Sie sind auf Seite 1von 12

Engineering Structures 33 (2011) 1698–1709

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Numerical procedure for the analysis of damaged polyester ropes


Juan Felipe Beltrán a , Eric B. Williamson b,∗
a
Department of Civil Engineering, University of Chile, Blanco Encalada 2002, Santiago, Chile
b
Department of Civil Engineering, The University of Texas at Austin, Austin, TX 78712, USA

article info abstract


Article history: In this paper, a numerical model capable of representing the impact of damage on the overall response
Received 11 July 2010 of jacketed polyester (PET) ropes is presented. This model considers rope behavior under axisymmetric
Received in revised form loading conditions and can account for the degradation of rope properties, including the effect of broken
5 January 2011
rope components, on overall rope response. To consider the confinement effect of rope jacketing on
Accepted 8 February 2011
Available online 11 March 2011
interior rope components, the rope jacket is assumed to behave like a thin-walled tube. If a rope has
broken rope components, the rope analysis is linearized by discretizing the rope into a series of two-
Keywords:
noded axial-torsional elements. The formulation of these elements accounts for material and geometric
Damage nonlinearities. Results of numerical simulations are provided to show how changes in the L/d ratio (rope
Synthetic-fiber rope length/rope diameter) and the initial state of rope damage influence the capacity, failure strain and
Analytical model stiffness of a damaged PET rope. Experimental data are used to validate the proposed model.
Rope failure © 2011 Elsevier Ltd. All rights reserved.
Rope capacity

1. Introduction these loading effects determine the overall rope response, which
can be expressed in terms of the rope extension and rotation [2].
Ropes are characterized by having a high axial strength For many years, steel wire ropes have been used extensively for
and stiffness in relation to their weight, combined with a low load bearing members mainly due to the strength offered by steel,
flexural stiffness. This combination is achieved by using a large coupled with the flexibility of rope construction, rope geometry,
number of components, each of which is continuous throughout and wire size that can be suited to a particular application. Over
a rope’s length. When loaded axially, each component provides the last few decades, however, the rope industry has been capable
tensile strength and stiffness, but when deformed in bending, the of producing fiber ropes with very high tenacity. Considering the
components have a low combined bending stiffness (provided many advancements made in fiber rope construction, synthetic-
their bending deformation is uncoupled from the axial response). fiber ropes can potentially replace steel wire ropes for certain
To facilitate handling, it is necessary to ensure that a rope has applications. The major difference between synthetic-fiber rope
some integrity as a structure, rather than being merely a set of and steel wire rope components is their strength-to-weight ratio.
parallel components. This characteristic is achieved by twisting the Synthetic-fiber ropes can be up to 10 times lighter than a steel wire
components together [1]. rope for a given member strength, reducing the weight (load) of the
A rope can be a critical load carrying member in many engi- system and installation costs [3]. Cases where synthetic-fiber ropes
neering applications, including cranes, lifts, mine hoisting, bridges, have replaced steel wire ropes are numerous and include the use
cableways, electrical conductors, offshore mooring systems, and of ropes for lifting equipment and materials at construction sites,
so on [1–3]. Different classes of ropes, suited for different pur- deep water mooring systems, anti-collision and protective nets,
poses, have a different number and arrangement of rope compo- heavy-lift helicopters used in military and commercial operations,
nents within a rope cross-section, and rope components can be and ship assist lines [3].
made from a variety of different materials. Although a rope is in- Each field of rope application has developed a specific body
tended primarily to carry a tensile load, the construction of a rope of knowledge, based on extensive testing and field experience,
is such that individual rope components are subjected to bending leading to empirical rules for each particular application. Unifying
and torsional moments, frictional and bearing loads, as well as ten- these empirical rules under some general mathematical and
sion. The magnitude and distribution of the stresses resulting from mechanics-based theory would allow a better understanding, and
in the long term, a better prediction of the mechanical behavior of
ropes than current methods allow. In addition, a unified modeling
∗ Corresponding author. approach can help reduce the need for expensive tests under a
E-mail address: ewilliamson@mail.utexas.edu (E.B. Williamson). variety of operating conditions. Thus, due to their extensive use
0141-0296/$ – see front matter © 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engstruct.2011.02.007
J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709 1699

Fig. 1. Cross-section of multi-level rope.

and the need to predict their behavior, several researchers have In order to compare the predicted damaged rope response with
presented analytical models that permit the calculation of rope available experimental data [18], two parameters—the number of
response based on the material properties and the geometrical broken rope components and the L/d (rope length/rope diameter)
arrangement of the rope components [4,5]. ratio—are parametrically evaluated using the proposed model to
In the particular case of synthetic-fiber ropes, one of the illustrate how these two variables affect the capacity, stiffness, and
major concerns of the rope industry is the impact of damage failure axial strain of a given damaged rope. In this particular study,
on overall rope performance. Damage to ropes is a complicated based on the results of prior research [17], degradation of rope
process that could depend on a variety of factors such as properties is represented by a power law for the damage index
strain range, abrasion, installation procedures, environmental evolution (see Appendix).
interaction, handling, creep, number of loading cycles, etc. ([6–9]
; among others). The process by which damage occurs can be 2. Numerical algorithm to simulate damaged rope response
represented through a degradation of the properties of individual
rope components, and it can also include the complete rupture
of one or more components. Information on the residual strength 2.1. Rope cross-section modeling
and deformation capacity of damaged polyester (PET) ropes is
important to the rope industry for developing guidelines for Full details of PET rope cross-section modeling are given
estimating allowable service life and determining whether or not elsewhere [11,15] and only the main concepts are presented below
damaged ropes can be retained in service. for completeness. In general, a synthetic-fiber rope is defined as a
The development of mathematical models to estimate rope re- structural member constructed by assembling all components in
sponse under a variety of loading conditions that can account for hierarchical order. In the terminology used for this research, levels
the degradation of mechanical properties has been a major focus refer to the number of structures in the hierarchy of components
of research in the last few decades [9,10]. In particular, Beltran and that exist in a rope, and layers refer to the number of concentric
Williamson [11,12] have developed a numerical model to simulate circles of components wrapped around the core. For example,
synthetic-fiber rope response under axisymmetric loading condi- Fig. 1 shows the cross-section of a five-level, two-layer rope, along
tions. This computational model, which neglects the effect of rope with the parameters needed to characterize the geometry of any
self-weight, builds on previous research conducted on steel wire individual rope component (helix radius (a), pitch distance (p),
ropes and synthetic-fiber ropes. Full details of the derivations are and helix angle (θ )). The structures considered in Fig. 1 are: fibers
given elsewhere ([13–15]; among others), but the main results are (j = 5), rope yarn (j = 4), strand (j = 3), sub-rope (j = 2), and rope
summarized here for convenience. Nonlinear strain–displacement (j = 1). Thus, the notation adopted for the current study indicates
relationships and potentially nonlinear material properties (see that level j includes all components in level j + 1, where the entire
Appendix for the case of PET rope components) of the components rope is defined as level one (j = 1). Furthermore, the core is always
that form a rope are considered in the model. The proposed model considered to be the first layer. In the context of Fig. 1, the six sub-
has the main feature that it accounts for the degradation of rope ropes around the core comprise the second layer of the rope.
properties, allowing for the possibility of complete rupture of one Rope components, in general, are subjected to radial and
or more components. Degradation of rope properties is captured circumferential normal contact forces as shown in Fig. 2. In this
by the computation of a damage index D that evolves during a figure, the gi+1,i and gi−1,i are the line inter-layer contact forces
rope’s strain history and can range from 0 to 1. A value of 0 cor- exerted on a rope component located in layer i by rope components
responds to the undamaged state of a rope component, and D = 1 of layer i + 1 and i − 1, respectively; pnc ,i is the line inter-component
indicates its complete rupture. For the particular case of PET ropes contact force exerted on a rope component located in layer i by the
under monotonically applied loads, it is assumed that the damage other components of the same layer i; and Xi is the is the radial line
index depends solely on the strain range experienced by the rope body force necessary to preserve the helical geometry of a rope
components, and static tension tests on small-scale PET ropes (de- component. For computational purposes, however, two types of
fined as ropes with overall diameters between 6 and 13 mm with cross-sectional arrangements of the components are considered:
corresponding maximum capacities that range between 25 kN to packing geometry and wedging geometry, which represent the
62 kN) [16] validate this assumption [11,17]. extreme cases of transverse deformation for the cross-sections of
The current paper provides a detailed description of the real ropes [15].
numerical procedure implemented to account for the presence of In packing geometry, it is assumed that the structures
broken rope components on overall rope response, which extends (i.e., fibers, rope yarn, etc.) used to form a particular rope
the capabilities of the original model. The potential confinement component are initially straight, circular in cross-section, and
effect of a rope jacket on rope components is also introduced transversely stiff, and a twist of a specified number of turns
in this paper, and it is modeled using thin-walled tube theory. is imparted to the central component. Contact between rope
1700 J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709

