Sunteți pe pagina 1din 14

International Journal of Computer Mathematics

Vol. 86, No. 2, February 2009, 301–313

Non-linear analysis of cable networks by FEM and


experimental validation
J.J. del Coz Díaza *, P.J. García Nietob , D. Castro Fresnoc and E. Blanco Fernándezc
a Department of Construction, University of Oviedo, EPSIG, Gijón, Asturias, Spain;
b Department of Mathematics, University of Oviedo, Sciences Faculty, Oviedo,
Asturias, Spain; c Department of Construction, University of Cantabria,
ETSIPCC, Santander, Spain

(Received 07 September 2007; revised version received 10 November 2007; accepted 05 January 2008)

This work studies the analysis of the resistant capacity of cable nets for the stabilization of slopes. Two tests
have been carried out, one with a distributed longitudinal load and the other with distributed transversal
load, in order to simulate in situ the working conditions of these systems. Tensile tests were also carried out
on the cable elements of the network in order to obtain the non-linear mechanical properties. On the one
hand, the proposed numerical procedure uses the finite element method (FEM) and it takes into account the
material and geometrical non-linearities due to the geometrical change in the cable net substructure. On
the other hand, the contour (boundary) beam is modelled by linear beam elements with contact between
them. The numerical results of the longitudinal and cross tests were simulated for different geometrical
configurations, generating convergent results with respect to strain and resistance, which permit the extrap-
olation from tested cable nets to untested simulated cable nets, maintaining a basic configuration of constant
parameters. The laboratory tests only provide information about the strain and maximum resistance, but
they do not establish a relationship between the values of stresses of each net element. These data have
been obtained through the computational simulation by FEM. A reliable model of the interaction of the
flexible contour beam with the cable network enables the achievement of more efficient solutions in the
design analysis. Finally, we compare the structural behaviour of the numerical and experimental results by
means of the equivalent elastic modulus and the equivalent Poisson ratio. Excellent agreement between the
predicted results by FEM and test observations was found. Besides, conclusions and suggested procedures
of calculation applied on the cable networks are given and numerical models to evaluate the stability of
the slope protection systems are presented.

Keywords: finite element modelling; wire mesh and cable net slope protection system; material and
geometrical non-linearities; soil stabilization; erosion control

2000 AMS Subject Classification: 74S05; 74C05; 35J60

1. Introduction

On gentle slopes, erosion blankets, both natural and man-made, or hydro-seeding can be employed
to hold soil and seed in place until vegetation gets established. These lightweight products may not

*Corresponding author. Email: juanjo@constru.uniovi.es

ISSN 0020-7160 print/ISSN 1029-0265 online


© 2009 Taylor & Francis
DOI: 10.1080/00207160801965339
http://www.informaworld.com
302 J.J. del Coz Díaz et al.

be enough to control erosion caused by high water velocities associated with steep slopes. Where
steeper slopes and high water velocities prevail, a wire mesh and a cable net slope protection
system may be used [5]. Wire mesh and cable net slope protection have been in use for more
than 50 years along highways to control rockfall on actively eroding slopes. The present work
studies these last flexible systems of slopes’ surface stabilization, the levelling of grounds and
loose materials by the finite element method (FEM) as well as the analysis and a subsequent
comparison with the experimental results. Effective control of erosion on highway slopes involves
assessing the erosive environment correctly, predicting the erosion resistance of materials before
their placement or exposure on the slopes, and using slope maintenance in order to minimize
slope erosion-resistance damage.
The combination of metallic meshes and/or cable networks with anchors [20,23] as one unique
system provides a system of surface stabilization and control of the erosion. Besides it is possible
that the specific and smooth correction of the slope as well as to get a solution integrated with the
environment by means of the combination of this system with the re-vegetation of the surface.
Thus, the system will be composed of a set of anchors [13], a control element of the erosion,
a surface element of support, all of them combined with the technique of re-vegetation of the
surface if it is necessary. The forces are transmitted through the anchors to the stable area of the
massif and, depending on the depth of this stable area, will be necessary anchors of smaller or
bigger length. The system can be reinforced with steel cables in order to allow the distribution
of loads. The fastening of the membrane to the anchors is carried out with special plates whose
geometry varies depending on the membrane type [15].
The system (see Figure 1) works like a lightly tightened continuous surface that receives from
the ground the loads due to the earth pressure on the cable networks and it transmits them to the
head of the anchors, which in turn transmits them to the stable area of the hillside or slope. The
grid of the anchors (separated between them a distance l), type of anchorage bar (with lengths LA
and LC ) and other characteristics such as the slope angle β, the sliding angle α and the thickness of
the unstable layer H , will depend on the geotechnical calculations carried out in each place. The
grid of anchors can have any dimension from a design point of view but it is convenient to adapt
it to the standard dimensions of the network cloths. The lengths of the inner anchors will depend
on the thickness of the altered or fractured area. Once the assembly is finished, it is provided with
a pre-stress to the whole system [2]. When the surface of the slope begins to move, its movement
will cause a pressure. When the pressure load and the pre-stress are equall, the system of anchors
will act as passive bolts.
Once the system object of the study has been described, the most important characteristics in
the numerical simulation will be discussed. Next, we shall carry out an analysis of the obtained
results and to conclude we shall do a comparison of numerical results with the experimental tests.
Finally, we shall analyse the causes of the possible discrepancies present between the experimental
and numerical results.

