You are on page 1of 10

Ocean Engineering 88 (2014) 204213

Contents lists available at ScienceDirect

Ocean Engineering
journal homepage: www.elsevier.com/locate/oceaneng

Installation and capacity of dynamically embedded plate anchors


as assessed through centrifuge tests
C.D. OLoughlin a,n, A.P. Blake a,1, M.D. Richardson b,2, M.F. Randolph a,3, C. Gaudin a,4
a
Centre for Offshore Foundation Systems, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia
b
Advanced Geomechanics, 5254 Monash Avenue, Nedlands, WA 6009, Australia

art ic l e i nf o a b s t r a c t

Article history: A dynamically embedded plate anchor (DEPLA) is a rocket or dart shaped anchor that comprises a
Received 27 February 2013 removable central shaft and a set of four ukes. Similar to other dynamically installed anchors, the
Accepted 22 June 2014 DEPLA penetrates to a target depth in soft seabed sediments by the kinetic energy obtained through
Available online 12 July 2014
free-fall in water and the self-weight of the anchor. In this paper DEPLA performance was assessed
Keywords: through a series of beam centrifuge tests conducted at 200 times earth's gravity. The results show that
Dynamic anchors the DEPLA exhibits similar behaviour to other dynamically installed anchors during installation, with tip
Plate anchors embedments of 1.62.8 times the anchor length. After anchor installation the central shaft of the DEPLA,
Kaolin termed a follower, is retrieved and reused for the next installation, leaving the DEPLA ukes vertically
Bearing capacity factors
embedded in the soil. The loaddisplacement response during follower retrieval is of interest, with
Centrifuge modelling
mobilisation of frictional and bearing resistance occurring at different rates. The load required to extract
the DEPLA follower is typically less than three times its dry weight. The vertically embedded DEPLA
ukes constitute the load bearing element as a circular or square plate. The keying and pullout response
of this anchor plate is similar to other vertically embedded plate anchors, with an initial stiff response as
the anchor begins to rotate, followed by a softer response as the rotation angle increases, and a nal stiff
response as the effective eccentricity of the padeye reduces and anchor capacity is fully mobilised.
For the padeye eccentricity ratios considered (0.380.63 times the plate breadth or diameter), the loss in
plate anchor embedment is between 0.50 and 0.66 times the corresponding plate breadth or diameter.
Finally, the bearing capacity factors determined experimentally are typically in the range 14.215.8 and
are higher than numerical solutions for at circular and square plates. This is considered to be due to the
cruciform uke arrangement which ensures that the failure surface extends to the edge of the
orthogonal ukes and mobilises more soil in the failure mechanism.
& 2014 Elsevier Ltd. All rights reserved.

1. Introduction plate anchors or suction caissons, although a number of other


anchoring systems such as suction embedded plate anchors (Zook
Floating oil and gas installations in deep water require anchor- and Keith, 2009) and dynamically installed anchors (Brando et al.,
ing systems that can withstand high components of vertical load, 2006; Zimmerman et al., 2009) have been used with effect in
whilst being economical to install. Deep water installations are recent years. Predicting plate anchor capacity for a given embed-
typically moored using either (drag embedded) vertically loaded ment is generally straightforward as it is a function of the local
undrained shear strength, the projected area of the plate and a
dimensionless bearing capacity factor, which, for deeply
n embedded plates, is typically in the range 11.714.3 depending
Corresponding author. Tel.: 61 8 6488 7326.
E-mail addresses: conleth.oloughlin@uwa.edu.au (C.D. OLoughlin), on plate geometry and roughness (Merield et al., 2003; Song and
anthony.blake@research.uwa.edu.au (A.P. Blake), Hu, 2005; Song et al., 2008). However, predicting the anchor
markr@ag.com.au (M.D. Richardson), mark.randolph@uwa.edu.au (M.F. Randolph), installation trajectory and hence nal embedment depth is more
christophe.gaudin@uwa.edu.au (C. Gaudin). challenging and requires accurate shear strength data over a wide
1
Tel.: 61 8 6488 3974.
2
Tel.: 61 8 9423 3300.
seabed footprint (Ehlers et al., 2004). Suction caissons are advan-
3
Tel.: 61 8 6488 3075. tageous in this regard as embedment is monitored and controlled,
4
Tel.: 61 8 6488 7289. although successful prediction of the ultimate holding capacity

http://dx.doi.org/10.1016/j.oceaneng.2014.06.020
0029-8018/& 2014 Elsevier Ltd. All rights reserved.
C.D. OLoughlin et al. / Ocean Engineering 88 (2014) 204213 205