b
a

c d

Fig. 2. (a) and (b) pattern of inter-layer/inter-component contact forces in an axially loaded rope element, (c) inter-layer contact forces for packing geometry, and
(d) inter-component contact forces for wedging geometry.

components in the same level is assumed to be only in the to the point at which this component resumes carrying its propor-
radial direction (inter-layer contact forces) as shown in Fig. 2(c). tionate share of axial load.
With wedging geometry, the components in the same level are Previously, the authors [12] used the recovery length concept
allowed to deform transversely and change their shape into a to evaluate and quantify the effect of rope component failure on
wedge or truncated wedge, which is the shape that would develop overall small-scale PET rope response. Details of the numerical
for deformable components. Radial contact is assumed to be algorithm used to compute damaged rope response, however,
negligible in comparison to the circumferential contact and is were not provided in the earlier paper. Furthermore, that initial
ignored for computational purposes. However, there exists an study ignored any potential confinement effects associated with
inter-component type of contact between rope components in the the rope jacket, and only the normal contact forces between rope
same layer for ropes with wedging geometry (Fig. 2(d)). The angle components were assumed to give rise to the inter-layer/inter-
βi is the complement of the subtended angle ψi , which is defined component interactions that permit development of admissible
as π/ni , where ni is the number of rope components in layer i. recovery lengths. Observations made from recent experimental
A multi-level rope could be modeled with different arrangements studies on large-scale PET damaged ropes [18] of parallel sub-
of packing and wedging geometry at different levels, generating a rope constructions suggest, however, that rope jacket confinement
mixed rope geometry. can induce normal contact forces between rope components.
In this context, large-scale refers to ropes with diameters that
2.2. Damaged rope analysis vary between 32 mm and 166 mm with corresponding specified
undamaged breaking strengths (SBS) of 35 tonnes (343.2 kN) and
As previously stated, damage in ropes can be represented 700 tonnes (6864.6 kN), respectively. Details of a model that
through a degradation of the properties of individual rope considers the effects of rope jacket confinement on the overall
components, and it can also include the complete rupture of one or response of large-scale damaged PET ropes and the numerical
more components. The modeling of rope degradation as a function implementation of the solution procedure are provided in the
of loading history is presented elsewhere [11,17]. Hence, the focus subsequent sections of this paper.
of this section is on the development of a numerical algorithm to
account for the effects of broken rope components on overall rope
behavior (breaking strain, breaking stress, and rope stiffness). 2.3. Confinement effect of rope jacket
Several experimental [18–23] and theoretical [12,24–28] stud-
ies have shown that if friction develops along rope components, a In general, a typical rope jacket has a double braid construc-
broken rope component is capable of supporting its total share of tion [29] as shown in Fig. 3. To account for the effect of the rope
load within a relatively short distance of the rupture location. This jacket on the radial pressure exerted on rope components, it is pro-
length is called the recovery length [25] and is defined as the dis- posed that the rope jacket can be modeled as a thin-walled tube
tance measured from the failure end of a broken rope component for the purposes of computing overall rope response. Raoof and
J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709 1701

is solid, by
2H
q= (1)
d
where d is the rope diameter, and the stress q is related to the
longitudinal stress in the rope jacket σL by the following expression
(Fig. 4(d)):
4σL t
q= (2)
d
Fig. 3. Rope jacket and sub-ropes of a parallel rope construction.
where t is the thickness of the rope jacket.
The stress q would be the radial stress if the rope’s cross-section
Huang [26] and Gjelsvik [30] studied the problem of determin- were solid and without voids between the rope components. If
ing the recovery length values of a fractured wire in a parallel- there are voids between rope components (packing geometry),
wire strand that was under external hydrostatic pressure exerted Gjelsvik [30] suggested a procedure to compute the compressive
by a prestressed wrapping or by intermittent cable bands. In the contact forces g2,1 exerted on the core based on an equivalent
current study, the works previously mentioned are extended and cross-section that makes the rope solid, (i.e., without the voids
then used to estimate the contact forces exerted by a rope jacket between rope components (Fig. 4(e))). The equivalent cross-
on rope components for the case of synthetic-fiber ropes, where section is obtained by intersecting the tangent lines at every
transverse deformation of the cross-section may occur according contact point between the core and the components wrapped
to the rope cross-sectional modeling techniques described previ- around it. For the rope cross-section shown in Fig. 4(a), the
ously (i.e., packing and wedging geometry). equivalent cross-section is a hexagon that circumscribes the core
For modeling purposes, a rope jacket can be assumed to produce cross-section (Fig. 4(b)). Thus, considering equilibrium in the
a uniform transverse compression in a rope. To illustrate the vertical direction (Fig. 4(e)), the compressive contact forces g2,1 are
procedure established to estimate this effect, consider first a one- given by
level, four-layer rope assuming inter-layer contact between the  
2 dc
rope components (i.e., packing geometry) (Fig. 4(a)). In the first g2,1 = H (3)
layer of this rope, there is one component (core), and in the
3 cos γ1
sin γi

