You are on page 1of 13

15

ARTICLE
Modelling the embedment process during offshore pipe-laying on
fine-grained soils
Z.J. Westgate, D.J. White, and M.F. Randolph

Abstract: Subsea pipelines are becoming an increasingly significant element of offshore hydrocarbon developments as exploration
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16

moves into deep-water environments further from shore. During the lay process, pipelines are subject to small amplitude vertical and
horizontal oscillations, driven by the sea state and lay vessel motions. Centrifuge model tests have been used to simulate these
small-amplitude lay effects, with varying degrees of idealization relative to the real lay process. In the soft soils found in deep water,
pipe embedment can exceed a diameter or more, thus significantly affecting the lateral pipe–soil interaction, axial resistance, and
thermal insulation. In this paper, results from centrifuge model tests are used to calibrate a model for calculating the dynamic
embedment of a subsea pipeline. The model uses elements of plasticity theory to capture the effects of combined vertical and
horizontal loading, and incorporates the softening of the surrounding soil as it is remoulded due to the pipeline motions. Influences
from the lay rate, lay geometry, and sea state are included in the calculation process. The model is compared with observed as-laid
pipeline embedment data from field surveys at three different offshore sites. Using site-specific soil parameters obtained from in situ
testing and idealized pipe loads and motions to represent the load and displacement patterns during offshore pipe-laying, respec-
tively, the model is shown to capture well the final as-laid embedment measured in the field surveys.

Key words: centrifuge modeling, clays, plasticity, offshore engineering, pipe-laying.

Résumé : Les pipelines en mer deviennent un élément de plus en plus significatif des projets de développement des hydrocarbures
For personal use only.

en mer, en raison de l'exploration qui se déplace dans des environnements d'eau profonde éloignés des côtes. Durant le processus de
placement, les pipelines sont soumis à des oscillations verticales et horizontales de faible amplitude, causées par l'état de la mer et les
mouvements des équipements de placement. Des essais par centrifugeuse ont été utilisés pour simuler ces effets de placement à faible
amplitude, en utilisant différents degrés d'idéalisation relativement au processus de placement réel. Dans les sols mous rencontrés en
eau profonde, le tuyau peut s'enfoncer d'un diamètre ou plus, ce qui affecte de façon significative l'interaction latérale tuyau-sol, la
résistance axiale et l'isolation thermique. Dans cet article, les résultats des essais par centrifugeuse sont utilisés pour calibrer un
modèle servant à calculer l'enfoncement dynamique d'un pipeline en mer. Le modèle utilise les éléments de la théorie de plasticité
pour intégrer les effets des charges verticales et horizontales combinées, et inclus l'adoucissement du sol environnant causé par le
remodelage du sol lors des mouvements du pipeline. Les effets du taux de placement, de la géométrie de placement et de l'état de la
mer sont inclus dans le processus de calcul. Le modèle est comparé à des données observées d'enfoncement de pipeline tel que placé;
ces données provenant de campagnes de terrain sur trois sites marins différents. L'utilisation de paramètres du sol spécifiques au site
obtenus, via des essais in situ et des charges de tuyau et mouvements idéalisés pour représenter respectivement les patrons de charges
et déplacement, permet que le modèle simule bien les mesures finales d'enfoncement obtenus dans les campagnes de terrain. [Traduit
par la Rédaction]

Mots-clés : modélisation par centrifuge, argiles, plasticité, ingénierie marine, placement de pipelines.

Introduction down zone due to the increasing restraint associated with pipe-
As offshore hydrocarbon developments have advanced into line embedment. Quantifying the influence of these dynamic
effects on the as-laid pipeline embedment is challenging.
deeper water, subsea pipelines have become a larger component
of project cost. The as-laid embedment of a subsea pipeline is an Static penetration of pipes
important design parameter, influencing the hydrodynamic re- The static bearing capacity of a pipeline on undrained soil is
sponse, lateral buckling, axial walking, and thermal insulation well established, based on plasticity solutions (Murff et al. 1989;
(Cathie et al. 2005; Bruton et al. 2006). During the lay process, the Randolph and White 2008a), finite element analyses (Aubeny et al.
pipeline is subject to vertical and horizontal oscillations that are 2005; Merifield et al. 2009; Chatterjee et al. 2012), and model test-
driven by the sea state and vessel motion. These oscillations, or ing (Dingle et al. 2008). Figure 2 shows the general shape of the
dynamic lay effects, cause the soil to soften and displace, resulting relationship between the normalized pipe embedment w/D and
in increased pipeline embedment. The length of pipeline–seabed the normalized pipe bearing capacity V/suD, where w is the pipe
contact where lay effects occur is termed the touchdown zone, invert embedment, D is the pipe diameter, V is the vertical pipe–
illustrated in Fig. 1. Near the front of the touchdown zone — soil load, and su is a characteristic undrained soil shear strength.
towards the lay vessel — the pipeline experiences cycles of seabed For a pipeline being laid from a vessel, the maximum value of V
separation, influenced by the severity of the sea state. The vertical within the touchdown zone is greater than the submerged pipe
loading and horizontal oscillations attenuate through the touch- weight W= due to the catenary effect. This stress concentration can

Received 17 May 2012. Accepted 4 October 2012.


Z.J. Westgate. Advanced Geomechanics, Houston, TX, USA.
D.J. White and M.F. Randolph. Centre for Offshore Foundation Systems, The University of Western Australia, 35 Stirling Highway, Perth, WA 6009, Australia.
Corresponding author: Z.J. Westgate (e-mail: zackw@ag.com.au).

Can. Geotech. J. 50: 15–27 (2013) dx.doi.org/10.1139/cgj-2012-0185 Published at www.nrcresearchpress.com/cgj on 22 October 2012.
16 Can. Geotech. J. Vol. 50, 2013

Fig. 1. Idealization of pipeline motions within touchdown zone during laying.


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16

Fig. 2. Pipe penetration response in fine-grained soil. which increase the vertical load, or pipe bearing pressure, as
shown by numerical simulations (Westgate et al. 2010b) and
field monitoring (Westgate et al. 2012). Combined loading reduces
the available vertical bearing capacity of the soil, as shown
through theoretical plasticity analysis for planar contacts (Green
1954) and also shallowly-embedded pipes under horizontal load-
For personal use only.

ing (Randolph and White 2008a) and axial loading (Yan et al. 2011).
Lastly, horizontal oscillations can displace the seabed soil later-
ally, creating a trench by lowering the seabed elevation adjacent
to the pipe, as shown via experimental simulations in a geotech-
nical centrifuge (Cheuk and White 2011) and via field observations
(Lund 2000; Westgate et al. 2010a).