must account for a number of factors including the caisson key to an orientation that is normal or near normal to the
geometry and padeye location, loading angle, time effects and direction of loading. In this way, the maximum projected area is
the integrity of the internal seal provided by the soil plug presented to the direction of loading, ensuring that maximum
(Randolph et al., 2011). Several disadvantages of these two anchor anchor capacity is achievable through bearing resistance. The
types have been mitigated by the suction embedded plate anchor installation and keying processes are shown schematically in
(SEPLA) described by Wilde et al. (2001). The SEPLA employs a Fig. 2.
suction caisson to install a vertically oriented plate anchor. After This paper provides DEPLA performance data through a series
installation the follower is removed (and reused) and the of centrifuge tests that were designed to quantify expected
embedded plate is rotated or keyed so that it becomes normal or embedment depths and plate anchor capacities.
near normal to the load applied by the mooring chain.
Anchor installation operations become increasingly more com-
plex, time-consuming and hence costly as water depth increases. 2. Centrifuge testing programme
However, this is not the case with dynamically installed anchors as
no external energy source or mechanical operation is required The centrifuge tests were carried out at 200g using the xed
during installation. Dynamically installed anchors are rocket or beam centrifuge at The University of Western Australia (UWA). The
torpedo shaped and are designed so that, after release from a UWA beam centrifuge is a 1.8 m radius Acutronic centrifuge with a
designated height above the seaoor, they will penetrate to a maximum payload of 200 kg at 200g (Randolph et al., 1991).
target depth in the seabed by the kinetic energy gained during
free-fall. Centrifuge model tests indicate that in normally con- 2.1. Model anchors and chain
solidated clay, expected penetration depths are 23 times the
anchor length and expected anchoring capacities are 35 times the The model DEPLAs used in the centrifuge tests comprised a set
anchor dry weight (OLoughlin et al., 2004). In contrast, expected of DEPLA ukes (which form the plate anchor after the follower is
anchoring capacities for vertically loaded plate anchors are typi- retrieved), a DEPLA follower and anchor lines connected to the
cally 3040 times the anchor dry weight in very soft clay. follower and plate anchor padeyes. An interchangeable modular
A new anchoring system, termed the dynamically embedded design for the follower (ellipsoidal tip, shaft, padeye) allowed
plate anchor (DEPLA), combines the advantages of dynamically various steel and aluminium sections to be combined to form
installed anchors and vertically loaded anchors in much the same different overall follower lengths and masses, although the fol-
way as the SEPLA combines the advantages of suction caissons lower diameter and ellipsoidal tip length remained constant at
with vertically loaded anchors. The DEPLA is a rocket or dart Df 6.0 mm and Ltip 11.4 mm (see Fig. 3). The overall length of
shaped anchor that comprises a removable central shaft or the follower (and hence height of the anchor) used in the
follower that may be fully or partially solid and a set of four centrifuge tests was either 51.5 mm, 76.0 mm or 101.5 mm, which
ukes (see Fig. 1). A stop cap at the upper end of the follower at 200g corresponds to equivalent prototype heights of 10.3 m,
prevents it from falling through the DEPLA sleeve and a shear pin 15.2 m and 20.3 m respectively.
connects the ukes to the follower. As with other dynamically The ukes were fabricated from 0.8 mm thick steel and the
installed anchors, the DEPLA penetrates to a target depth in the sleeve from 0.75 mm thick copper (rather than steel) to permit
seabed by the kinetic energy obtained through free-fall. After soldering of the ukes to the sleeve without inducing high
embedment the follower line is tensioned, which causes the shear temperatures. The outer diameter of the sleeve is 7.8 mm, which,
pin to part allowing the follower to be retrieved for the next together with the sleeve wall thickness of 0.75 mm, gave a
installation, whilst leaving the anchor ukes vertically embedded nominal clearance of 0.15 mm between the 6.0 mm diameter
in the seabed. The embedded anchor ukes constitute the load follower and the sleeve. The DEPLA ukes that formed the plate
bearing element as a plate anchor. A mooring line attached to the anchor element were either circular or square, with the latter
embedded ukes is then tensioned, causing the ukes to rotate or oriented in a diamond shape on the DEPLA sleeve (see Fig. 3).

Fig. 1. Dynamically embedded plate anchor concept.


206 C.D. OLoughlin et al. / Ocean Engineering 88 (2014) 204213

Fig. 2. DEPLA keying stages: 1, DEPLA installed through self-weight & kinetic energy obtained during freefall; 2, follower removed and reused shear pin connection parts
leaving DEPLA ukes embedded in soil; 3, DEPLA ukes keyed into position; and 4, nal DEPLA orientation under load.

Fig. 3. 1:200 reduced scale DEPLAs.

Table 1
Model anchors.

Anchor ref. Flukes Lf (mm) Lf/Df Hs (mm) D or B (mm) D/Lf or B/Lf e (mm) e/B or e/D Mass (g)

Follower Flukes (plate) Total

A1 Circular 76.00 12.67 28.97 30.00 0.39 13.50 0.45 10.46 11.00 21.46
A2 Circular 76.00 12.67 21.21 22.60 0.35 9.80 0.43 10.46 6.09 16.55
A3 Circular 76.00 12.67 36.78 37.60 0.49 17.30 0.46 10.46 16.12 26.58
A4 Circular 51.50 8.58 12.81 16.00 0.31 6.00 0.38 9.29 2.91 12.20
A5 Circular 101.5 16.92 28.97 30.00 0.30 13.50 0.45 12.04 11.00 23.04
A6 Square 76.00 12.67 30.00 26.73 0.35 16.78 0.63 10.46 10.75 21.21

The diameter, D, or breadth, B, of the DEPLA ukes varied between plates and e/B 0.63 for the square plate. Details of the six model
16 mm and 37.6 mm, which corresponds to 3.27.5 m in equiva- anchors employed for the centrifuge tests are summarised in
lent prototype scale. The plate anchor padeye was located at an Table 1 and shown in Fig. 3.
eccentricity, e, from the central axis of the plate, such that the Both the follower and plate anchor lines were modelled using
eccentricity ratio was in the range e/D0.380.46 for the circular uncoated stainless steel wire of 0.90 mm diameter (180 mm in
C.D. OLoughlin et al. / Ocean Engineering 88 (2014) 204213 207

Table 2
Kaolin clay properties (after Stewart, 1992).