d
second, third, and fourth layers, there are six, twelve and eighteen i=1
components, respectively, wound around the core (Fig. 4(a)). The
where dc is the diameter of the core and the angle γi is given in
compression forces exerted by the rope jacket on rope components
Fig. 4(e). The procedure just described can be used to obtain the
of the fourth layer are transmitted radially from components of the value of the compressive contact forces g2,1 for a different number
fourth layer to components of the first layer (core). Hence, the core of rope components wrapped around the core.
is subjected to six symmetrically distributed compressive contact LeClair [31] presents a study in which the contact geometry
forces g2,1 per unit length (Fig. 4(b)). between helical layers is established and in which equilibrium is
It is assumed that the rope jacket develops longitudinal and enforced in the radial direction at contact points. In this model,
circumferential stresses according to the assumptions of thin- the radial force is transmitted from the outer layers to the inner
walled tube behavior (i.e., a thin-walled cylindrical pressure layers, accounting for the distance between points of contact along
vessel). If H is the circumferential line force in the rope jacket the length of a helical rope component. This distance between
acting at a given cross-section (Fig. 4(c)), the nominal compressive contact points depends upon both the helix angle and the radius
radial stress q can be computed, assuming the rope cross-section of components in the layers under consideration. Thus, for i ≥

a b c

e
d

Fig. 4. (a) Rope cross-section; (b) Normal forces acting on the core; (c) and (d) circumferential force and longitudinal stress acting in the rope jacket with thin-walled tube
behavior, respectively; (e) compressive stress acting on the core.
1702 J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709

2, expressions for the contact line forces (gi−1,i (see Fig. 2(c)))
transmitted from the outer layers to the inner layers can be
estimated as
 
j−1
nj pi cos αi cos αk
−ml ∏
gi−1,i = Xi + Xj (4)
n p cos αj
j=i+1 i j k=i+1
cos αk
where ml is the number of layers in the rope and nl , pl and Xl are
the number of rope components, the pitch distance (which can be
used to determine the helix angle), and the radial line body force
of a rope component in layer l, respectively. The angles αl and αl Fig. 5. Two-noded axial-torsional element.
are defined as tan−1 (pl /2π (al − rl )) and tan−1 (pl /2π (al + rl )),
respectively, where al is the helix radius and rl is the radius of a
a
rope component in layer l [32].
Eq. (4) can be used to evaluate the contact line force g1,2 due
solely to the compression forces, grj,ml , exerted by the rope jacket
(rj) on the rope components of the outermost layer (ml) of a rope.
Setting the value of Xml equal to grj,ml and eliminating the values
of Xi for i = 2, . . . , ml − 1, an expression for g1,2 is obtained.
Rope components that belong to layer two are wound around a b
central straight core; thus, the inter-layer contact between layers
one and two describes a helical path. As such, equilibrium in the
radial direction requires that the contact line forces g1,2 and g2,1 are
related by g1,2 = g2,1 cos(θ2 ), where θ2 is the helix angle associated
with the rope components in the second layer. Therefore, the
contact line force g2,1 can be estimated in terms of the compressive
contact force grj,ml as follows:
ml−1 
∏ cos αk Fig. 6. Damaged rope discretizations: (a) failure located at one end of the rope and
(nml p2 cos α2 ) cos αk (b) failure located around the midspan of the rope.
k=3
g2,1 = grj,ml . (5)
cos θ2 n2 pml cos αml length values for a general multi-level rope with packing, wedging,
Hence, or mixed geometry, considering a general nonlinear slip-stick
friction model, are not presented in this paper due to space
cos θ2 n2 pml cos αml limitations; readers are encouraged, however, to consult the work
grj,ml =  g2,1 . (6)
by Beltran and Williamson [12] for a detailed description of these
ml−1
∏ cos αk
(nml p2 cos α2 ) cos αk equations. To simplify the analysis, the gradual increase in load
k=3
carried by broken rope components is ignored. Thus, only the full
For the case of a rope cross-section modeled using wedging geom- value of axial load at distances outside admissible recovery length
etry, by assumption, there are no voids between rope components values are taken into consideration. This assumption is based
and there is no radial contact between rope components of con- on parametric studies in which the gradual (nonlinear) increase
tiguous layers. The circumferential line force H in the rope jacket in tension of a broken rope component for a particular rope
is equilibrated by the rope components located in the outermost construction was computed, considering different positions within
layer of the rope. Thus, using the same notation as before, the com- the rope cross-section and different failure axial strain values of
pressive contact force grj,ml exerted by the rope jacket on every rope the broken rope components. These studies, which assume that
component of the outermost layer (ml) can be estimated as frictional forces are linearly dependent on the normal forces (i.e., a
2H π Coloumb friction model), show that most of the tension load is
grj,ml = (7) accumulated toward the end of the recovery length interval [0, rl]
nml (nonlinear profile) or the recovery length values are relatively
where nml is the number of components in the outermost layer short [33].
of the rope being analyzed. Once the compressive contact force The recovery length of an individual broken rope component
grj,ml exerted by the rope jacket on every rope component of the is computed for a prescribed deformation level of the entire
outermost layer (ml) is obtained, the radial force transmitted from rope (ε, ζ ), where ε is the rope axial strain and ζ is the rate of
the outer layers to the inner layers and/or the contact force exerted twist. Associated with the rope axial strain ε is the rope axial
on components that belong to lower levels can be estimated by displacement u, and associated with the rate of twist of the
considering equilibrium in the radial direction as explained by rope is the rope axial rotation φ . If the recovery length value
Beltran and Williamson [12]. is admissible (i.e., within the physical length constraints of the
rope being analyzed), the problem is linearized by discretizing
2.4. Numerical model of damaged rope response the rope along its length into axial-torsional elements (Si ), with
element lengths LSi , based on the value of the recovery lengths
The proposed numerical model developed to estimate the effect computed. Each element includes two degrees of freedom at each
of broken rope components on overall rope response considers, as node: axial displacement (ui ) and axial rotation (φi ). Associated
an initial step, whether or not broken rope components can take up with these degrees of freedom are the corresponding axial force
their appropriate share of axial load by considering the equilibrium Ti and torsional moment Mi (Fig. 5).
equations that govern rope behavior and by computing their In Fig. 6, two possible rope discretizations are presented along
recovery lengths. The equations that allow computing recovery with the corresponding axial-torsional element cross-sections,
J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709 1703

