nc = bearing capacity factor for strip footing = 5.14
sc = shape factor = 1 + 1/5.14 ≈ 1.2
⎧1 + 0.4 D B for D B ≤ 1
dc = depth factor = ⎨
⎩1 + 0.4 tan (D B) for D B > 1
−1
OTC 23521 3

and p′o is as per Equation (1).


Apart from updating the notation for undrained shear strength (from cu to su), ISO 19905-1 provides considerable freedom
in selecting the representative su to be used in Equation (1): “an evaluation should be made of the sampling method, the
laboratory test type and the field experience regarding the prediction and observations of spudcan penetrations.” There is a
subsequent note, however, that “for typical Gulf of Mexico shear strength gradients and spudcan dimensions, spudcan
penetrations in clay are well predicted by selecting su as the average over a depth of B/2 below the widest cross-section.”
Thus the approach used in SNAME 5-5A is maintained – see the definition of cu in Equation (2) – but with a warning that
this field-based experience may not be directly applicable to other regions. Selection of the representative strength by
averaging over a depth of B/2 (i.e. one footing radius) follows Endley, et al. (1981) and Young, et al. (1984).
A more significant feature of Equation (1), from ISO 19905-1, is the adoption of the shape and depth factors proposed by
Skempton (1951). The formulation in Equation (2), from SNAME 5-5A, uses shape and depth factors from Brinch Hansen
(1970), with multiplicative superposition as proposed by Vesic (1975). The twin equations for the depth factor in Equation
(2) give rise to a small (~6%) discontinuity at D/B = 1, but this apparent oversight by Brinch Hansen (subsequently
reproduced by Vesic) can easily be rectified if necessary (Martin, 1994; Houlsby & Martin, 2003; Dean, 2008).
The move from Vesic (1975) to Skempton (1951) as the default calculation method is in line with current industry
practice. In SNAME 5-5A, Skempton’s method is only mentioned in passing in the commentary section, along with several
other calculation methods for vertical bearing capacity on clay, notably those that incorporate a direct treatment of increasing
strength with depth (Davis & Booker, 1973; Salençon & Matar, 1983; Houlsby & Wroth, 1983). In ISO 19905-1, Skempton’s
method is given prominence because it is widely used in the jack-up industry and has a long-established reputation as a
sufficiently accurate method for soft, single-layer clays (Young, et al., 1984; Menzies & Roper, 2008).
ISO 19905-1 also describes (but does not explicitly recommend) an alternative calculation method based on tables of
theoretical bearing capacity factors for embedded conical footings. Unlike the semi-empirical methods in Equations (1) and
(2), the theoretical factors take direct account of axisymmetric geometry, embedment depth, cone apex angle, cone
roughness, and increasing undrained shear strength with depth. The tables appearing in ISO 19905-1 were compiled at
Oxford University using the program FIELDS, developed by Prof. G.T. Houlsby. They originally appeared in the thesis of
Martin (1994), were reproduced in the original (1994) version of SNAME 5-5A, and were later published by Houlsby &
Martin (2003) with some additional discussion and curve-fits.
Each of the tabulated factors is derived from a numerical analysis using the axisymmetric slip-line method with the
Tresca yield criterion, for a particular combination of cone apex angle (β = 30° to 180°), cone roughness (α = 0 to 1),
embedment ratio (D/B = 0 to 2.5) and dimensionless rate of strength increase with depth (ρB/sum = 0 to 5). Figure 1 shows a
typical slip-line field. It is important to realize that to facilitate construction of the solution, the cavity above the conical
footing is assumed to be occupied by a rigid but smooth-sided cylindrical shaft (Houslby & Martin, 2003). A stress field
obtained in this way is not strictly applicable when there is an unsupported sidewall above footing level. It does, however,
provide a valid lower bound plasticity solution for the case where the cavity is completely filled with soil (assuming that the
soil in the cavity has the same strength profile as the surrounding soil – this is of course unlikely to be the case for the
backfilled soil above a continuously penetrating spudcan footing).
Figure 2(a) shows that for a flat, rough, circular footing embedded in clay with uniform strength (ρ = 0), the Skempton
(1951) method from ISO 19905-1 is significantly more conservative (by up to 20%) than the Vesic (1975) method from
SNAME 5-5A. The theoretical bearing capacity factors of Houlsby & Martin (2003) lie close to or slightly above the values
from Skempton’s formula (maximum difference 7% at D/B = 1).
Figure 2(b) shows a comparison of the various methods for a case more typical of a spudcan penetrating soft normally
consolidated clay: β = 150°, α = 1, ρB/sum = 5. For the Skempton and Vesic methods, the bearing capacity was determined
using the strength at depth D + B/4 (see the discussion following Equation (2)), but normalization was performed using the
strength at depth D, denoted suo, to allow direct comparison with the theoretical bearing capacity factors. The Vesic (1975)
method from SNAME 5-5A again provides the least conservative predictions. The curve plotted from the tables of Houlsby
& Martin (2003) is more conservative (by up to 19% at D/B = 2.5) than that obtained using the Skempton (1951) method
from ISO 19905-1. This is consistent with the findings of Menzies & Roper (2008), who compared predictions from both of
these methods (and others) with actual leg penetration records from the Gulf of Mexico.
The usage of the Houlsby & Martin (2003) tables in ISO 19905-1 is essentially unchanged from that in SNAME 5-5A,
except that no separate bearing capacity factor N′c is introduced; it is simply stated that the tabulated values “provide a
theoretical lower bound to the total bearing factor Nc.sc.dc to apply to the shear strength at the spudcan base level, suo.”
In ISO 19905-1 the available vertical bearing capacity during preloading is denoted VL. The available capacity is
determined by adjusting the gross capacity for the beneficial effect of soil buoyancy, and – if applicable – the detrimental
effect of backfill:
⎧Q + BS with no backfill
VL = ⎨ V (3)
⎩Q V − WBF + BS with backfill

where
4 OTC 23521

WBF = submerged weight of backfill


BS = soil buoyancy of spudcan below bearing area = γ′VD
This definition avoids any potential confusion over the need to include soil buoyancy, even when there is no backfill. The
volume of the protruding underside of the spudcan is given a separate symbol, VD, to distinguish it from Vspud, the total
volume of the spudcan that is displacing soil. ISO 19905-1 clarifies these definitions with a drawing (Figure 3). Note
however that in the key of this figure, the terms “partial backfill” (B) and “full backfill” (C) are used in an unusual way,
referring to the degree of coverage of the top surface of the spudcan, not the degree of filling of the cavity above the spudcan.
Equation (3) from ISO 19905-1 is an elaboration of the corresponding expression in SNAME 5-5A, which reads
VLo = FV − Fo′ A + γ′V (4)
where
Fo′ = effective overburden pressure due to backflow at depth D
V = volume of soil displaced by spudcan
This equation is accompanied by a note in SNAME 5-5A that “the terms −Fo′ A + γ′V should always be considered together.”
This is potentially misleading, however, because even if there is no backflow (such that Fo′ A = 0 ) it is still necessary to
account for the buoyancy of the soil below depth D that is displaced by the protruding underside of the spudcan, which gives
a non-zero γ′V term.
In ISO 19905-1 it is the gross vertical bearing capacity, QV, rather than the available capacity, VL, that forms the basis of
the yield interaction surfaces used for combined V-H-M loading. This change of emphasis from SNAME 5-5A, where the
available capacity VLo is used to size the V-H-M yield surfaces, is highlighted by the inclusion in ISO 19905-1 of a new
equation that clarifies the force balance at the point of maximum preload (subscript = o):
Q Vo = VLo + WBF,o − B S (5)

This is simply a rearranged version of Equation (3). It is recognized, however, that the initial gross capacity established by
preloading may increase (such that QV > QVo) if any subsequent actions such as additional backfilling during operation, or
yielding of the leeward spducan(s) during a storm, result in additional leg penetration.

Backfill in Clay
Perhaps the most significant new feature of the ISO 19905-1 treatment of bearing capacity on clay is the introduction of a
rational method for predicting the depth at which a flow-round failure mechanism will become preferential for a penetrating
spudcan, leading to backfilling of the cavity above (Figure 3). Before outlining the new method, however, it is useful to recall
the existing provisions in SNAME 5-5A.
Unlike ISO 19905-1, SNAME 5-5A uses the term “backflow” in a generic sense, to describe any soil deposited on top of
the spudcan, regardless of the plastic flow or collapse mechanism responsible. In SNAME 5-5A, the main mechanism is
envisaged to be collapse of the cavity walls, with instability predicted for penetrations
Nc us
D≥ (6)
γ′
where
N = stability factor for excavation with vertical sides
cus = undrained shear strength at depth D/2
The main text recommends the plane-strain stability factors of Meyerhof (1972), and states that these should be conservative
for cylindrical excavations. The commentary provides an additional reference to the axisymmetric stability factors of Britto &
Kusakabe (1983), describing these as less conservative but potentially more appropriate, particularly for normally
consolidated strength profiles.
With respect to backflow, the SNAME 5-5A commentary also provides the following caveat: “It should be noted that the
expression in [Equation (6)] is based on static hole stability. In reality, during penetration of the spudcan the soil will
probably flow along the spudcan upwards on to the top of the spudcan. Hence, the hole stability derived from the expression
provided in [Equation (6)] may be too optimistic.” There is no specific guidance, however, for predicting the depth at which
backflow will be initiated by a continuously-penetrating spudcan. The main text does acknowledge that “in very soft clays
complete back-flow may occur”, yet this is inconsistent with Equation (6), which predicts large (mostly > 10 m) stable cavity
depths for a range of soft clay sites in the Gulf of Mexico (Menzies & Roper, 2008). Indeed, for jack-ups operating in the
Gulf of Mexico and other soft clay regions it is common practice to ignore Equation (6) – which actually turns out to be
unconservative – and instead assume complete backfilling of the cavity above the spudcan (Endley, et al., 1981; Young, et
al., 1984; Menzies & Roper, 2008).
OTC 23521 5