Liquid limit, LL (%) 61


Plastic limit, PL (%) 27
Specic gravity, Gs 2.6
Angle of internal friction, 0 (deg) 23
Voids ratio at p0 1 kPa on critical state line, ecs 2.14
Slope of normal consolidation line, 0.205
Slope of swelling line, 0.044
Strength ratio, su = 0v 0.19
Coefcient of vertical consolidation, cv (m2/yr)
4.5
(at OCR 1 and 0v 150 kPa)

prototype scale) and a tensile capacity of 660 N. This line was


selected due to its exibility, high tensile strength and resistance
to stretching. The follower line was attached to the padeye located
Fig. 4. Undrained shear strength proles.
at the top of the follower, whereas the plate anchor line was
connected to the plate anchor padeye located on one of the anchor
ukes. The other extremity of each anchor line was connected via 2.3. Experimental arrangement and testing procedures
a 1.7 kN load cell to the actuator that effected the follower
recovery and the plate anchor pullout. The model DEPLAs were installed in-ight by allowing them to
free-fall through the elevated acceleration eld from a drop
2.2. Soil properties height of 300 mm. This approach has been adopted for the
installation of dynamically installed torpedo piles in centrifuge
Normally consolidated clay samples were prepared by mixing tests (Richardson et al., 2009; OLoughlin et al., 2013) and has
commercially available kaolin clay powder with water to form a resulted in anchor impact velocities of the order of 30 m/s at 200g,
slurry at 120% (i.e., twice the liquid limit, see Table 2) and similar to that expected in the eld. The rotation of the centrifuge
subsequently de-aired under a vacuum close to 100 kPa. The clay requires the anchors to be installed through a guide to prevent
slurry was carefully transferred into a rectangular strongbox, with lateral movement of the anchor during free-fall due to Coriolis
internal dimensions 650 mm long, 390 mm wide and 325 mm acceleration. The anchor installation guide (shown in Fig. 5) was
deep, before self-weight consolidation in the centrifuge at 200g for fabricated from PVC and featured a 1.0 mm wide slot to accom-
approximately four days to create a normally consolidated clay modate one of the four DEPLA ukes and a 8.3 mm diameter
sample. A 10 mm sand layer at the base of the clay permitted channel machined into the front face of the guide to accommodate
drainage towards the base of the sample and a 510 mm layer of the DEPLA sleeve (7.8 mm diameter). Two PVC rails with an
water was maintained at the sample surface to ensure saturation. internal circular prole of 8.3 mm diameter were attached to the
Additional clay slurry was added to the sample during consolida- guide with brackets to prevent the model DEPLA falling away from
tion to achieve the required sample height of 230 mm. Pore the guide. The rails permitted an open slot over the length of the
pressure transducers placed at various depths within the sample guide, allowing unobstructed access for the follower and plate
allowed for the measurement of excess pore water pressures and anchor lines. Anchor release was achieved in-ight by a resistor
these, together with T-bar penetration tests, served to indicate the which, when supplied with current, heated and subsequently
degree of consolidation of the sample. burned through an anchor release cord attached to a DEPLA uke.
Soil characterisation tests were performed using a T-bar penet- This eliminated the need for a shear pin, since the release cord
rometer (Stewart and Randolph, 1991) from which a continuous connected to the DEPLA uke ensured that the follower and ukes
prole of the undrained shear strength, su, was derived using the stayed together prior to release. The anchor velocity was measured
commonly adopted T-bar factor, NT-bar 10.5. The T-bar tests were using multiple photoemitterreceiver pairs (PERPs) positioned at
conducted at a velocity of 1 mm/s, which corresponds to a 10 mm intervals along the guide. The DEPLA uke travelling
dimensionless velocity, V vdT-bar/cv in the range 32113 (where through the guide slot during free-fall would temporarily interrupt
v is the velocity, dT-bar the T-bar diameter and cv the vertical each PERP signal and the time between the trailing edges of
coefcient of consolidation, varying from 1.4 to 5 m2/yr for 0v in consecutive PERP signals together with the xed PERP spacing of
the range 10300 kPa), thus ensuring undrained conditions 10 mm allowed the instantaneous velocity of the DEPLA to be
(House et al., 2001). determined at various points above the clay surface.
T-bar penetration tests were conducted during the latter stages The testing arrangement is shown in Fig. 6a. Before each DEPLA
of consolidation and before, during and after the anchor tests to installation the model DEPLA was positioned within the installa-
observe any changes in strength with time. Typical shear strength tion guide such that the follower tip was 300 mm above the clay
proles from Samples 1 and 2 conducted before and after testing surface (apart from one test with a 200 mm drop height). The
are provided in Fig. 4. Centrifuge strength data tend to deviate follower line was connected to the actuator and the plate anchor
from the expected linear prole for normally consolidated clays in line was temporarily connected to a xed point at the base of the
the lower third of the sample. This effect is due to the slight actuator with sufcient slack such that it would not impede the
increase in radial acceleration level and effective unit weight with ight of the DEPLA during installation. Each DEPLA test was
increasing sample depth. The variation in radial acceleration and conducted at 200g and comprised the following stages:
effective unit weight has therefore been accounted for in the
theoretical strength prole in Fig. 4, which was obtained using an 1. Anchor installation. The model DEPLA was released by burning
undrained shear strength ratio, su = 0v 0:22. Although the theore- through the temporary release cord (as described previously)
tical and measured proles deviate beyond 140 mm, the theore- and the DEPLA travelled vertically through the installation
tical prole provides a good approximation of the strength over guide before impacting the clay surface. After anchor release
the range of interest for the DEPLA plate (up to 103 mm). the uke passing the highest PERP in the installation guide
208 C.D. OLoughlin et al. / Ocean Engineering 88 (2014) 204213

Fig. 5. Anchor installation guide.