2. Mathematical model

The FEM is used to model the cable network and the boundary structure. Approximation of the
structure by a discrete number of finite size straight and curved elements, connected at nodes, is a
quite natural simulation since the original cable network consists of turns made of elements and
curves. According to stiffness properties, the structural system is divided into several substructures
(one per wire). Therefore, the resolution of this problem implies the simultaneous study of two non-
linearities [3]: (1) material non-linearity (elastoplastic behaviour in this case), and (2) geometrical
non-linearity or large displacements. Thus, the cable network is modelled by curved three-noded
International Journal of Computer Mathematics 303

Figure 1. Cable networks: a scheme of the system depending on the membrane type [15].

elements with the following assumptions:


• This cable element can resist tension, bending, shear and torsion. It has six degrees of freedom
per node.
• Behaviour of the cable material between any nodes is non-linear and elastoplastic.
• Geometrical non-linearity due to large displacements is accounted for.
The most straightforward way to create the mathematical model of a cable network with a flexural
contour structure is to include the cable and the beam elements in a common system of equations
in general matrix form. If load P is applied at the nodes of the original undeflected structure, an
unbalanced resultant force is produced [1]:

R = F − P, (1)
where F is the vector of internal forces at nodes and P is the vector of external forces at nodes.
R is a non-linear function with respect to the displacements of the structure. In order to find
304 J.J. del Coz Díaz et al.

the equilibrium state geometry, Equation (1) has to be solved, using relevant numerical methods
[4,10]. According to the Newton-Raphson method we calculate the correction of the solution at
every iteration cycle [25]:
 −1 i
di = Ki R, (2)
where is the stiffness matrix of the structure. The new geometry after cycle i is determined by the
vector of nodal co-ordinates:
ui+1 = ui + d i . (3)
Nodal displacements produce elasto-plastic elongation [19,24] of the cable elements and change
of the tension and bending forces according to axial and flexural stiffness. Therefore stiffness
matrix K and unbalanced force R have to be updated for every iteration cycle according to the
current geometry. Convergence of the described method is usually rapid near the equilibrium
state, but may be disturbed if elements of the stiffness matrix vary significantly or if the initial
geometry is inappropriate.

2.1 Plasticity

Plasticity refers to non-recoverable deformation and non-unique stress paths in contrast to non-
linear elasticity, where the entire load-deflection path is unique and the strains are recovered
on load removal. The mathematical theory of plasticity is of a phenomenological nature on the
macroscopic scale, and the objective of the theory is to provide a theoretical description of the
relationship between stress and strain for a material that exhibits an elasto-plastic response [12].
The plastic behaviour is characterized by irreversibility of stress paths and the development of
permanent (i.e. non-recoverable) deformation (or strain), known as yielding (or plastic flow). A
hardening plastic material model provides a refinement of the ideal plastic material model. In this
model, it is assumed that the yield stress depends on some parameter κ (e.g. plastic strain ε p ),
called the hardening parameter. The general yield criterion is expressed in the form [17,26]:
 
F σij , κ = 0. (4)

This yield criterion can be viewed as a surface in the stress space, with the position of the surface
dependent on the instantaneous value of the hardening parameter κ. Since any yield criterion
should be independent of the orientation of the coordinate system used, F should be a function of
the stress invariants only. Experimental observations indicate that plastic deformation in metals
is independent of hydrostatic pressure. Therefore, F must be a function of the stress invariants of
the deviatoric stress tensor σ  [16,26]:
  1   1   
F J2 , J3 , κ = 0, J2 = σ σ , J3 = σ σ σ . (5)
2 ij ij 3 ij jk ki
Two of the most commonly used yield criteria are given next, the Tresca yield criterion:

F = 2σ̄ cos θ − Y (κ) = 0, σ̄ = J2 ; (6)

and the Huber-von Mises yield criterion:



F = 3 J2 − Y (κ) = 0, (7)

where Y is the yield stress from uniaxial tests, θ is the angle between the line of pure shear and
the principal stress σ1 , and σ̄ = J2 is called the effective stress.
International Journal of Computer Mathematics 305

After initial yielding, the stress level at which further plastic deformation occurs may be depen-
dent on the current degree of plastic straining, known as strain hardening [14]. Thus, the yield
surface will vary (i.e. expand) at each stage of plastic deformation. When the yield surface is
independent of the degree of plasticity, the material is said to be ideally (or perfectly) plastic. If
the subsequent yield surfaces are a uniform expansion of the original yield surface, the hardening
model is said to be isotropic. On the other hand, if the subsequent yield surfaces preserve their
shape and orientation but translate in the stress space, kinematic hardening is said to take place.
Consider the uniaxial stress–strain traditional curve. The behaviour is initially linear elastic with
slope E (Young’s modulus) until onset of yielding at the uniaxial yield stress σY . Thereafter, the
material response is elasto-plastic with the local tangent to the curve, ET , called the elasto-plastic
tangent modulus, continually changing.
At some stress level σ in the plastic range, if the load is increased to induce a stress of dσ ,
it results in a corresponding strain dε. This increment of strain contains two parts: elastic dε e
(recoverable) and plastic dε p (non-recoverable) [16,22]:

dσ dσ
dε = dεe + dε p , dεe = , = ET . (8)
E dε
The strain-hardening parameter, H , is defined by [11]:

dσ (dσ/dε) ET
H = = = . (9)
dε p 1 − (dε e /dε) 1 − (ET /E)

The element stiffness for the linear elastic portion is, say [K e ] [7,22]:
 xb
[K e ] = [B]T [D e ][B] dx, (10)
xa

where [D e ] is the linear elasticity matrix (D e = E for the uniaxial case). When the element deforms
plastically, [D e ] reflects the decreased stiffness. This is computed, for uniaxial material behaviour,
by the following procedure. The increment in load dF causes an incremental displacement du:

du = he dεxx = he (dε e + dε p ) , dF = A dσ = Ae H dεp , (11)

where he is the length and Ae the area of cross-section of the element. The effective stiffness is
[7,22]:
 
dF Ae H dεp E Ae E
E =
ep
= = 1− . (12)
du he (dε e + dε p ) he (E + H )
The element stiffness for the plastic range becomes [22],
 xb
ep

K = [B]T D ep [B] dx, (13)


xa

where [D ep ] is the material stiffness in the plastic range. For the uniaxial case D ep = E ep .
Equation (10) is valid when σ < σY and Equation (13) is valid for σ > σY . Note that dσ = σ − σY
when σ > σY .

2.2 Large displacements

Whether the displacements (or strains) are large or small, equilibrium conditions between inter-
nal and external ‘forces’ have to be satisfied. Thus, if the displacements are prescribed in the
306 J.J. del Coz Díaz et al.

usual manner by a finite number of nodal parameters a , we can obtain the necessary equilibrium
equations using the virtual work principle [6,21,27]:

a) =
 ( B̄ T σ dV − f = 0, (14)
V

where  once again represents the sum of external and internal generalized forces, and in which
B̄ is defined from the strain definition ε as:

dε = B̄ d
a. (15)

The bar suffix has now been added for, if displacements are large, the strains depend non-linearly
on displacement, and the matrix B̄ is now dependent on a . We see that it can be conveniently
write:
B̄ = B0 + BL (
a ), (16)

in which B0 is the same matrix as in linear infinitesimal strain analysis and only BL depends on
the displacement. In general, BL will be found to be a linear function of such displacements.
Clearly the solution of Equation (14) will have to be approached iteratively. If, for instance,
the Newton–Raphson process is to be adopted we have to find the relation between d a and d.
Thus taking appropriate variations of Equation (14) with respect to da we have [6,22]:
 
d = dB̄ σ dV +
T
B̄ T dσ dV = KT d
a (17)
V V

and using equation dσ = D dε and Equation (15) it is obtained:

dσ = D dε = D B̄ d
a

and taking into account the Equation (16), it is verified that dB̄ = dBL . Therefore,

d = dBLT σ dV + K d
a = Kσ d
a + K d
a, (18)
V

where

T
K= B D B dV = K0 + KL ,
V

in which K0 represents the usual, small displacements stiffness matrix and the matrix KL is due
to the large displacements, and are given by refs. [6,21,27]:
 