ISO 19905-1 uses more precise terminology than SNAME 5-5A: the overall term “backfill” refers to the combination of
“backflow” (plastic flow-round onto the top of the spudcan during preloading) and “infill” (cavity wall collapse during or
after preloading, or gradual sediment deposition). Different notation is used to distinguish between the submerged weights of
the backfill accumulated during preloading (WBF,o) and after preloading (WBF,A). The distinction is important because the first
results in additional plastic penetration during preloading, see Equation (5), whereas the second is likely to be reacted
elastically (inside the V-H-M yield surface) during operation.
In recent investigations of spudcan penetration in clay using centrifuge model tests and continuous-penetration finite
element analyses, Hossain, et al. (2006) and Hossain & Randolph (2009a) identified three distinct mechanisms of soil flow
around the advancing spudcan: (a) outward and upward flow leading to surface heave and formation of a cavity above the
spudcan; (b) gradual flow back into the cavity; and (c) fully localised flow around the embedded spudcan with the cavity
unchanging (see Figure 4). At a certain stage of penetration, soil backflow is initiated and this occurs not because of
instability of the open cavity (as envisaged in SNAME 5-5A) but because of a preferential local flow mechanism for soil
displaced by the downward-moving spudcan. The continual backflow gradually develops a seal above the spudcan and limits
the cavity depth. The subsequent backflow continues below the limiting cavity depth and above the advancing spudcan, i.e.
the initial cavity is not filled up by the backflow process – see Figure 4(c).
As such, the cavity stability approach recommended in SNAME 5-5A (via Equation (6) and associated curves for
determining N) is replaced in ISO 19905-1 by the equation proposed by Hossain & Randolph (2009a):
⎛ ρ⎞
⎜ 1− ⎟⎟
H cav S ⎛ s ⎞ ⎜⎝ γ′ ⎠
= S0.55 − where S = ⎜⎜ um ⎟⎟ (7)
B 4 ⎝ γ′B ⎠
The corresponding design chart from ISO 19905-1 is shown in Figure 5. Although this plot is labelled “Experimental data
and curve-fit”, in fact only a few of the data points (labelled a) are experimental; the remainder (labelled b and c) are derived
from continuous-penetration finite element analyses.
Equation (7) is applicable to a single-layer clay with a clear profile of increasing strength with depth. For multi-layer
clays with moderate changes of strength, ISO 19905-1 provides the following equation from Hossain, et al. (2006):
0.55
H cav ⎛ s uH ⎞ 1⎛s ⎞
= ⎜⎜ ⎟⎟ − ⎜⎜ uH ⎟⎟ (8)
B ⎝ γ ′B ⎠ 4 ⎝ γ ′B ⎠
where suH is the shear strength at the backflow depth, Hcav (see Figure 5). Iteration is required to establish consistent values
for Hcav/B and suH.
Equations (7) and (8) predict that in very soft normally consolidated clay with nearly zero strength intercept at the
mudline, backflow occurs immediately after the full diameter of the spudcan has penetrated into the ground, i.e. complete
backflow is likely to occur. On the other hand, in firm to stiff clay with a significant mudline strength intercept, where
spudcan penetration is expected to be small, the possibility of backflow within that penetration depth diminishes.
Any soil flowing back into the cavity created by spudcan penetration primarily affects the bearing response by (partially)
negating the surcharge component of the gross capacity, QV, by the submerged weight of the backfill, WBF (see Equations (5)
and (6)). It would also be expected that, following the onset of backflow, the bearing capacity factor Nc should increase
because the failure mechanism must now pass through the backfilled soil (though this soil is likely to be heavily remolded).
Hossain, et al. (2006) found that a fully localised flow-round mechanism, as shown in Figure 4(c), occurs at a penetration of
0.7B to 1.0B beyond the onset of backflow.
In addition to affecting the available vertical bearing capacity during preloading, the degree of backfilling influences the
embedment condition of the spudcan and hence the resistance to uplift, horizontal and moment loads. Appropriate
modifications of the V-H-M yield surface (by means of an ‘adhesion envelope’) are discussed below.
Although only their formulae for cavity depth (Equations (7) and (8) respectively) are included in ISO 19905-1, Hossain
& Randolph (2009a) and Hossain, et al. (2006) had presented these in the context of complete proposed methods for
predicting the complete load–penetration response of a spudcan in clay, and while further research and comparisons with
field data for all the analysis methods would appear to be warranted, some discussion of their details here seems appropriate.
Soil strength non-homogeneity and spudcan base roughness were taken into account, and the approaches they developed
account for the evolving soil failure mechanisms (see Figure 4). In particular, the approach after the depth of backflow (D >
Hcav) captures the soil backflow above the advancing spudcan (see Figures 4(b) and 4(c)).
Hossain & Randolph (2009c) also investigated the effects of strain-rate dependent shear strength and gradual remolding,
in an attempt to model real soil behavior more closely. A 15~20% reduction is suggested to apply on the computed vertical
penetration resistance using solutions on ideal non-softening soils. Hossain & Randolph (2008) carried out large deformation
finite element analyses of the 13 case histories presented by Menzies & Roper (2008), and for those cases accounting for the
combined effect of strain rate and softening produced an excellent match with measured field data.
6 OTC 23521

Bearing Capacity in Sand


Evaluating the penetration of a spudcan in silica sand in ISO 19905-1 is based on a drained bearing capacity calculation, a
reasonable assumption as spudcan preloading is usually performed sufficiently slowly to ensure that fully drained conditions
will occur. The drained ultimate vertical bearing capacity of the spudcan is expressed as
Q V = γ ′N γ πB3 / 8 + p′o d q N q πB 2 / 4 (9)

Where γ′ is the effective unit weight of the sand, B is the effective diameter in contact with the soil and Nγ is a dimensionless
bearing capacity factor calculated for the axisymmetric case. Only if the spudcan penetrates past its widest point and the
overburden of soil above this point creates an effective surcharge p′o does the second term lead to additional bearing
capacity. Nq is a dimensionless bearing capacity factor and dq is the depth factor for drained soils, which is assumed as
d q = 1 + 2 tan φ′(1 − sin φ′)2 tan −1 (D B) , where D is the greatest depth of maximum cross-sectional spudcan bearing area.
Theoretical values of Nγ and Nq for a flat, rough circular footing and for soil friction angles from 20° to 40° are directly
provided in the in ISO 19905-1 document. These were calculated with the slip-line method using the ABC program of Dr
C.M. Martin of Oxford University (Martin, 2003). Further lower bound solutions of Nγ that account for circular conical
foundations of angles between 60 and 180° (flat plate), a range of roughness ( tan δ tan φ′ of 0.6 to 1 and where δ is the soil-
steel interface friction angle) and angles of friction of 20° to 40° are also provided in tabulated form in the Appendix of ISO
19905-1 (though these were calculated using the program FIELDS). Although detailed bearing capacity values are provided
it is noted in ISO 19905-1 that the dominant parameter in a penetration analysis in sand is the assumed value of the friction
angle. For instance, a “1° change in φ′ gives at least a 20% change in Nγ” (ISO 19905-1, 2011). Some guidance to choosing
an appropriate φ′ value is provided in the Appendix of ISO 19905-1.
While acknowledging that prediction of the penetration of spudcans in calcareous sands is “likely to be less accurate than
for silica sands”, some limited guidance is provided in ISO 19905-1. One method provided for uncemented calcareous sand is
the use of a bearing modulus, as proposed by Randolph, et al. (1993) and Finnie & Randolph (1994). The method is based on
the results of a series of centrifuge experiments of model footings which indicated that the vertical bearing capacity increases
linearly with depth and an estimation of the bearing pressure can be performed as a function of the overburden pressure rather
than the self-weight. Amongst other reference sources, the formulas of Yamamoto, et al. (2008, 2009) for calculating bearing
capacity based on a compressional deformation mechanism and a punching shear pattern are discussed in the context of
predicting the response of shallow footings on compressible sands. This guidance is an additional component from that
available in SNAME 5-5A.