Fig. 6. Experimental arrangement and testing procedure: (a) prior to installation; (b) DEPLA embedment; (c) follower retrieval; and (d) anchor keying and pullout.

triggered the high speed data acquisition and PERP data were extracted by driving the actuator vertically away from the
logged at 100 kHz. sample surface at a rate of 0.3 mm/s; this rate was selected so
2. Follower recovery. The actuator was driven horizontally 43 mm that the dimensionless velocity, V vDf/cv (where v is the
(offset between the vertical axes of the actuator and the extraction velocity, Df is the follower diameter and cv is the
installation guide) such that the vertical axis of the actuator vertical coefcient of consolidation quantied previously)
was directly above the anchor drop site. The follower was exceeded 10, so that the response is primarily undrained
C.D. OLoughlin et al. / Ocean Engineering 88 (2014) 204213 209

(House et al., 2001). Follower extraction was continued until 3.1. Impact velocity
video footage conrmed that the follower was free from the
clay surface. Anchor velocities determined from the measured PERP data are
3. Exchange anchor line. Immediately after follower extraction the shown in Fig. 7 together with theoretical velocity proles gener-
centrifuge was stopped for approximately ve minutes to ated using the equations of motion whilst accounting for the
disconnect the follower line from the load cell and to connect radially varying acceleration eld in the centrifuge. The theoretical
the plate anchor line. The centrifuge was then restarted and a velocity proles have been adjusted to t the PERP data; this
reconsolidation period of just over 13 min (equivalent to one adjustment is necessary to account for friction along the anchor
prototype year at 200g) was permitted before commencing the guide interface, which varies in accordance with the contact area
anchor pullout test. The DEPLA plate embedment depth between the plate anchor and the guide. The reduction of the
remains constant during this process. theoretical velocity for the three tests shown in Fig. 7 is in the
4. Plate anchor keying and pullout. The plate anchor keying and range 816%, which is typical for the entire dataset. This is
pullout were performed continuously at an actuator rate of consistent with a kinetic friction coefcient, k 0.170.36,
0.1 mm/s such that V 410 (for the lowest plate anchor dia- between the anchor and the guide, which is reasonable for a
meter, D 16 mm) to ensure undrained conditions. Each anchor plasticmetal interface. As shown in Fig. 7, the impact velocity is
pullout was vertical, causing the plate anchor to key from the taken from the adjusted velocity prole at the clay surface. The
initial vertical orientation to a nal horizontal orientation, impact velocities achieved in the centrifuge tests (for the 300 mm
before loading to failure. Pullout of the plate anchor was drop height) are in the narrow range 27.530.0 m/s, which is
continued until the plate was fully extracted from the clay. typical for a drop height of 300 mm (Richardson, 2008) and is at
the upper end of the range from eld experience of vi in the range
24.527.0 m/s (Lieng et al., 2010) and 1627 m/s (Brando et al.,
The presence of a radial acceleration eld limited anchor 2006).
installation sites to the longitudinal centreline of the beam
centrifuge strongbox. To minimise interaction effects, installation 3.2. Embedment depth
sites were located at least 20 DEPLA follower diameters (i.e.,
120 mm) from rigid boundaries and adjacent installation sites The DEPLA tip embedment (ze,tip) was determined using the
were separated by at least 12 DEPLA follower diameters (i.e., approach adopted by Richardson (2008) (see Fig. 8), whereby ze,tip
72 mm). In addition to the eight dynamic installation tests, two is calculated from
further installations were conducted in which the DEPLA was ze;tip zchain  zLC  zslack Lf 1
jacked at a constant penetration rate of 1 mm/s at 1g using a
custom-built 6 mm diameter shaft connected to the follower where zLC is the initial vertical position of the actuator load cell
padeye. The centrifuge was then accelerated to 200g and the above the sample surface, zchain is the length of the follower chain,
anchor keying and pullout were conducted in accordance with Lf is the follower length and zslack is the amount of slack removed
Stage 4 described above. As dynamic installation is known to from the follower chain prior to the onset of a signicant tensile
reduce short term anchor capacity by reducing the effective load during follower extraction. The vertical installation guide
stresses in the soil surrounding the embedded anchor located above the centerline of the soil sample prevents the
(Richardson et al., 2009), these static installation tests provided a anchor tilting during the freefall stage. Although there is potential
basis for examining the effect of dynamic installation on the for the anchor to deviate slightly from a vertical trajectory for the
keying and capacity of the plate anchor. remaining one anchor length (on average) embedment, previous
centrifuge studies reported by OLoughlin et al. (2009) report a
maximum inclination of 31 for dynamically installed anchors in
normally consolidated kaolin clay. In the case of the DEPLA, a 31
inclination would have little bearing on the capacity mobilisation
3. Results and discussion of the plate, other than a negligible error in determining the
embedment depth using Eq. (1).
Table 3 summarises the main results from the 10 centrifuge DEPLA tip embedments were in the range 144186 mm, or 1.6
tests considered here. The results are discussed in further detail in 2.8 times the length of the DEPLA follower, where the lower and
the following sections. upper bounds correspond to the longest and shortest DEPLA

Table 3
Summary of test results.