 
K0 = B0T D B0 dV , KL = B0T D BL + BLT D BL + BLT D B0 dV . (19)
V V

To summarize, Equation (18) can be expressed globally as [26]:

d = (K0 + Kσ + KL ) d
a = KT d
a, (20)

where KT represents the total, tangential stiffness, matrix. Newton-type iteration can once more
applied precisely in order to solve the final non-linear problem.
International Journal of Computer Mathematics 307

2.3 Multibody coupling by joints

One of the major areas of non-linear analysis is the solution of problems in which separate bodies
or structures may come in contact with each other. Several methods have been developed to handle
such problems and, in this paper, the multibody coupling has been adopted [6]. Often it is desirable
to have two (or more) rigid bodies connected in some specified manner. For example, in our case,
each turn in the cable network is hooked to another. Both turns are treated as flexible non-linear
bodies and it is necessary to consider coupling among them in order to obtain the global structural
behaviour of the network. This type of interconnection is commonly referred to as a joint. Joints
may be constructed by a combination of two types of simple constraints: translational constraints
and rotational constraints.
When generating our model, we define the relationships among different degrees of freedom
by using elements to connect the nodes. However, we need to model distinctive features, such as
the special internodal connections, which cannot be adequately described with elements. In this
way, we can establish special associations among nodal degrees of freedom by using coupling. In
this work, we use a joint which is a linear connection where one body may freely move around the
other but relative translation is prevented (see Figure 2 below). Thus each turn may not translate
relative to another in any direction. If a full translation constraint is imposed a simple relation
may be introduced as [6,9]:
Cj = x(a) − x(b) = 0,
 (21)
where a and b denote two rigid bodies. Thus, addition of the Lagrange multiplier constraint

j = λ Tj x(a) − x(b) (22)

imposes the spherical joint condition. It is necessary only to define the location for the spherical
joint in the reference configuration. Denoting this as X  j (which is common to the two bodies)
and introducing the rigid motion yields a constraint in terms of the rigid body positions as [6]:

j = λ Tj r(a) +
 (a) X j − R (a) − r(b) −
 (b) X j − R (b) , (23)

where R is some reference point in the undeformed bodies a and b, r r is the position vector
position of the same point in the deformed bodies and  is the matrix rotation of each body.
The variation and subsequent linearization of this relation yields the contribution to the residual
and tangent matrix for each body, respectively. If the translation constraint is restricted to be in one
direction with respect to, say, body a it is necessary to track this direction and write the constraint
accordingly.
To accomplish this, the specific direction of the body a in the reference configuration is required.
This may be computed by defining two points in space X  1 and X
 2 from which a unit vector V is
defined by
2 − X
X 1
V =  . (24)
  1 
X2 − X
The direction of this vector in the current configuration, v, may be obtained using the rigid rotation
for body a
 (a) V .
v = (25)
A constraint can now be introduced into the variational problem as ref. [6]:
 T 
j = λj V T  (a)  (a) X
r(a) +  j − R (a) − r(b) −
 (b) X j − R (b) , (26)
308 J.J. del Coz Díaz et al.

Figure 2. Finite-element model and detail of von Mises stresses: (a) couplings among turns; (b) detail of the wire mesh;
(c) von Mises stresses for the direct tensile test; and (d) von Mises stresses for the cross tensile test.

where, owing to the fact there is only a single constraint direction, the Lagrange multiplier is a
 j denotes the reference position where the constraint is imposed.
scalar λj and, again, X
The above constraints may also be imposed by using a penalty function [8,9]. The most
direct form is to perturb each Lagrange multiplier form by a penalty term. Accordingly, for
each constraint we write the variational problem as [8,22]:

1 2
j = λj C j − λ , (27)
2 · kj j

where that the limit kj → ∞ yields exact satisfaction of the constraint. Use of a large kj and
variation with respect to λj gives
 
1
δλj C j − λj = 0 (28)
kj
International Journal of Computer Mathematics 309

and may easily be solved for the Lagrange multiplier as

λj = kj · Cj , (29)

which when substituted back into Equation (26) gives the classical form

kj
2
j = Cj . (30)
2

We will recognize that Equation (27) is a mixed problem, whereas, Equation (30) is irreducible.
An augmented Lagrangian form is also possible for contact problems [8,9].

