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

NUMERICAL ANALYSIS OF TORPEDO ANCHORS

H. Sabetamal, M. Nazem, and J.P. Carter

ARC Centre of Excellence for Geotechnical Science and Engineering, The University of Newcastle,
Newcastle, NSW, Australia

ABSTRACT: This paper presents the development of a numerical framework based on the
finite element method and its application in the analysis of torpedo anchors. The procedure is
based on a mixture theory for the dynamic behaviour of saturated porous media. The
nonlinear behaviour of the solid phase of soil is represented by the Modified Cam Clay
material model and the interface between the soil and the structure is modelled by a mortar
segment-to-segment frictional contact method. An Arbitrary Lagrangian-Eulerian (ALE)
method is adopted to avoid mesh distortion throughout the numerical simulation. The
generalised- method is utilised to integrate the governing equations of motion in the time
domain. Results obtained from the installation phase of a torpedo anchor reveal that the
anchor decelerates at a constant rate during most of its penetration. Analysis results show a
typical distribution of excess pore-water pressure during free falling installation, having
higher magnitudes at the face and lower magnitudes along the shaft. The computational
results for the setup phase indicate that for soil elements located within a radial distance of
approximately one diameter from the centreline of the torpedo, 90% of consolidation takes
place in a few days after installation, depending on the value of soil permeability.

1 INTRODUCTION
Torpedo anchors have proven to be promising systems for anchoring taut mooring lines of
floating offshore oil and gas exploration and production units due to their relatively easy
installation process. The kinetic energy of a torpedo anchor attained by gravity throughout
free fall through the water column provides the required dynamic penetration force, making it
much more practical and cost-effective than other offshore structures such as suction piles,
driven piles, drilled and grouted piles, and drag embedment anchors. The torpedo usually
consists of a pipe pile (12 to 18m in length, 0.76 to 1.07m in diameter) filled with scrap metal
and concrete, close ended and fitted with a conical tip and sometimes including fins at the top
end which provide stability during free fall. The impact velocity of the torpedo piles reported
by Medeiros (2002) varied between 10 to 22 m/sec, for hanging heights from which free fall
commenced between 30 to 150 m as measured from the seabed, and the penetration depth
usually varied between 8 m and 22 m.
The first step in the analysis of a torpedo anchor involves simulation of the installation
phase in order to predict the penetration depth, soil resistance, and the development of excess
pore water pressure. However, in the majority of research works devoted to the analysis of
deep penetration anchors (DPA) the effect of installation on pullout or lateral capacity of the
anchor is ignored. In other words, in most analyses conducted to date deep foundation
systems are whished in place, with no effort to model the installation phase and hence a