Penetration in Layered Soils


Methods for calculating the bearing capacity of a spudcan penetrating a layered system are provided in ISO 19905-1, though
there is little difference between advice previously provided by SNAME and now by ISO.
For stratification where the strengths of the layers do not vary significantly, use of average soil strengths and the general
shear method for single layers is recommended. When there is a substantial difference, methods for calculating the squeezing
(soft overlaying strong) or punch-through (strong overlaying soft) of a penetrating spudcan are provided.
Although there has been a series of recent research publications on the potential for punch-through failure (sand over
clay: Teh, et al. 2008, 2009, 2010; Lee, et al., 2009; stiff over soft clay: Edwards & Potts, 2004; Hossain & Randolph, 2009b;
Hossain & Randolph, 2010a,b), the primary methods provided in the SNAME industry guidelines are essentially repeated in
ISO 19905-1. These new methods were not considered by Panel 4 to be the primary recommendation as they were just
recently published and have yet to become acceptably established, even though they have shown to provide superior
performance at least for simplified centrifuge conditions. Further verification of these against offshore penetration data could
prove valuable and is highly recommended.
For instance, the methods provided for calculating the peak penetration resistance of a spudcan on sand overlaying clay
are the loading spreading method and the punching shear mechanism developed by Hanna & Meyerhof (1980). Both were
developed for a flat footing wished into place at one particular depth. In the former the load is assumed to be projected
through the upper sand layer at an angle of spread in the range of 1:5 to 1:3 to an imaginary footing of increased bearing area
at the sand-clay boundary. In the latter, the punching shear failure mechanism is assumed to comprise a truncated cone in the
sand layer being depressed into the underlying clay. However, for calculation purposes the simpler vertical shear plane is
used and the peak resistance Qpeak is calculated following the punching shear method as
2HA
Q peak = Q u , b − AHγ′s + (γ′s H + 2p′0 )K s tan φ (10)
B
and where the punching shear coefficient Ks is related to the normalized shear strength of the underlying clay layer through
the ratio of the bearing capacity of clay to sand and effective angle of internal friction of sand, and is suggested to derive
from the chart proposed by Meyerhof & Hanna (1978).
OTC 23521 7

Both the punching shear and projected area methods were originally developed for shallow wished-in-place footings
(though with modification for application to spudcans by SNAME). However, the mechanisms are significantly different
from post-failure observations of four centrifuge tests performed by Craig & Chua (1990) and the more recent observations
of Teh, et al. (2008). In the latter, digital images were captured continuously by installing a half-spudcan against a transparent
window. Analysed using particle image velocimetry (PIV) coupled with close range photogrammetry corrections, the change
in spudcan failure mechanism with penetration depth (and at stress conditions of equivalent similitude to the offshore case)
was observed. The calculation method of Teh, et al. (2009), based on this research, is briefly described in the Appendix of
ISO 19905-1.
The method for calculating punch-thorough in stiff over soft clays is essentially as in SNAME 5-5A, and follows the
approach of Brown & Meyerhof (1969). However, from an extensive investigation, Hossain & Randolph (2010a,b) showed
that the failure modes assumed by the recommended approaches are not consistent with those observed from centrifuge tests
(again from half-spudcan tests against a transparent window and subsequent quantification of the captured images through
PIV) and large deformation finite element analyses. Severe punch-through was associated with vertically downward soil
displacements beneath the spudcan in the upper layer, involving punching shear, with clear shear planes in the shape of a
truncated cone forming in the upper layer below the spudcan. A concise approach for assessing profiles of spudcan
penetration resistance through stiff-over-soft clay was reported by Hossain & Randolph (2009b), which is referenced in the
Appendix.
On the reverse stratification i.e. soft soil overlying stronger material, the squeezing is suggested to calculate using Brown
& Meyerhof’s (1969) and Vesic’s (1975) factors, but adjusted for embedment depth by applying a depth factor, following the
semi-empirical approach of Skempton (1951).
A bottom up approach for multilayered system is advocated in ISO 19905-1. Firstly the bearing capacity of a spudcan at
the top of the bottom two layers is calculated using the two layer methods provided. These bottom two layers are then treated
as one, and the calculation repeated for the layer above. Current centrifuge observations of soil failure patterns (from
captured images and results from PIV) and load penetration responses provided new insight into spudcan behaviour in
multilayered soils with interbedded stiff clay, carbonate sand and silica sand layer (Hossain & Randolph, 2011; Hossain, et
al., 2011). For instance, a plug depth beneath the advancing spudcan may be accumulated trapping and carrying down soil
even from a softer layer.

Yield Interaction
During operation of a jack-up unit, the foundations experience complex load paths as a result of environmental and inertial
loads acting on the superstructure (legs and hull). Each spudcan is subjected to six load components (vertical, 2 × horizontal,
2 × moment, torsional) but for assessment of the ultimate limit state it is usual to consider a simplified scenario involving just
three load components (vertical, horizontal, moment). These loads are assumed to be co-planar and are denoted V, H, and M.
In ISO 19905-1, assessments of foundation capacity under combined V-H-M loading are based on the concept of a yield
interaction surface. For simplicity, it is assumed that all soil types and spudcan geometries can be treated within a unified
framework, based on a yield surface with the general equation
2 2 2 2
⎡ FH ⎤ ⎡ FM ⎤ ⎡ FV ⎤ ⎡ FV ⎤ ⎡ FV ⎤ ⎡ FV ⎤
⎢ ⎥ +⎢ ⎥ − 16(1 − a )⎢ ⎥ ⎢1 − ⎥ − 4a ⎢ ⎥ ⎢1 − ⎥=0 (11)
⎣ QH ⎦ ⎣ QM ⎦ ⎣ QV ⎦ ⎣ QV ⎦ ⎣ QV ⎦ ⎣ QV ⎦
where
a = shape parameter (0 to 1)
FV, FH, FM = applied loads
QV, QH, QM = maximum load capacities
For the extreme values of a, the general equation simplifies to
2 2 2 2
⎡ FH ⎤ ⎡ FM ⎤ ⎡ FV ⎤ ⎡ FV ⎤
⎢ ⎥ +⎢ ⎥ − 16⎢ ⎥ ⎢1 − ⎥ = 0 (a = 0 : parabola in V) (12a)
⎣ QH ⎦ ⎣ QM ⎦ ⎣ QV ⎦ ⎣ QV ⎦
2 2
⎡ FH ⎤ ⎡ FM ⎤ ⎡ FV ⎤ ⎡ FV ⎤
⎢ ⎥ +⎢ ⎥ − 4⎢ ⎥ ⎢1 − ⎥ = 0 (a = 1 : ellipse in V) (12b)
Q
⎣ H⎦ ⎣ M⎦ Q ⎣ QV ⎦⎣ QV ⎦
Intermediate values of a result in V-H and V-M cross-sections that lie between the two extremes, as shown in Figure 6. On
planes of constant V, the H-M cross-section is always elliptical, regardless of the value of a.
The applied loads FH and FM in Equation (11) are simply the horizontal and moment loads arising from the structural
analysis of the jack-up unit (referred to in ISO 19905-1 as the “assessment load case”). Care is needed with the definition of
FV, since for consistency with QV (which is taken as the gross vertical bearing capacity, see below) it must reflect the total
vertical load felt by the spudcan; this in turn is the total vertical force that needs to be reacted by the soil beneath the spudcan.
8 OTC 23521

The applied load FV is therefore given by the vertical force from the structural analysis, plus the submerged weight of backfill
(if applicable), less the effect of soil buoyancy. This is clarified by the following equation in ISO 19905-1:
⎧Vst − BS with no backfill
FV = ⎨ (13)
⎩Vst + WBF, o + WBF, A − BS with backfill

Apart from Vst, which is the structural vertical load, all other terms on the right-hand side have been defined and discussed
previously in the sections on bearing capacity and backfill.
With regard to the capacity terms in Equation (11), the vertical capacity QV is fundamental. ISO 19905-1 emphasizes that
QV is the gross vertical bearing capacity of the spudcan, given by (for example) Equation (1) for clay and Equation (9) for
silica sand. Unless the spudcan undergoes additional plastic penetration during operation, QV remains equal to the value
established at the point of maximum preload, namely QVo. It is important to recall, however, that the preloading process does
not provide a direct measurement of the gross capacity QVo. Instead it is the available capacity, VLo, that is recorded by the rig
operator as the apparent ‘maximum applied preload’. For a deeply penetrated spudcan in soft clay, partial or complete
backfilling is likely to occur, so QVo will be significantly greater than VLo – see Equation (5). On dense silica sand, a spudcan
is unlikely to penetrate to its maximum diameter, so in this case the backfill term will be zero and QVo will be slightly less
than VLo as a result of soil buoyancy.
ISO 19905-1 provides various formulae for the evaluation of QH and QM in Equation (11). The expressions provided for
clay (where deep embedment and backfilling may occur) are more complicated than those for sand (where embedment is
unlikely to be an issue). With the exception of the horizontal capacity QH at shallow embedment on clay, all horizontal and
moment capacities are expressed as multiples of the vertical bearing capacity. Importantly, however, the relevant multipliers
are not applied to the gross vertical capacity QV, nor to the available vertical capacity VL, but to a third quantity denoted QVnet
and defined as the “net ultimate vertical foundation capacity”:
Q Vnet = Q V − p′o πB 2 4 (14)
Comparing this with Equation (1), for clay, it can be seen that QVnet omits the ‘surcharge’ component of the gross bearing
capacity, but retains the ‘cohesive’ component. This can be defended as a conservative and rational adjustment, and in fact
the same approach has been used previously by Martin & Houlsby (1999) when calculating QH (though not QM). In ISO
19905-1 it is justified on the basis that “For clay, the net vertical bearing capacity is used because the weight of soil on top of
the spudcan does not affect the horizontal and moment capacities.” This wording could be improved, however, because p′o is
the effective overburden pressure at depth D, which is present even if there is no soil on top of the spudcan. On sand, the
adjustment in Equation (14) is less important because it only comes into play when a spudcan has penetrated to its maximum
diameter and become embedded (such that p′o > 0 ). Even then, comparison with Equation (9) shows that the effect is likely
to be small because dqNq is much greater than 1 for typical friction angles. The ISO 19905-1 justification for the use of QVnet
rather than QV on sand is that “the use of net capacity is conservative because it neglects the increase in capacity due to the
weight of any soil on top of the spudcan which would have a beneficial effect on the horizontal and moment capacities.”
In SNAME 5-5A, the V-H-M yield interaction surface used in Step 2b of the foundation stability assessment is
2 2 2 2
⎡ FHM ⎤ ⎡ FM ⎤ ⎡ FVHM ⎤ ⎡ FVHM ⎤
⎢ ⎥ +⎢ ⎥ − 16⎢ ⎥ ⎢1 − ⎥ =0 (15)
⎣ H Lo ⎦ ⎣ M Lo ⎦ ⎣ VLo ⎦ ⎣ VLo ⎦

