Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

China Ocean Eng., Vol. 30, No. 6, pp.

942 – 953
© 2016 Chinese Ocean Engineering Society and Springer-Verlag Berlin Heidelberg
DOI 10.1007/s13344-016-0061-3, ISSN 0890-5487

Study on Interaction Between Soil and Anchor Chain with Finite


Element Method*

LI Sa (李 飒)a, 1, XU Bao-zhao (徐保照)b,


WU Yun-zhou (吴蕴洲)c and LI Zhong-gang (李忠刚)d
a
The Civil Engineering Department, Tianjin University, Tianjin 300072, China
b
Dongjiang Construction and Development Company of Tianjin Port, Tianjin 300456, China
c
Investigation and Design Institute of Water Resources and Hydropower Liaoning Province,
Shenyang 110006, China
d
China Shipbuilding Industry Corporation, Beijing 100097, China

(Received 17 September 2014; received revised form 3 July 2015; accepted 18 September 2015)

ABSTRACT
With the development of offshore engineering, deeply embedded anchors are needed to be penetrated to appreciable
depth and attached at the pad-eye. The interaction between anchor chain and soil is a very complex process and has not
been thoroughly understood yet. In this paper, the finite element method (FEM) was used to study the interaction of
soil-chain system. Results of the analysis show that when the attachment point is at a shallow depth, the load-development
characteristics of the chain from FEM are in good agreement with that from the model tests and theoretical analysis. But
with the depth increment, the results are different obviously in different methods. This phenomenon is resulted from a
variety of reasons, and the plastic zone around the chain was studied to try finding the mechanism behind it. It could be seen
that the plastic zone extended in different modes at different depths of attachment points. The interaction between the soil
and anchor chain makes the load acting on the anchor decrease, but the soil disturbed surrounding the chain increases the
anchor failure possibility. When the anchor bearing capacity is evaluated, these two factors should be considered properly
at the same time.

Key words: anchor chain; clay; interaction; finite element; the load-development characteristics

1. Introduction

Deeply embedded anchors, such as pile anchors, suction anchors and vertically loaded anchors, are
penetrated to appreciable depth and attached at the pad-eye. It should be noted that the position of the
pad-eye below the mud line introduces a pull-out force on the anchor that has to be resisted by the soil
resistance on the anchor. A good design of the anchor requires an accurate prediction of the loading from
the mooring line, including any soil-chain interaction. At present, the interaction between the mooring
lines and the seabed (soil) at the anchor location is often overlooked but can have an impact on the
mooring performance of a floating system (Wales et al., 2011).
The interaction between anchor chain and soil is a very complex process and has not been

* This study was financially supported by the State Key Program of National Natural Science of China (Grant No. 51239008).
1 Corresponding author. E-mail: lisa@tju.edu.cn
LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953 943

thoroughly understood yet. The method provided by Degenkamp and Dutta (1989) is considered as the
basic theory (Miedema et al., 2007). Degenkamp and Dutta (1989) verified the chain tension at the pad-
eye by a model test and recommended a method to calculate the chain tension at the anchor point and
the chain configuration inside the soil. They pointed out that the entire embedded chain configuration
can be assumed to be a summation of the discrete chain elements, as shown in Fig. 1.
In Fig. 1, f is the soil frictional resistance per unit chain length; q is the soil normal/bearing
resistance per unit chain length; w is the effective chain weight in soil per unit chain length; T1 is the
chain tension at the top of the chain element; T2 is the chain tension at the bottom of chain element; s
is the chain-element length; 1 is the chain angle at the top of the chain element; and 2 is the chain
angle at the bottom of chain element. For each element, the value of T2 and 2 can be determined with
the known values of T1 , 1 , w , f and q . To determine f and q , the tangential movement has
been assumed to cause an uncoupled sliding resistance, independent of the normal soil resistances.

Fig. 1. Free-body diagram of the chain element inside soil.