621
perfect interface between the anchor system and the surrounding soil is assumed. The initial
stress state of the soil is normally estimated based on the submerged unit weight, the lateral
earth pressure coefficient at rest, and assumes zero excess pore-water pressure. Estimation of
penetration depth of the anchor usually relies on the theoretical framework developed by
True (1976).
Sturm and Andresen (2010) presented a finite element model for torpedo anchors using
the commercial software package Abaqus, with a user defined contact algorithm and an
Updated Lagrangian (UL) formulation. However, they simulated the installation process
quasi-statically with a constant penetration rate and neglecting any inertia effects. They used
the Tresca material model to simulate soil behaviour and evaluated the excess pore pressure
using knowledge of the mean stress distribution and shear strain in the soil. Raie (2009)
developed an alternative procedure based on Computational Fluid Dynamics (CFD) to predict
the embedment depth as well as the installation effects, including shear distributions on the
soil-anchor interface and the soil state parameters. This method is based on the principles of
fluid dynamics where stress at any point of the media is equal to the pressure at that point
independent of the direction, i.e., the vertical and horizontal stresses on soil elements are
assumed to be identical, with this being an unrealistic assumption for soil.
The second stage in the simulation of torpedo anchors is the set up analysis. With
knowledge of the effective stresses and the excess pore water pressures, the set up analysis
can be performed by reconsolidation of the soil in the vicinity of the anchor. In DPA systems
excess pore pressure is generated due to two main factors: shearing of soil during installation
and increase in total stress due to the vertical and mostly radial soil volume changes. The
excess pore water pressures result in lower frictional resistance, which leads to lower pullout
capacity of the anchor. As the soil consolidates, the pullout capacity of the anchor increases
due to dissipation of excess pore water pressures and corresponding increases with time of
the effective stresses.
The third and final step of the analysis includes estimation of the pullout capacity of the
anchor. The finite element method (FEM) and the API (American Petroleum Institute, 2002)
method are the two common techniques for estimating the holding capacity of torpedo
anchors. The API method takes advantage of the conventional theory of pile bearing capacity
based on the total stresses, and predicts the undrained holding capacity of DPA in cohesive
soils. Randolph et al. (2005) applied this method to estimate the capacity of anchors
embedded in calcareous sand. Richardson et al. (2009) employed the API method to study
the skin friction ratio and fluke effects on the holding capacity by comparing its results with
experimental results obtained by centrifuge tests. Modelling the soil as a Drucker-Prager
material and assuming the anchor to be wished in place, Sousa et al. (2010) employed the
FEM to evaluate the long term load capacity of a typical torpedo anchor subjected to vertical
and inclined loads.
A brief survey in the literature reveals that the analysis of dynamically penetrating objects
needs further research to realistically model and evaluate their behaviour. Most available
studies are based on experimental or approximate analytical solutions, and the current FEM
simulations are generally based on a displacement formulation (neglecting the pore water
pressures) and consider simplifying assumptions in the modelling. A more realistic model
must incorporate pore fluid pressure development along with deformations, velocities and
accelerations to facilitate a thorough understanding of soil response. Moreover, by providing
the initial undrained or partially drained distributions of pore pressure, subsequent dissipation
can be investigated. Such problems require a fully coupled analysis that takes into account
the interaction between soil and pore fluid by incorporating the effect of the transient flow of
the pore fluid through the inter-connected voids of the solid skeleton.
In this paper, a computational framework based on the FEM is developed and employed to
simulate torpedo anchor systems. The procedure is based on a mixture theory for the

622
dynamic behaviour of saturated porous media. The nonlinear behaviour of the solid phase of
soil is represented by the Modified Cam Clay material model and the interface between the
soil and the structure is modelled by a mortar segment-to-segment frictional contact method.
The Arbitrary Lagrangian-Eulerian (ALE) method is adopted to avoid mesh distortion
throughout the numerical simulation. The generalised- method is utilised to integrate the
governing equations of motion in the time domain. Energy absorbing boundaries are used to
model the radiation of waves towards infinity at the truncated finite element mesh
boundaries. An automatic time stepping procedure for the dynamic consolidation algorithm
is also employed to increase the efficiency of the method. In the following, we briefly
explain the governing equations, the time-integration method, energy absorbing boundaries,
modelling of contact, large deformation, and the strategy of mesh refinement. Then, a
numerical example is presented to illustrate the effectiveness and utility of the proposed
approach. Note that for brevity we only consider the analysis of the installation phase and the
subsequent consolidation phase in this paper. The analysis of pullout behavior will be the
subject of future studies.

2 FINITE ELEMENT FORMULATION


A saturated porous medium can be considered as a two phase mixture composed of a solid
constituent as well as a fluid constituent in which the two phases in the assemblage interact
with each other and affect the overall behavior of the medium. Here, a continuum approach
based on the theory of mixtures (Truesdell & Touppin, 1960) is employed to derive the
governing equations using the concept of volume fraction (Morland, 1972). The resulting
equations guarantee the fulfillment of the local balance relations, for each individual
constituent, as well as the balance relations of the entire mixture. The global finite element
equations for a two phase saturated porous media might be written in the following matrix
form