∂ G2 M (λ + δλ, η) − M (λ − δλ, η)
[ ]
where the broken rope components have been identified with an
kηλ = = (10b)
‘‘X’’. Also shown in the figure are the incremental changes in the ∂λ 2δλ
degrees of freedom and nodal forces for each lengthwise element.
∂ G1 T (λ, η + δη) − T (λ, η − δη)
[ ]
The two possible discretizations are based upon the location of the kλη = = (10c)
failed rope component(s): (a) the failure region is located near one ∂η 2δη
end of the rope (i.e., at a splice location), or (b) the failure region
∂ G2 M (λ, η + δη) − M (λ, η − δη)
[ ]
is located near the rope midspan. In both possible discretizations, kηη = = (10d)
the length LS2 represents the damage length; for case (a) its value ∂η 2δη
is equal to rl and for case (b) its value is equal to 2 · rl. The proposed where the centered differentiation formula is used to evaluate the
model captures the effect of having a weakened cross-section stiffness coefficients, T and M are the axial force and torsional
acting over a localized region characterized by the recovery length. moment, respectively, and the values of the perturbations δλ
Based on equilibrium considerations, the response of an individual and δη are found using an iterative process where variations are
axial-torsional element can be described in terms of only two considered until two consecutive values of the stiffness coefficients
independent degrees of freedom (referred to as the ‘‘essential set’’). are sufficiently close based upon a prescribed tolerance. The values
Hence, based on the degrees of freedom defined for the element of T and M are computed using the formulation described in [11].
shown in Fig. 5, only the axial deformation (u3 − u1 ) and the axial The 4 × 4 tangent stiffness matrix [Ksr ]i of each axial-torsional
rotational deformation (φ4 − φ2 ) are needed. The element forces element Si , associated with the degrees of freedom u1 , φ2 , u3 and
associated with the essential set are taken as T3 and M4 . φ4 (referred to as the ‘‘complete set’’) described in Fig. 5, can be
The tangent stiffness matrix of each axial-torsional element Si obtained through consideration of equilibrium. Hence, with the
is computed, and then each tangent stiffness matrix is assembled 2 × 2 element stiffness matrix [ksr ]i computed, [Ksr ]i can be
to obtain the tangent stiffness matrix of the entire rope. For each calculated as follows:
axial-torsional element, a 2 × 2 stiffness matrix is computed having [ ]
[ksr ]i −[ksr ]i
the following form: [Ksr ]i = . (11)
−[ksr ]i [ksr ]i
[ ]
kλλ kλη A displacement control analysis is carried out to obtain overall
[ksr ]i = (8)
kηλ kηη damaged rope response. Increments in rope axial strain, dεj , and
i

where the subscripts λ and η are associated with the axial rope axial rotation per unit length, dζj , are specified for the entire
strain and the rate of twist, respectively, of the cross-section rope in the jth incremental step of the analysis. The aim of the
of element Si . No closed-form expressions exist to compute the analysis is then to obtain the values of the increments of axial
displacements (dui )j and axial rotations (dφi )j of the individual
stiffness coefficients due to the nonlinear geometry and nonlinear
axial-torsional rope elements, as well as the increments in the
constitutive response of the rope components ([11] and Appendix).
external loads (axial forces and torsional moments), required to
Accordingly, a numerical procedure is needed to compute these
produce the specified increment in the deformation level (dεj , dζj )
terms. Thus, if there exists a function Gi = (G1 , G2 )i such that
of the rope. These quantities are obtained by solving the following
(Ti , Mi ) = Gi (λ, η)i , the stiffness coefficients can be obtained
linear system of equations:
by computing the gradient of the function Gi . Accordingly, the
stiffness matrix (Eq. (8)) of each axial-torsional element has the [Kr ]j−1 {du}j = {dP }j (12)
following form:
where [Kr ]j−1 is the rope tangent stiffness matrix that is computed
 ∂G ∂ G1  by considering the rope configuration of the (j − 1)th incremental
1
step, which is obtained by assembling the tangent stiffness
[ksr ]i =  ∂∂λ ∂η 
∂ G2  . (9)

G2 matrices of each axial-torsional element computed from Eq. (11).
∂λ ∂η The value of [Kr ]j−1 is used to compute the rope configuration
i
at the jth incremental step because at the end of the (j −
As stated before, the function Gi does not admit closed-form 1)th incremental step the rope has reached an equilibrium
expressions for any element Si . Thus, the partial derivates of configuration so that all the stiffness coefficients are known; {du}j
Eq. (9) are evaluated using a numerical scheme by inducing small is the increment in axial displacements and axial rotations, and
perturbations (δλ, δη)i to a prescribed level of deformation (λ, η)i {dP }j is the increment in axial forces and torsional moments
for each axial-torsional element Si . These induced perturbations in the jth incremental step. Eq. (12) represents the linearized
are applied separately (i.e., small perturbations are assumed for solution of damaged rope behavior for each increment in the
one variable while holding the other perturbation equal to zero), deformation level of the rope being analyzed. Because of the
generating two different perturbed deformations. Because the nonlinearity of the system properties, unbalanced forces are
computational model considers small increments in axial strain generated due to linearization; hence, an incremental-iterative
and axial rotation per unit length (δλ and δη, respectively), numerical procedure is used to compute the overall damaged rope
the standard engineering strain definition is used for each response. Therefore, Eq. (12) can be recast in the following form:
element Si . Using the damage-dependant constitutive equations
corresponding to the current level of rope deformation (Appendix), [Kr ]jk−1 {du}kj = {dQ }kj + {R }kj −1 (13)
the increments in axial force and twisting moment on the
where the superscript k represents the iterative step, [Kr ]kj −1
axial-torsional elements due to axial deformations and twisting
represents the tangent stiffness at the end of the (k − 1)th iteration,
deformations are then computed. These computed forces and
{dQ }kj is the increment in the external load vector of the kth
moments, along with the forces and moments associated with the
unperturbed deformation level (ε, ζ )i , are used to compute the iteration, and {R }kj −1 is the unbalanced load vector (axial forces
stiffness coefficients. Accordingly, the stiffness coefficients of the and torsional moments) that represents the imbalance between
matrix [ksr ]i can be evaluated, for any element Si , as follows: the existing internal and external loads at the end of the (k − 1)th
iteration.
∂ G1 T (λ + δλ, η) − T (λ − δλ, η)
[ ]
To illustrate the solution and the form of the components in
kλλ = = (10a)
∂λ 2δλ Eq. (13), consider a rope in which two of its seven components have
1704 J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709