Existing models for assessing dynamic pipe embedment


Quantifying dynamic lay effects is complicated by the nature of
the lay process, where the presence of each mechanism, and its
relative influence, changes as an element of pipeline moves
through the touchdown zone. Experimental studies have sup-
ported a tentative model in which dynamic pipe embedment can
be assessed on a cycle-by-cycle basis (Cheuk and White 2011). The
model links progressive soil softening to cumulative horizontal
pipe displacements, at a rate related to the strength degradation
measured in cyclic T-bar tests. The progressive embedment of the
be estimated from the lay tension, pipe bending stiffness, and the pipe is then assessed from the combined vertical–horizontal bear-
seabed response (Pesce et al. 1998; Lenci and Callegari 2005; ing capacity for the present soil strength. Similar strain-based
Randolph and White 2008b; Palmer 2009). strength degradation models are common in offshore foundation
design (e.g., Andersen 1991), and combined loading models have
Dynamic lay effects been proposed for monotonically loaded pipelines in fine-grained
Dynamic lay effects increase the as-laid embedment beyond soils (e.g., Hodder and Cassidy 2010). However, the complexity of
simply the penetration under the maximum static pipe–soil load. deformations around an oscillating pipe makes such models cum-
These mechanisms include: bersome for assessing as-laid pipeline embedment. Models for
• increased vertical pipe–soil load from the vertical oscillations cyclic pipe–soil interaction purely in the vertical plane have been
of the pipeline catenary; developed for steel catenary riser (SCR) behaviour (Bridge et al.
• reduced soil strength due to remoulding and disturbance; 2004; Clukey et al. 2005; Aubeny et al. 2008; Randolph and
• reduced vertical bearing capacity for combined vertical– Quiggin 2009). However, these models do not include any influ-
horizontal loading; and ence of horizontal motions. More general models that incorporate
• generation of a trench around the pipe, locally lowering the the effects of combined vertical and horizontal pipe motions on
seabed. pipe embedment have not been developed.

Pipeline oscillations in the touchdown zone soften the soil and Objective and focus of study
also entrain water, which leads to a further reduction in strength. The objective here was to investigate pipe embedment on fine-
Studies of purely vertical pipe movements have shown that both grained soils by systematically isolating each of the dynamic lay
the rate and magnitude of soil softening are increased signifi- effect mechanisms through simulations in a geotechnical centri-
cantly due to water entrainment (Gaudin and White 2009). In fuge. Different types of dynamic lay processes were simulated,
rough sea states, transient reductions in lay tension can occur, with the aim of separating the effects listed previously. The results

Published by NRC Research Press


Westgate et al. 17

of the simulations were then back-analysed to calibrate a modi- Fig. 3. Penetration resistance degradation during T-bar cycling in
fied version of the cycle-by-cycle model for pipe embedment pro- kaolin clay.
posed by Cheuk and White (2011), providing a framework that
quantifies the different effects separately. The modified model
has been evaluated against as-laid pipeline embedment data from
field surveys, based on idealizations of the vertical loads and hor-
izontal motions of the pipeline at the seabed (Westgate et al.
2010b, 2012).

Experimental test programme


Test apparatus
Centrifuge model tests were performed using the beam centri-
fuge at The University of Western Australia (Randolph et al. 1991).
A 20 mm diameter by 120 mm length model pipe was tested at a
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16

centrifuge acceleration of 25g, resulting in a prototype pipe diam-


eter of 0.5 m. The equivalent prototype depth of the soil model
was 3.0 m. Vertical and horizontal actuators controlled the pipe
motions through an instrumented loading arm capable of record-
ing the vertical and horizontal load applied to the pipe. Data
acquisition and control was through in-house software capable of
modelling sophisticated sequences of pipe motions (Gaudin and
White 2009; De Catania et al. 2010).

Soil sample preparation


The soil sample was prepared by mixing kaolin powder with
water to approximately twice its liquid limit (⬃120% water con-
tent), then pouring it into the strongbox over a 10 mm (model
scale) sand drainage layer covered with a geotextile. The sample
was spun at 100g in the centrifuge for 2 days, after which the Table 1. Centrifuge test sample: T-bar resistance degradation
For personal use only.

upper 10 mm (model scale) of the sample was removed and the parameters.
sample was spun at a lower acceleration of 25g, creating a lightly
Parameter (field scale) Value
overconsolidated clay sample with a small mudline strength
intercept. Deep remoulding degradation factor, ␦rem-deep 0.45
Deep remoulding degradation cycle number, 3
Strength characterization N95,rem-deep
A 5 mm diameter, 30 mm long T-bar penetrometer was used to Deep structure degradation factor, ␦str-deep 0.14
characterize the strength of the soil after the preparation was Deep structure degradation cycle number, 0.75
complete. The undrained shear strength, su, was determined from N95,str-deep
the T-bar penetration resistance, qT-bar, using a T-bar bearing fac- Surface remoulding degradation factor, ␦rem-surface 0.25
tor, NT-bar, of 10.5 (Stewart and Randolph 1994) with an adjustment Surface remoulding degradation cycle 4
to this factor applied near to the soil surface, following White number, N95,rem-surface
et al. (2010). The resulting initial strength profile was idealized as Surface structure degradation factor, ␦str-surface 0.15
a linear variation with mudline strength intercept, sum, and Surface structure degradation cycle number, 0.75
strength gradient, ksu, as N95,str-surface

[1] su-ini ⫽ 0.8 ⫹ 3z kPa


su-op
[2] ␦op ⫽ ⫽ ␦rem ⫹ (1 ⫺ ␦rem ⫺ ␦str)e⫺3N/N95,rem
su-ini
where z is the depth below the soil surface in metres. Cyclic se-
⫹ ␦stre⫺3N/N95,str
quences of T-bar penetration and extraction were performed to
quantify the softening response of the soil. Deep cycles were per-
formed where the T-bar remained buried below the soil surface by
several diameters, and surface cycles were performed where the where subscripts op and ini denote operative and initial, respec-
T-bar was fully extracted from the soil during each cycle. The rate tively; N is the T-bar cycle number; and N95,rem and N95,str are
of T-bar penetration and extraction was sufficient to ensure un- defined in the next paragraph. ␦rem is the fully remoulded
drained soil conditions, with velocity, v, equal to 1 mm/s, normal- strength degradation factor, defined as
ized as vdT-bar/cv (with dT-bar being the T-bar diameter equal to
5 mm and cv the consolidation coefficient of the soil not greater 1
[3] ␦rem ⫽
than 0.1 mm2/s) (e.g., Randolph and Hope 2004). St,cyc
The degradation in strength during cyclic T-bar testing can be
modelled using an exponential decay function linked to the num-
ber of cycles, which is a proxy for the accumulated shear strain and ␦str is the component of strength degradation due to the loss
(Einav and Randolph 2005; Zhou and Randolph 2009). The model of soil structure, defined as
may include an additional more brittle component of degradation
that can be attributed to the structure of the soil (Randolph et al.
2007). The strength degradation response can therefore be ex- 1
[4] ␦str ⫽ 1 ⫺
pressed as St,in-out

Published by NRC Research Press


18 Can. Geotech. J. Vol. 50, 2013

Table 2. Centrifuge pipe embedment test details.