This is essentially the same as Equation (12a), the parabolic specialization of Equation (11), except that the vertical load
capacity is taken as VLo (the available vertical bearing capacity) rather than QV (the gross vertical bearing capacity). This
important distinction has already been discussed with reference to Equation (5). For consistency with VLo, the applied vertical
load FVHM in Equation (15) must also exclude the effects of backfill and buoyancy; it is thus given by the structural vertical
load alone.
The ‘cigar-shaped’ V-H-M yield interaction surfaces in Equations (11), (12) and (15) are all specialized forms of the
general function proposed by Martin (1994), which allows additional adjustment of the shape along the V axis, as well as the
possibility of a rotated ellipse in the H-M plane (see also Houlsby & Martin, 2000). While these refinements are useful for
curve-fitting to specific sets of experimental data, the full equation would be unnecessarily complex for use in routine
assessments. The suggestion of using a single parameter to interpolate between limiting parabolic and elliptical shapes is due
to Templeton, et al. (2005) and Templeton (2006).

Horizontal and Moment Capacity


Clay. At shallow embedment on undrained clay, ISO 19905-1 gives the horizontal capacity as
Q H,shallow = s uo A + (s uo + s u ,l )A s (16)

where
OTC 23521 9

suo = undrained shear strength at depth D


A = bearing area = πB2/4
su,l = undrained shear strength at tip level
As = horizontally projected area in contact with soil
The first term is the basic (flat footing) sliding resistance, while the second term accounts for passive resistance. At deep
embedment (D/B ≥ 2.5) the horizontal capacity is taken to be a multiple of QVnet:
⎛ s ⎞⎛ A ⎞
Q H ,deep = ⎜⎜1 + u ,a ⎟⎟⎜ 0.11 + 0.39 s ⎟Q Vnet (17)
⎝ s uo ⎠⎝ A ⎠

where
su,a = undrained strength of disturbed/remolded backfill above spudcan
This equation is a modified version of an expression proposed by Templeton (2009), based on a parametric study using 3D
finite element analysis. To appreciate the significance of this new guidance, curves of the horizontal capacity multiplier
QH,deep/QVnet are plotted in Figure 7(a). These show that for a spudcan with a large projected area ratio, covered by backfill
that is assumed to retain a significant fraction of the undisturbed strength, the capacity multiplier can be quite high. For
example, As/A = 0.4 and su,a/suo = 0.5 gives QH,deep/QVnet = 0.4, compared with the conservative value of 0.11 that is obtained
by ignoring all backfill strength and passive resistance.
In SNAME 5-5A, the following equation is given for the horizontal capacity HLo to be used in the yield interaction
surface of Equation (15):
H Lo = c uo A + (c uo + c ul )A s (18)
Apart from notational differences, Equation (16) is identical to Equation (18), so for a shallow spudcan ISO 19905-1 gives
the same horizontal capacity as SNAME 5-5A. This is not the case at deep embedment, however. Whereas ISO 19905-1
adopts a new expression, namely Equation (17), SNAME 5-5A continues to use Equation (18). This makes a direct
comparison difficult, but if complete backfilling is assumed and soil buoyancy is neglected, such that VLo ≈ 9cuoA, and if cul ≈
cuo in Equation (18), it can be shown that SNAME 5-5A implies
⎛ A ⎞
H Lo = ⎜ 0.11 + 0.22 s ⎟VLo (19)
⎝ A⎠
for a deeply embedded spudcan. If it is further assumed that VLo (SNAME) is approximately equal to QVnet (ISO), which is
reasonable for conditions of deep embedment and complete backfilling, it is possible to plot Equation (19) in Figure 7(a). It is
clear that the use of ISO 19905-1 to calculate horizontal load capacity at depth will be significantly less conservative than the
use of SNAME 5-5A, but this is believed to be justified on the basis of numerical analysis by Templeton (2009).
The moment capacities given in ISO 19905-1 for shallow and deep (D/B ≥ 2.5) spudcans on clay can be expressed in the
form
Q M,shallow = 0.1BQVnet (20)

and
⎡ s ⎛ H ⎞⎤
Q M ,deep = ⎢0.15 + 0.025 u ,a ⎜1 − cav ⎟⎥ BQ Vnet (21)
⎣ s uo ⎝ D ⎠⎦

Curves of the moment capacity multiplier QM,deep/BQVnet are plotted in Figure 7(b). With the most optimistic assumptions, i.e.
complete backfilling and no degradation of backfill strength (Hcav/D = 0, su,a/suo = 1), the moment capacity multiplier is 0.175.
Somewhat surprisingly, this is only 17% greater than the value of 0.15 obtained with the most pessimistic assumptions
(Hcav/D = 1 and/or su,a/suo = 0). The basis of Equation (21) is not discussed in ISO 19905-1.
In SNAME 5-5A, the moment capacity MLo is assumed to be independent of embedment, and is given by
M Lo = 0.1BVLo (22)
For the case of a surface spudcan, where QVnet and VLo are identical, Equations (20) and (22) show that ISO 19905-1 gives
the same moment capacity as SNAME 5-5A. This is not the case at deep embedment, where ISO 19905-1 uses Equation (21)
but SNAME 5-5A continues to use Equation (22). Assuming once more that VLo (SNAME) is approximately equal to QVnet
(ISO), which is reasonable for conditions of deep embedment and complete backfilling, it is possible to compare Equations
(21) and (22) on the same axes in Figure 7(b). This shows that ISO 19905-1 permits significantly enhanced predictions of
moment capacity at depth, compared with SNAME 5-5A. It would be useful if a future edition of ISO 19905-1 provided
some indication of the basis for the new approach.
10 OTC 23521

To handle intermediate embedments in clay, various interpolation formulae are used in ISO 19905-1 to ensure that the
horizontal and moment capacities undergo a smooth transition from QH,shallow and QM,shallow to QH,deep and QM,deep. In addition,
the overall shape of the yield interaction surface in Equation (11) undergoes a gradual change from parabolic at shallow
embedment (a = 0 when D/B = 0) to elliptical at deep embedment (a = 1 when D/B ≥ 2.5). The basis for this is primarily the
3D finite element work reported by Templeton, et al. (2005) and Templeton (2006), which investigated the V-M yield
interaction locus for a spudcan embedded in clay at various embedment ratios (D/B = 0.3, 1.2 and 1.7).
ISO 19905-1 provides an additional clause allowing the V-H-M yield surface to modified in the region of low vertical
load: “The yield surface in the region 0 < FV/QV < 0.5 (typically applicable to windward legs) can be replaced by an adhesion
envelope that provides additional horizontal and moment capacity due to spudcan-soil adhesion.” The modification is
essentially the same as that in SNAME 5-5A, apart from some enhancements necessitated by the introduction of the shape
parameter, a, in the expression for the main V-H-M yield surface (Equation (11)).

Sand. In ISO 19905-1, the horizontal and moment capacities QH and QM for spudcans on sand are calculated as
Q H = 0.12Q Vnet (23)

Q M = 0.075BQ Vnet (24)


The shape parameter for use in Equation (11) is fixed at a = 0, i.e. parabolic cross-sections in the V-H and V-M planes. The
capacity multipliers for horizontal and moment loading have been carried over directly from SNAME 5-5A, where the
corresponding equations are
H Lo = 0.12VLo (25)

M Lo = 0.075BVLo (26)
The multipliers are empirical, and have appeared in SNAME 5-5A since Revision 2 (2002). They are based on the results of
model tests by various researchers including Tan (1990), Gottardi & Butterfield (1993), Gottardi, et al. (1999) and Byrne &
Houlsby (2001). Other recent model testing studies cited in ISO 19905-1 (Bienen, et al., 2006; Cassidy, 2007) have not
suggested any need to adopt different multipliers.
In the capacity equations above, the comparison between ISO 19905-1 and SNAME 5-5A is not quite like-for-like
because QVnet is not identical to VLo. However, because a spudcan on sand typically fails to penetrate to its maximum
diameter, all terms related to surcharge ( p′o ) and backfill weight (WBF) reduce to zero. Equations (5) and (14) then show that
the net and available capacities differ only by the soil buoyancy term: Q Vnet = VLo − BS .