failed around the rope midspan and their recovery length values In addition, the values of {δ(dQ3 ), δ(dQ4 ), δ(dQ5 ), δ(dQ6 )}kj also
are admissible. This damaged rope is discretized into three axial- vanish in Eq. (15) for k ≥ 2 due to equilibrium considerations.
torsional elements as shown in Fig. 6(b). For this case, the system Hence, the variation in the increments in axial displacements and
of equations presented in Eq. (13) has the following form axial rotations, {δ(du3 ), δ(dφ4 ), δ(du5 ), δ(dφ6 )}kj , are obtained by
solving the following system of equations:
 k−1  du1 k dQ1 k R1 k−1
k11 ··· k18
δ(du3 ) k
 k−1
dφ 2 
 0  
dQ2  R2  k33 ··· k36 R3
 .. .. ..  
 . . .   ..  =  ..  +  ..  . (14)  .. .. ..  δ(dφ4 ) = R4  .
     
. . .  . . .   δ(du5 )  R  (16)
k81 ··· k88 j 5
dφ 8 j dQ8 j R8 j k63 ··· k66 j δ(dφ6 ) R6 j
j

For this example, the unknown increments in displacements, The variation in the increments of the external axial loads
rotations, axial loads, and torsional moments for each incremental and the torsional moments {δ(dQ1 ), δ(dQ2 ), δ(dQ7 ), δ(dQ8 )}kj are
step of the analysis of Eq. (13) are du3 , dφ4 , du5 , dφ6 and computed based on the first, second, seventh and eighth equations
dQ1 , dQ2 , dQ7 , dQ8 , respectively. The quantities du1 = −du7 and of the system of equations represented by Eq. (15). In general, the
dφ2 = −dφ8 are known because, as stated before, a displacement- total displacements, rotations, external axial loads, and torsional
based analysis is performed, and dQ3 , dQ4 , dQ5 , and dQ6 vanish due moments in the jth incremental step of the analysis can be updated
to equilibrium considerations. In this particular application, the as follows:
tangent stiffness [Kr ]jk−1 is kept constant during iterations, and it is mi
updated at the end of each increment. For the first incremental step −
{U }kj = {U }j−1 + {du}1j + {δ(du)}kj (17a)
(j = 1) of the analysis, the tangent stiffness [Kr ]j=1 is computed for
k=2
a small level of deformation (λ, η)i , on the order of 1E-4 for each
mi
element Si . −
At the beginning of each incremental step (i.e., k = 1), the {Q }kj = {Q }j−1 + {dQ }1j + {δ(dQ )}kj (17b)
system is in equilibrium. As such, the unbalanced load vector {R }0j k=2

vanishes and, as stated before, the increments in displacement where {δ(du)}kj is the vector of the variation of the increments
and rotation of the entire rope are prescribed. At the discretized in axial displacements and axial rotations for the kth iteration,
level of the axial-torsional elements, the unknown increments {δ(dQ )}kj is the vector of the variation of the increments in external
in displacements and rotations {du3 , dφ4 , du5 , dφ6 }1j are first axial loads and torsional moments for the kth iteration, and mi is
computed using the third through sixth equations of the system the total number of iterations performed during the jth analysis
of equations represented by Eq. (14) using a Gaussian elimination step.
scheme. Having computed these quantities, the increments in Once the total axial displacements and axial rotations {U }kj
external axial loads and torsional moments {dQ1 , dQ2 , dQ7 , dQ8 }1j are computed, axial strain {λ}kj and rate of twist {η}kj for
are computed based on the first, second, seventh and eighth each axial-torsional element Si are computed, as before, using
equations of the system of equations given by Eq. (14). The the standard engineering strain definition. Likewise, using the
total displacements, rotations, external axial loads, and torsional damage-dependant constitutive equations corresponding to the
moments can be updated as {U }1j = {U }j−1 + {du}1j and {Q }1j = current level of rope deformation (Appendix), the internal axial
{Q }j−1 + {dQ }1j where {U }1j and {Q }1j are the vectors of the current force {T }kj and twisting moment {M }kj on each axial-torsional
axial displacements and axial rotations and external axial loads and element due to axial deformations and twisting deformations are
torsional moments, respectively, of the rope; {U }j−1 and {Q }j−1 computed. Then, the unbalanced load vector {R }kj is computed in
are the vectors of total axial displacements and axial rotations and a similar manner as explained in the case of {R }1j . If the norm of
total external axial loads and torsional moments, respectively, at
the end of the (j − 1)th increment.
{R }kj is less than a prescribed tolerance, the iteration process is
stopped and a new analysis is performed for the (j + 1)th increment
Axial strain {λ}1j and rate of twist {η}1j for each element Si are
in axial deformation (dε(j+1) ) and axial rotation (dζ(j+1) ) of the
computed using the standard engineering strain definition. Using
rope. Otherwise, new iterations are performed until equilibrium
the damage-dependant constitutive equations corresponding to
is achieved.
the current level of rope deformation (Appendix), the internal
axial force {T }1j and twisting moment {M }1j on each axial-torsional
3. Numerical implementation
element due to axial and twisting strains are computed. Then, the
unbalanced load vector {R }1j is computed as the difference between
In this section, two example ropes are analyzed using the
the internal and the external loads. If the norm of the unbalanced
methodology described in the previous sections. Based on the
load vector {R }1j (i.e.,‖Rj1 ‖) is greater than a prescribed tolerance,
model formulation, some of the variables that may affect damaged
the rope is not in equilibrium, and iterations on the variables
rope response are the number of broken rope components and
du3 , dφ4 , du5 , dφ6 , dQ1 , dQ2 , dQ7 , and dQ8 are needed. For k ≥ 2,
their distribution around a rope cross-section, the L/d (rope
the following system of equations is solved:
length/rope diameter) ratio, the location of a damaged cross-
 0  δ(du1 ) k R1 k−1 δ(dQ1 )k section relative to the rope length, the potential confinement
k11 ··· k18 exerted by a rope jacket, and the rope cross-sectional modeling
 .. δ(dφ2 ) R2  δ(dQ2 )
.. ..   .  = .  + .  technique (i.e., packing or wedging geometry) [33]. Due to space
 . . .  .  .  .  (15)
limitations and availability of experimental data [18], only the L/d
. . .

