You are on page 1of 8

The 12

th
International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India


Large Deformation FE Analysis of Plate Anchor Keying in Clay
Yuxia Hu, Zhenhe Song
Department of Civil Engineering, Curtin University of Technology, GPO BOX U1987, Perth 6845,WA, Australia
Keywords: Plate anchor, Anchor keying, Large deformation FE analysis, Anchor loss in embedment
ABSTRACT: In this paper, large deformation finite element analyses of plate anchor keying in clay were
performed. The anchor rotation was analysed using RITSS (Remeshing and Interpolation Technique with Small
Strain model) method. The RITSS method was incorporated into the AFENA finite element package to simulate
continuous anchor rotation during initial pullout. The effects of anchor thickness, anchor padeye eccentricity,
anchor submerged weight and soil non-homogeneity on anchor keying were investigated. The numerical results
were compared with existing centrifuge test data. The study shows that the absence of a chain tightening phase
in FE analysis contributes to the 12% less loss in anchor embedment during keying. The incomplete anchor
rotation in centrifuge tests contributes to 23% less anchor capacity. All different effects on anchor keying are
summarised into one formula to estimate anchor loss in embedment during keying.
1 Introduction
In recent years, oil and gas mining has moved into increasingly deeper water in search of undeveloped fields. For
water depths in excess of 500 m, conventional platforms are replaced by floating facilities, anchored to the
seabed using catenary or taut-wire moorings. The latter type of mooring imparts significant vertical loading to the
anchor, and consequently many different types of anchoring system have been developed (Dove et al. 1998;
Wilde et al. 2001). The SEPLA (Suction Embedded Plate Anchor) is one such system, which comprises a plate
anchor that is penetrated vertically using a caisson, and subsequently rotated by applying the anchoring force at
the anchor padeye until the plate becomes perpendicular to the applied force. This process is schematically
illustrated in Fig. 1. It has been conceived to combine the advantage of suction caissons (known penetration
depth and geographical location) and vertical loaded anchors (geotechnical efficiency and low cost).

Figure 1. Keying processes for the Suction Embedded Plate Anchor (SEPLA)
The anchor pullout and rotation process, before the full capacity of an anchor is reached, is commonly referred as
keying. During keying, the anchor moves upwards, thus embedment depth will reduce as the plate rotates
during pullout. As offshore clay deposits are typically characterised by an increasing strength profile with depth,
any loss in embedment will correspond to a non-recoverable loss in potential anchor capacity.
Reports on the loss of embedment of vertically installed anchors during keying show a large range. US Naval Civil
Engineering Laboratory guidelines (Rocker 1985) proposed that the loss of embedment during anchor keying was
twice the anchor breadth in cohesive soils, whilst recognising that the loss of embedment is also a function of
anchor geometry, soil type, soil sensitivity and duration of time between penetration and keying. Wilde et al.
(2001) reported in situ full scale and reduced scale onshore and offshore test results for SEPLAs in clay. Soil
3299