Vertical lay effects Horizontal lay effects
Pipe bearing pressure,
Number of
V/D (kPa)
vertical Horizontal pipe Number of Simulated dynamic embedment
Test ID Vmin/D Vavg/D Vmax/D cycles displacement, u/D horizontal cycles mechanism(s)
A — — 12 — — — —
B <0 2.8 5.7 120 — — Soil softening with water entrainment
C 5.7 5.7 5.7 — ±0.05 420 Horizontal loading
D 3.5 5.7 8.0 280 ±0.05 140 Combined vertical and horizontal
loading
E <0 2.2 4.4 1200 ±0.05 1200 Combined vertical and horizontal
loading with water entrainment
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16

The cyclic sensitivity, St,cyc (in terms of resistances), is the ratio Fig. 4. Example pipe-laying simulation results (test D) showing
of the initial penetration resistance to the stable penetration re- (a) vertical load and horizontal displacement path and (b) vertical
sistance after many cycles. St,in-out is the ratio of the initial pene- and horizontal bearing pressure.
tration resistance to the initial extraction resistance. By
definition, the cycle number for loss of structure, N95,str, is 0.75,
corresponding to the first cycle of penetration and extraction and
noting that the initial penetration represents cycle number
N = 0.25 (Randolph et al. 2007). The cycle number, N95,rem, for 95%
resistance degradation is obtained by fitting the measured degra-
dation response. Example fits to the T-bar test data in the kaolin
clay centrifuge sample, for both deeply cycled and surface-cycled
tests, are shown in Fig. 3, with the corresponding degradation
parameters listed in Table 1. As indicated, the degradation re-
sponses are markedly different for deep and surface cycling and
For personal use only.

can also vary significantly for different soil types, as shown by


Gaudin and White (2009) for high-plasticity clay and carbonate
silt.

Testing programme
The testing programme comprised a baseline test (test A) in
which the pipe was statically penetrated to a depth of one diam-
eter, as well as four dynamic pipe-laying simulations (tests B to E),
in which specific embedment mechanisms were sequentially in-
troduced. For each dynamic simulation, the pipe was initially
penetrated statically to approximately one-third of the pipe diam-
eter. A relatively heavy pipe was modelled to generate significant
dynamic embedment under the small amplitude horizontal oscil-
lations. Axial oscillations were not considered in this study and it
is expected that the effects of axial loading on the vertical bearing
capacity during pipe-laying are minor compared to the relatively
larger horizontal oscillations.
The total number of vertical or horizontal cycles varied from
120 to 1200, which represents the typical range of oscillations that
a pipe will be subjected to as it passes through the touchdown
zone while being laid, with the upper value representing a situa-
tion where a period of downtime delays laying. In all tests the rate
of initial monotonic penetration was equal to 0.1 mm/s, which
ensured undrained conditions during monotonic penetration
(i.e., vD/cv > 10, Randolph and Hope 2004). The pipe displacement
rate during cycling ranged from 0.25 to 3 mm/s, or cyclic periods,
tcyc, ranging from about 7–16 s. This corresponds to dimensionless
time generally less than T = 0.001, where T = tcyccv/D2, which re-
sults in a limit to the dissipation of excess pore pressures gener-
ated during cyclic loading (Randolph and Hope 2004; Westgate
et al. 2012).
The test details are presented in Table 2. Test B targeted soil
softening and water entrainment during large-amplitude vertical
pipe motions. The pipe in this test was fully extracted from the minimizing water entrainment. In test C, the pipe was held under
soil during each cycle and was penetrated to a target vertical load a constant vertical load of V/D = 5.7 kPa. Test D varied from test C
of V/D = 5.7 kPa, all in the absence of horizontal motions. Tests C by including vertical load cycles, with an average value equal to
and D targeted soil softening and combined vertical and horizon- the static vertical load in test C. Within each cycle of horizontal
tal loading. In both of these tests the pipe remained in contact motion, the pipe was subjected to two cycles of vertical loading
with the soil during cycles of fixed-amplitude horizontal motions, between 3.5 and 8.0 kPa, representing a peak-to-peak amplitude of

Published by NRC Research Press


Westgate et al. 19

Fig. 5. Images of soil remoulding following (a) static pipe penetration Table 3. Calculation model parameters for centrifuge back-analysis.
during test A and (b) dynamic pipe penetration during test E. Parameter Value Comments
Mudline strength intercept, 0.8 Fit to T-bar test
sum (kPa)
Intact shear strength 3 Fit to T-bar test
gradient, ksu (kPa/m)
Discounting factor, r 4 Fit to test data
Effective unit weight, ␥= 6 Assumed based on typical
(kN/m3) values
Fitting parameter (to 7 Fit to static pipe penetration
strength profiles), a response
Fitting parameter (to 0.3 Fit to static pipe penetration
strength profiles), b response
Parallel point, (V/Vmax)pp 0.5 Assumed constant for
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16

simplicity
Flow rule parameter, ␭ 1.5 Assumed constant for
simplicity

Fig. 6. Example yield surface for pipe–soil interaction.


For personal use only.

horizontal displacement. The resulting cyclic accumulation of


pipe embedment and increase in horizontal resistance during test
D is shown in Fig. 4b. All imposed oscillations — both vertical load
and horizontal displacement — were sinusoidal in time.
Test E was the most onerous laying simulation, targeting soil
softening and water entrainment combined with cyclic vertical
loading and horizontal displacement. The pipe in this test under-
went one cycle of extraction from the soil and vertical loading to
a load of only V/D = 4.4 kPa during each cycle of horizontal motion.
The change in soil consistency during the modelled lay effects is
shown in Fig. 5. While the soil breaks into blocks of intact mate-
rial for a statically penetrated pipe (test A, Fig. 5a), in test E the soil
surrounding the pipe degrades into slurry due to the extremely
onerous cyclic motions (Fig. 5b). Tests B, C, and D represent inter-
mediate cases relative to these two extremes.

Modelling dynamic pipe embedment


Before presenting the remaining test results, the model used to
interpret and back-analyse the results is described. The model is
based on the one presented by Cheuk and White (2011), modified
to allow tracking of the full pipe displacement throughout cyclic
vertical loading and horizontal oscillations of varying amplitude;
about 75% of the static V/D. This pattern of loading for test D is by contrast the original model was limited to tracking pipe em-
presented in Fig. 4a, which shows that the experimental data bedment on a cycle-by-cycle basis. The modelling approach uti-
consistently follow the demanded profiles of vertical load and lizes theoretical V–H (V, vertical; H, horizontal) yield surfaces for

Published by NRC Research Press


20 Can. Geotech. J. Vol. 50, 2013

Table 4. List of centrifuge back-analyses performed.