Stiffness and Stiffness Reduction


Initial stiffness. In order to determine the level of combined loads on the spudcans and the overall response of the jack-up,
an assumption of the spudcan stiffness must be incorporated into the structural analysis of the jack-up during a storm loading
analysis. As stated in ISO 19905-1, the “foundation analysis under time-varying loading requires knowledge of the load-
deflection behaviour of the soil. This is usually described by spring stiffnesses in the vertical, horizontal and rotational
modes.”
The new ISO recommends that the initial stiffnesses of the spudcan be estimated from the solutions for a rigid circular
plate on an elastic half-space, but with modifications to account for spudcan embedment, combined with an appropriate shear
moduli representing the stress level under the spudcan. Dimensionless elastic stiffness coefficients are provided in tables with
the modifications for embedment from the finite element study of a thin plate footing in an isotropic elastic and homogenous
soil by Bell (1991). These coefficients remain limited to the two conditions of no-backflow or compete-backflow, and
furthermore, no advice on inclusion of the cross coupling of rotational and horizontal stiffness is provided. However, users
are proposed the option of developing their own “continuum model” if required.
The initial stiffness, however, is more sensitive to the soil shear modulus assumed than the dimensionless coefficients,
and an appropriate value is still one of the most difficult parameters to establish. Because the value should represent typical
conditions under the spudcan, as the mobilised shear stiffness of soil is strongly dependent on the shear strain, the
recommendations now incorporated in ISO 19905-1 are based on findings of back-analysis of case records of jack-up
platforms in the North Sea (Cassidy, et al. 2002, Noble Denton Europe & Oxford University 2006). These have replaced the
original recommendations in SNAME 5-5A, which were based on small scale laboratory and centrifuge experiments.
The site records contained the dynamic behaviour and environmental loading conditions of three jack-ups, the Santa Fe’s
Magellan, Monitor and Galaxy-1, as reported in Temperton, et al. (1997) and Nelson, et al. (2000). Using this data in a
project sponsored by the IADC, Noble Denton and Oxford University conducted the back analysis of eight sites of varying
soil conditions (three clay and five sand sites) and water depth (between 28 and 98 m). The jack-ups had been subjected to
substantial storms, with significant wave height Hs of 4.1-9.85 m (Nelson, et al. 2000, Cassidy, et al. 2002). For each storm,
horizontal deck movements (evaluated by double integration of the accelerometer measurements) of the jack-up and recorded
sea-state and wind speed and direction were used to compare with numerical simulations. A suite of random time domain
OTC 23521 11

analyses were performed for each site with adjustments in soil stiffness made, until the best fit stiffness levels were derived
and the representative shear modulus calculated.
From the recommendations of this study (Cassidy, et al. 2002, Noble Denton Europe & Oxford University 2006) the
following formulations for the shear modulus for clay and sand seabeds are provided in ISO 19905-1. The shear modulus for
clay can be determined by
G = Irsu (27)
where su is the undrained shear strength measured at 0.15 diameters below the reference point of the spudcan (taken at the
level at which the maximum diameter is reached). Ir is the rigidity index and can be calculated from
G 600
Ir = = (28)
s u R OC 0.25

where ROC is the overconsolidation ratio of the clay.


In sands the shear modulus can be estimated by
0.5
G ⎛V ⎞
= j⎜⎜ sw ⎟⎟ (29)
pa ⎝ Ap a ⎠
where Vsw is the gross vertical spudcan reaction inclusive of backfill under still water conditions (the reaction that would be
obtained if the jack-up were supported on an infinitely rigid foundation, plus the reaction due to the submerged weight of any
backfill on the spudcan, less the submerged weight of soil displaced by the spudcan below D, the greatest depth of maximum
cross-sectional spudcan bearing area below the sea floor), A is the spudcan contact area and pa is atmospheric pressure
(typically taken as 101.3 kPa). The recommended value for the dimensionless stiffness factor j for a relative density DR (in
percent) is
⎛ D ⎞
j = 230⎜ 0.9 + R ⎟ (30)
⎝ 500 ⎠
For selection of an appropriate shear modulus in layered soils reference to Ueshita & Meyerhof (1967) is provided in ISO
19905-1. For deeply embedded spudcans where the truss work legs are providing additional stiffness, the method described
by Brekke, et al. (1989) is provided as a reference to be followed. In this method the leg is assumed to be acting like a pile
with appropriate load deflection P-y curves to be defined.

Reduction in stiffness. ISO 19905-1 states that “the reduction in stiffness as the spudcan reactions approach or exceed the
yield surface shall be included in the analysis” and allows for “different approaches to determining the softening of the
stiffnesses.” The most comprehensive approach is to contain the reduction of stiffness within the soil model, such as within
the theoretical framework of work-hardening plasticity theory as encapsulated by the so called force-resultant models for
spudcan-soil interaction (e.g. Schotman, 1989; Van Langen et al, 1999; Martin & Houlsby, 1999, 2001; Cassidy & Houlsby,
1999; Houlsby & Cassidy, 2002; Cassidy, et al. 2002; Bienen, et al., 2006).
If access to such models is not available, ISO 19905-1 provides provision to reduce the rotational secant stiffness with a
reduction formula. In this approach the vertical and horizontal stiffness remain unchanged. The rotational stiffness initially
assumed is reduced (after the analysis) based on the distance of the highest combined loads to the yield capacity envelope.
The analysis must be rerun until the reduced stiffness is compatible with the reduction formulation provided. This is known
as a Step 2b check, with full details of the various levels of acceptance checks provided in a latter section of this paper. The
method of degrading the rotational stiffness is essentially the same as SNAME 5-5A, but the formulation is different. In the
original version of SNAME, separate stiffness degradation functions were originally prescribed for sand and for clay, but in
Revision 2 a new single formula was introduced for both sand and clay. This single formula was seen as a compromise for
sand and clay, which better approximated some available data, but the single formula was quite tedious, and it did not allow
recovery of an underlying moment-rotation relation, as did the original formulae. In the new ISO code the rotation stiffness
reduction factor is given by:

fr =
(1 − n )rf (31)
ln ((1 − n.rf ) (1 − rf ))
with the equation variable n controlling the rate of rotational stiffness degradation as a function of the failure ratio, rf, which
is the measure of the proximity of the combined footing loads to the yield surface. As stated in the ISO document: “The
parameter, n, accommodates spudcan rotation resistance curves with various degrees of curvature change. In practice the
value of this parameter should be set to suit the best available data (either empirical or analytical) applicable to the jack-up
and site. Finite element analysis for the Gulf of Mexico clay indicates the range of n to be −0.25 to −1.0, with n = −0.5
12 OTC 23521

providing the best overall representation.” However, the recommendation for sites with limited data is for n to be set at zero
and the equation to be simplified to
− rf
fr = (32)
ln (1 − rf )
Certain particular values of the parameter, n, result in simple formulae with special interest.
• The value, n = 0, (used for the last formula above) results in the original SNAME formula for sand, which is related to an
underlying exponential moment-rotation relation.
• The value, n = +1, results in the original SNAME formula for clay, which is related to an underlying hyperbolic moment-
rotation relation.
• The value, n = −1, results in the form related to an underlying hyperbolic tangent moment-rotation relation – with a
relatively sharp curvature which may be suitable for overconsolidated conditions.
More details of this new reduction formulation, its derivation and applicability, can be found in Templeton (2007). An
application of this ISO 19905-1 reduction in a pushover analysis of a jack-up in clay, with comparison to the same analysis
with a force-resultant plasticity model, can be found in Purwana, et al. (2012).

Acceptance Checks
ISO 19905-1 framework for foundation stability assessment has evolved from the SNAME 5-5A recommended methodology
with some revisions made for consistency and more details added for clarity. In this framework, Level 1, 2, 3 checks are
introduced in order of increasing complexity and reducing conservatism. Figure 8 illustrates the general approach of the
foundation stability assessment and the principle of each level of checks. A higher level of check, associated with more
sophisticated modeling of the foundation, can be performed if a lower level fails to meet the acceptance criteria.
It is also important to note that the partial resistance factor (being greater than or equal to unity) in ISO 19905-1
foundation acceptance check is applied as a denominator to be consistent with LRFD framework. This is in contrast to
SNAME 5-5A in which the partial resistance factor (being less than unity) serves as a multiplier to the ultimate resistance
although both approaches essentially produce similar factored resistance. In more detail, each level of check is described
below.

Level 1: Preload check. Level 1 check, which is often called “preload check” in SNAME 5-5A, serves as a quick means of
check using the simplest foundation stability and response analysis model, i.e. pinned conditions for all the spudcans. In this
simplified check, the bearing capacity and sliding capacity (in terms of the V-H capacity envelope) are not explicitly involved
in the check although they are used as the basis for deriving the safe envelope. The check involves the assessment of the
vertical storm reaction within the restriction of a limiting horizontal reaction. Using different formulation from that in
SNAME 5-5A, the limiting horizontal reaction in ISO 19905-1 ensures the safe (allowable) envelope is within the foundation
capacity.
The Level 1 check comprises a Step 1a check for the leeward leg and a Step 1b for the windward leg. In Step 1a check,
the maximum vertical reaction (in gross term) of the leeward leg is assessed against the “maximum allowable” reaction being
the factored pre-load capability taking due account of backfill. The preload resistance factor of 1.10 is adopted giving a
comparable safety factor with that in SNAME 5-5A (1/1.10 = 0.91). In combination with Step 1a, Step 1b is used to check
the minimum vertical reaction of the windward leg. Unlike in SNAME where at this level of check the sliding line must be
first established, the vertical reaction shall not be less than a “minimum allowable” vertical reaction that is defined solely as a
function of the gross ultimate vertical foundation capacity QV in ISO 19905-1. With the same level of safety margin as in
Step 1a, the limiting vertical reaction in Step 1b is conservative and valid for sliding friction angle equal to or greater than
25°. For friction angles less than 25° in which case the sliding line is likely to intersect the vertical-horizontal bearing
capacity envelope, the sliding check in Step 2a (to be described later) should be exercised.