sensitivity was in the range 1.8 4.0 for the different test sites and the loss of embedment during keying was 0.5 -
1.7 times the anchor breadth, with lower embedment losses corresponding to higher soil sensitivities.
The effect of loading eccentricity on the keying process was studied by OLoughlin et al. (2006) in centrifuge.
Plate anchor displacement was quantified through a series of digitally captured images of the clay-Perspex
interface. The results showed a high impact of loading eccentricity on anchor loss in embedment, i.e. when
eccentricity ratio, e/B > 1 the loss in embedment is no greater than 0.1B; when e/B < 1 the loss in embedment
increases in a linear fashion to ~1.5B at e/B = 0.17.
The suction caisson installation effect was investigated by Gaudin et al. (2006). The centrifuge tests were
conducted in Kaolin clay. They found that for square anchors with e/B = 0.66, the loss of embedment during
keying for jacked-in anchors is in the range 1.3 - 1.5B, reducing to 0.9 1.3B for suction installed anchors. In
addition, load inclination angles indicate a strong effect on anchor loss in embedment during keying.
As outlined above, the current field and laboratory experimental database shows a wide range in loss of
embedment during anchor keying. This paper, using large deformation FE analysis, studies the factors that could
affect anchor keying. These factors include loading eccentricity, anchor submerged unit weight , pullout angle,
soil shear strength profile, anchor thickness and roughness. The FE results will be validated against existing
centrifuge test data.
2 Numerical Setup
Numerical analyses were conducted using the adaptive RITSS (Remeshing and Interpolation Technique with
Small Strain model) approach, which was implemented in a finite element package, AFENA (Carter and Balaam
1990). The RITSS approach (Hu and Randolph 1998a; Hu and Randolph 1998b) was used to simulate
continuous rotation during anchor keying.
Although modelling of square and rectangular anchor rotation is a 3D problem, to reduce computational time and
to simplify the problem, a two dimensional strip plate anchor pre-embedded vertically in clay was analysed. The
interface between the soil and the anchor plate was assumed to be rough, and as the plate anchors are
considered to be deeply embedded after installation (i.e. the soil failure zone does not extend to the soil surface),
a full attachment between anchor and soil is assumed. Although the majority of FE analysis deals with an anchor
without a shank, analyses with and without shank were performed to indicate the shank effect on FE analysis.



(a) After installation

(b) During keying
Figure 2 Loading conditions of anchor in the Finite Element analyses
The anchor loading eccentricity (e) is measured from the anchor padeye to the front face of the anchor plate. The
pullout force (F) is initially applied vertically to the anchor padeye through a vertically installed anchor chain (or
mooring line in Fig. 2a). When a pullout force (F) is applied to the anchor padeye, the anchor starts to rotate (Fig.
2b) and the resultant loading system is expressed as horizontal force (F
H
), vertical force (F
V
) and moment (M)
about the anchor centre below:
cos cos f F F
a H
+ = (1)

<
>
=
0 0
0 sin ' sin
V
V a a
V
F for
F for f W F
F

(2)
3300


[ ]

<
>
=
0 0
0 sin ' ) 90 ( sin
M for
M for e W e f e F
M
w a f a

(3)
where
a
is the angle of force F at the padeye to the horizontal (initially
a
= 90 in Fig. 2a); is the initial pullout
angle from pulley to anchor padeye;
0
is the chain angle (to the horizontal) at the soil surface and is the plate
anchor inclination to the horizontal; W
a
is the submerged anchor weight in soil, which relates to the soil bulk unit
weight
s
= 17 kN/m
3
and steel anchor unit weight
a
= 77 kN/m
3
. The eccentricity (e
w
) of W
a
is generated by the
combined weight of the plate and shank. The shank friction (f) acts in the opposite direction to anchor movement,
which is approximated as parallel to the anchor plate and located with eccentricity (e
f
) from the front face of the
anchor.
In simulating the anchor keying process, the forces developed along the anchor chain (particularly when the
anchor is not pulled out vertically) is also included. For an inclined anchor pullout, when the anchor chain slides
and cuts through the soil, an inverted catenary shape is formed, and this generates significant frictional capacity
along the length of the chain (Neubecker and Randolph 1995). The analytical solution proposed by Neubecker
and Randolph (1995), which relates the chain orientation, the chain tension and the chain bearing resistance per
unit length, is used in the present study to estimate the chain profile at any given stage during keying. Thus the
chain tension force at the anchor padeye can be estimated using:
Q H Qdz
F
H
z
a
=

=0
2
0
2
) (
2

(4)
where
0
is the chain angle at the soil surface, F is the chain tension at the (padeye) attachment point at depth H,
Q is the chain tension at a depth z and
Q
is the average bearing resistance (per unit length of chain) over the
depth from soil surface (z = 0) to the padeye embedment depth H.
The initial pullout angle at the padeye was set up vertically (Fig. 2a,
a
= 90). After the first step of remeshing in
the FE analysis, the position of the anchor and the whole domain was updated according to the anchor and chain
displacements. The new interaction point between the soil surface and chain element system can then be
calculated. Hence the new
0
can be used to calculate the new pullout angle
a
from Eq. 4. This updating process
is repeated until the keying process is complete. The following steps summarise this procedure:

Step 1: Set up the initial force F at the padeye vertically (
a
= 90);
Step 2: Use Equations 1 to 3 to calculate the resultant loading forces and moment applied to the anchor;
Step 3: Conduct incremental small-strain FE analyses;
Step 4: Update anchor location and chain profile;
Step 5: Calculate new
a
using Equation 4;
Step 6: Apply a new force F with the updated
a
;
Step 7: Stop if the anchor ultimate bearing capacity is reached, otherwise go to Step 2.
3 Numerical Results
The results of large deformation FE analysis were firstly compared with existing centrifuge test data. The
centrifuge plate anchor tests were performed in transparent soil, thus anchor rotational behaviour can be
recorded and measured (Song et al., 2006). After the validation of the large deformation FE analysis, an
extensive parametric study was carried out to examine the effects of various factors on the anchor keying
process. These factors include soil strength non-homogeneity, anchor padeye eccentricity, soil shear strength,
submerged anchor weight, loading inclination. Anchor thickness and roughness.
3.1 Validation of large deformation FE analysis
The validation of large deformation FE analysis was conducted by simulating the transparent soil centrifuge test
numerically. The details of centrifuge test and transparent soil sample setups can be found in Song et al. (2006).
The rotational behaviour of the anchor was determined by careful examination of the digital photos captured
during the test. According to the transparent soil test, the numerical analysis was set up as soil bulk unit weight
s

= 9.23 kN/m
3
, the anchor unit weight for steel
a
= 77 kN/m
3
and the undrained shear strength of soil is s
u
= 18
kPa. A square anchor 4 m by 4 m in prototype was tested in the centrifuge; while a strip anchor of 4 m in breath
was simulated in the large deformation FE analysis.

Figure 3 shows the development of anchor pullout capacity against anchor chain displacement. From the
centrifuge test data, it is apparent that the pullout process can be divided into four phases: (1) Chain tightening (1
to 2 for = 60 and 1 to 2 for = 90); (2) Half way anchor rotation (2 to 3 for = 60 and 2 to 3 for = 90); (3)
Full rotation and pullout capacity development (3 to 4 and 3 to 4); (4) Steady pullout (4 to 5 and 4 to 5).
However, the FE results show the absence of the first phase: chain tightening. This is because in centrifuge test,
3301


the anchor chain was loosely installed with the anchor. Thus, the chain had to be tightened first before the pullout
force could be transferred to the anchor padeye. On the contrary, in FE analysis with = 90
o
pullout (Fig. 3a), the
pullout force was acting on the anchor padeye from the start of the analysis. Thus the anchor pullout response
starts from anchor rotation, which is equivalent to the point 2 in Figure 3a. For the anchor with = 60
o
pullout (Fig.
3b), the anchor tightening is acting together with the chain cutting through the soil in the centrifuge test. However,
the numerical result only shows the response of chain cutting through soil without chain tightening. From both
cases in Fig. 3, it can be seen that, when anchor reaches its ultimate capacity, the chain displacement in the
centrifuge test is about 12% higher than the one in the FE analysis, which should be mainly due to the absence of
chain tightening phase in the FE analysis.
The lower ultimate anchor capacities in the centrifuge test are also evident in Fig. 3. This might be due to two
aspects: (1) the anchor final orientation () is about 20
o
in the centrifuge test and 0
o
in the FE results; (2) soil
remoulded strength could be effective during anchor keying. The first aspect, incomplete anchor rotation, might
be the major one and can be seen in Fig. 4a. The ultimate capacity in the centrifuge test is 77% of that from FE
analysis.