k81 ··· k88 j
δ(dφ8 ) j R8 j δ(dQ8 ) j ratio and the number and distribution of broken rope components
around a rope cross-section are evaluated in the current paper
where the operator δ() is associated with the variations of to determine their influence on the capacity, stiffness, and failure
the increments in axial displacements, axial rotations, external strain of a given damaged PET rope. The L/d ratios selected in
axial loads and torsional moments. As stated previously, the the testing program reported by Ward, et al. [18] vary between
values of {du1 , dφ2 , du7 , dφ8 }j are prescribed; thus, the values of 40 and 1000 and consider an initial damage state in which
{δ(du1 ), δ(dφ2 ), δ(du7 ), δ(dφ8 )}kj vanish in Eq. (15) for k ≥ 2. 10% of the cross-section is cut to simulate what may happen
J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709 1705

during installation and handling. While the data reported by Ward a b


et al. [18] serve as a basis for the analysis results presented
below, the primary intent of including this information is to
illustrate the implementation of the proposed numerical modeling
technique and to demonstrate how the model can be used to
evaluate damaged rope performance. As such, the numerical
results consider a broader variation in the L/d ratio (from 40 to
3000) and initial damage state (from 3% to 20%) than was reported
in the experimental testing program.
The numerical simulations and comparisons with experimental
data presented in this study consider two different types of
rope construction, with corresponding damaged PET rope cross-
sections shown in Fig. 7. These ropes were selected to correspond
to two of the specimens evaluated in a recent experimental testing
program [18] on large-scale, polyester-jacketed ropes (DPJR) with Fig. 7. Damaged (10%) rope cross-section (a) DPJR1 and (b) DPJR2.
a specified breaking strength (SBS) of 35 Tonnes (343 kN). Both
ropes have a parallel sub-rope construction, which means that the capacity, is comprised of the same PET material that is used to
sub-ropes are not twisted together in a helical fashion. Each sub- form the components within the rope, and has a thickness that
rope itself, however, is built by twisting together a certain number is approximately equal to 2% of the rope diameter. It is also
of strands. For analysis purposes, and to simulate the testing assumed that the rope jacket is completely bound to the rope
program previously mentioned, damage is inflicted to each rope at (i.e., there is no slip between the rope components and the jacket).
the midspan location and at the strand level prior to loading. This Ayres [35] established that, for undamaged ropes, the jacket can
damage is simulated by cutting (i.e., specifying a damage index D = account for up to 8% of the theoretical rope strength. Because the
1 (see Appendix)) a certain number of strands to reduce the cross- actual value is not known and has not been measured, a small
sectional area by a prescribed percentage. All damage is inflicted to value, relative to the overall rope capacity, is assumed for the
adjacent sub-ropes near the exterior of the rope to simulate surface rope jacket capacity, which is an assumption that is consistent
damage. For modeling purposes, the rope is analyzed as a two- with the research reported by Ayres [35]. Furthermore, results
level rope: the first level of the rope (sub-rope) is treated as having from parametric studies using different values for the assumed
a packing geometry, and the second level of the rope (strand) axial capacity of the jacket show small differences in the overall
is modeled using wedging geometry if no initial damage exists computed response of a rope [33]. Only for those cases in which
and packing geometry otherwise. This assumption is based on the jacket strength is ignored or is selected to be an unrealistically
results from simulations of small-scale PET damaged rope behavior large value do the results deviate from those presented below.
presented by Beltran and Williamson [17] and due to the fact Recovery length values are computed assuming a Coulomb
that if damage is present at the strand level, undamaged strands friction model with a friction coefficient equal to 0.1 ([36,37]) and
can migrate to a new a position because of the space available a small level of rope deformation (10% of the specified failure
from voids between undamaged strands (i.e., the packing geometry axial strain). Based on these selected parameters, the recovery
assumption). For each rope analyzed, the complete hierarchical length of the broken rope components that belong to the rope-
structure is given. Thus, the response of each strand is based on cross section depicted in Fig. 7(a) (DPJR1) is equal to 820 mm,
the constitutive properties of the rope yarns that comprise the and the recovery length of the broken rope components that
strand. This information is available from test results that give belong to the rope-cross section depicted in Fig. 7(b) (DPJR2) is
tensile load as a function of the axial strain, where the breaking equal to 615 mm. These results indicate that damage propagates
axial strain is 13% and the breaking axial load is specified for each over a longer length in DPJR1 than in DPJR2. This behavior is
rope construction [34]. due to the fact that the strands of DPJR1 are stiffer than the
According to work by Beltran and Williamson [12], which strands of DPJR2 (Fig. 7), which induces a higher axial load level
ignores the presence of a rope jacket, damaged components in the strands of DPJR1 relative to DPJR2 for the same deformation
within ropes that have a parallel sub-rope construction do not level (10%). This effect is not compensated by the fact that the
develop contact forces as a result of rope axial deformations; pressure exerted by the rope jacket of DPJR1 is distributed over
hence, no frictional forces exist between components. Under such fewer components than DPJR2 (i.e., the confining pressure per
conditions, broken rope components do not develop admissible component for DPJR1 is greater than DPJR2) and the rope jacket
recovery length values, and damaged rope response is computed of DPJR1 is thicker than the rope jacket of DPJR2, thereby inducing
ignoring the contribution of broken rope components. As stated higher pressure on its rope components. Despite these mitigating
previously (Section 2.2), however, Ward et al. [18] suggest that factors, the recovery length in DPJR1 is longer than the recovery
rope jacket confinement can induce normal contact forces between length in DPJR2. Consequently, for the case of DPJR1 with an L/d
rope components. Thus, broken rope components can eventually ratio equal to 40, the recovery length on each side of the centrally
develop admissible recovery length values and start carrying load, located failure location is greater than the total rope length, making
which indicates that they can contribute to overall rope response. the recovery length value inadmissible. In this case, damaged
To study this potential non-uniform strain distribution along the rope response is computed by using a net area effect model. This
length of a damaged rope, the confinement exerted on a rope by model ignores the contribution of broken rope components on
the jacket is considered using the thin-walled tube approximation overall rope response after they fail. Accordingly, the stiffness
described in Section 2.3. As such, a damaged rope analysis is carried and strength loss of a rope are proportional to the area that the
out by discretizing the length of the rope into axial-torsional broken rope components represent with respect to the total rope
elements, where the lengths and the number of rope components cross-sectional area. Furthermore, with the net area effect model,
for each axial-torsional element are based on the value of the the rope breaking strain remains unchanged with respect to the
recovery lengths computed. initially undamaged rope failure strain [11,16].
For each rope analyzed, it is assumed that the rope jacket For damaged ropes with an L/d ratio equal to 290 and 1000,
has an axial capacity that is 3% of the initially undamaged rope damage lengths of the ropes are approximately 15% and 4% of the
1706 J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709