Ms 0 U Cs 0 U K ep L U F u
+ + = (1)
M LT
0 P

S P 0

H P F p
f

where U, , and P denote the vectors of displacement, velocity, acceleration and pore
water pressure, respectively, and M s , M f , C and K ep are the solid mass, fluid mass, damping,
and stiffness matrices, respectively. L, H, and S represent, respectively, the coupling matrix,
the fluid flow matrix, and the compressibility matrix, and F u and F p are the vectors of
external nodal forces. The system of second-order ordinary differential equations in (1) is
usually solved using direct time integration methods. The selected time integration scheme
should possess some form of numerical dissipation capability to attenuate the high frequency
modes. Meanwhile, it should also allow the accurate capture of the low frequency behaviour
of the system, so that it appears in the solution without attenuation. Among the various
implicit time integration schemes we have used the generalised- method, knowing that it is a
second order accurate scheme and preserves the accuracy while providing the requisite
numerical damping. In addition, it allows the analyst to control the amount of numerical
dissipation at high frequencies.
The analysis of a torpedo anchor comprises a rapid installation phase as well as a
subsequent consolidation phase in which the first phase requires relatively small time
increments (to capture the rapid dynamic response) in comparison with the second phase
which demands larger time steps. Typically, the size of the time steps in the consolidation
phase can be up to 10,000 larger that the size of the increments in the installation phase.
Therefore, an important characteristic of an efficient time stepping method is to adaptively
choose small increments in the transient phase, while increasing the size of time steps as the

623
solution becomes smoother in the second phase. To devise such an automatic time stepping
scheme, we use two different local truncation error estimators. One of the error estimators is
based on Taylors series and the other one takes advantage of a family of 4 stage-one step
algorithms of Thomas and Gladwell (1988). A number of coarse time steps are defined at the
beginning, which are then automatically sub-incremented into a number of smaller time
increments if necessary. In order to increase the time increments, the estimated error has to
be smaller than a percentage of a user defined tolerance. If the error is larger than this
quantity, but smaller than the tolerance, the time increment will be held constant to minimise
the number of rejected time sub-increments.

3 INTERFACE MODELLING
The so called node to segment (NTS) discretisation method is widely used to analyse large
sliding and large deformation problems of contact mechanics. However, it has been
highlighted that this approach cannot pass the contact patch test (Papadopoulos & Taylor,
1990), as coupling with higher order elements is not feasible without losing accuracy of
displacements and stresses in the contact area. This can be a particular disadvantage in soil
mechanics where higher order shape functions are often used to improve the accuracy and to
avoid mesh locking. Oscillation of the contact force predicted by the NTS technique due to
the non-smooth surface of the low order elements is another deficiency of this method. In
order to overcome these issues, the mortar segment to segment approach has been developed
which allows the interpolation functions of the contact elements to be of order n. (Wriggers,
2006). We use a contact algorithm based on the mortar method to model soil structure
interactions in the torpedo problem. The interactions at a frictional contact interface for a
two phase saturated porous media generally arise from the contact traction as well as the fluid
flow. Therefore, when contact occurs between two deformable bodies, constraints have to be
established to fulfill the continuity of the contact traction as well as the fluid flow across the
interface. This means that in addition to applying constraints on displacements, two other
constraints on the Darcy velocity and the pore fluid pressure must be enforced in order to
guarantee, respectively, the linear momentum balance of the fluid phase and the conservation
of mixture mass. In a frictional contact element, two conditions, stick and slip, are
distinguished on the basis of the level of interface frictional force in comparison with the
Coulombs frictional force represented by

fs
= tT t N (2)

where tT and t N are, respectively, tangential and normal effective stress components of the
total traction at the contact interface, and denotes the friction coefficient at the interface. In
order to differentiate between stick and slip cases, we use the concept of a moving cone
(Wriggers, 2006), which is a relatively efficient methodology in deriving the contact
kinematics. For the coupled consolidation formulation presented in Section 2, Darcys
velocity does not appear explicitly in the formulation, so that pore pressures at the points of
contact are used to derive the corresponding normal effective stress. Meanwhile, a constraint
is applied to fulfil the balance condition of the pore pressure at the contact interface where
necessary. Furthermore, the penalty method is used to enforce the contact constraints.