Test Anchor Installation Impact Tip Follower Initial anchor Normalised Final anchor Plate Undrained Net plate Bearing
no. ref. method velocity, embedment extraction embedment, loss in plate embedment, embedment shear anchor capacity
vi (m/s) depth, load (N) ze,plate (mm) anchor ze,plate (mm) ratio, strength at capacity, factor,
ze,tip (mm) embedment, ze,plate/D or peak Fv,net (N) Nc
ze,plate/D or ze,plate/B capacity,
ze,plate/B su,p (kPa)

1 A1 Dynamic 23.87 155.4 69.99 95.4 0.55 79.0 2.63 21.5 238.44 15.8
2 A1 Dynamic 30.04 161.4 78.37 101.4 0.59 83.7 2.79 24.1 297.70 17.3
3 A6 Dynamic 27.53 162.4 61.79 102.9 0.61 86.5 3.24 23.5 233.31 14.5
4 A2 Dynamic 28.50 156.9 65.12 93.0 0.64 78.5 3.47 23.5 137.78 14.9
5 A4 Dynamic 29.17 145.9 40.84 102.3 0.57 93.1 5.82 29.3 71.10 14.8
6 A5 Dynamic 29.37 186.4 75.93 100.9 0.55 84.4 2.81 24.5 220.62 12.8
7 A6 Dynamic 29.40 144.4 60.74 84.9 0.63 68.0 2.54 20.0 204.34 14.9
8 A3 Dynamic 29.17 150.6 45.55 94.5 0.66 69.6 1.85 21.7 378.51 15.8
9 A1 Static 160.6 100.5 0.50 85.4 2.85 23.0 229.66 14.2
10 A6 Static 157.0 97.5 0.53 83.4 3.12 23.8 241.94 14.9
210 C.D. OLoughlin et al. / Ocean Engineering 88 (2014) 204213

followers respectively (51.5 mm and 101.5 mm). The majority of although it may be due to secondary frictional resistance that
the test data relate to the 76 mm follower (L/d 12.7), for which occurs as the follower emerges from the upper end of the DEPLA
ze,tip is in the range 144162 mm, or 1.92.1 times the follower sleeve. The average displacement to Peak 1 was one follower
length. This is in good agreement with reported eld experience of diameter, with a range of 0.71.3 times the follower diameter
1.92.4 times the anchor length for a 79 t dynamically installed arising from uncertainties regarding the onset of follower move-
anchor with L/d 10.8 and 4 wide ukes (Lieng et al., 2010). ment. Interestingly the displacement between Peak 1 and Peak 3 is
also approximately one follower diameter.
The magnitude of the Peak 1 capacity ranged from 2.2 to
3.3. Follower extraction 3.2 times the dry weight of the follower, with capacity increasing
with follower length and hence frictional surface area. These
Fig. 9 presents typical load versus displacement data for the capacity ratios are lower than those reported for geometrically
extraction of the DEPLA follower. The extraction response was similar dynamically installed anchors (e.g., 35, OLoughlin et al.,
characterised by a sharp increase in load after the follower line 2004), mainly because the DEPLA sleeve reduces the frictional
slack had been taken up, towards an initial maximum (Peak 1) surface area of the follower by between 30% and 50% depending
followed by a sudden drop in load and a subsequent gradual on the model anchor conguration. Furthermore, minimal recon-
increase towards a secondary maximum (Peak 2) of lower magni- solidation was permitted before extraction of the follower (limited
tude than Peak 1. In some instances the load increased rapidly by the minimum time required to adjust the actuator position and
towards a further maximum (Peak 3), similar to Peak 1, before take up slack in the follower line). Thus the measured follower
reducing suddenly again. The Peak 1 and Peak 2 responses have extraction resistance represents short term capacity that is likely
been observed during extraction of geometrically similar dynami- to be 2030% of the long term capacity (Richardson et al., 2009).
cally installed anchors (Richardson et al., 2009), and is considered
to be due to the different rates of mobilisation for frictional and
bearing resistance as observed for suction caissons (Jeanjean et al., 3.4. Plate anchor keying and capacity
2006). It appears that high, but brittle, frictional resistance along
the follower shaft (either against soil or the DEPLA sleeve) is Example plate anchor keying and capacity curves are provided
mobilised rst (Peak 1), before a more gradual mobilisation of in Fig. 10 for a square and circular plate anchor installed both
bearing resistance at the follower padeye (Peak 2). The reason for dynamically and statically. The keying and anchor pullout
Peak 3 and its sporadic occurrence is not wholly understood, response is characterised by three distinct stages. After recovering

Fig. 7. DEPLA velocity proles during centrifuge free-fall. Fig. 9. Loaddisplacement response during follower retrieval.

Fig. 8. Determining DEPLA tip embedments.


C.D. OLoughlin et al. / Ocean Engineering 88 (2014) 204213 211

Fig. 10. Loaddisplacement response during anchor keying and pullout. Fig. 11. Dimensionless loaddisplacement response during all anchor keying and
pullout tests.