3. Geometrical model

Several geometrical models were analysed in this work. In order to carry out precise numerical
simulations, it is necessary to reproduce accurately both the geometrical and mechanical char-
acteristics of the system object of study. In the present work, the main characteristics of meshes
are: (1) geometry of the mesh – rhomboidal; (2) size of each rhombus – inscribed circle of 65 mm;
(3) dimensions of each rhombus – 143 × 83 mm (±2%); (4) surface of each rhombus – 2625 mm2 ;
(5) number of turns in horizontal direction – 12 turns/m; (6) number of meshes in vertical
direction – 7; meshes/m; (8) diameter of the wire – 3.0 mm; (9) nominal resistance – 1770 N/mm2 ;
and (10) material – high strength steel.

4. Finite-element model and analysis

Based on the geometrical model of a piece of cable network (see Figures 2a and 2b) the finite-
element model was built, following a four-step process. Firstly, the definition of mechanical
properties base on real tests. Secondly, the selection of the element types, formulations and
physical properties. Thirdly, the geometrical model was meshed and finally, loads and boundary
conditions were applied.

4.1 Mechanical properties

The mechanical properties of the high strength steel used, obtained by means of experimental
direct tensile tests, are:

• Young’s modulus (MPa): 196,906.


• Stress (MPa) versus strain: (0.00–0.00; 1101.328–0.0056; 1482.037–0.0093; 1794.777–0.0139;
1826.427–0.0300).

From these results, it is necessary to take into account that in an analysis of plasticity with large
strains, the used calculation program is fed on the material’s constants of a true stress curve while
for the analysis of small strains it makes use of an engineering stress–strain curve. On the one
hand, since in small strains the engineering strain and the logarithmic (true) strain are practically
identical, the true stresses and the logarithmic strains can be employed in general applications. On
the other hand, since in this simulation important strains are reached, it is necessary to transform the
engineering strains obtained in the tests into logarithmic strains. For that, the following formulas
310 J.J. del Coz Díaz et al.

are used:
 
Strain: εtrue = Ln 1 + εeng ; (31)
 
Stress: σtrue = σeng 1 + εeng . (32)

4.2 Element types and meshing

The element used in this study is the BEAM189 [18], since, due to its characteristics, it is the
most appropriate for the problem. It is a three-dimensional element suitable for analysing slender
to moderately stubby/thick beam structures. This element is based on Timoshenko beam theory
and shear deformation effects are included. This finite element is a quadratic (three-node) beam
element with six degrees of freedom at each node and it includes stress stiffness terms. These
include translations in the x, y, and z directions and rotations about the x, y, and z directions.
This element is well suited for large rotation and large strain non-linear applications as in this
work. In this work a extremely fine mesh has been used in order to obtain good accuracy. This
mesh has an element size of 1 mm, giving rise to a mesh of approximately 70,000 elements.

4.3 Loads and boundary conditions

Two different boundary conditions have been applied in the numerical model. On the one hand,
in order to simulate the direct tensile test, several boundary conditions have to be imposed in the
finite-element model. Firstly, at the lower side of the mesh, displacements in directions x, y and
z, are constrained. Secondly, at the left and right sides, displacements in directions x and z are
constrained in order to allow the displacement in vertical direction y. Finally, at the upper side, the
displacement in z direction is constrained and it is imposed a displacement in y direction of 60 mm.
On the other hand, the cross tensile test requires different boundary conditions. In the first
place, at the upper and lower sides, displacements in directions z and y are constrained, allowing
the displacement in direction x. Secondly, at the left side, displacements in directions x and z are
constrained. Finally, at the right side, a displacement in direction x of 60 mm is imposed.