Level 2: Bearing capacity and sliding capacity check. In Level 2 check, a more complex assessment method can be
undertaken or should be undertaken in case the Level 1 check or its restrictions fail in which the spudcan reactions are
checked against the bearing capacity and sliding capacity in terms of the V-H envelope. There are some major differences
from SNAME 5-5A in Level 2 check implementation. In ISO 19905-1, a single formula is used consistently to construct both
the yield surface (for fixity iteration when applicable) and V-H bearing capacity envelope. The equation is expressed mainly
as a function of QV, QH, QM and applicable for both sand and clay. Apart from the consistency of foundation capacity used in
modeling the soil-structure interaction (fixity) and assessing the foundation stability, the simple and closed-form equation
makes it easier to construct the envelope. For a given preload base reaction, sensitivity to the input spudcan penetration depth
and contact diameter to ensure proper construction of the envelope, which is often encountered in constructing the bearing
capacity formula in SNAME 5-5A, can be avoided.
Another major change included in ISO 19905-1 is the method for constructing the factored foundation capacity. In the
development of ISO, it was realized that different analysts were using a different scaling origin when implementing SNAME
with some taking zero as the origin while others used the still water reaction. In order to ensure the scaling is performed in a
OTC 23521 13

consistent way, the factored capacity is now established by scaling down the unfactored capacity with respect to zero net
reaction (FH = 0, FV = WBF – BS). A resistance factor of 1.10 for V-H bearing capacity is used in ISO 19905-1 to achieve a
similar level of safety to what is being implemented in SNAME 5-5A and as a result of the change of method to constructing
factored foundation capacity. Equally importantly, this resistance factor also helps ensure the philosophy of the check in that
a higher level check results in a reduced conservatism. Unlike in SNAME, the use of a single resistance factor for bearing
capacity regardless of spudcan contact condition is intended to avoid a step change in the resulting factored resistance at
transitional points between partial and full contact conditions. For the sliding resistance, the factored sliding capacity is
developed by scaling only the horizontal coordinate of the unfactored capacity as is implemented in SNAME. The same
resistance factor for sliding capacity as in SNAME is adopted. The inclusion of soil passive resistance is only applicable for
the sliding resistance and is generally not allowed for the bearing capacity of sand in ISO.
ISO 19905-1 also introduces a new approach for calculating the utilization of foundation capacity and sliding capacity.
This in no way will change the assessment results, in terms of “pass” or “fail”, but will ensure consistent interpretation of the
utilization check. Although the origin of the vector is arbitrary, for consistency and to help produce a meaningful value of the
resulting utilization the origin of the vectors should be taken from 0.5QV / γR,VH . This applies to both the bearing capacity and
sliding capacity checks as illustrated in Figure 9. It is worth noting that SNAME 5-5A uses the still water reaction as the
origin of the vectors for the bearing capacity utilization while the sliding capacity utilization is determined as the ratio
between the horizontal reaction and the horizontal capacity under the corresponding vertical reaction.
In the ISO 19905-1 Level 2 check, there are a total of three sub-level checks as opposed to two in SNAME 5-5A. As in
SNAME, Step 2a is applicable for analyses with pinned conditions for all the spudcans while fixity conditions with non-
linear rotational stiffness is adopted in Step 2b. Step 2c is added in ISO as a higher level check in which full non-linear
stiffness (both translational and rotational) or non-linear load-displacement models can be used. No additional foundation
stability check is required in Step 2c since the foundation capacity has been implicitly assessed through the compliance with
unfactored yield surface. The resulting displacement, however, needs to be checked in Level 3 (Step 3a). The combination of
Step 2c and Step 3a resembles Step 3 check of SNAME (also defined as “displacement check”).

Level 3: Displacement check. A new development in ISO 19905-1 foundation check framework includes the more elaborate
Level 3 check. Unlike in Level 2 check in which the assessment is focused on identifying the utilization level of the
foundation capacity, the foundation and structural assessments are essentially coupled in Level 3 check. In ISO, more
guidance and details of Level 3 applications are provided.
In Step 3a check, spudcan additional penetration resulting from an overloading situation in Level 2 is checked against the
structural tolerance. A simplified method for determining the additional penetration is provided. The guidance placed in ISO
in this respect is essentially the same as what has been used in practice. The first step in practicing Step 3a check is to
identify the “equivalent” preload level which would be required to expand the unfactored V-H bearing capacity (used in
Level 2) such that the factored envelope fits the factored storm reactions. The leg penetration curve is then used to predict the
additional penetration associated with the “equivalent” preload level for which the structural tolerance should then be
sufficient to accept the effects of this displacement
In addition to the bearing capacity failure, a simplified framework for further check is also given in Step 3a check in the
case of sliding failure. Where a sliding “failure” of the windward leg is encountered in Level 2, the redistribution of loads can
be investigated in Level 3 by considering the horizontal reaction of the windward leg to be limited by the allowable sliding
capacity. The consequences of “allowing” the windward leg sliding to the foundation and structure should be evaluated.
Step 3b constitutes the most rigorous approach in case all the lower level checks are not satisfied. It consists of numerical
analysis or finite-element approach of the complete jack-up structure and non-linear coupled foundation in which large
displacement effects are accounted for. The direct consequences of any foundation displacement to the structure is inherently
included and simulated in the model.

Additional considerations
This paper has concentrated on the most important details of the new and improved foundation assessment provisions of ISO
19905-1. This ISO Standard includes a large number of other new provisions which could not be detailed here. A few,
however, merit some mention. Notable among those are provisions related to foundation cyclic and dynamic behavior,
namely those for fatigue, earthquakes and damping.
The stiffnesses of the spudcan foundations are a function of the soil properties, the strain amplitudes and loading history.
As a consequence, the foundation modeling should consider upper and lower bound stiffnesses as appropriate. Typically the
fatigue assessment of the spudcan and lower part of the leg requires the use of upper bound stiffness while the fatigue
assessment for the upper leg and the leg-to-hull interface requires lower bound stiffness. Although the foundation stiffness
varies as a function of the reactions beneath the spudcan, the variation is unlikely to be of significance except, possibly, for
low-cycle fatigue. Foundation stiffnesses are generally assumed to be linear in smaller sea states. A check of non-linearity
should be performed to validate this assumption for higher sea states.
In the case of assessment for earthquake conditions, ISO 19905-1 states “When fixity brings the structural natural period
closer to the excitation frequency, the inclusion of foundation fixity can amplify the response and shall therefore be
14 OTC 23521

considered.” This marks the first time that a standard or recommended practice has required the inclusion of fixity in jack-up
assessment for earthquakes.
ISO 19905-1 also provides substantially more guidance related to foundation damping than was previously given in
SNAME 5-5A. Both radiation damping and hysteretic damping are explicitly discussed. In the case of radiation damping in
earthquake assessments, specific recommendations are made based on the work of Lysmer and Richart (1966). In the case of
assessments even for storm loading, 19905-1 provides, “Foundation hysteretic damping can, in certain situations, increase the
2% small strain foundation damping … and is discussed further in ISO/TR 19905-2.” The current draft of 19905-2
(Technical Reference), in turn, provides a specific method and procedure for determination of hysteretic damping based on
Templeton & Lewis (2011).

Summary
This paper has discussed important new foundation assessment provisions of ISO 19905-1, Site-Specific Assessment of
Mobile Jack-Up Units. The new foundation assessment requirements of ISO 19905-1 benefit from a fresh approach,
comprehensive reformulation and incorporation of substantial new research, particularly over the past two decades. Notable
among the new or improved provisions include:
• Site Investigations
o Significantly expanded guidance for geophysical and geotechnical site surveys
• Backfill of the Hole above Spudcan
o A new distinction between the backflow and infill components of backfill
o A new method for determination of the backflow component of backfill
o Improved specific accounting for the effects of weight and strength of the backfill soil
• Yield Surface Models
o Changes to the yield surface formulation to allow better accuracy for deep penetrations
o Clarification of the yield surface formulae by the use of terms and symbols consistent with other technical
disciplines
o Additional clarification of yield surface formulae via better specification of the capacity components involved
• Bearing Capacity Formulation
o Incorporation of a new bearing capacity method
o Clarification of the distinction between net and gross bearing capacities
o New horizontal capacity for deep penetrations in clay
o New moment capacity for deep penetrations in clay
• Spudcan Foundation Stiffness
o Verification of the new initial stiffness recommendations based on back-calculation of monitored jack-up units
offshore
o Increases in initial stiffness for clay via improved recommendations for modulus of rigidity
o Unification of stiffness reduction formulation via a single generalized form for use with all soil types
• Foundation Acceptance Check
o Logical and technical improvements in the checking process for foundation acceptance
Not only are the new foundation assessment requirements of ISO 19905-1 better grounded in current research, they are less
restrictive and more realistic while remaining reliably conservative.
There has of course been further research published, but in a time frame too late for it to be considered accepted by all
practitioners. It was therefore not included in this ISO 19905-1 document. However, further verification of this, and indeed
the current ISO 19905-1 recommendations, with monitored installation, in-situ behavior in storms and retrieval is considered
essential for the continual improvement of this ISO standard.