the slack in the plate anchor line, the load, Fv, starts to increase as adjusting the peak force by the submerged weight of the anchor
the anchor starts to key (Stage 1). During Stage 2 Fv begins to (Ws), is also indicated in Fig. 11, and is compared with theoretical
plateau and then reduce slightly. Similar observations were made bearing capacity factors later in the paper.
by Blake et al. (2010), but not by Gaudin et al. (2006) and Song The vertical displacement of the anchor padeye, dv, is in the
et al. (2006) in their respective centrifuge studies on square plates. range 0.961.12 times the diameter for circular plates and 1.16
The reason for this slight reduction is not understood, but may be 1.26 times the breadth for square plates. However dv represents
attributed to local softening of the soil as the plate rotates. As the displacement of the anchor padeye rather than the loss in plate
keying progresses the effective eccentricity of the padeye reduces, anchor embedment, ze,plate, which may be calculated as
which requires higher padeye tension to maintain sufcient ze;plate dv  e sin , where is the load inclination (to the
moment on the plate to continue rotation. This increase in tension horizontal) at the anchor padeye. Hence for a vertical pullout
is shown by Stage 3, which continues until the projected plate area ze,plate is in the range 0.500.68 times the diameter for circular
reaches its maximum value, at which point a peak capacity is plates and 0.530.63 times the breadth for square plates. This is
observed before reducing again as the plate displaces towards the shown in Fig. 12 together with other centrifuge data for square,
soil surface through soil of reducing strength. It is noteworthy that rectangular (L/B 2) and strip anchors. The horizontal axis on
the dynamic installation process does not appear to have a Fig. 12 combines the two main parameters that inuence the
discernable effect on either the plate anchor keying response or keying response, padeye eccentricity and plate thickness (and
the magnitude of the peak load. hence anchor weight). Also shown in Fig. 12 are numerical data
The anchor keying process gives rise to two effects. Firstly, the reported by Wang et al. (2011), who showed that embedment loss
soil in the immediate vicinity of the plate will be partially increases with increasing shear strength for a given eccentricity
remoulded, and secondly, the anchor embedment depth will and anchor thickness to a limiting value at a normalised shear
reduce as the plate rotates and translates vertically (Randolph strength, su = 0a B  0:75. Wang et al. (2011) approximate this max-
et al., 2005). Blake et al. (2010) demonstrated that the reduction in imum loss in embedment for square anchors as
soil strength due to the former effect reduces plate anchor capacity
by about 33%, with the capacity recoverable in due course through ze; max 0:144
2
consolidation. However, the loss of embedment due to the latter B e=Bt=B0:2 1:15
effect is more signicant as any loss in embedment will corre-
spond to a non-recoverable loss in potential anchor capacity (for Eq. (2) is also shown in Fig. 12 and represents the theoretical
offshore clay deposits that are typically characterised by an maximum loss in embedment for a given anchor eccentricity and
increasing strength prole with depth). thickness. Experimental data for the DEPLA square plates are in
The loss in plate anchor embedment during keying is demon- reasonable agreement with Eq. (2), ranging between 94% and 113%
strated by Fig. 11, which plots dimensionless loaddisplacement of ze,max calculated using Eq. (2). Applying Eq. (2) to the circular
response for all tests. Loads are normalised by the area of the plate, DEPLA plate, but with D instead of B, reduces the agreement to
A, and the undrained shear strength at the estimated anchor 6179%. The poorer agreement for the circular plate data is not
embedment at peak capacity, su,p, which was selected using the unexpected as Wang et al. (2011) note that the maximum embed-
T-bar test data most relevant to each DEPLA test. The vertical ment loss differs for different plate shapes. Fig. 12 indicates that
displacement of the anchor padeye is normalised by either the the loss in DEPLA plate embedment during keying could be
diameter (D) or breadth (B) of the plate anchor (for circular and reduced if either the eccentricity or anchor thickness was
square plates respectively). The occasional sharp drops in the increased (relative to the plate diameter or breadth), or if the soil
response curves are considered to be due to minor mechanical was weaker (relative to the initial moment applied to the anchor
effects such as adjustments of the anchor line at the padeye. In the at the onset of keying). The most inuential of these factors is the
main, the loaddisplacement responses in Fig. 11 are typical of those eccentricity (OLoughlin et al., 2006; Song et al., 2009; Wang et al.,
observed experimentally for plate anchors (e.g., Song et al., 2006; 2011) and the embedment loss for the DEPLA should be lower if
Gaudin et al., 2006; Blake et al., 2010), with an initial stiff response the padeye eccentricity was increased (albeit that this would
as the anchor begins to rotate, followed by a softer response as the require a different plate geometry). Whilst an increase in plate
rotation angle increases, and a nal stiff response as the anchor thickness may be attractive in that it should reduce embedment
capacity is fully mobilised. The range of (FvWs)/Asu,p values, loss during keying and hence the loss in potential anchor capacity,
212 C.D. OLoughlin et al. / Ocean Engineering 88 (2014) 204213

Fig. 12. Loss in anchor embedment during keying. Experimental: DEPLA, circular (exp); DEPLA, square (exp); OLoughlin et al. (2006), strip (exp); OLoughlin
et al. (2006), L/B=2 (exp); Gaudin et al., (2009), strip (exp); Blake et al., (2010), square (exp); Song et al, (2009), square (exp). Numerical, Wang et al. (2011):
su = 0a B 0:01; su = 0a B 0:025; su = 0a B 0:05; su = 0a B 0:1; su = 0a B 0:015; su = 0a B 0:035; su = 0a B 0:075; su = 0a B 0:2.

this may be offset by the cost implications of increasing the


thickness of the plate.
Experimental bearing capacity factors for the DEPLA are calcu-
lated using