4 ABSORBING BOUNDARY
Another challenge in dynamic FE analyses is to cope with spurious wave reflection from
artificial boundaries of the problem domain. Due to the fairly high wave velocity in soils and
rocks, modelling a large portion of the domain does not seem to be an effective way to

624
proceed as the waves reflected from the boundaries will probably have enough time to return
to the area of interest. Therefore, absorbing boundary conditions should be facilitated to
absorb the outgoing waves. Generally, two types of bulk waves, including the dilatational
waves (P- waves) as well as the shear waves (S- waves), appear in a saturated porous media
(Biot, 1956). The dilatational waves can be decoupled into two waves, P1 and P2. The P1
waves propagate faster and attenuate slower compared to the P2 waves, known as Biots slow
wave. Shear waves are transmitted only in the solid constituent and are mainly governed by
its shear stiffness while the propagation of acoustic waves essentially depends on the
frequency of excitation, the hydraulic permeability and the mechanical properties of the
constituent materials (Corapcioglu &Tuncay,1996; Straughan, 2008). For the dynamic
problems concerned in this study, two types of the bulk waves, viz., P2 waves and S waves
are predominant. This is mainly due to the conditions of low-frequency excitations and the
relatively low permeabilities present in the problem. Under such circumstances, very low
relative motions between the solid matrix and the viscous pore fluid are likely and body
waves are mostly transmitted via the structure of the solid skeleton. Hence, a local
transmitting boundary such as the standard viscous boundary of Lysmer & Kuhlemeyer
(1969) can be utilised to ensure the absorption of the arriving elastic energy. However, as the
standard viscous boundary embodies no static stiffness, it is not able to model a static
problem and rigid body movement would occur for low frequencies. Alternatively, we use
the cone boundary of Kellezi (2000) which consists of both dashpots and springs. The
stiffness of the springs varies linearly along each boundary based on the radial distances of
the boundary node from the source of excitation.

5 LARGE DEFORMATION AND MESH REFINEMENT


Penetration problems involve large deformations as well as rigid body rotations, which affect
the soil stiffness and its permeability, and such special effects may not be simply disregarded
in the analysis. The theory of large deformation has widely been used in the framework of
the FEM to analyse large deformation problems of geomechanics on the basis of Lagrangian
approaches. The effects of material and geometrical nonlinearities in the theory of
consolidation were first presented by Small et al. (1976) and Carter et al. (1979),
respectively. Dynamics of porous media at finite strain has also been elaborated and applied
by others based on the theory of mixture (e.g., see Chao et al, 2004). The Updated-
Lagrangian (UL) method is a more prevalent approach in handling geometrical nonlinearities
comparing to the Total-Lagrangian (TL) method, especially where contact kinematics are
invoked and formulated, since the UL methodology uses the last equilibrium configuration
to evaluate the state variables. However, the Lagrangian methods are prone to failure and
numerical errors where the finite element mesh undergoes excessive distortion, and are more
likely to end up with a negative Jacobian of an individual element. This is due to the fact that
the motion of the body and the mesh are the same in the Lagrangian approaches, viz., a given
node remains coincident with the same material point throughout the analysis. However,
numerical difficulties associated with excessive element distortion can be circumvented by
combining the merits of the UL method and an Eulerian approach. In the literature, this
strategy has provided two similar methods known as the Arbitrary Lagrangian-Eulerian
(ALE) method and the Coupled Eulerian Lagrangian (CEL) method. In this study, we use the
ALE technique and the mesh refinement strategy presented by Nazem et al. (2006) and
Nazem et al. (2008). In this method the analysis comprises two steps; an UL step in order to
obtain the material displacements, and an Eulerian step to find the mesh displacements. The
separation of grid and material displacement and solving them in two different steps is
usually referred as the decoupled ALE method. Following these steps, mesh refinement is
performed and then all kinematic and static variables are transferred between the two meshes.