0
2
4
6
8
10
12
14
0 2 4 6 8 10 12
Chain Pullout distance in prototype (m)
q/s
u
1'
2'
3'
4'
5'
FE result
Centrifuge test data

(a) = 90
o

0
2
4
6
8
10
12
14
0 2 4 6 8
Chain Pullout distance in prototype (m)
q/s
u
1
2
3
5
FE result
Centrifuge test data
4

(b) = 60
o

Figure 3 Anchor capacity responses during keying
Figure 4 shows anchor rotational behaviour against anchor vertical displacement (or loss in embedment, z
e
)
during keying with and without the shank effect. The anchor vertical displacement is measured at the anchor
centre. As can be seen in Fig. 4a, the numerical analysis with anchor shank effect is in better agreement with the
centrifuge test data. The shank weight and friction have positive effect on anchor keying with reduced loss in
embedment.
Figure 4b plots the orientation of the plate anchor against the normalised loss in anchor embedment (z
e
/B) for
the inclined pullout plate anchor ( = 60
o
). The numerical result, with anchor shank weight and friction included,
agrees very well with the centrifuge test data. Thus, the anchor shank is included in all the other FE analysis,
unless specified. This is also true for the numerical results shown in Fig. 3.
0
10
20
30
40
50
60
70
80
90
100
0 0.1 0.2 0.3 0.4 0.5 0.6
z
e
/B
O
r
i
e
n
t
a
t
i
o
n

o
f

p
l
a
t
e

a
n
c
h
o
r

)
FE result without shank
FE result with shank
Centrifuge test data

(a) = 90
o

0
10
20
30
40
50
60
70
80
90
100
0 0.1 0.2 0.3 0.4 0.5
z
e
/B
O
r
i
e
n
t
a
t
i
o
n

o
f

p
l
a
t
e

a
n
c
h
o
r

(

)
FE result without shank
FE result with shank
Centrifuge test data

(b) = 60
o

Figure 4 Numerical simulation of transparent soil test
4 Factors Affecting Anchor Keying
Large deformation analyses were conducted to simulate the continuous movement of the plate anchor during
keying with different parameters that could affect anchor keying process. The parameters considered in this study
3302


include soil non-homogeneity, anchor padeye eccentricity, soil shear strength, submerged anchor weight, loading
inclination, anchor thickness and roughness.
4.1 Effect of Soil Non-Homogeneity
To study the effect of soil strength non-homogeneity on the anchor keying process, FE analyses were conducted
whereby a 4 m breath (B = 4 m) strip rough plate anchor was embedded in both uniform and normally
consolidated (NC) clays. The bulk unit weight of the soil was
s
= 17 kN/m
3
and the anchor unit weight was
considered as
a
= 77 kN/m
3
. The anchor padeye eccentricity ratio was e/B = 0.625 and the initial embedment
ratio H
i
/B was 3. The soil-anchor interface was assumed rough. The undrained shear strength of the uniform soil
was s
u
= 8.4 kPa whereas the undrained shear strength of the NC soil was s
u
= 0.7z kPa, where z is the soil
depth in metres. At the initial embedment depth of H
i
= 3B = 12 m, the undrained shear strength for both soils at
the anchor centre are the same at s
ui
= 8.4 kPa. Soil non-homogeneity around the plate was defined as (s
ut

s
ub
)/s
ui
, where s
ut
and sub are undrained shear strengths at the top and bottom edges of the vertical plate. This
can also be expressed as kB/s
ui
, where kB/s
ui
= 0 for uniform soil and kB/s
ui
= B/H
i
= 0.33 in NC soil for a plate
anchor embedded 3 times its breadth.
The anchor keying responses during vertical pullout ( = 90) are displayed in Figure 5, which shows the variation
in plate anchor inclination () as the anchor loses embedment during keying (z
e
). In these analyses, the anchor
shank weight and friction were not considered in order to isolate the effect of the strength heterogeneity. It is
evident from Fig. 5 that the anchor rotational behaviour is not influenced by the soil strength profiles during the
first 40 of rotation ( = 90 - 50). When the anchor orientation angle, , is less than 50, the anchor in NC clay
rotates slightly faster than in uniform clay. However, the difference is minimal. The final losses of embedment are
the same for both a homogenous and a heterogeneous strength profile. As the plate anchor is commonly
embedded at least 3 times its breadth in practice, the soil non-homogeneity will be 0.33 or less. Therefore, the
soil non-homogeneity will have minimal effect on the anchor keying process.
0
10
20
30
40
50
60
70
80
90
100
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
z
e
/B
O
r
i
e
n
t
a
t
i
o
n