Using this assumption, the frictional resistance F is written as:


F  ( Et d ) f . (1)
And the normal soil resistance Q is written as:
Q  ( En d ) q , (2)
where, Et d is the effective chain width in sliding; En d is the effective chain width in bearing; En and
Et are multipliers to give the effective widths in the normal and tangential directions, respectively. They
are determined based on the type of chain. The chains generally used in offshore mooring applications
are of the stud-link type, with a link length 6 times the nominal chain diameter (6d chain), where d is
the nominal chain diameter. The parameters En and Et can be determined by calculating the effective
areas in bearing and shearing per pair of links, and dividing these areas by the effective length of the pair.
In fact, Et d and En d are the effective bearing area of frictional resistance and normal soil resistance,
and the effective areas of chain in sliding and bearing are modeled the same as the width of an equivalent
944 LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953

strip footing,
The value of q is evaluated using the formula given by Skempton (1951) for the ultimate net soil
resistance of a strip footing in clay:
q  N c Su , (3)
where S u is the undrained shear strength of the clay, and N c is the capacity bearing factor for clay.
For the estimation of bearing capacity factor ( N c ), theoretical solutions are available in the literature.
The most popular of them is the Prandtl (1920) solution in which the analysis of the bearing capacity of
the strip footing on the surface (Z= 0, Z is the embedded depth of foundation) showed the value of the
capacity bearing factor:
N c = 2+= 5.14.
In the analysis, the failure mechanism assumes that the footing pushes in front of itself a wedge of
clay, which, in turn, pushes the adjacent material sideways and upwards.
Different researchers used the different assumptions to determine N c , and the value of N c
changed with the assumptions. How to decide the value of N c is a problem that many researchers have
been interested in until now.
Degenkamp and Dutta (1989) used the formula given by Skempton (1951),
 h 
N c  5.14  1  0.2  (with the maximum N c = 7.6),
 Bb 

where, Bb (= En d ) is the width of the footing, and h is the depth of footing.


It can be seen that although the values of various design parameters such as the effective chain
width in sliding ( Et d ), the effective chain width in bearing ( En d ), and the capacity bearing factor in
clay ( N c ) have been suggested, little experimental proof has confirmed their validity (Miedema et al.,
2007).
The value of f is,
f   Su , (4)
where  is a reduction factor, and   1 for soft clay.
Neubecker and Randolph (1995a) also used these formulations to establish the relationship of the
force between anchor chain and soil, but they derived closed-form expressions for both the load
development and chain profile. These expressions simplify the procedure for estimating the load and
inclination of an embedded chain at some connection points in the soil. Neubecker and O’Neill (2004)
presented normalized charts that allowed for a quick and simple evaluation of chain slippage versus
load, as a function of chain size and soil strength profile. Wang L. Z. et al. (2010) considered the three-
dimensional characteristics of soil and chain to present a method to predict the behavior of an
embedded chain.
All work mentioned above is based on Degenkamp and Dutta’s method, and try to estimate the
profile or frictional capacity of embedded anchor chains. During this procedure, it can be seen that the
LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953 945

accuracy of F and Q are governed by the parameters Et , En , and N c respectively, which are
difficult to be determined accurately.
The interaction between anchor chain and soil is so complex that the finite element method
(FEM) is a good choice to deal with it. Currently, the FEM method has been used widely to analyze the
deep penetration anchor behavior. FEM has many advantages, including spatially varying soil
properties, advanced nonlinear and anisotropic constitutive models, complex geometries, to name a
few. The major aspects that are typically used in the design of foundations for offshore structures are:
(1) bearing capacity and sliding resistance; (2) installation aspects; (3) foundation stiffness; (4)
consolidation and settlements; (5) soil reactions against the structure; and (6) soil-structure interaction
(SSI) (Andresen et al., 2010). Although many researchers have used FEM to do many analyses about
offshore foundation (Yu et al., 2009; O’Neill et al., 2003; Cao and Alhayari, 2011; Wang D. et al.,
2010; Dickin and Laman, 2007; Silva-González et al., 2013), most of them focused on the interaction
between anchor and soil. There are few studies about the chain–soil interaction with FEM.
This paper utilizes a 3D large deformation FEM to study the behavior of embedded chain in soil
in 3D space. The embedded depth of attachment point effecting on the load-development characteristics
of the chain, named   T0 / Ta (see Fig. 2), is studied, and the plastic zone surrounding the chain in the
soil is investigated.

Fig. 2. Profile of the mooring line.