625
6 NUMERICAL EXAPMLE
The computational scheme explained here has been implemented into SNAC, a finite element
code developed over many years by the geomechanics group at the University of Newcastle,
Australia. SNAC was used to analyse the free falling penetration of a rigid finless torpedo
anchor into a saturated soil layer. The aim of the analysis was to study the total penetration
depth of the anchor, the mobilised soil resistance, the deceleration characteristics of the
anchor during its installation phase, generation of the excess pore-water pressures in the
surrounding soil, and its subsequent dissipation.
6.1 The installation of a torpedo anchor.
The geometry of the torpedo anchor, the finite element mesh containing 3433 axisymmetric
triangular elements and 7079 nodal points, the boundary conditions, and the soil properties
are presented in Figure 1. The material properties shown in Figure 1 are defined as K 0 : the
coefcient of earth pressure at rest, : the slope of the normal compression line (NCL) in the
space of the logarithmic mean stress ( ln p ) versus the void ratio e, : the slope of the
unloading-reloading line in the ln p e space, e 0 : the intercept of the NCL on the e axis
when ln p = 0 , OCR: the over-consolidation ratio of the soil, ': the drained friction angle,
': the Poissons ratio, k: the coefficient of permeability, and : the unit weight of the soil.
The radial thickness of the soil elements underneath the anchor is equal to a third of the
anchor shaft radius. The submerged weight of the anchor, W, its diameter, D, and its length,
L, are assumed to be, respectively, 40 kN, 0.6 m, and 6.8 m. To avoid further material
nonlinearity, the shear strength increase due to strain rate effects has been ignored in this
example. The use of the mortar type method in the finite element contact model facilitates a
curved surface between the torpedo and the soil due to quadratic shape functions. Therefore,
the finite element discretisation of the pile does not include any sharp corners, thus reducing
the numerical oscillations in the soil response. The coefficient of friction at the interface, ,
is 0.20 and a penalty parameter equal to 1 106 is used to enforce the contact constraints. The
Modified Cam Clay soil model with a rounded Mohr Coulomb failure surface in the
deviatoric plane was utilised to study the dynamic soil response. The plastic flow was
governed by an associated flow rule and, for the stress states that are not at the critical state,
plastic straining entails volumetric hardening of the material.
The simulation was started by applying a body force loading due to the self-weight of the
soil. Then an overburden pressure of p 0 = 50kPa was applied to the surface of the soil layer
over a long period of time to allow dissipation of excess pore pressures. After generating a
non-zero stress field, the location of the yield surface at each integration point in the finite
element mesh was adjusted according to the initial effective stresses and the value of the
overconsolidation ratio (OCR). Finally, the torpedo was deployed and allowed to impact the
soil vertically at an initial velocity of 15 m/sec, noting that the impact velocity is assumed to
be less than its terminal velocity in water.

626
Torpedo Anchor

D =0.6m
W= 40kN
= 0.2
K0 = 0.67
= 0.25
= 0.05
e0 = 1.8
OCR = 2
= 25
= 0.3
k =10-8 m/sec
= 20 k/m3

Figure 1. Finite element model of torpedo anchor

The predicted total dynamic soil resistance versus penetration, normalised by D, is plotted
in Figure 1. According to Figure 1, the predicted response is reasonably free of oscillation.
The energy absorbing boundaries, the numerical damping of the time-integration algorithm
and, more importantly, incorporation of higher order mortar contact elements have all
contributed to obtaining such a smooth curve. Numerical analyses have shown that
smoothing the torpedo surface via the use of higher order contact elements probably makes
the major contribution in decreasing these oscillations, while the effect of absorbing
boundaries is further manifested when the analysis proceeded to the consolidation stage.

600
Total dynamic soil resistance (kPa)

500

400

300

200

100

0
0 2 4 6 8 10 12 14
Penetration / D

Figure 2. Total dynamic soil resistance profile

627
The deceleration characteristics of the torpedo can be investigated by plotting the velocity
variations versus time, as depicted in Figure 3a, as well as by plotting velocity variations
versus normalised penetration depth, as shown in Figure 3b. It is observed that the torpedo
slightly accelerates at the early stages of penetration, mainly due to two reasons; (a) the sum
of soil resistance and the frictional forces at the interface between the soil and torpedo is
initially less than the weight of the torpedo, (b) the impact velocity of torpedo is less than its
terminal velocity in water, allowing the torpedo to continue to accelerate even after the initial
impact. The torpedo then starts to decelerate at an approximately linear rate, as shown in
Figure 3a. This trend is consistent with the soil resistance profile plotted in Figure 2. Figure
3b indicates that the deceleration occurs after the torpedo has penetrated ~0.5D into the soil
layer, and the predicted depth of installation is 13.4D.