o
f

p
l
a
t
e

a
n
c
h
o
r

)
s
u
= 8.4 kPa (kB/s
ui
=0)
s
u
= 0.7z kPa (kB/s
ui
=0.33)

Figure 5 Soil non-homogeneity effect on anchor keying
0
10
20
30
40
50
60
70
80
90
100
0 0.5 1 1.5 2 2.5 3
O
r
i
e
n
t
a
t
i
o
n

o
f

p
l
a
t
e

a
n
c
h
o
r

)
z
e
/B
Centrifuge test data (O'Loughlin et al. 2006)
FE results
e/B=1
e/B=0.5
e/B=0.17

Figure 6 Padeye eccentricity effect on anchor keying

4.2 Effect of Anchor Padeye Eccentricity
In order to investigate the effect of anchor padeye eccentricity ratio, numerical analyses were conducted with
varying loading eccentricity. Three eccentricity ratios were considered; e/B = 1, 0.5 and 0.17, to allow comparison
with the centrifuge test data reported by OLoughlin et al. (2006). In the centrifuge tests strip anchors (with
equivalent prototype dimensions B = 3 m and t = 0.2 m) were pre-embedded in NC Kaolin clay at an initial
embedment ratio of H
i
/B = 3. The undrained shear strength was determined using a T-bar penetrometer to give
an average s
u
= 0.7z kPa over the depth of the sample. The plate anchor was pulled out vertically with a rigid
shaft connected to the anchor padeye at e/B ratios of 1, 0.5 and 0.17, and the length of the strip anchor was
equal to the width of the testing chamber to ensure that the anchor remained in contact with the front Perspex
panel. The Perspex panel was digitally photographed, to facilitate observation and quantification of the keying
process. Experimental limitations restricted the final plate anchor orientation to = 20, rather than the expected
= 0, which is the same as the ones seen in Fig. 4.
Figure 6 displays the centrifuge test and FE results. With eccentricity ratios e/B = 1 and 0.5, the FE results agree
well with the centrifuge test data. However, with an eccentricity ratio e/B = 0.17, the anchor in numerical analysis
rotates much faster than the one in centrifuge test. This may be because that, in centrifuge test, the friction
experienced by the anchor ends, in contact with Perspex panel, is more profound with reduced loading
eccentricity, hence this slowed down the anchor rotation.
3303


4.3 Effects of Anchor Roughness and Shear Strength
In order to investigate the effects of anchor roughness and soil shear strength, numerical analyses were
conducted with four eccentricity ratios e/B = 0.4, 0.5, 0.625 and 1 combined with fully smooth and rough soil-
anchor interfaces. The soil was modelled as a uniform soil with undrained shear strength s
u
= 8.4kPa, 20kPa
and100kPa with soil bulk unit weight
s
= 17 kN/m
3
and steel anchor unit weight
a
= 77 kN/m
3
.
Figure 7 shows the effect of the soil-anchor interface with s
u
= 20 kPa. It can be seen that the effect of anchor
roughness is minimal with large eccentricity of e/B 0.625. However, with lower eccentricity of e/B < 0.625, the
rough anchor rotates more slowly than a smooth anchor. Since a field anchor is normally designed with e/B > 0.5,
thus the roughness effect can be ignored.