F v;net
Nc 3
Asu;p

where Fv,net is the peak load minus the submerged weight of the
plate anchor in soil, A is the area of the plate and su,p is the
undrained shear strength at the estimated anchor embedment at
peak capacity as dened previously. Experimentally derived Nc
factors are shown in Fig. 13 and are typically in the range
Nc 14.215.8. The average experimental Nc 15.0 for both circular
and square plates is in good agreement with Nc 15.3 back
analysed from reduced scale DEPLA eld trials (Blake and
OLoughlin, 2012). Corresponding solutions for rough at plate
anchors are lower, e.g., Nc 13.11 for an innitely thin circular Fig. 13. Comparison of experimental and numerical bearing capacity factors.
plate (Martin and Randolph, 2001), Nc 13.70 for a circular plate
with thickness equal to 5% of the diameter (Song et al., 2008) and
Nc 14.33 for a square plate with thickness equal to 5% of the 4. Conclusions
diameter (Wang et al., 2010). However, as noted by Zhang et al.
(2012), soil is forced to take a deeper and longer route to ow This paper has presented experimental data from a series of
around thicker plates, which mobilises more soil in the failure beam centrifuge tests undertaken to assess the penetration and
mechanism and results in higher bearing capacity factors. The capacity response of dynamically embedded plate anchors in
DEPLA cruciform uke arrangement ensures that the failure sur- normally consolidated clay. The main ndings from this study
face extends to at least half the plate diameter (D/2) from the are summarised in the following.
plate, which should result in a higher bearing capacity factor for
the DEPLA compared with a at plate. From a design perspective, a) At mudline impact velocities of approximately 30 m/s, DEPLA
anchor capacity would be based on the (higher) strength at the tip embedment depths are in the range 1.62.8 times the
initial plate location. The average experimental capacity factor length of the DEPLA follower, which is in good agreement with
determined in this way is Nc 11.8, 21% lower than the average reported eld experience of 1.92.4 times the anchor length for
Nc 15 based on the strength at the anchor embedment at peak a 79 t dynamically installed anchor with L/d 10.8 and 4 wide
capacity. ukes (Lieng et al., 2010).
The nal plate embedment of the anchor (at peak anchor b) Extraction of the DEPLA follower is characterised by an initial
capacity) is in the range 1.85.8 times the plate diameter or sharp increase in pullout resistance corresponding with fric-
breadth, with lower values corresponding with larger plates and tional resistance along the follower shaft, followed by reduc-
vice versa. The lower bound of this range is still sufciently high to tion in soil (and mechanical) friction and then a more gradual
mobilise a deep failure mechanism (Song and Hu, 2005; Song increase in pullout resistance corresponding with bearing
et al., 2008; Wang et al., 2010). This is supported by the observa- resistance at the follower padeye. In some instances a second
tion that the experimental Nc values do not reduce over this range frictional capacity peak was observed at approximately one
of normalised embedment, indicating that the DEPLA mobilises a follower diameter after the rst frictional capacity peak. This is
deep failure mechanism at relatively shallow plate embedment considered likely to arise from frictional resistance as the
ratios. follower emerges from the top of the DEPLA sleeve. Follower
C.D. OLoughlin et al. / Ocean Engineering 88 (2014) 204213 213