Time (sec) Penetration / D


0.0 0.2 0.4 0.6 0.8 1.0 0 2 4 6 8 10 12 14
0 0
Velocity (m/sec)

Velocity (m/sec)
-4 -4

-8 -8

-12 -12

-16 -16

(a) (b)

Figure 3. a- Velocity versus penetration; b - Velocity versus normalised penetration

5.2 Pore pressure generation and dissipation


To investigate the generation of excess pore pressures, we monitor the development of excess
pore pressures at seven different points, located at depth of 5D and radial distances of 0.0,
0.17D, 0.33D, 0.5D, 0.67D, 1.5D, and 4.0D, throughout the installation process. The excess
pore pressures developed at these seven points are plotted versus time in Figure 4, noting that
the total installation time is 0.85 s. According to this figure, the excess pore water pressures
experience a relatively steady state following a peak value in each case. Typically, the
magnitude of excess pore water pressure at a point in the soil increases as the pile advances
towards it, but once the pile passes that location the magnitude decreases sharply, and
approaches the steady state value. The sudden drop in the excess pore water pressure is
significant for the soil elements located within a radial distance of 1D from the pile shaft. It
is also observed that the installation of pile only influences a region ~3D in the radial
direction and ~1D in the vertical direction, measured from the pile tip. A contour plot of the
excess pore water pressures developed at the end of pile installation is depicted in Figure 5.
When the full embedment depth is achieved and the torpedo comes to rest, the
computational process automatically proceeds to the setup analysis and consolidation of the
soil is permitted. The adopted automatic time stepping algorithm chooses very small time
steps at the beginning of the set up analysis when high pore water pressure gradients occur,
and then it gradually increases the time steps based on the estimated errors.

628
300

250
Excess pore water pressure (kPa)
r=0.17D
200
r=0.33D
150 r=0.50D

100 r=0.67D r=0.0


r=1.50D
50

0 r=4.0D

-50
0.0 0.2 0.4 0.6 0.8
Time (sec)

Figure 4. Excess pore water pressure evolution at depth 5D throughout the installation phase

This process is continued until the generated excess pore water pressures dissipate
entirely. Figure 6 depicts the dissipation curves for the soil elements located at depth of 5D
and different radial distances measured from the pile tip. It is observed that a degree of
consolidation of 90% (the degree of consolidation is equal to 1 minus the ratio between the
current excess pore water pressure and the initial excess pore water pressure) for elements
within 1D radial distance from the torpedo shaft is achieved 13~33 days after installation.
The same degree of consolidation takes place within 33~72 days for elements between 1D
and 1.83D. Figure 6 also shows that a degree of consolidation of ~96% is attained for the
entire affected zone at depth 5D after ~260 days. It is concluded that the most of the pullout
capacity of the torpedo anchor (soil resistance) is available much earlier than the completion
of consolidation as the most of the excess pore pressure dissipates within a matter of days or
weeks.

7 CONCLUSION
A fully coupled dynamic consolidation procedure for analysing torpedo anchors, taking into
account inertia effects, large deformation and the flow of pore water through the soil, was
presented in this study. The evolution of soil resistance as well as the development of excess
pore water pressures in the saturated soil due to the free falling penetration of a torpedo
anchor was studied. The nonlinear behaviour of the solid soil skeleton was predicted by the
Modified Cam Clay (MCC) soil model, and the mortar frictional contact method was used to
model soil-structure interaction. The numerical results showed that for the lightly over
consolidated soil simulated by the MCC material model, excess pore water pressures are
generated in the soil surrounding the pile tip and its shaft. The magnitude of excess pore
water pressure first increases when the pile tip is above or at the level of the point of interest
in the soil, then decreases once the pile tip has moved below the evaluation point, and finally
approaches a steady value at the end of the installation phase.

629
Figure 5. Excess Pore-water pressure contour at the end of installation

120

100 r=0.0
Excess pore water pressure (kPa)

80
r=0.17D

r=0.33D
60
r=0.50D
r=0.67D
40 r=1.0D
r=1.50D
r=1.83D
20

r=3.0D
0
1.E-05 1.E-04 1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03

Time (Day)

Figure 6. Excess pore water pressure dissipation versus time for elements at depth 5D