0
10
20
30
40
50
60
70
80
90
100
0 0.2 0.4 0.6 0.8 1 1.2
z
e
/B
O
r
i
t
a
t
i
o
n

o
f

p
l
a
t
e

a
n
c
h
o
r

)
e/B=0.4 Rough
e/B=0.4 Smooth
e/B=0.5 Rough
e/B=0.5 Smooth
e/B=0.625 Rough
e/B=0.625 Smooth
e/B=1.5 Rough
e/B=1.5 Smooth

Figure 7 Anchor roughness effect
0
10
20
30
40
50
60
70
80
90
100
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
z
e
/B
O
r
i
e
n
t
a
t
i
o
n

o
f

p
l
a
t
e

a
n
c
h
o
r

)
Rough
Smooth
Rough
Smooth
Rough
Smooth
s
u
=100kPa
s
u
=100kPa
s
u
=20kPa
s
u
=20kPa
s
u
=8.4kPa
s
u
=8.4kPa

Figure 8 Soil strength effect (e/B = 0.625)


Figure 8 shows the effect of soil shear strength on an anchor with e/B = 0.625. The loss of anchor embedment
during keying increases with increasing soil shear strength. The anchor roughness effect is minimal for medium
and soft soils. The loss of anchor embedment ratio, z
e
/B is 0.45 for s
u
= 8.4 kPa, 0.55 for s
u
= 20 kPa and 0.6
(smooth)/0.65 (rough) for s
u
= 100 kPa. A linear interpolation for other soil strength (s
u
) should be reasonable due
to the small range in z
e
/B for a large range of s
u
.
4.4 Anchor Thickness and Weight Effect
The loss of embedment during anchor keying is examined numerically by considering different plate thickness
ratios (t/B = 0.1, 0.067, 0.05), in conjunction with various loading eccentricity ratios (e/B = 0.3, 0.4, 0.5, 0.75, 1,
1.5). The soil was modelled as a uniform soil with an undrained shear strength s
u
= 20 kPa, soil bulk unit weight
s
= 17 kN/m
3
and steel anchor unit weight
a
= 77 kN/m
3
.
The numerical results of losses in embedment during anchor keying are summarised in Fig. 9. The anchor
thickness has no effect for anchor with t/B 0.067. However, a thicker anchor loses less embedment during
anchor keying for anchor with t/B > 0.067. This is due to the anchor weight can resist initial vertical force (F
v
= 0 in
Eq. 2), hence the moment induced by the applied force (Eq. 3) drives anchor rotation first. This effect can also be
seen in Fig. 10. It is clear that when anchor weight is increased by increasing its relative unit weight (
a
), the loss
in anchor embedment during keying is reduced, i.e. z
e
/B reduces from ~0.8 to ~0.4 when
a
increases from 0 to
70 kN/m
3
.

0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
e/B

z
e
/
B
t/B=0.05
t/B=0.067
t/B=0.1

Figure 9 Anchor thickness effect
0
10
20
30
40
50
60
70
80
90
100
0 0.2 0.4 0.6 0.8 1
z
e
/B
O
r
i
e
n
t
a
t
i
o
n

o
f

p
l
a
t
e

a
n
c
h
o
r

)
'
s
= 0 kN/m
3

a
= 0 kN/m
3

a
'=0 kN/m
3
'
s
=17 kN/m
3

a
=77 kN/m
3

a
'=60 kN/m
3
'
s
= 7 kN/m
3

a
=77 kN/m
3

a
'=70 kN/m
3

Figure 10 Relative anchor unit weight effect


3304


Figure 9 also shows that when anchor padeye eccentricity ratios e/B > 0.5, the eccentricity ratio has minimal
effect on the loss of anchor embedment (z
e
/B), which ranges between 0.2 and 0.4. However, when e/B < 0.5,
the loss of anchor embedment increases dramatically with decreasing eccentricity ratio. Thus, the anchor padeye
eccentricity ratio e/B > 0.5 is recommended in practice.
To include all the factors affecting anchor keying, the FE results of the loss in anchor embedment during keying
are summarised in Fig. 11. Both anchor padeye eccentricity (e) and anchor thickness (t) are normalised by
anchor breath (B). The submerged anchor unit weight (
a
) is normalised by anchor thickness and soil undrained
shear strength. The upper bound above the fitted line provides a more conservative estimation for design, which
can be expressed as for vertical pullout anchors ( = 90
o
):
1 . 0
'
2 . 0
3 . 0

u
a
e
s
t
B
t
B
e
B
z

(5)