corresponding overall rope length, respectively, which indicates a


that these recovery lengths are admissible. For the case of DPJR
2 and the selected values of the L/d ratio (40, 290, and 1000), all
recovery length values are admissible and correspond to damage
lengths of approximately 96%, 13%, and 3.5%, respectively, of the
corresponding overall rope length.
Fig. 8(a) shows the predicted response of DPJR1 as a function
of the L/d ratio for an assumed level of initial damage equal
to 10% of the total cross-sectional area. In this study, consistent
with the experimental data [18], the L/d ratio changes by varying
the rope length L and holding the rope diameter d fixed. For
reference, the predicted response of the initially undamaged
rope is also provided in the figure. All analyses account for the
degradation of rope properties due to axial deformation according b
to the damage-dependent constitutive model presented in the
Appendix. Numerical simulations show that as the value of the L/d
ratio increases, the failure axial strain decreases, damaged rope
response approaches the undamaged case, and the rope capacity
values appear to be unchanged. These results suggest that as rope
length increases, the impact of initial damage on overall rope
response becomes less important. In Fig. 8(b), the dependence
of the damaged rope response on the initial damage state for
an assumed value of L/d equal to 80 is evaluated. This value
of L/d was selected to clearly distinguish the response curves
of DPJR1 considering different initial damage states. For greater
L/d values, the aforementioned response curves converge to one
Fig. 8. Damaged DPJR1 response (a) for different values of L/d and 10% of initial
another as explained in the subsequent discussion. As the initial damage and (b) for different values of initial damage and a value of L/d = 80.
damage state increases, the damaged rope capacity decreases and
both the failure axial strain and the rope stiffness decrease. The
response trends identified in Fig. 8 can be extended to evaluate a
ropes with other combinations of L/d ratio and initial damage.
Thus, considering that the damaged rope stiffness approaches the
undamaged case as the L/d ratio increases (Fig. 8(a)) and that the
proposed model suggests (Fig. 8(a) and further discussion below)
that the damaged rope capacity does not depend strongly on the
L/d ratio, the analysis results presented in Fig. 8(b) would approach
those of the initially undamaged rope response curve if the value
of the L/d ratio were increased. However, rope failure strain would
be different for each curve considering its initial damage state—as
the initial damage state increases, rope failure strain gets smaller.
These factors are considered further in the discussion below.
In Fig. 9, the predicted rope capacities and failure strains of
DPJR1 for varying levels of initial damage are plotted as a function b
of the L/d ratio. The predicted rope capacities (Fig. 9(a)) do not
show a strong dependence on the L/d ratio for all the initial damage
states considered, which means that the capacity reduction in
damaged ropes depends primarily on the initial loss of cross-
sectional area. This behavior is a direct result of the assumptions
used to formulate the proposed analysis model. Based on these
assumptions and the fact that loss of symmetry due to rope
component failures is neglected, the damaged rope capacity is
controlled by the capacity of the weakened cross-section of the
rope that acts over a localized region represented by an individual
axial-torsional element as explained in Section 2.4. The predicted
failure axial strains for a given level of initial damage (Fig. 9(b)),
however, do show a dependence on the L/d ratio, with a decreasing Fig. 9. Variation of (a) failure axial load and (b) failure axial strain with respect to
rate of change as the L/d ratio increases. Eventually, for all cases L/d ratio for different initial rope damage states of rope DPJR1.
studied, the failure axial strains reach a minimum value for large
values of the L/d ratio, and this minimum value depends upon the axial-torsional elements if the value of its recovery length is
initial level of damage. In addition, the rate of convergence to this admissible, where the central element (Fig. 6(b)) is the localized
minimum value is faster for smaller values of initial damage. region where strain and damage concentrate along the rope
In evaluating the computed results associated with DPJR1 length [12]. Considering that the value of the recovery length is
(Figs. 8 and 9), it is important to note that the recovery length constant for a given distribution of cross-sectional damage, as
is independent of the rope length and solely depends on the the L/d ratio gets larger, the percentage of the damaged length
rope cross-sectional model [12]. Based on the formulation of the with respect to the overall rope length gets smaller, and the sum
proposed model, rope DPJR1 is discretized into three two-noded of the lengths of the axial-torsional elements that consider the
J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709 1707

a damage state corresponding to 10% of the cross-sectional area of


each rope, which was the initial damage state value considered in
the testing program [18]. Not only are the values of the failure load
and failure strain compared, their dependence on the L/d ratio is
also provided. For each rope tested, only the values of the rope
breaking strain and rope breaking load were reported [18]. No
information was given about the gage length used to compute the
values of rope breaking strain obtained from the rope elongation
measurements.
Failure axial loads for both ropes analyzed (Fig. 10(a)) are
accurately predicted by the proposed numerical model (within a
range of −5% to +10% of the experimental values). Experimental
data indicate that for DPJR1, the failure axial load increases (about
7%) as the L/d ratio increases from 40 to 560, but it remains nearly
constant for values of L/dequal to 560 and 1000. For the case
b of DPJR2, the failure axial load also increases (about 5%) as the
L/d ratio increases from 40 to 290, but then it decreases (about
18%) as the value of L/dincreases from 290 to 1000. Because the
experimental results between DPJR1 and DPJR2 are not consistent
and do not show a clear trend in how their capacity depends
upon the L/d ratio, the computed results, which show limited
dependence on the L/d ratio, are considered to be reasonable.
As shown in Fig. 10(b), predicted values for the rope failure
axial strain overestimate the experimental values for the two cases
analyzed. Specifically, experimental values for the failure axial
strain are overestimated within a range of 30% to 40% for DPJR1
and 40% to 60% for DPJR2 relative to the experimental values over
the entire range of L/d ratios considered. In Fig. 10(b), as expected,
Fig. 10. Comparisons between experimental and predicted values of (a) failure DPJR2 reaches its minimum failure axial strain value for a smaller
axial load and (b) failure axial strain considering a 10% of initial damage state. value of rope length L than DPJR1 because the ratio of damaged
length to overall rope length is smaller for DPJR2 than it is for DPJR1
contribution of all rope components in the overall rope response as discussed previously. Though the predicted values of the failure
(LS1 and LS3 in Fig. 6(b)) approaches the actual rope length. axial strain overestimate the experimentally measured values,
Hence, the damaged rope stiffness approaches the stiffness of both the test data and simulation results show the same trend—
the initially undamaged rope and, as previously stated, the rope failure axial strain decreases as the L/d ratio increases. Despite the
failure strain converges to a minimum value. Accordingly, for ropes fact that the predicted results are significantly greater than the
with a given damage state in which the damage length is a small measured values, these estimates are better than values predicted
percentage of the rope length, the reduction of the failure axial by other models previously reported in the research literature
strain with respect to the undamaged case can be considered to be (e.g., net area effect model) that do not show any dependence on
its minimum value. Likewise, it should also be noted that overall the L/d ratio and do not account for damage. Furthermore, it is
damaged rope response shows a similar dependency on the axial unclear as to how the failure axial strains were measured during
capacity of the rope jacket, which determines the recovery length the testing program because, even for the initially undamaged case,
values. Thus, if jacket axial capacity is increased for a given number the measured values reported by Ward et al. [18] (approximately
of broken rope components, the recovery length becomes shorter, 8% for the undamaged case) are below the failure axial strain of
indicating that the percentage of damaged length compared to the the PET fibers (13%). Considering that the helical nature of the rope
overall rope length becomes smaller; hence, the same performance geometry does not have a significant impact on the distribution
trends as identified above are observed [33]. In addition, as the of rope component strains because both rope geometries have
initial loss of rope cross-sectional area increases for a given L/d parallel sub-rope constructions and the helix angle at the strand
ratio, the predicted failure axial strain decreases. This result is level is less than 10°, the rope failure axial strain is expected to
due to the fact that as the initial loss of rope cross-sectional area be larger than what was reported. Hence, the differences between
gets larger, the localized damage region of the rope defined by the the measured and predicted values may be attributable to several
recovery length also increases [12]. For example, based on Fig. 9(b), factors, and additional research is needed to identify the potential
the analyzed rope (DPJR1) with initial damage states of 3%, 10%, sources of error.
and 20% has a reduction in its failure axial strain by approximately
7%, 15%, and 24%, respectively, relative to the undamaged case 4. Conclusions
(rope failure strain equal to 0.13) for a value of the L/d ratio equal
to 3000. In this paper, a detailed description of a model for predicting
In Fig. 10, comparisons between experimental data and the effect of broken rope components on overall rope response
predicted values of the failure axial load and failure axial strain was presented. The proposed model depends on the ability of
are presented for DPJR1 and DPJR2. Both of these ropes (and all broken rope components to pick up their proportionate share
the ropes tested in the study reported by Ward [18]) were made of axial loads over a distance measured from the failure region,
of polyester fibers obtained from a single manufacturer. Thus, which is referred to as the recovery length. The model assumes
for computational purposes, the same constitutive function that axisymmetric response and ignores any loss of symmetry due to
describes the tensile load as a function of the axial strain of the the failure of any rope components. If the recovery length value
polyester textile yarns was used for all the ropes analyzed in this is admissible, a damaged rope is discretized along its length into
paper (see Appendix). The results shown in Fig. 10 are for an initial two-noded axial-torsional elements, which provides a means for
1708 J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709