2. Numerical Model and 3D Large Deformation FE Method

Fig. 2 is a schematic diagram of a penetration anchor. It can be seen that there are three parts in this
system, anchor, soil, and anchor chain. The soil discussed in this paper is clay. It is simulated using an
elasto-plastic model with a DP yield criterion. DP material model describes materials with
pressure-dependent inelastic behavior. It contains a dependence on hydrostatic stress. To describe the
behavior of soils, it is necessary to define the plasticity limit in which the multidimensional state of stress
is generalized by the yield surface that defines the area separating elastic and plastic zones in the material
body. The necessary input parameters of ANSYS software for this model include modulus of elasticity
E , Poisson coefficient  , material density  , angle of internal friction  , cohesion value c , and
946 LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953

dilatancy angle  . The modulus of elasticity E is also defined as the initial tangential modulus.
Twenty-node hexahedral elements are selected to simulate the soil.
A 3D spar element is selected to simulate the anchor chain. It has the unique feature of a bilinear
stiffness matrix resulting in a uniaxial tension-only (or compression-only) element. With the
tension-only option, the stiffness is removed if the element goes into compression (simulating a slack
cable or slack chain condition). It has three degrees of freedom at each node: translations in the nodal x,
y, and z directions. No bending stiffness is included in the tension-only (cable) option. Stress stiffening
and large deflection capabilities are available. The element is defined by two nodes, the cross-sectional
area, an initial strain or gap, and the isotropic material properties.
To establish the anchor chain model, the first thing is to decide the form of the chain in the soil.
Because the chain is buried in the soil, it is difficult to determine the initial form of the chain under the
effect of gravity with ANSYS automatically. In this paper, the initial form of the anchor chain is
assumed as an inverse catenary as others did (Neubecker and Randolph, 1995a, 1995b). It can be
determined by the reverse catenary equation of the embedded line in the uniform soil, and expressed as
follows:
  z  z  sin  0 z 
x  xa  B   cos2  0    sin 2  0     arcsin  sin 2  0   0  ; (5)
  B  B 2 B 
za . (6)
B
sin 2  a  sin 2  0
The meanings of the symbols in the equations are shown in Fig. 2. If the embedded line is tangential
to the theoretical seafloor, i.e., at the embedment point, θ0 = 0 in the uniform soil, it can be expressed as
follows (Liu et al., 2013; Li, 2010):
 z z z
x  xa  B   1    arcsin ; (7)
 B  B  B 
z
B  2a . (8)
sin  a
Based on Eqs. (7) and (8), the initial form of the anchor chain is obtained. The length of the anchor
chain can be calculated by
2 za . (9)
L
sin  a
The anchor chain is simulated by 3D spar elements with the length of 1.0 m to define a continuous
line. In order to use line-to-surface contact element, the nodes must be entered in the right sequence. The
element should not be used for static convergence applications where the final solution is known to be a
taut structure but a slack condition is possible while iterating to a final converged. Therefore, a very
small initial strain of 0.001 is applied before loading. The material property parameter needed to be input
in this analysis is the Young’s modulus of the chain and it can be determined based on the material of
the chain.
The contact between anchor and soil is the key step in this simulation. The geometric contact should
be established at first. The soil element should be established firstly and divided into different parts as
LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953 947

shown in Fig. 3a. The size of soil should be large enough to eliminate the boundary effect on the anchor
chain’s behavior. The shadow area is the place where the anchor chain is located. In this area, the anchor
chain elements and the soil elements have the common points (Fig. 3b). The anchorchain interaction is
simulated by using line-to-surface contact element. It is located on the surface of 3D spar elements. The
isotropic Coulomb friction is used to describe the interaction between anchor chain and soil. The
coefficient of friction is 0.5 in clay. To simplify the problem and reduce computational time, the anchor
is simulated as a fixed point, which means that the end of anchor chain in the soil does not move during
the loading applied.
The present numerical analyses are conducted with the ANSYS. The incremental full Newton
Raphson iterative solution procedure is used in order to account for both large deformation effects and
material plasticity.

Fig. 3. Anchor chain element in the soil.

3. Verification of the Model

Verification procedure is the primary means of assessing accuracy in computational simulations.