5. Non-linear analysis of cable networks

A non-linear analysis was performed taking into account the geometrical and material non-
linearities: large displacements and plasticity. The solution controls were also adjusted to improve
convergence. Thus the parameter time was set to the value of 1, corresponding to the value of
maximum displacement applied (60 mm), the geometrical non-linearity was activated, the inertial
effects were not included, the number of equilibrium iterations was specified and the convergence
tolerance values of displacements were delimited as well as the time step for the analysis. In this
work, we have adopted the following values: (1) maximum number of equilibrium iterations – 250;
(2) convergence tolerance values of displacements – 0.5%; (3) starting time step – 0.001 (relative
to one); and (4) minimum time step – 0.00001 (relative to one). Finally, in order to get convergence,
the total number of iterations was about 160.

6. Analysis of results and discussion

Next, we show the von Mises stresses for the direct and cross tensile tests (see Figures 2c and
2d), corresponding to the last load step with a 60 mm displacement.
International Journal of Computer Mathematics 311

6.1 Comparison with experimental results

The previous numerical results were compared with the experimental results obtained in the
laboratory of the Cantabria University. Graphs representing both the numerical and experimen-
tal results are shown in Figures 3a and 3b, in order to get the longitudinal and transversal
equivalent elastic moduli by means of a linear fitting from data. On the one hand, values
obtained of the longitudinal equivalent elastic modulus are 1690.6 kN/m (numerical simula-
tion) and 1886.3 kN/m (experimental test) from the direct tensile test. On the other hand, values
obtained of the transverse equivalent elastic modulus are 319.9 kN/m (numerical simulation) and
222.47 kN/m (experimental test) from the cross tensile test.
In order to simulate the membrane behaviour of the cable networks, it is necessary to calculate
the relationship between the longitudinal and transversal stresses for the two types of tests: direct
tensile test and cross tensile test. The quotient of the transversal stress divided by the longitudinal
stress are named equivalent Poisson’s ratio:

σy
xy =
μeq . (33)
σx

eq
Figures 3c and 3d show the μxy calculated by linear fitting from experimental and numerical data
for the direct tensile test and cross tensile test. On the one hand, values obtained of the longitudinal
equivalent Poisson’s ratio are 0.33 (numerical simulation) and 0.217 (experimental test) from the
direct tensile test. On the other hand, values obtained of the transverse equivalent Poisson’s ratio
are 0.4437 (numerical simulation) and 0.4721 (experimental test) from the cross tensile test.

Figure 3. Comparison with experimental results: (a) stress versus strain by FEM and experimental test for the direct
tensile test; (b) stress versus strain by FEM and experimental test for the cross-test; (c) equivalent Poisson’s ratio for the
direct tensile test; and (d) equivalent Poisson’s ratio for the cross-tensile test.
312 J.J. del Coz Díaz et al.

7. Conclusions

A method for modelling static behaviour of cable networks has been developed and verified
here. The equilibrium equations of elasto-plastic network are derived in the incremental form.
The procedure can serve as an alternative tool in order to avoid expensive physical tests in the
laboratory of different configurations and geometries of cable networks. From the results obtained,
the following conclusions can be drawn:
In the first place, the FEM has been shown as suitable tool in the modelling and analysis
of singular structures, such as the complex structural behaviour of cable networks with strong
non-linearities.
The definition of the cable network geometry is very cumbersome using a finite-elements
analysis program. For this reason, a three-dimensional parameter design program was used in
order to design the turn appropriately as well as the entire assembly of the cable network.
We have compared the numerical results with the experimental ones. The following aspects are
observed:
• A good agreement between the longitudinal equivalent moduli for both techniques: numerical
and experimental.
• A small deviation in the value of the transversal equivalent moduli, due to the residual stresses
[12,16] in turns and the flexibility of the test device.
• A good performance is observed with respect to the equivalent Poisson’s ratios.
In conclusion, the comparison between both methods prove the finite-element analysis as a reliable
tool to get quite accurate results in a reasonable amount of time, which allows the designers to
evaluate and optimize the design prior to manufacture and prototype testing.
Finally, in view of the obtained results for both tests, it can be considered that a numerical
simulation like this one can provide accurately results that will help us to understand the behaviour
of these stabilization systems. In future works, it would be necessary to take into account other
local phenomena in turns such as the effects due to residual stresses, the hysteresis of materials, etc.

Acknowledgements
We gratefully acknowledge the financial support provided by the Construction Technology Research Group (GITECO)
at the University of Cantabria, Malla Talud Cantabria Ltd. (MTC) and SODERCAN. We thank Swanson Analysis Inc.
for the use of ANSYS University Intermediate program. We also thank the Construction and Production Engineering
Department of the University of Oviedo for the computational support.