Back-analysis case
Flow rule and Flow rule and
Base case base case optimized
Parameter softening softening softening
Mudline strength intercept, sum (kPa) 0.8 0.8 0.8
Intact shear strength gradient, 3 3 3
ksu (kPa/m)
Discounting factor, r Not used (0) Not used (0) 4
Effective unit weight, ␥' (kN/m3) 6 6 6
Bearing capacity parameter (modified 7 7 7
to suit strength profile), a
Bearing capacity parameter (modified 0.3 0.3 0.3
to suit strength profile,) b
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16

Parallel point, (V/Vmax)pp Not used 0.5 0.5


Flow rule parameter, ␭ Not used 1.5 1.5

partially embedded pipes, combined with a soil-softening model Vmax


calibrated to cyclic T-bar tests. Full details are provided in Cheuk [7] ⫽ Ncsu-op ⫹ Nb␥ w
D
and White (2011) and are summarized below, together with the
proposed modifications. The model parameters are listed in
Table 3. where su-op is the operative undrained shear strength at depth w
(eq. [2]), ␥= is the soil effective unit weight, and w is the pipe invert
Model description embedment. Nc and Nb are bearing capacity factors for soil
The dynamic embedment process is modelled using features of strength and soil buoyancy, respectively. Nc can be approximated
yield surfaces for V–H bearing capacity in undrained conditions, as
an example of which is shown in Fig. 6 (Merifield et al. 2008). Yield
For personal use only.

( wD )
b
surfaces have been derived analytically based on plastic limit anal- [8] Nc ⫽ a
ysis and through numerical modelling, and depend on pipe
roughness, strength heterogeneity, bonding of the pipe–soil inter-
face, and pipe embedment (Randolph and White 2008a). The bear- where fitting parameters a and b capture the influence of soil
ing capacity under purely vertical load is defined as Vmax. The strength heterogeneity, pipe roughness, pipe penetration rate,
parallel point, (V/Vmax)pp, corresponds to the vertical load level and strain softening (Aubeny et al. 2005; Merifield et al. 2009;
that results in purely horizontal movement at failure, and can be Chatterjee et al. 2012). Nb can be written as
assessed from the failure envelopes by invoking normality. (V/
Vmax)pp = 0 for fully bonded or deeply buried pipes, and increases fbAs ␲D
to (V/Vmax)pp = 0.5 for an unbonded pipe on the soil surface. In this [9] Nb ⫽ ≤
paper, (V/Vmax)pp is taken as a constant value of 0.5, reflecting the Dw 4w
unbonded condition. The basis for this assumption is that there is
unlikely to be significant tension sustainable at the pipe surface, where the buoyancy factor fb captures the influence of the soil
due to soil softening and the tendency for weak surface soils to be heave adjacent to the pipe, with a value of 1.5 being commonly
carried downward by the pipe during laying. adopted (Merifield et al. 2009), and As is the area of the pipe within
For the undrained conditions being modelled here, the assump- the soil.
tion of normality (and the neglect of elastic behaviour) allows the The effect of the softened soil strength on the bearing capacity
slope of the yield surface to be used to define the pipe trajectory due to remoulding can be assessed using a modified form of eq. [2],
during horizontal motion. The parameter that controls the direc- substituting the cumulative pipe movement for the number of
tion of the pipe motion at failure, ␭, varies approximately linearly T-bar cycles, so that the operative strength is expressed as
with load level, V/Vmax, in the range 0.2 < V/Vmax < 0.8 (Cheuk and
White 2011) allowing the direction of pipe movement to be writ- [10] ␦op ⫽ ␦rem ⫹ (1 ⫺ ␦rem ⫺ ␦str)e⫺3␣N(rem) ⫹ ␦stre⫺3␣N(str)
ten as

where ␣N accounts for the cumulative pipe displacement, ex-


[5] ( dw
du )
f
⫽␭
关VV max
⫺ 共 兲兴
V
Vmax pp
pressed as

⌺(s/D)
[11] ␣N ⫽ ␨r
(⌺s/D)95
where u is the horizontal pipe displacement, leading to incre-
ments of pipe embedment given by
where s/D is the total pipe displacement per cycle, for w > 0 (when
⌬w dw ⌬u w < 0 the pipe is above the soil, so it is not remoulding it). The sum
[6]
D
⫽ ( )( )
du f D of the pipe displacement required to cause 95% strength degrada-
tion, (⌺s/D)95, is analogous to the parameter N95 from the cyclic
T-bar test. Zhou and Randolph (2009) showed that the vertical
where ⌬u/D is the incremental horizontal pipe displacement. extent of the failure mechanism around a penetrating T-bar is ap-
The vertical bearing capacity of the soil is affected by the soil proximately 1.5 diameters, so it can be assumed that the remoulding
strength and soil buoyancy, and is expressed as during a single pass of a T-bar penetrometer is equal to the remould-

Published by NRC Research Press


Westgate et al. 21

Fig. 7. Calculations using base case softening law versus observed pipe embedment for (a) test B, (b) test E, (c) test C, and (d) test D.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16
For personal use only.

ing caused by a cumulative pipe movement of 1.5 D. As one cycle soil softening in the calculation of embedment, i.e., the effects of
involves two passes of the T-bar, (⌺s/D)95 = 3N95 (Cheuk and White horizontal pipe movement on the pipe penetration were ignored.
2011). A discounting parameter, ␨, is included in eq. [11] to reflect Instead, eqs. [10]–[12] were used to assess the operative soil
that, as the pipe embeds, the soil below the oscillating pipe is less strength around the pipe throughout the lay process, and the
affected by the earlier oscillations and so the cumulative damage resulting embedment was calculated based on Vmax, from eqs.
is lower than if all cycles had remoulded the same zone of soil. The [7]–[9]. This calculation used either the deep or surface T-bar pa-
depth-discounting parameter, ␨, is defined as rameters (Table 1) without any empirical adjustment to the soft-
ening response, i.e., r = 0 in eq. [11]. This back-analysis was done for
wavg 0.5(w ⫹ wstatic) all four lay simulations.
[12] ␨⫽ ⫽
w w Second, a base-case comparison was performed using both soil
softening and the flow rule (to add the increasing embedment
where wavg and wstatic represent the average and static, or initial, that results from combined V–H loading), with either the deep or
values of w. surface T-bar parameters and r = 0 (i.e., no depth-discounting ef-
A fitting parameter r is included to moderate the depth- fect). This back-analysis was performed for tests C and D, where
discounting effect (eq. [11]), as a simplification to the more com- the pipe remained in contact with the soil (so V/Vmax was always
plicated procedure of tracking the cumulative degradation defined).
through different soil horizons over the depth of embedment. Lastly, an optimized comparison was performed for tests C and
D using both soil softening and the flow rule, based on the average
Back-analyses using modified embedment model of the deep and surface T-bar parameters (Table 1), and with the
The modified model performance was assessed in three ways. parameter r calibrated to the observed embedment. The adopted
First, a base case comparison was performed that only considered parameters in each of these back-analyses are listed in Table 4.