Acknowledgment
The new and improved foundation assessment methodologies would not have been possible without the tireless effort put
forth by the current and former members of Panel 4, consisting of experienced geotechnical practitioners in the industry and
well-respected researchers in academia, and the support and understanding of their organizations in recognizing the
importance of this International Standard. Contributions from participating oil companies and organizations such as OGP
(International Association of Oil and Gas Producers) and IADC (International Association of Drilling Contractors) Jack-up
Committee are very much appreciated. Finally, the authors thank WG7 Convenor Mike Hoyle, WG7 Editing Review Panel
and SC7 Committee for their guidance, support and patience.

References
Bell, R.W. (1991) The analysis of offshore foundations subjected to combined loading. M.Sc. Thesis, University of Oxford.
Bienen, B., Byrne, B.W., Houlsby, G.T. & Cassidy, M.J. (2006). Investigating six degrees of freedom loading of shallow foundations on
sand. Géotechnique, 56(6), 367-379.
Brekke, J. N., Murff, J. D., Campbell, R. B. and Lamb, W. C. (1989) Calibration of jack-up leg foundation model using full scale structural
measurements. OTC 6127, Proceedings, 21st Offshore Technology Conference, Houston, 49-58.
OTC 23521 15

Brinch Hansen, J. (1970). A revised and extended formula for bearing capacity. Bulletin No. 28, Danish Geotechnical Institute,
Copenhagen, 5-11.
Britto, A.M. & Kusakabe, O. (1983). Stability of axisymmetric excavations in clays. J. Geotech. Eng., ASCE, 109(5), 666-681.
Brown, J. D. and Meyerhof, G.G. (1969) Experimental study of Bearing Capacity in Layered Soils. Proceedings, 7th ICSMFE. Vol. 2.
Byrne, B.W. & Houlsby, G.T. (2001). Observations of footing behaviour on loose carbonate sands. Géotechnique, 51(5), 463-466.
Cassidy, M.J. (2007). Experimental observations of the combined loading behaviour of circular footings on loose silica sand.
Géotechnique, 57(4), 397-401.
Cassidy, M.J. and Houlsby, G.T. (1999) On the modelling of foundations fro jack-up units on sand. OTC 10995, Proceedings, 31st
Offshore Tech. Conf., Houston.
Cassidy, M.J., Byrne, B.W. and Houlsby, G.T. (2002) Modelling the behaviour of circular footings under combined loading on loose
carbonate sand. Géotechnique, Vol. 52, No. 10, pp. 705-712.
Cassidy, M.J., Houlsby, G.T., Hoyle, M.J.R. and Marcom, M.A. (2002) Determining appropriate stiffness levels for spudcan foundations
using jack-up case records. Proceedings, 21st Int. Conf. On Offshore Mechanics and Arctic Engineering (OMAE)., Oslo, Norway .
Craig, W. H. and Chua, K., (1990) Deep penetration of spudcan foundations on sand and clay. Géotechnique , Vol. 40, No. 4, pp. 541-556.
Davis, E.H. & Booker, J.R. (1973). The effect of increasing strength with depth on the bearing capacity of clays. Géotechnique, 23(4), 551-
563.
Dean, E.T.R. (2008). Consistent preload calculations for jackup penetration in clays. Candidan Geotechnical Journal, 45, 705-714.
Edwards, D.H. and Potts, D.M. (2004) The bearing capacity of a circular footing under ‘punch through’ failure. Proceedings, Int. Symp.
Num. Models in Geomech., (NUMOG), Ottawa, 493-498.
Endley, S.N., Rapoport, V., Thompson, P.J. & Baglioni, V.P. (1981). Prediction of jack-up rig footing penetration. OTC 4144,
Proceedings, Offshore Technology Conf., Houston.
Finnie, I.M. and Randolph, M.F. (1994) Bearing response of shallow foundations in uncemented calcareous soil, Proceedings, Int. Conf.
Centrifuge ’94, Singapore, pp. 535-540.
Gottardi, G. & Butterfield, R. (1993). On the bearing capacity of surface footings on sand under general planar loads. Soils and
Foundations, 33(3), 68-79.
Gottardi, G., Houlsby, G.T. & Butterfield, R. (1999). The plastic response of circular footings on sand under general planar loading.
Géotechnique, 49(4), 453-470.
Hanna, A.M. and Meyerhof, G.G.(1980) Design Charts for Ultimate Bearing Capacity of Foundations on Sand Overlying Soft Clay.
Canadian Geotechnical Journal Vol. 17.
Hossain, M.S. & Randolph, M.F. (2008). Overview of spudcan performance on clays: current research and SNAME. Proceedings, 2nd
Jack-up Asia Conf. and Exhibition, Singapore.
Hossain, M.S. & Randolph, M.F. (2009a). New mechanism-based design approach for spudcan foundations on single layer clay. J.
Geotech. Geoenv. Eng., ASCE, 135(9), 1264-1274.
Hossain, M.S. & Randolph, M.F. (2009b). New mechanism-based design approach for spudcan foundations on stiff-over soft-clay.
OTC19907, Proceedings, Offshore Technology Conference, Houston.
Hossain M.S. & Randolph M.F. (2009c). Effect of strain rate and strain softening on the penetration resistance of spudcan foundations on
clay. Int. J. Geomechanics, ASCE, 9(3), 122-132.
Hossain, M.S. & Randolph, M.F. (2010a). Deep-penetrating spudcan foundations on layered clays: centrifuge tests. Géotechnique, 60(3),
157-170.
Hossain, M.S. & Randolph, M.F. (2010b). Deep-penetrating spudcan foundations on layered clays: numerical analysis. Géotechnique,
60(3), 171-184.
Hossain, M.S. & Randolph, M.F. (2011). Spudcan foundations on multi-layered soils with interbedded sand and stiff clay layers.
Proceedings, 21st Int. Offshore and Polar Engineering Conference, Maui, Hawaii 2, 463-470.
Hossain, M.S., Randolph, M.F., Hu, Y. & White, D.J. (2006). Cavity stability and bearing capacity of spudcan foundations on clay.
OTC17770, Proceedings, Offshore Technology Conf., Houston.
Hossain, M.S., Randolph, M.F. & Saunier, Y.N. (2011). Spudcan deep penetration in multi-layered fine-grained soils. Int. J. Physical
Modelling in Geotechnics, 11(3), 100-115.
Houlsby, G.T. and Cassidy. M.J. (2002) A plasticity model for the behaviour of footings on sand under combined loading, Géotechnique,
Vol. 52, No. 2, pp. 117-129.
Houlsby, G.T. & Martin, C.M. (2003). Undrained bearing capacity factors for conical footings on clay. Géotechnique, 53(5), 513-520.
Houlsby, G.T. & Wroth, C.P. (1983). Calculation of Stresses on Shallow Penetrometers and Footings. Proceedings, IUTAM/IUGG Symp.
on Seabed Mechanics, Newcastle upon Tyne, 107-112.
ISO 19905-1 (2012). Site-specific assessment of mobile offshore units – Part 1: Jack-ups. International Organization for Standardization, to
be issued.
Lee, K.K., Randolph, M.F. and Cassidy, M.J. (2009) New simplified conceptual model for spudcan foundations on sand overlying clay
soils. OTC20012, Proceedings, 41st Offshore Technology Conference, Houston.
Lysmer, J. and Richart, F. E., (1966) Dynamic Response of Footings to Vertical Loading, J. Soil Mec. and Found. Div., 92(SM1) 65-91.
Martin, C.M. (1994). Physical and numerical modelling of offshore foundations under combined loads. DPhil thesis, University of Oxford.
Martin, C.M. (2003) User guide for ABC-Analysis of Bearing Capacity. Department of Engineering Science, University of Oxford, Report
No. OUEL 2261/03, 2003, available from http://www-civil.eng.ox.ac.uk/people/cmm/software/abc/
Martin, C.M. & Houlsby, G.T. (1999). Jackup units on clay: structural analysis with realistic modelling of spudcan behaviour. OTC 10996,
Proceedings, Offshore Technology Conf. Houston.
Martin, C.M. & Houlsby, G.T. (2000). Combined loading of spudcan foundations on clay: laboratory tests. Géotechnique, 50(4), 325-337.
Martin, C.M. & Houlsby, G.T. (2001). Combined loading of spudcan foundations on clay: numerical modelling. Géotechnique, 51(8), 687-
699.
16 OTC 23521