References

[1] A.A. Atai and A. Mioduchowski, Equilibrium analysis of elasto-plastic cable nets, Comput. Struct. 66 (1988),
pp. 163–171.
[2] M.R. Barnes, Form-finding and analysis of prestressed nets and membranes, Comput. Struct. 30 (3) (1988),
pp. 685–695.
[3] K. Bathe, Finite Element Procedures, Englewood Cliffs, Prentice-Hall, New York, 1996.
[4] D. Braess, Finite Elements: Theory, Fast Solvers, and Applications in Solid Mechanics, Cambridge University Press,
New York, 2001.
[5] H.A. Buchholdt, An Introduction to Cable Roof Structures, Thomas Telford, London, 1999.
[6] T. Chandrupatla and A. Belegundu, Introduction to Finite Elements in Engineering, Englewood Cliffs, Prentice-Hall,
New Jersey, 1991.
[7] R.D. Cook D.S. Malkus, M.E. Plesha and R.J. Witt, Concepts and Applications of Finite Element Analysis, Wiley,
New York, 2001.
[8] J.J. del Coz Díaz, P.J. García Nieto, C. Betegón Biempica, and G. Fernández Rougeot, Non-linear analysis of
unbolted base plates by the FEM and experimental validation, Thin-Walled Struct. 44 (2006), pp. 529–541.
[9] J.J. del Coz Díaz, F. Rodrígez Mazón, P.J. García Nieto, and F.J. Suarez, Design and finite element analysis of a wet
cycle cement rotary kiln, Finite Elem. Anal. Des. 39 (2002), pp. 17–42.
International Journal of Computer Mathematics 313

[10] P. Deuflhard and A. Hohmann, Numerical Analysis in Modern Scientific Computing, Springer-Verlag, New York,
2005.
[11] W. Han and B. Daya Reddy, Plasticity: Mathematical Theory and Numerical Analysis, Springer-Verlag, New York,
2006.
[12] R. Hill, The Mathematical Theory of Plasticity, Oxford University Press, New York, 1998.
[13] L. Hobst and J. Zajic, Anchoring in Rock and Soil, Elsevier Scientific Publishing Company, New York, 1983.
[14] L.M. Kachanov, Fundamentals of the Theory of Plasticity, Dover Publications, New York, 2004.
[15] J.W. Leonard, Tension Structures: Behaviour and Analysis, McGraw-Hill, New York, 1988.
[16] V.A. Lubarda, Elastoplasticity Theory, CRC Press, Boca Raton, 2001.
[17] J. Lubliner, Plasticity Theory, McMillan Publishing Company, New York, 1998.
[18] E. Madenci and I. Guven, The Finite Element Method and Applications in Engineering Using ANSYS, Springer-
Verlag, New York, 2005.
[19] J.L. Meek, Elasto-plastic analysis of cable net structures, Proceedings of IASS-ASCE International Symposium,
Atlanta, USA, 1994, pp. 781–790.
[20] B. Muhunthan S. Shu, N. Sasiharan, O. Al Hattamleh, T.C. Badger, S.M. Lowell and J. Duffy, Analysis and design
of wiremesh/cable net slope protection, Res. Rep. WA-RD 612.1, Washington State Department of Transportation,
2005.
[21] D.R.J. Owen and E. Hinton, Finite Elements in Plasticity: Theory and Practice, Pineridge Press Limited, West Cross,
Swansea, 1980.
[22] J.N. Reddy, An Introduction to Nonlinear Finite Element Analysis, Oxford University Press, New York, 2004.
[23] S. Shu, B. Muhunthan, T.C. Badger and R. Grandorff, Load testing of anchors for wire mesh and cable net rockfall
slope protection systems, Eng. Geol. 79 (2005), pp. 162–176.
[24] J.C. Simo and T.J.R. Hughes, Computational Inelasticity, Springer-Verlag, New York, 1998.
[25] J. Stoer and R. Bulirsch, Introduction to Numerical Analysis, Springer-Verlag, New York, 2004.
[26] L. Zhang, Engineering Plasticity and Impact, World Scientific Publishing Company, New York, 2002.
[27] O.C. Zienkiewicz and R.L. Taylor, The Finite Element Method: Solid and Fluid Mechanics and Non-linearity,
McGraw-Hill Book Company, London, 1991.

S-ar putea să vă placă și