Published by NRC Research Press


22 Can. Geotech. J. Vol. 50, 2013

Fig. 8. Calculations using flow rule and base case soil-softening law Fig. 9. Calculations using flow rule and optimized soil-softening
versus observed pipe embedment for (a) test C and (b) test D. law versus observed pipe embedment for (a) test C and (b) test D.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16
For personal use only.

the deep T-bar degradation parameters, suggesting that the soft-


ening model alone (excluding any element of water entrainment)
The bearing capacity was approximated using eq. [8] by fitting
is sufficient for simulating dynamic embedment during vertical
parameters a and b to the pipe penetration response from the
pipe motions in isolation. However, with continued cycling some
static penetration case (test A), resulting in a = 7 and b = 0.3. This
element of water entrainment may develop, as the pipe embed-
simplified penetration resistance profile approaches Nc = 10.5 at a ment gradually continues towards that calculated using the sur-
pipe embedment of about 3D, and is consistent with the expected face T-bar degradation parameters.
profile for this particular soil strength to weight ratio (su/␥=D) For test E, the calculated rate of embedment is very rapid,
(White et al. 2010). with the combined vertical and horizontal oscillations causing
the operative strength to reach the fully remoulded value (ex-
Back-analysis 1: Soil softening only cluding water entrainment) within a small number of cycles
Comparisons of the observed pipe embedment and the calcu- (Fig. 7b). In contrast to test B, the observed embedment rapidly
lated pipe embedment using only soil softening are shown in exceeds the calculated embedment using the deep degradation
Fig. 7. In these calculations, the embedment reached in each cycle response, and continues to increase towards the calculated em-
is calculated from the softened soil strength (using eqs. [7]–[12]) bedment using the surface degradation response (i.e., the mini-
and the maximum vertical load in that cycle. mum credible remoulded strength), suggesting a significant
The rate of pipe embedment is slightly overestimated during water entrainment component. The observed embedment even-
the large-amplitude vertical displacements of test B (Fig. 7a), sug- tually exceeds the surface degradation response at very high cycle
gesting that a portion of the pipe displacement near the soil sur- numbers (N > 1000), reflecting the additional destabilizing influ-
face does not contribute to strain accumulation. The observed ence of the horizontal loading, which is neglected in this back-
embedment eventually reaches the calculated embedment using analysis case. Due to the complexity of the pattern of pipe loads

Published by NRC Research Press


Westgate et al. 23

Fig. 10. Calculations of observed pipe trajectory for (a) test C and (b) test D.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16
For personal use only.

and motions in test E, which do not lend themselves to simulation within a single zone of soil, but is progressively transferred to new
using the proposed model, no further back-analyses were per- soil that is encountered at greater depth as the pipe moves down-
formed for this test. wards. This is the depth-discounting effect modelled by eq. [12],
For test C, the observed pipe embedment exceeds the calcu- and the “r” parameter in eq. [11].
lated embedment based on softening alone, even when using
the surface degradation parameters (Fig. 7c). The observed pipe Back-analysis 2: Soil softening and combined loading
embedment for test D is similar to that of test C (Fig. 7d), but the To improve on the first set of back-analyses, the influence of the
calculated values are greater due to the higher vertical load. combined V–H loading was introduced (via eqs. [5] and [6]), while
From these comparisons, two clear observations can be made. retaining r = 0 (i.e., no depth-discounting is applied to the cumu-
First, the observed embedment for tests with combined horizon- lative softening). The comparisons of the observed pipe embed-
tal and vertical loading generally exceeds the calculated embed- ment and the calculated pipe embedment using both soil
ment at the end of each test, showing that soil softening alone is softening and the flow rule are shown in Fig. 8. In both tests C and
not sufficient to explain the observed high embedments; an addi- D, the calculated rate of embedment still remains too high com-
tional embedment mechanism must be present. Second, the ob- pared to the observed response, but the final embedment is better
served rate of embedment is slower than calculated by direct predicted, progressing towards a steady-state embedment be-
analogy with T-bar tests. This supports the explanation that the tween the embedment calculated using the deep and surface
remoulding action of the pipe movement is not concentrated degradation responses. This shows that inclusion of the com-

Published by NRC Research Press


24 Can. Geotech. J. Vol. 50, 2013

bined loading effect improves the accuracy of the predicted Fig. 11. Effect of seabed stiffness on pipe–soil contact stress through
final embedment. touchdown zone (TDZ). zw, water depth.
A further improvement in the prediction accuracy is achieved
by using the average of the deep and surface degradation param-
eters in eq. [10], coupled with a depth-discounting adjustment
using r = 4 in eq. [11] (Figs. 9a, 9b). The depth-discounting effect
does not change the final embedment, only the rate at which this
is approached. This further improvement appears to indicate that
the depth-discounting aspect of the model is necessary to scale
from the softening rate observed in the T-bar test to the rate
occurring during the lay process.
The pipe trajectory during embedment in tests C and D is shown
in Fig. 10. The general shapes of the two pipe trajectories are
captured using the optimized model, even though the rate of pipe
embedment in test C is overpredicted during the initial cycles
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16

(Fig. 10a). For test D, which included a varying vertical load, the
pipe trajectory shows greater nonlinearity, and the model, due to
the assumptions of constant (V/Vmax)pp and ␭, does not accurately
capture the detailed trajectory (Fig. 10b).