Menzies, D. & Roper, R. (2008). Comparison of jackup rig spudcan penetration methods in clay. OTC 19545, Proceedings, Offshore
Technology Conf., Houston.
Meyerhof, G.G. (1972). Stability of slurry trench cuts in saturated clay. Proceedings, Speciality Conf. on Performance of Earth and Earth
Supported Structures, ASCE, 1451-1466.
Meyerhof, G.G. and Hanna, A.M. (1978) Ultimate bearing capacity of foundations on layered soils under inclined load. Canadian
Geotechnical Journal 15, No. 4, 565-572.
Nelson, K., Smith, P., Hoyle, M., Stoner, R., Versavel, T. (2000) Jack-up response measurements and the under-prediction of spud-can
fixity by SNAME 5-5 A. OTC 12074, Proceedings, Annual Offshore Tech. Conf., Houston.
Noble Denton Europe and Oxford University (2006) The calibration of SNAME spudcan footing equations with field data. Report No.
L19073/NDE/mjrh, Rev 5, London, November 2006, available from: http://www.nodent.co.uk/iadc/fixity/
Purwana, O.A., Perry, M.J., Quah, C.k., Cassidy, M.J. (2012) Comparison of ISO 19905-1 framework and plasticity based spudcan model
for jack-up foundation assessments. OTC 23225, Proceedings, 44th Offshore Technology Conference, Houston, USA.
Randolph, M.F., Finnie, I.M. and Joer, H. (1993) Performance of shallow and deep foundations on calcareous soul. Proceedings, Symp. On
Foundations of Difficult Soils, Kagoshima, Japan.
Salençon, J. & Matar, M. (1983). Bearing capacity of circular shallow foundations. In Foundation Engineering, 1, 159-168. Presses de
l'Ecole Nationale des Ponts et Chaussées.
Schotman, G.J.M. (1989). ‘The effects of displacements on the stability of jackup spudcan foundations’. OTC 6026, Proceedings, 21st
Offshore Technology Conf., Houston.
Skempton, A.W. (1951). The bearing capacity of clays. Proc. Building Research Congress, London, 1, 180-189.
SNAME 5-5A (1994 and subsequent editions). Guidelines for site-specific assessment of mobile jack-up units. Technical & Research
Bulletin 5-5A, Society of Naval Architects and Marine Engineers.
Tan, F.S.C. (1990). Centrifuge and theoretical modelling of conical footings on sand. PhD thesis, University of Cambridge.
Teh, K.L., Cassidy, M.J., Leung, C.F., Chow, Y.K., Randolph, M. F. and Quah, C.K. (2008) Revealing the bearing capacity mechanisms
of a penetrating spudcan through sand overlying clay. Géotechnique, Vol. 58, Issue 10, pp. 793-804.
Teh, K.L., Leung, C.F., Chow, Y.K. and Cassidy, M.J. (2010) Centrifuge model study of spudcan penetration in sand overlying clay.
Géotechnique, Vol. 60, No. 11, pp. 825-842.
Teh, K.L., Leung, C.F., Chow, Y.K. and Handidjaja, P. (2009) Prediction of punch-through for spudcan penetration in sand overlying clay.
OTC20060, Proceedings, 41st Offshore Technology Conference, Houston.
Temperton, I., Stoner, R.W.P., Springett, C.N. (1997) Measured jack-up fixity: analysis of instrumentation data from three North Sea jack-
up units and correlation to site assessment procedures. Proceedings, Int. Conf. Jack-Up Platform Design, Construction and Operation,
City University, London.
Templeton, J.S., III, Brekke, J.N. & Lewis, D.R. (2005). Spud can fixity in clay, final findings of a study for IADC. Proceedings, 10th Int.
Conf., The Jack-up Platform, City University, London.
Templeton, J.S., III (2006). Jackup foundation performance in clay. OTC 18367, Proceedings, Offshore Technology Conf., Houston.
Templeton, J. S., III (2007) Spud Can Fixity in Clay, Findings from Additional Work in a Study for IADC, Proceedings, Eleventh
International Conference, The Jackup Platform, City University, London.
Templeton, J.S., III (2009). Spud can fixity in clay, further results from a study for IADC. Proceedings, 12th Int. Conf., The Jack-up
Platform, City University, London.
Templeton, J. S., III, and Lewis, D. R. (2011) Hysteretic Damping in Jack-up Dynamics. Proceedings, Thirteenth International Conference,
The Jack-up Platform, City University, London.
Ueshita, K. and Meyerhof, G.G. (1967) Surface Displacement of Multilayer Soil Systems. J. Soil Mechanics and Foundations Division,
ASCE Vol. 93, No. 5.
Van Langen, H., Wong, P.C., and Dean, E.T.R. (1999) Formation and validation of a theoretical model for jack-up foundation load-
displacement analysis. Marine Structures 12(4): 215-230.
Vesic, A.S. (1975). Bearing capacity of shallow foundations. In Foundation Engineering Handbook (ed. H.F. Winterkorn & H.Y. Fang),
121-147. Van Nostrand.
Yamamoto, N., Randolph, M.F. and Einav, I. (2008) Simple formulas for the response of shallow foundtions on compressible sands. Int. J.
for Geomechanics, ASCE 8:4 pp 230-230.
Yamamoto, N., Randolph, M.F. and Einav, I. (2009) A numerical study of the effect of foundation size for a wide range of sands. Journal
of Geotechnical and Geoenvironmental Engineering, ASCE., 135(1): pp. 37-45.
Young, A.G., Remmes, B.D. & Meyer, B.J. (1984). Foundation performance of offshore jack-up drilling rigs. J. Geotech. Engng Div.,
ASCE, 110(7), 841-859.
OTC 23521 17

Figure 1. Typical slip-line field for embedded conical footing in clay (β = 150°, α = 0.8, D/B = 0.5, ρB/sum = 5)

(a) β = 180°, α = 1, ρB/sum = 0

(b) β = 150°, α = 1, ρB/sum = 5

Figure 2. Bearing capacity factors (Ncscdc) for embedded conical footings in clay
18 OTC 23521

Figure 3. Definition of spudcan volumes in ISO 19905-1

Figure 4. Backfill mechanisms for spudcan penetrating clay (Hossain, et al., 2006)
OTC 23521 19

Figure 5. Estimation of limiting cavity depth, Hcav, in ISO 19905-1

Figure 6. Effect of shape parameter, a, on yield interaction surface


20 OTC 23521

(a) Horizontal capacity

(b) Moment capacity

Figure 7. Capacity multipliers for deeply embedded spudcan in clay


OTC 23521 21

Perform foundation assessment


See Sections 9 and A.9.3.6

Step 1a
Perform preload
check, see A.9.3.6.2, OK
and Step 1b sliding
check, see A.9.3.6.3

Not OK

Step 2a
Perform bearing
check, see A.9.3.6.4. OK
Uses Step 1b sliding
check, see A.9.3.6.3

Not OK
Perform structural analysis assuming
degrading moment fixity with elastic
vertical and horizontal springs.
See A.9.3.4.2.3 and Figure A.10.4-2

Step 2b
see A.9.3.6.5
Perform bearing and OK
sliding checks. Uses
A.9.3.6.4 and A.9.3.6.3.

Not OK

Perform structural analysis Step 3


assuming full non-linear Perform foundation OK
foundation stiffnesses. check on all legs.
See A.9.3.4.2.4 See A.9.3.6.6

Figure 8. Approach to foundation acceptance check Not OK

Foundation NOT shown to Foundation


be acceptable acceptable
Figure 8. Approach to foundation acceptance check
22 OTC 23521

V
U = | (FH,FV) - (FH,FV)ORG | / | QVH,f - (FH,FV)ORG |
QV

1
2

QVH,f

(FH,FV)

(FH,FV)ORG =
0,5 QV / γR,VH

(FH,FV)

(FH,FV)
QVH,f
4

3
QVH,f

H
(Figure 9a) Sand

V
1 5
QV
U = | (FH,FV) - (FH,FV)ORG | / | QVH,f - (FH,FV)ORG |
2
(QV+BS)/γR,VH-BS

QVH,f

(FH,FV)

(FH,FV,)ORG =
0,5 QV / γR,VH

(FH,FV)

QVH,f

4 3

H
- BS QH/γR,Hfc QH

(Figure 9b) Clay with spudcan buoyancy and no backfill


OTC 23521 23

V
5
1
QV
U = | (FH,FV) - (FH,FV)ORG | / | QVH,f - (FH,FV)ORG |
2
(QV-WBF,o+BS)/γR,VH
+WBF,o-BS
QVH,f

(FH,FV)

(FH,FV,)ORG =
0,5 QV / γR,VH

(FH,FV)

QVH,f

WBF,o-BS
4 3

H
QH/γR,Hfc QH

(Figure 9c) Clay with spudcan buoyancy and backfill

Key
1 vertical-horizontal foundation capacity
2 factored vertical-horizontal foundation capacity (coordinates multiplied by 1/γR,VH) relative
to the scaling origin as defined
3 sliding capacity
4 factored sliding capacity (unfactored horizontal sliding capacity coordinate multiplied by
1/γR,Hfc)
5 Vectors indicating origin for construction of the factored V-H bearing capacity envelope

|…| represents the vector magnitude


H horizontal reaction or horizontal capacity
QV gross ultimate vertical foundation capacity (with zero horizontal load)
point where the vector originating from (FH,FV)ORG and passing through (FH,FV) intersects
QVH,f the factored vertical-horizontal capacity surface derived by dividing the coordinates of the
applicable surface by the resistance factor γR,VH
U utilization for environmental response point (Fv,FH)
V vertical reaction or vertical capacity
γR,VH partial resistance factor for foundation (bearing) capacity
γR,Hfc partial resistance factor for horizontal (sliding) capacity

Figure 9. Vertical-horizontal foundation capacity envelopes

You might also like