0.00
0.50
1.00
1.50
2.00
2.50
3.00
3.50
4.00
0 0.5 1 1.5 2
(e/B)(t/B)
0.2
(t
a
'/s
u
)
0.1

z
e
/
B
Numerical (t/B=0.05)
Numerical (t/B=0.067)
Numerical (t/B=0.1)
Numerical (t/B=0.05)
Numerical (t/B=0.067)
Numerical (t/B=0.1)
Numerical (t/B=0.05)
Numerical (t/B=0.067)
s
u
=100kPa
s
u
=100kPa
s
u
=100kPa
s
u
=20kPa
s
u
=20kPa
s
u
=20kPa
s
u
=8.4kPa
s
u
=8.4kPa
Fitted curve
Upper bound
Practical curve

Figure 11 Anchor keying design curves ( = 90o)
0
10
20
30
40
50
60
70
80
90
100
0 0.1 0.2 0.3 0.4 0.5 0.6
z
e
/B
O
r
i
e
n
t
a
t
i
o
n

o
f

p
l
a
t
e

a
n
c
h
o
r

)
= 60
= 90
= 45
= 30

Figure 12 Anchor pullout angle effect

Since the loss in anchor embedment during keying is under-estimated by 12% in FE analysis due to the
absence of the chain tightening phase, a revised formula including this factor is proposed as follows for design
purposes:
1 . 0
'
2 . 0
34 . 0

u
a
e
s
t
B
t
B
e
B
z

(6)

4.5 Effect of Pullout Angle
Figure 12 plots the pullout angle () effect on anchor keying. The pullout angle is in the range = 30 - 90. The
anchor submerged unit weight was set as
a
= 60 kN/m
3
(equivalent to steel anchor in kaolin clay) and the anchor
eccentricity ratio as e/B = 0.625. The anchor shank effect was ignored in these analyses to reduce computation
complexity and computation time. Thus, the results of losses in anchor embedment are expected to be higher
than those with the anchor shank considered (Fig. 4a). It is apparent that the loss in anchor embedment (z
e
/B)
during keying decreases with decreasing pullout angle (), since less rotation is needed for an anchor with lower
pullout angle. It is also observed that the loss in anchor embedment decreases linearly with decreasing pullout
angle for in the range of 30 - 90. The gradient (k

) is found to be 0.005 in the formula below:



C k
B
z
e
+ =

(7)
where is pullout angle in degree, C

is a constant, which can be calculated once the loss in anchor embedment


for = 90
o
is known either from physical test or from Eq. 6.
5 Conclusion
The keying behaviour of vertically installed plate anchor has been investigated in this paper. Large deformation
finite element (FE) analyses were conducted and validated against centrifuge test data. In FE analysis, adaptive
RITSS (remeshing and interpolation technique with small strain) approach was used to simulate continuous
rotation of the plate anchor during keying. Various factors that could affect anchor keying behaviours were
3305