Simulating as-laid pipeline embedment


The agreement between the centrifuge test simulations and
the calculated pipe embedment using the modified model is
encouraging, given the complex nature of the pipe–soil inter-
action during the simulations. In this section, the modified where W= is the submerged pipeline weight. During pipe-laying,
model is used to simulate pipeline-laying in the field using an the catenary-induced force changes due to changes in the pipe
idealized pattern of constant vertical loading and cyclic hori- tension, resulting in transient increases in V that can exceed the
zontal pipe displacement through the touchdown zone. The static Vmax value. Numerical simulations based on field conditions
results are then compared with the measured embedment from have shown that the dynamic peak catenary force, Vdyn, can attain
For personal use only.

as-laid pipeline surveys. nearly twice the value of Vmax in severe sea states (Westgate et al.
2010a, 2010b).
Modelling the pipeline loads and motions As with Vmax, the dynamic peak catenary force, Vdyn, also de-
During pipe-laying, the patterns of pipeline loads and motions pends on the seabed stiffness, which itself depends on the
are highly irregular. These vertical loads and horizontal oscilla- achieved embedment. It is therefore necessary to use an iterative
tions can be modeled numerically, but require some site-specific process. For a given ratio, Vdyn/Vmax, Vdyn was first calculated from
details such as the sea state and lay vessel response. For the cur- the static Vmax value for an estimated soil stiffness (eq. [14]). The
rent analysis, the changing vertical load and horizontal displace- as-laid embedment was then calculated, and used to update k,
ment through the touchdown zone has been idealized in a simple Vmax, Vdyn, and hence the embedment, and the process repeated
manner, as described below. until convergence.
Static profiles of the pipe–soil catenary force can be calculated It is assumed that the normalized horizontal pipe displacement
theoretically (e.g., Lenci and Callegari 2005; Palmer 2009). How- amplitude, u/D, attenuates to zero at the end of the touchdown
ever, these profiles depend on the seabed secant stiffness, which zone (where V stabilizes at the pipe weight), and can be idealized
varies through the touchdown zone due to the nonlinearity of the using a sinusoidal relationship linked to the period of pipe mo-
load–penetration response. The effect of the changing seabed tion, tcyc. In this study, the horizontal motion was modelled as
stiffness for a typical deep-water pipeline is shown in Fig. 11, with

关 共 兲兴 共 兲
stiffness, k, defined as the pipe–soil contact force, V, divided by the umax ␤
u t t
pipe embedment, w. As the stiffness reduces, two effects are ob- [15] ⫽ ⫺ sin 1⫺
D D tcyc tTDZ
served: (i) the location of the peak force moves further away from
the touchdown point and (ii) the length of the touchdown zone
increases, which increases the number of motions that an ele- where umax/D is the maximum horizontal displacement ampli-
ment of pipe is subjected to during laying. A parameter to quan- tude occurring at the front of the touchdown zone (i.e., the vessel
tify the length of the touchdown zone is the characteristic length, end) and ␤ is an attenuation factor. Based on numerical simula-
␭TDZ, (Pesce et al. 1998), calculated as tions of pipe-laying presented in Westgate et al. (2010a), ␤ was
taken as 3. The time, t, in eq. [15] represents the time during which

冪T
EI a pipe element passes through the touchdown zone. The total
[13] ␭TDZ ⫽
0 time in the touchdown zone, tTDZ, is calculated from the lay rate
and the touchdown zone length, ␭TDZ.
where E is the elastic modulus of the pipeline, I is the second
Summary of field surveys
moment of area, and T0 is the bottom tension in the pipeline.
Three sets of as-laid field survey data were back-analysed and
For the example pipeline in Fig. 11, ␭TDZ captures the length
linked to the variation in lay conditions and soil parameters along
of the touchdown zone under the initially higher stiffness values.
the different pipeline routes. The field surveys represent a range
Randolph and White (2008b) relate this parameter to the static
of pipe-laying conditions across three major offshore oil and gas
peak catenary force, Vmax, along the touchdown zone for a given
producing regions. The pipelines at these sites range from 0.33 to
seabed stiffness, expressed as
0.63 m in diameter, with submerged pipe weights between 0.22

关 兴
and 0.95 kN/m, laid in fine-grained sediments with undrained
Vmax (␭TDZ
2
)k 0.25
shear strength gradients between 4 and 38 kPa/m (and negligible
[14] ⫽ 0.6 ⫹ 0.4
W' T0 strengths at the mudline).

Published by NRC Research Press


Westgate et al. 25

Table 5. Site-specific pipeline and soil properties for field study comparisons.
Site A (Westgate et al. 2010a) Site B (Westgate et al. 2010b) Site C (Westgate et al. 2012)
Parameter A1 A2 A3 B1 B2 B3 C1 C2 C3
Outer pipe diameter, D (m) 0.39 0.39 0.39 0.33 0.33 0.33 0.63 0.63 0.63
Submerged bearing 0.56 0.56 0.56 1.4 1.4 1.4 1.5 1.5 1.5
pressure, W=/D (kPa)
Bending rigidity, EI (MN/m2) 37 37 37 44 44 44 230 230 230
Water depth, zw (m) 140 140 140 1350 1350 1350 600 600 600
Lay angle, ␾ (°) 68 68 68 84 84 84 64 64 64
Bottom tension, T0 (kN) 19 19 19 73 73 73 447 447 447
Significant wave height, 0.3 to 4 0.3 to 4 0.3 to 4 0.7 to 2.5 0.7 to 2.5 0.7 to 2.5 1.5 to 2.6 1.5 to 2.6 1.5 to 2.6
Hs (m)
Wave/swell period, Tp (s) 5 5 5 13 13 13 10 10 10
Lay rate (km/day) 2.5 2.5 2.5 1 1 1 2 2 2
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16

Number of cycles in 306 306 306 192 192 192 98 98 98


touchdown zone, NTDZ
Assumed maximum 0.05 0.25 0.5 0.001 0.025 0.05 0.05 0.25 0.5
horizontal oscillation
amplitude, umax/D
Dynamic amplification 1 1.25 1.5 1 1.25 1.5 1 1.25 1.5
factor, Vdyn/Vmax
Soil type Silty clay Silty clay Silty clay High-plasticity High-plasticity High-plasticity Carbonate Carbonate Carbonate
clay clay clay silt silt silt
Shear strength gradient, 10 10 10 20 20 20 10 10 10
su/z (kPa)
Strength degradation 0.33 0.33 0.33 0.25 0.25 0.25 0.14 0.14 0.14
factor, ␦rem
Degradation cycle number, 4 4 4 3 3 3 2.5 2.5 2.5
N95
For personal use only.