pullout resistances quantied using the initial capacity peak prole. In: Proceedings of the Offshore Technology Conference, OTC 18007.
were in the range 2.23.2 times the dry weight of the follower. Houston, Texas.
Lieng, J.T., Tjelta, T.I., Skaugset, K., 2010. Installation of two prototype deep
c) The anchor keying and pullout response is typical of that for penetrating anchors at the Gja Field in the North Sea. In: Proceedings of the
plate anchors, with an initial stiff response as the anchor begins Offshore Technology Conference, OTC 20758. Houston, Texas.
to rotate, followed by a softer response as the rotation angle Martin, C.M., Randolph, M.F., 2001. Application of the lower bound and the upper
bound theorems of plasticity to collapse of circular foundations. In: Proceedings
increases, and a nal stiff response as the effective eccentricity
of the 10th International Conference of the International Association for
of the padeye reduces and anchor capacity is fully mobilised. Computer Methods and Advances in Geomechanics, vol. 2. Tucson, pp. 1417
The dynamic installation process does not appear to have a 1428.
Merield, R.S., Lyamin, A.V., Sloan, S.W., Yu, H.S., 2003. Three-dimensional lower
discernable effect on either the plate anchor keying response or
bound solutions for stability of plate anchors in clay. J. Geotech. Geoenviron.
the magnitude of the peak load. Eng. 129 (3), 243253.
d) The loss in plate anchor embedment during keying is in the OLoughlin, C.D., Randolph, M.F., Richardson, M.D., 2004. Experimental and theore-
range 0.500.68 times the diameter for circular plates and tical studies of deep penetrating anchors. In: Proceedings of the Offshore
Technology Conference, OTC 16841. Houston, Texas.
0.530.63 times the breadth for square plates. The square plate OLoughlin, C.D., Lowmass, A., Gaudin, C., Randolph, M.F., 2006. Physical modelling
embedment loss data are within 94115% of the value calcu- to assess keying characteristics of plate anchors. In: Proceedings of the
lated using formulations proposed by Wang et al. (2011) for 6th International Conference on Physical Modelling in Geotechnics, vol. 1,
pp. 659665.
square plates, and may be reduced further by increasing the
OLoughlin, C.D., Richardson, M.D. and Randolph, M.F. (2009). Centrifuge tests on
padeye eccentricity. dynamically installed anchors. In: Proceedings of the 28th International
e) Experimental bearing capacity factors are typically in the range Conference on Ocean, Offshore and Arctic Engineering, OMAE2009-80238,
Nc 14.215.8, with an average Nc 15.0 for both square and Honolulu, Hawaii.
OLoughlin, C.D., Richardson, M., Randolph, M.F., Gaudin, C., 2013. Dynamic anchor
circular plates, which is higher than existing numerical solu- embedment in clay. Gotechnique 63 (11), 909919.
tions for rough at circular and square plates. The higher Randolph, M.F., Jewell, R.J., Stone, K.J.L., Brown, T.A., 1991. Establishing a new
bearing capacity factor for the DEPLA is considered to be due centrifuge facility. In: Proceedings of the International Conference on Centri-
fuge Modelling, Centrifuge 91. Boulder, Colorado, pp. 39.
to the cruciform uke arrangement which ensures that the Randolph, M.F., Cassidy, M.J., Gourvenec, S., Erbrich, C.T., 2005. Challenges of
failure surface extends to the edge of the orthogonal ukes. offshore geotechnical engineering. In: Proceedings of the 16th International
Conference on Soil Mechanics and Geotechnical Engineering, vol. 1. Osaka,
Japan, pp. 123176.
Randolph, M.F., Gaudin, G., Gourvenec, S.M., White, D.J., Boylan, N., Cassidy, M.J.,
2011. Recent advances in offshore geotechnics for deep water oil and gas
Acknowledgements developments. Ocean Eng. 38 (7), 818834.
Richardson, M.D., OLoughlin, C.D., Randolph, M.F., Gaudin, C., 2009. Setup follow-
ing installation of dynamic anchors in normally consolidated clayJ. Geotech.
This work forms part of the activities of the Centre for Offshore
Geoenviron. Eng. ASCE 135 (4)487496.
Foundation Systems (COFS), currently supported as a node of the Richardson, M.D., 2008. Dynamically Installed Anchors for Floating Offshore
Australian Research Council Centre of Excellence for Geotechnical Structures (Ph.D. thesis). The University of Western Australia.
Science and Engineering and as a Centre of Excellence by the Song, Z., Hu, Y., 2005. Vertical pullout behaviour of plate anchors in uniform clay.
In: Proceedings of the 1st International Symposium on Frontiers in Offshore
Lloyd's Register Foundation. The Lloyd's Register Foundation Geotechnics, ISFOG05. Perth, Australia, pp. 205211.
invests in science, engineering and technology for public benet, Song, Z., Hu, Y., Wang, D., OLoughlin, C.D., 2006. Pullout capacity and rotational
worldwide. The experimental work was funded by Enterprise behaviour of square anchors in kaolin clay and transparent soil. In: Proceedings
of the 6th International Conference on Physical Modelling in Geotechnics, vol.
Ireland (Grant no. CFTD-2008-318) through their Commercialisa-
2. Hong Kong, pp. 13251331.
tion Fund Technology Development Programme. Song, Z., Hu, Y., Randolph, M.F., 2008. Numerical simulation of vertical pullout of
plate anchors in clay. J. Geotech. Geoenviron. Eng. ASCE 134 (6), 866875.
Song, Z., Hu, Y., OLoughlin, C.D., Randolph, M.F., 2009. Loss in anchor embedment
References during plate anchor keying in clay. J. Geotech. Geoenviron. Eng. ASCE 135 (10),
14751485.
Blake, A.P., OLoughlin, C.D., Gaudin, C., 2010. Setup following keying of plate Stewart, D.P., 1992. Lateral Loading of Piled Bridge Abutments due to Embankment
anchors assessed through centrifuge tests in kaolin clay. In: Proceedings of the Construction (Ph.D. thesis). The University of Western Australia.
2nd International Symposium on Frontiers in Offshore Geotechnics, ISFOG2010. Stewart, D.P., Randolph, M.F., 1991. A new site investigation tool for the centrifuge.
Perth, Australia, pp. 705710. In: Proceedings of the International Conference on Centrifuge Modelling,
Blake, A.P., OLoughlin, C.D., 2012. Field testing of a reduced scale dynamically Centrifuge 91. Boulder, Colorado, pp. 531538.
embedded plate anchor. In: Proceedings of the 7th International Conference Wang, D, Hu, Y., Randolph, M.F., 2010. Three-dimensional large deformation nite
Offshore Site Investigation and Geotechnics. London, pp. 621628. element analysis of plate anchors in uniform clay. J. Geotech. Geoenviron. Eng.
Brando, F.E.N., Henriques, C.C.D., Arajo, J.B., Ferreira, O.C.G., Amaral, C.D.S., 2006. ASCE 136 (2), 355365.
Albacora Leste eld development FPSO P-50 mooring system concept and Wang, D, Hu, Y., Randolph, M.F., 2011. Keying of rectangular plate anchors in
installation. In: Proceedings of the Offshore Technology Conference, OTC 18243. normally consolidated clays. J. Geotech. Geoenviron. Eng. ASCE 137 (12),
Houston, Texas. 12441253.
Ehlers, C.J., Young, A.G., Chen, J.H., 2004. Technology assessment of deepwater Wilde, B., Treu, H., Fulton, T., 2001. Field testing of suction embedded plate anchors.
anchors. In: Proceedings of the Offshore Technology Conference, OTC 16840. In: Proceedings of the 11th International Offshore and Polar Engineering
Houston, Texas. Conference, pp. 544551.
Gaudin, C., OLoughlin, C.D., Randolph, M.F., Lowmass, A., 2006. Inuence of the Zhang, Y., Bienen, B., Cassidy, M.J., Gourvenec, S., 2012. Undrained bearing capacity
installation process on the performance of suction embedded plate anchors. of deeply buried at circular footings under general loading. J. Geotech.
Gotechnique 56 (6), 389391. Geoenviron. Eng. ASCE 138 (3), 385397.
Gaudin, C., Tham, K.H., Ouahsine, S., 2009. Keying of plate anchors in NC clay under Zimmerman, E.H., Smith, M.W., Shelton, J.T., 2009. Efcient gravity installed anchor
inclined loading. J. Offshore Polar Eng. 19 (2), 18. for deepwater mooring. In: Proceedings of the Offshore Technology Conference,
House, A.R., Oliveira, J.R. M.S., Randolph, M.F., 2001. Evaluating the coefcient of OTC 20117. Houston, Texas.
consolidation using penetration tests. Int. J. Phys. Model. Geotech. 1 (3), 1725. Zook, J.R. and Keith, A.L. (2009) Improvements in Efciency for Subsea Operations
Jeanjean, P., Znidarcic, D., Phillips, R., Ko, H.Y., Pster, S., Schroeder, K., 2006. in Deepwater Angola. In: Proceedings of the SPE/IADC Drilling Conference and
Centrifuge testing on suction anchors: doublewall, stiff clays, and layered soil Exhibition, SPE-119651-MS. Amsterdam, The Netherlands.

You might also like