investigated.
When FE results were compared with existing centrifuge test data, it is found that the absence of a chain
tightening phase in FE analysis contributes to the 12% less loss of embedment during anchor keying. The anchor
capacity in a centrifuge test is 77% of that from FE analysis due to the incomplete anchor rotation and remoulded
soil in the centrifuge test.
The non-homogeneity of a soil profile shows minimal effect on the anchor keying process due to deep
embedment after anchor installation. Anchor loading eccentricity ratio (e/B) needs to be higher than 0.5 to
minimise the loss in anchor embedment during keying. When anchor thickness or anchor submerged unit weight
is increased, the loss in anchor embedment is reduced, since the initial weight can resist anchor initial vertical
movement and drives anchor rotation first.
When FE results considering all factors are summarised, a unique design formula for loss in anchor embedment
under vertical pullout (Eq. 6) is proposed using the normalised anchor thickness, anchor loading eccentricity and
submerged anchor unit weight. This design formula is also adjusted for the 12% underestimation of loss in anchor
embedment in FE analysis for design purposes.
When anchor pullout angle reduces, the loss in anchor embedment in keying is reduced as well. The reduction of
loss in anchor embedment is found to be linearly decreasing with the pullout angle with a constant gradient k

=
0.005 for kaolin clay. Once the loss in anchor embedment for a vertically pullout anchor is known, the loss in
anchor embedment with any pullout angles can be estimated using Eq. 7.
The large range of loss in anchor embedment during keying in the previous reports is believed to be due to the
difficult quantification of the anchor chain tightening effect. The lack of detailed description of the influence factors
in some previous reports makes the comparison impossible. More research is needed to study the factors not
included here.
6 Acknowledgements
The research presented here is supported by Australian Research Council through the ARC discovery grant
scheme (DP0344019). This support is gratefully acknowledged.
7 References
Carter, J. P., and Balaam, N. (1990). "AFENA user's manual." Geotechnical Research Centre, The University of Sydney.
Dove, P., Treu, H., and Wilde, B. "Suction embedded plate anchor (SEPLA): a new anchoring solution for ultra-deepwater mooring." Proc. Deep
Offshore Tech. Conf., New Orlean, USA.
Finnie, I. M. S., and Randolph, M. F. "Punch-through and liquefaction induced failure of shallow foundations on calcareous sediments." Proc. Int.
Conf. on Behaviour of Offshore Structures, Boston, USA., 217-230.
Gaudin, C., O'Loughlin, C. D., Randolph, M. F., and Lowmass, A. C. (2006). "Influence of the installation process on the performance of suction
embedded plate anchors." Gotechnique, 56(6), 381-391.
Hu, Y., and Randolph, M. F. (1998a). "H-adaptive FE analysis of elasto-plastic non-homogeneous soil with large deformation." Computers and
Geotechnics, 23(1), 61-83.
Hu, Y., and Randolph, M. F. (1998b). "A practical numerical approach for large deformation problems of soil." International Journal for Numerical
and Analytical Methods in Geomechanics, 22(5), 327-350.
Neubecker, S. R., and Randolph, M. F. (1995). "Profile and frictional capacity of embedded anchor chains." Journal of Geotechnical Engineering,
121(11), 797-803.
O'Loughlin, C. D., Lowmass, A., Gaudin, C., and Randolph, M. F. "Physical modelling to assess keying characteristics of plate anchors."
International Conference on Physical Modelling in Geotechnics 2006, Hong Kong.
Rocker, K. (1985). Handbook for Marine Geotechnical Engineering, US Naval Civil Engineering Laboratory, Port Hueneme, California, USA.
Song, Z., Hu, Y., Wang, D., and O'Loughlin, C. D. 2006"Pullout Capacity and Rotational Behaviour of Square Anchors in Kaolin Clay and
Transparent Soil." International Conference on Physical Modelling in Geotechnics, Hong Kong, P.R. China, 4-6 August, 1325-1331.
Stewart, D. P., and Randolph, M. F. (1994). "T-bar penetration testing in soft clay." J Geotech. Engr. Div., ASCE, 120(12), 2230-2235.
White, D. J., Randolph, M. F., and Thompson, B. (2005). "An image based deformation measurement system for the geotechnical centrifuge."
Int. J. Physical Modelling in Geotechnics, 5(3), 1-12.
Wilde, B., Treu, H., and Fulton, T. "Field testing of suction embedded plate anchors." 11th (2001) International Offshore and Polar Engineering
Conference, Jun 17-22 2001, Stavanger, 544-551.


3306

You might also like