630
REFERENCES
API RP 2A, (2002), Recommended Practice for Planning, Designing and Constructing Fixed
Offshore Platforms. 21st Edition, American Petroleum Institute, Washington, D.C.
Biot, M.A. (1956), The theory of propagation of elastic waves in a fluid-saturated porous
solid. J. Acous. Soc. Amer., 28, 168-191.
Carter, J.P, Booker JR, Small JC. (1979), The analysis of finite elasto-plastic consolidation.
Int. J. Numer .Anal. Meth .Geomech., 3,107129.
Chao, L., Borja, R.I., Regueiro, R.A. (2004) Dynamics of porous media at finite strain.
Comput. Methods. Appl. Mech. Engrg., 193, 38373870
Corapcioglu, M.Y., Tuncay, K. (1996), Chapter 5- propagation of waves in porous media.
Advances in porous media., Vol 3, Elsevier, New York., 361440
Kellezi, L. (2000), Local transmitting boundaries for transient elastic analysis. Soil
Dynamics and Earthquake Engineering,. Vol.19 (7), 533547
Lysmer, J. & R. L. Kuhlemeyer. (1969), Finite dynamic model for infinite media. J.
Engrg. Mech. Div., ASCE, Vol. 95, No. EM4 859-877.
Medeiros Jr, C. J. (2002), Low Cost Anchor System for Flexible Risers in Deep Waters,
Proceedings of '34 Annual Offshore Technology Conference, Houston, paper No. OTC
14151.
Morland, L. W. (1972), A simple constitutive theory for a fluid-saturated porous solid. J.
Geoph. Res., 77, 890900.
Nazem, M., Sheng, D., Carter, J.P. (2006), Stress integration and mesh refinement in
numerical solutions to large deformations in geomechanics. Int. J. Numer .Meth .Engng.
65,1002-1027.
Nazem, M., Sheng, D., Carter, J.P. & Sloan, S.W. (2008), Arbitrary-Lagrangian-Eulerian
method for large-deformation consolidation problems in geomechanics. Int. J. Num.
Anal. Meth. Geomech., Vol. 32, 1023-1050.
Papadopoulos, P. & Taylor, R.L. (1990), A mixed formulation for the finite element solution
of contact problems Technical Report UCB/SEMM Report 90/18, University of
California at Berkeley.
Raie, M. & Tassoulas, J. (2009). Installation of Torpedo Anchors: Numerical Modeling. J.
Geotech. Geoenviron. Eng., Vol. 135(12), 18051813.
Randolph, M; Richardson, M.D; OLoughlin, C. D. (2005), The geotechnical performance
of Deep Penetrating Anchors in calcareous sand, Proceedings of the International
Symposium on Frontiers in Offshore Geotechnics (IS-FOG 2005), Perth, WA, Australia
Richardson, M. D., OLoughlin, C. D; Randolph, M. F; C. Gaudin. (2009), Setup Following
Installation of Dynamic Anchors in Normally Consolidated Clay. J. Geotech.
Geoenviron. EngC
Small, J.C., Booker, J.R., Davis, E.H. (1976), Elasto-plastic consolidation of soil.
International Journal of Solids and Structures., 12, 431-448.
Sousa, J.R.M., Cristiano, S., de Aguiar., Ellwanger, G.B., Elisabeth, C.P., Foppa, D.,
Medeiros, J.C. (2010), Undrained Load Capacity of Torpedo Anchors Embedded in
Cohesive Soils. J. Offshore Mech. Arct. Eng., Vol. 133(2), 021102-021114
Straughan, B., (2008), Stability and wave motion in porous media, applied mathematical
sciences, vol 165. Springer, New York
Sturm, H & Andresen, L. (2010), Large deformation analysis of the installation of dynamic
anchor, Proceedings of 7th European Conference on Numerical Methods in
Geotechnical Engineering (NUMGE '10), pp. 255260.
Thomas, R. M. & Gladwell, I. (1988), Variable-order variable-step algorithms for second-
order systems. Part 1: The methods Int. J. Numer. Meth. Engng., 26, 3953.
True, D. G. (1976), Undrained Vertical Penetration into Ocean Bottom Soils, PhD
Dissertation, University of California, Berkeley, California.

631
Wriggers. P. (2006), Computational Contact Mechanics. 2nd ed. Springer, Heidelberg.

632
All in-text references underlined in blue are linked to publications on ResearchGate, letting you access and read them immediately.

You might also like