The FE model should be verified by comparing it with hand calculation, test data or other known sources
before it is used to do analysis. The results of a series of pull-through tests of chain in clay performed by
Degenkamp and Dutta (1989) are used to do this work. The primary aim of these tests was to use
measured values of chain tension at the attachment point ( Ta ) and chain inclination at the attachment
point (  a ) to determine the effective-width parameters En and Et of the chain by comparing the results
with the numerical solution. Therefore, during the experiment, at regular intervals, the horizontal pulling
force (PULLF, T0 ), and the horizontal and vertical lug forces (FHM and FVM, Ta  FVM 2  FHM 2 )
were measured. All these data are used to check the FE model in this analysis.
The saturated clay used for the tests had an undrained shear strength of 4.52 kPa, and the wet unit
weight of the clay was 18.8 kN/m3. During the test, the chain was pulled at a speed of approximately
0.002 m/s. This pulling speed was believed to be adequate for the generation of undrained loading
conditions in clay-type soil. The parameters used in FE were based on the model tests. Poisson’s ratio
  0.49 and the friction and dilation angles     0 were used to simulate undrained conditions.
948 LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953

The ratio of Young’s modulus to soil shear strength was taken as E / S u  500 .
Three chains, having nominal diameters (d) of 6.4 mm, 9.5 mm, and 16 mm, were used for the tests,
and the chain with d =9.5 mm was selected to do analysis in the FE model. The weight of the chains was
20.0 N/m.
In the FE model, the PULLF measured in the model tests is applied at the mud line on the anchor
chain. The loading at the fixed point regarded as the force at the pile lug can be obtained by FEM, and it
is compared with the measured data (resultant force at lug) in the tests. At the same time, the
load-development characteristics of the chain, or   T0 / Ta obtained from the FE analysis is also
compared with that from tests. The results are shown in Tables 1 and 2.

Table 1 Comparison of the force at the lug in the model test and FE analysis
Depth of lug Horizontal force at the top of
Case No. Resultant force at lug in Resultant force at lug in
from soil embedded chain in the tests and
the test (N) FE analysis (N)
surface (m) FE analysis (N)
Case 6.1 0.4065 708 404 414
Case 7.1 0.2040 672 364 361
Case 8.1 0.6120 1268 658 671
Case 6.4 0.4065 6148 5447 5346
Case 7.4 0.2040 7799 7036 6723
Case 8.4 0.6120 5831 4875 5070

Table 2 Comparison of   T0 / Ta in the model test, calculated value and FE analysis


Calculated
Case No. In the tests FE analysis
value
Case 7.1 1.84 1.83 1.86
Case 6.1 1.75 1.69 1.71
Case 8.1 1.93 1.92 1.89
Case 8.4 1.20 1.20 1.15
Case 6.4 1.13 1.12 1.15
Case 7.4 1.11 1.10 1.16

Neubecker and Randolph (1995a, 1995b) did the similar work. First, they used closed-form
expressions to calculate  as follows:

  T0 / Ta  e 2/ T *
, (10)
where T0 and Ta are shown in Fig. 2.   F / Q , and T is a normalized tension given by
*

Ta
T*  , (11)
DQ
where D is subsoil depth of attachment, and Q is the average bearing resistance per unit length of
chain. The derivation of formula mentioned above can be seen in Neubecker and O’Neill (2004). Then,
they compared the load-development characteristics of the chain, or  with the experimental values.
The results are also shown in Table 2.
Table 2 shows the comparison of the results. It can be seen that the results from FE analysis are in
good agreement with those from the model tests and from Neubecker and Randolph (1995a, 1995b). The
LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953 949

model can be used to do further analysis.

4. Results and Discussion

4.1 Effects of the Depth of the Attachment Point


Analysis of the performance of an embedded anchor chain is important because many studies have
shown that the pulling force applied at the mud line can be much larger than that at the attachment point.
It is said that  is the key parameter that many researchers concerned. The change of  with the
depth of the attachment point was studied.
Two methods are used to do the analysis. One is the FEM, and the other is the Neubecker and
Randolph’s method. Select Case 6.4 (see Table 1) in Degenkamp and Dutta’s model tests as the basic
calculation condition. It means that all the parameters used in the FE analysis and Neubecker and
Randolph’s method are the same as those in the model test except the depth of the attachment point in the
soil. Fig. 4 shows the change of  with the increment of the depth of the attachment point in the soil. It
can be seen that  from the FE is smaller about 40% than the calculated value when the attachment
point is located at 14.3 m.