Discounting parameter, r 4 4 4 4 4 4 4 4 4
Effective unit weight, 7.5 7.5 7.5 3 3 3 5 5 5
␥= (kN/m3)
Operative strength factor, 0.69 0.45 0.37 0.98 0.81 0.73 0.64 0.32 0.21
␦op (eq. [10])

Calculations of pipe embedment using the soil softening and tion, given by ␦rem, although in the most severe lay conditions
combined loading model components proposed in this paper, ␦op⬃␦rem.
including the discounting parameter “r” in eq. [11], were per- As ␦op is considered directly within the calculations, and is linked
formed for each site using the pipeline properties, soil conditions, quantitatively to the sea states and pipe motion inputs, it is possible
and model parameters representative of the site-specific lay con- to identify the relative influences of the soil-softening embedment
ditions, which are listed in Table 5. Three calculations of the em- mechanism and the combined vertical and horizontal loading. In
bedment process were performed for each site, representing the relatively benign conditions, where pipe motions are small, the soil
range in observed sea state. The parameters that are linked to the strength is only slightly degraded (␦op is close to 1), but the combined
sea state are (i) the maximum horizontal displacement amplitude, loading effect leads to significantly greater embedment than if the
umax/D, and (ii) the ratio of the maximum dynamic vertical load to
pipe were simply placed statically onto the seabed.
the maximum static vertical load, Vdyn/Vmax. The adopted values
It is important to emphasize that this comparison is limited to
for each site were based on video footage of pipe motions in the
typical pipe-laying conditions in fine-grained soils. Conditions
touchdown zone and numerical simulations of pipe-laying. Full
back-analyses of the as-laid embedment at each site are detailed in that were absent from the centrifuge simulations that may lead to
Westgate et al. (2010a, 2010b, 2012). further dynamic embedment processes include (i) longer down-
time events during pipe-laying where a larger (N ⬎⬎ 1000) number
Comparisons of as-laid embedment to calculated of cycles is imposed, causing the growth of a trench around the
embedment pipe; (ii) extreme lateral movements (e.g., of an SCR during a
The ranges of calculated embedment values are compared with storm) in which the pipe can translate laterally to a new touch-
histograms of as-laid embedment for each site in Fig. 12. In all down alignment, re-commencing the embedment process; (iii) ep-
cases, the agreement between the calculated embedment and the isodes of reconsolidation prior to further cyclic loading, such as
observed embedment is excellent, capturing both the average em- the whole-life response at an SCR touchdown zone; (iv) dynamic
bedment (when using the best-estimate parameters) and the embedment in partially drained conditions, reducing the level of
range of embedment (when using the extreme minima and max- soil softening; and (v) scouring erosion of soil due to the pumping
ima parameters). action of the oscillating pipeline.
As well as providing the as-laid embedment of the pipeline, the Notwithstanding these limitations, the model represents a signif-
calculation process also leads to an estimate of the operative soil icant improvement to the assessment of as-laid pipeline embedment
strength around the pipe at the end of laying, through eqs. [10]–[12]. in fine-grained soils. The model could potentially be implemented as
The final degree of remoulding associated with the operative a pipe–seabed element response, for use in numerical simulations of
strength is given by ␦op, which is shown in the bottom row of pipe-laying and SCR response. Currently, models for this behavior do
Table 5. In general, ␦op is greater than the fully remoulded condi- not incorporate explicitly the soil softening and combined loading

Published by NRC Research Press


26 Can. Geotech. J. Vol. 50, 2013

Fig. 12. Comparisons of calculated and observed pipe embedment for Conclusions
(a) site A, (b) site B, and (c) site C.
This paper has presented data from centrifuge model simula-
tions of pipeline-laying in soft fine-grained soils. The simulations
comprised different combinations of isolated or concurrent verti-
cal and horizontal lay effects, which represent varying degrees of
realism relative to the pipe-laying process occurring in the field.
The results of the centrifuge tests were used to calibrate a simple
model for dynamic pipe embedment, which is a modified version
of a cycle-by-cycle model originally proposed by Cheuk and White
(2011).
The model was then compared with as-laid field survey data
from three locations that comprised a variety of lay conditions,
pipeline properties, and soil conditions representative of typical
offshore pipe-laying in fine-grained soils. Realistic estimates were
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16

made of the pipe motions and loading at the seabed, quantified


from detailed field observations and numerical modelling of the
lay process. The model was shown to capture the average and
range of observed as-laid embedment at each site, illustrating the
approximate contributions from soil softening and combined ver-
tical and horizontal loading on the as-laid embedment.
The model provides a simple approach for assessing as-laid
pipeline embedment in fine-grained soils that explicitly accounts
for dynamic lay effects, linked to the lay rate and sea state. This
type of model could be incorporated into a pipe–seabed interac-
tion element to be used in numerical modelling of pipe-laying and
SCR behaviour, encapsulating the effects of soil softening and
combined loading behaviour in the touchdown zone.

Acknowledgments
For personal use only.

The work forms part of the activities of the Centre for Offshore
Foundation Systems (COFS), currently supported as a node of the
Australian Research Council (ARC) Centre of Excellence for
Geotechnical Science and Engineering, in partnership with The
Lloyd's Register Educational Trust. Support for the second and
third authors from the ARC Future Fellowships and Discovery
programs is also acknowledged. The first author is supported by
an international postgraduate research scholarship and a univer-
sity postgraduate award from The University of Western Austra-
lia. The authors would like to acknowledge the support of Mr. Paul
Brunning, of Subsea7, during this work.

References
Andersen, K.H. 1991. Foundation design of offshore gravity structures. In Cyclic
loading of soils: From theory to design. Edited by M.P. O'Reilly and S.F. Brown.
Blackie, Glasgow. pp. 122–173.
Aubeny, C.P., Shi, H., and Murff, J.D. 2005. Collapse loads for a cylinder embed-
ded in trench in cohesive soil. International Journal of Geomechanics, 5(4):
320–325. doi:10.1061/(ASCE)1532-3641(2005)5:4(320).
Aubeny, C.P., Gaudin, C., and Randolph, M.F. 2008. Cyclic tests of model pipe in
kaolin. In Proceedings of the Offshore Technology Conference, Houston. Pa-
per OTC19494.
Bridge, C., Laver, K., Clukey, E., and Evans, T. 2004. Steel catenary riser touch-
down point vertical interaction model. In Proceedings of the Offshore Tech-
nology Conference, Houston. Paper OTC16628.
Bruton, D.A.S., White, D.J., Cheuk, C.Y., Bolton, M.D., and Carr, M.C. 2006. Pipe-
soil interaction behaviour during lateral buckling, including large amplitude
cyclic displacement tests by the Safebuck JIP. In Proceedings of the Offshore
Technology Conference, Houston. Paper OTC17944.
Cathie, D.N., Jaeck, C., Ballard, J.-C., and Wintgens, J.-F. 2005. Pipeline
geotechnics – state of the art. In Proceedings of the 1st International Sympo-
sium on Frontiers in Offshore Geotechnics, Perth. pp. 95–114.
Chatterjee, S., Randolph, M.F., and White, D.J. 2012. The effects of penetration
rate and strain softening on the vertical penetration resistance of seabed
pipelines. Géotechnique, 62(7); 573–582. doi:10.1680/geot.10.P.075.
Cheuk, C.Y., and White, D.J. 2011. Modelling the dynamic embedment of seabed
pipelines. Géotechnique, 61(1): 39–57. doi:10.1680/geot.8.P.148.
Clukey, E.C., Haustermans, L., and Dyvik, R. 2005. Model tests to simulate riser-
soil interaction effects in touchdown point region. In Proceedings of the 1st
International Symposium on Frontiers in Offshore Geotechnics, Perth. pp.
651–658.
behaviour (Randolph and Quiggin 2009). However, both of these De Catania, S., Breen, J., Gaudin, C., and White, D.J. 2010. Development of a
effects are shown in this study to be important, having a signifi- multiple-axis actuator control system. In Proceedings of the 7th International
cant effect on the pipe embedment process. Conference on Physical Modelling in Geotechnics, 1, Zurich. pp. 325–330.