Fig. 4. Change of  with depth.

In order to explain the difference of  in two methods, the bearing capacity factor N c is back
analyzed from the FE and calculated from the Neubecker and Randolph’s method, respectively.
From Eqs. (1), (2), (3), (4), (10) and (11), Eq. (12) is shown as:
Ta En N
2
ln 2   c2 D . (12)
2 Su dEt Nc
It can be seen that if  , Ta , S u , En , Et and d are all known, the relationship of the depth of the
attachment point (D) and N c N c2 can be obtained. From FE analysis, the values of  and the chain
tension at attachment point ( Ta ) can be obtained, and d , S u , En , and Et are used the same value as that
in the Neubecker and Randolph’s method, and then we can obtain the relationship between D and
N c N c2 .
Moreover, the relationship between D and N c N c2 can also be calculated from the assumption of
950 LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953

Neubecker and Randolph (1995a, 1995b). It is said that the bearing capacity factor N c is assumed to
rise from 5.14 at the seabed to 7.6 at the depth of 2.4En d , so the average bearing capacity factor over
the depth (for D  2.4 En d ) is ,
N c  7.6  2.95( En d / D) . (13)
2
Fig. 5 shows this relationship based on different methods. It can be seen that N c N c
back
analized from the FE analysis is smaller than that from Neubecker and Randolph (1995a, 1995b). It
means that the maximum value of N c =7.6 can be underestimated by Neubecker and Randolph (1995a,
1995b) in the deeper soil.

Fig. 5. Change of N c / N c2 with depth.

4.2 Disturbed Area Surrounding the Anchor Chain


The soil behavior around the anchor is also studied in order to explain the difference of  in
different methods. As we know, the main effect of soil-anchor chain interaction behavior can be
expressed in terms of a redistribution of stresses in subsoil and forces in the anchor chain. Differential
stress field can change in the soil, compared with calculations with no account of the interaction of the
soil-chain, and anchor chain elements can acquire new forces due to the differential stress field. Fig. 6a
shows the plastic zone on the soil surface of the model which is established based on Case 6.4. The depth
of the attachment point is 0.4065 m, see Table 1. Here the plastic zone is expressed by the stress state
ratio. If the stress state ratio is larger than 1, the equivalent stress is larger than the yielding stress. It can
be seen that because the depth of the attachment point in the soil is very shallow, the plastic zone almost
appears in the total area where the anchor chain passes through. Fig. 6b is the plot of the plastic zone on
the soil surface of the model with the attachment point located at 14.3 m under the soil. It can be seen that
the plastic zone is just around the point where the anchor chain comes out of the soil.
Fig. 7 gives the plastic zone at different sections with different distances away from the attachment
point, from which it can be seen clearly that the plastic zone in the soil is like a “dumbbell”. Near the
attachment point, the plastic zone is larger, and with the increment of distance away from the attached
point, the plastic zone becomes smaller, and the largest area appears at the point where the anchor chain
comes up out of the ground. The shape of the plastic zone in the soil is shown in Fig. 8, in which
LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953 951

d1>d3>d2.

Fig. 6. Plastic zone at the soil surface where the attachment point at 0.4 m and 14.3 m deep in the soil.

Fig. 7. Development of the plastic zone along the anchor chain.

Fig. 8. Profile of the plastic zone around the anchor chain.

The distribution of the plastic zone may imply different soil failures at different depths of the
attachment point. When the depth of the attachment point is deep, the force is transferred by the friction
between the anchor chain and soil mainly. But when the depth of the attachment point is shallow, in
addition to the friction between the anchor chain and soil, more soil around the anchor chain is
motivated. It means that the range of interaction affected is large. Neubecker and Randolph’s method is
952 LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953

well agreement with this mechanism. But with the increment of the depth of the attachment point, the
effective overburden pressure increases, soil around the chain is difficult to be motivated, and more
loading transfers to the anchor. That is why  from the Neubecker and Randolph’s method is larger
than that from FE analysis. The mechanism of the soil failure around the anchor chain is different at
different depths of the attachment point.

5. Conclusion

The interaction between soil and anchor chain is studied by FEM in this paper.  obtained from
the FEM is consistent with that from the model tests and theoretical analysis when the attachment point
is penetrated at a shallow depth. But with the attachment point depth increasing, the difference appears.
In this paper, when the depth of the attachment point is 14.3 m in the soil,  from FEM is about 40%
smaller than that from Neubecker and Randolph’s method which means that N c could be underestimated.
The analysis shows that the shape of the plastic zone around the chain changes with the location of
the attachment point in the soil. It could imply that the soil failure is in different modes around the chain.
It also shows that there is an area with little strength surrounding the chain. Although the interaction
between the soil and anchor chain makes the load acting on the anchor decrease, the soil disturbed
surrounding the chain increases the failure possibility of the anchor. When the anchor bearing capacity is
evaluated, these two factors should be considered properly at the same time.

References
Andresen, L., Petter Jostad, H. and Andersen, K. H., 2010. Finite element analyses applied in design of foundations
and anchors for offshore structures, International Journal of Geomechanics, 11(6): 417430.
Cao, J. C. and Alhayari, S., 2011. Study on behavior of VErtically Loaded Plate Anchors (VELPA) in soft clay,
Proceedings of the 21st International Offshore and Polar Engineering Conference, Maui, Hawaii, USA.
Degenkamp, G. and Dutta, A., 1989. Soil resistances to embedded anchor chain in soft clay, J. Geotech. Eng.-ASCE,
115(10): 14201438.
Dickin, E. A. and Laman, M., 2007. Uplift response of strip anchors in cohesionless soil, Adv. Eng. Softw., 38(8-9):
618625.
Li, Y., 2010. Reverse Catenary Equation of the Drag Line and Its Application to the Kinematic Model of Drag
Anchors, MSc. Thesis, Tianjin University, China. (in Chinese)
Liu, H., Liu, C., Zhao, Y. and Wang, C., 2013. Reverse catenary equation of the embedded installation line and
application to the kinematic model for drag anchors, Appl. Ocean Res., 43, 8087.
Miedema, S. A., Lagers, G. H. G. and Kerkvliet, J., 2007. An overview of drag embedded anchor holding capacity
for dredging and offshore applications, Proceedings of World Dredging Conference XVIII, WODA, Orlando,
USA.
Neubecker, S. R. and O’Neill, M. P., 2004. Study of chain slippage for embedded anchors, Proceedings of Offshore
Technology Conference, Houston, Texas, OTC-16445-MS.
Neubecker, S. R. and Randolph, M. F., 1995a. Performance of embedded anchor chains and consequences for
anchor design, Proceedings of Offshore Technology Conference, Houston, Texas, OTC-7712-MS.
Neubecker, S. R. and Randolph, M. F., 1995b. Profile and frictional capacity of embedded anchor chains, J.
LI Sa et al. / China Ocean Eng., 30(6), 2016, 942  953 953

Geotech. Eng.-ASCE, 121(11): 797804.


O’Neill, M. P., Bransby, M. F. and Randolph, M. F., 2003. Drag anchor fluke soil interaction in clays, Can. Geotech.
J., 40(1): 7894.
Skempton, A. W., 1951. The bearing capacity of clays, Proceedings of Building Research Congress, England.
Silva-González, F., Heredia-Zavoni, E., Valle-Molina, C., Sánchez-Moreno, J. and Gilbert, R. B., 2013. Reliability
study of suction caissons for catenary and taut-leg mooring systems, Struct. Saf., 45, 5970.
Wang, L. Z., Guo, Z. and Yuan, F., 2010. Quasi-static three-dimension analysis of suction anchor mooring system,
Ocean Eng., 37(13): 1127–1138.
Wales, S., Sincock, P. L. and Santosa, M., 2011. Impact of catenary embedment on the mooring performance of a
deep water floating production unit, Proceedings of the 21st International Offshore and Polar Engineering
Conference, Maui, Hawaii, USA.
Wang, D, Hu, Y. and Randolph, M. F., 2010. Three-dimensional large deformation finite-element analysis of plate
anchors in uniform clay, J. Geotech. Geoenviron., 136(2): 355365.
Yu, L., Liu, J., Kong, X. and Hu, Y., 2009. Three-dimensional numerical analysis of the keying of vertically installed
plate anchors in clay, Comput. Geotech., 36(4): 558567.

You might also like