Published by NRC Research Press


Westgate et al. 27

Dingle, H.R.C., White, D.J., and Gaudin, C. 2008. Mechanisms of pipe embed- Randolph, M.F., and Quiggin, P. 2009. Non-linear hysteretic seabed model for
ment and lateral breakout on soft clay. Canadian Geotechnical Journal, 45(5): catenary pipeline contact. In Proceedings of the International Conference on
636–652. doi:10.1139/T08-009. Offshore Mechanical and Arctic Engineering, Honolulu. Paper OMAE2009-
Einav, I., and Randolph, M.F. 2005. Combining upper bound and strain path 79259.
methods for evaluating penetration resistance. International Journal for Nu- Randolph, M.F., and White, D.J. 2008a. Upper-bound yield envelopes for pipe-
merical Methods in Engineering, 63: 1995–2016. doi:10.1002/nme.1350. lines at shallow embedment in clay. Géotechnique, 58(4): 297–301. doi:
Gaudin, C., and White, D.J. 2009. New centrifuge modelling techniques for 10.1680/geot.2008.58.4.297.
investigating seabed pipeline behavior. In Proceedings of the 17th Interna- Randolph, M.F., and White, D.J. 2008b. Pipeline embedment in deep water:
tional Conference on Soil Mechanics and Geotechnical Engineering, Alexan- processes and quantitative assessment. In Proceedings of the Offshore Tech-
dria, Egypt. 5–9 October. nology Conference, Houston, Tex., 5–8 May. OTC19128.
Green, A.P. 1954. The plastic yielding of metal junctions due to combined shear Randolph, M.F., Jewell, R.J., Stone, K.J.L., and Brown, T.A. 1991. Establishing a
and pressure. Journal of the Mechanics and Physics of Solids, 2: 197–211. new centrifuge facility. In Proceedings of the International Conference on
doi:10.1016/0022-5096(54)90025-3. Centrifuge Modelling, Boulder, Colo. pp. 2–9.
Hodder, M.S., and Cassidy, M.J. 2010. A plasticity model for predicting the verti- Randolph, M.F., Low, H.E., and Zhou, H. 2007. In situ testing for design of pipe-
cal and lateral behaviour of pipelines in clay soils. Géotechnique, 60(4): line and anchoring systems. In Proceedings of the 6th International Confer-
247–263. doi:10.1680/geot.8.P.055. ence on Offshore Site Investigation and Geotechnics, Society for Underwater
Lenci, S., and Callegari, M. 2005. Simple analytical models for the J-lay problem. Technology, London. pp. 251–262.
Acta Mechanica, 178: 23–39. doi:10.1007/s00707-005-0239-x. Stewart, D.P., and Randolph, M.F. 1994. T-Bar penetration testing in soft clay.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by 91.106.74.120 on 05/04/16

Lund, K.M. 2000. Effect of increase in pipe penetration from installation. In Journal of the Geotechnical Engineering Division, ASCE, 120(12): 2230–2235.
Proceedings of the Conference on Offshore, Marine and Arctic Engineering, doi:10.1061/(ASCE)0733-9410(1994)120:12(2230).
New Orleans, February. Westgate, Z.J., Randolph, M.F., White, D.J., and Li, S. 2010a. The influence of sea
Merifield, R., White, D.J., and Randolph, M.F. 2008. The ultimate undrained state on as-laid pipeline embedment: a case study. Applied Ocean Research,
resistance of partially embedded pipelines. Géotechnique, 58(6): 461–470. 32(3): 321–331. doi:10.1016/j.apor.2009.12.004.
doi:10.1680/geot.2008.58.6.461. Westgate, Z.J., White, D.J., Randolph, M.F., and Brunning, P. 2010b. Pipeline
Merifield, R.S., White, D.J., and Randolph, M.F. 2009. Effect of surface heave on laying and embedment in soft fine-grained soils: field observations and nu-
response of partially embedded pipelines on clay. Journal of Geotechnical merical simulations. In Proceedings of the Offshore Technology Conference.
and Geoenvironmental Engineering, 135(6): 819–829. doi:10.1061/(ASCE)GT. Paper OTC20407.
1943-5606.0000070. Westgate, Z.J., White, D.J., and Randolph, M.F. 2012. Field observations of as-laid
Murff, J.D., Wagner, D.A., and Randolph, M.F. 1989. Pipe penetration in cohesive pipeline embedment in carbonate sediments. Géotechnique, 62(9): 787–798.
soil. Géotechnique, 39(2): 213–229. doi:10.1680/geot.1989.39.2.213. doi:10.1680/geot.12.OG.001.
Palmer, A. 2009. Touchdown indentation of the seabed. Applied Ocean Re- White, D.J., Gaudin, C., Boylan, N., and Zhou, H. 2010. Interpretation of T-bar
search, 30(3): 235–238. doi:10.1016/j.apor.2008.09.004. penetrometer tests at shallow embedment and in very soft soils. Canadian
Pesce, C.P., Aranha, J.A.P., and Martins, C.A. 1998. The soil rigidity effect in the Geotechnical Journal, 47(2): 218–229. doi:10.1139/T09-096.
touchdown boundary layer of a catenary riser: Static problem. In Proceedings Yan, Y., White, D.J., and Randolph, M.F. 2011. Penetration resistance and stiffness
of the 8th International Offshore and Polar Engineering Conference, Mon- factors for hemispherical and toroidal penetrometers in uniform clay. Inter-
tréal. pp. 207–213. national Journal of Geomechanics, 11(4): 263–275. doi:10.1061/(ASCE)GM.1943-
For personal use only.

Randolph, M.F., and Hope, S. 2004. Effect of cone velocity on cone resistance and 5622.0000096.
excess pore pressures. In Proceedings of the International Symposium on Zhou, H., and Randolph, M.F. 2009. Numerical investigations into cycling of
Engineering Practice and Performance of Soft Deposits, Osaka, Yodogawa full-flow penetrometers in soft clay. Géotechnique, 59(10): 801–812. doi:
Kogisha Co. Ltd. pp. 147–152. 10.1680/geot.7.00200.

Published by NRC Research Press

You might also like