representing a weakened cross-section acting over a localized assumed to be a polynomial function of the normal strain up to
region. The developed model takes into account the geometric the fifth degree and has the following form:
nonlinearity of a rope’s geometry and the potential nonlinearity 
εε 2 ε
    3 
of the constitutive response of the components that form a rope. In
σd (ε, ε , . . .) = (1 − D) a1
2
+ a3+ a2
addition, the potential confinement effect of a rope jacket on rope εb
εb εb
components is included in the proposed model by assuming the
 5 
ε ε
 4
rope jacket behaves like a thin-walled tube.
Parametric studies were performed to show how two selected + a4 + a5 σb (18)
εb εb
parameters, the initial loss of rope cross-sectional area and the
L/d ratio, influence the predicted values of axial load capacity and where ε is the axial strain experienced by the rope component,
failure axial strain of two PET ropes. Results showed that rope εb is the normal strain at which the rope component reaches its
capacity decreased as the initial loss of rope cross-sectional area maximum normal (σb ) stress under monotonic loading, and D is
increased, but it did not show a strong dependence on the L/d ratio a damage index that can vary from 0 to 1. The coefficients ai are
for all the initial damage states considered. The reduction of rope constitutive parameters that are chosen to provide a best fit to
capacity was mainly proportional to the loss of cross-sectional area measured data for rope components that belong to the lowest
due to rope component failures (net area effect model). Conversely, hierarchical level of a rope.
failure axial strain decreased and showed a decreasing rate of Under the assumption that the strain range experienced by rope
change as both the initial loss of rope cross-sectional area and the components is the main source of damage under static tension
L/d ratio increased. Damaged rope stiffness tended to converge to loading conditions, Beltran and Williamson [17] proposed that the
the initial undamaged rope stiffness as the value of the L/d ratio evolution of the damage index D can be computed as
became larger, and the rate of convergence increased as the initial
η−ηu
damage state of the rope decreased. D(η) = DI + α1 ηβ1 + (1 − α1 ηuβ1 − DI )e θ1 (19)
Experimental values obtained from static capacity tests on
different types of PET rope constructions were used to evaluate where the value of DI is the initial damage that a rope component
the proposed model. Predicted values of the failure axial load could have before being loaded, the coefficients α1 and β1 are
of an initially damaged rope showed good correlation with the constitutive parameters that can be obtained from experimental
experimental data for the ropes analyzed. Experimental values, data, θ1 is a constant much smaller than 1 (θ1 ≪ 1), and η is
however, did not show a clear trend in their dependency on the a dimensionless variable defined as (ε − εt )/εb , where εt is the
L/d ratio. For the two ropes analyzed in this study, the proposed threshold strain that must be exceeded before damage occurs.
numerical model overestimated the rope failure axial strains. The The values of the damage parameters assumed for the analyses
experimental and predicted values of the rope failure axial strain, presented in this paper are: α1 = 0.12, β1 = 0.84, εt = 0.04 and
however, showed a similar trend in their dependence on the L/d θ1 = 0.001, which come from the results of extensive parameter
ratio: as the value of the L/d ratio increased, the value of failure studies on small-scale PET ropes [17].
axial strain decreased.
The proposed model can be used to perform parametric studies References
to evaluate how other variables, such as location of the damaged
cross-section along the rope length, confinement exerted by the [1] Chaplin CR. Torsional failure of wire rope mooring line during installation in
deep water. Eng Fail Anal 1998;6:67–82.
presence of a rope jacket, and rope cross-sectional models (i.e.,
[2] Utting WS, Jones N. A survey of literature on the behaviour of wire ropes. Wire
packing or wedging geometry), influence the overall behavior Ind 1984;51(109):623–9.
of damaged ropes. The authors have validated the proposed [3] Foster GP. Advantages of fiber rope over wire rope. J Ind Text 2002;32:67–75.
model using experimental data obtained from large-scale tests [4] Cardou A, Jolicoeur C. Mechanical models of helical strands. Appl Mech Rev
1997;50(1):1–14.
on PET damaged ropes, and a forthcoming paper will describe [5] Ghoreishi S, Cartraud P, Davies P, Messager T. Analytical modeling of synthetic
these results. While the current model can be used to obtain a fiber ropes subjected to axial loads. Part I: a new continuum model for
reasonable estimate of rope behavior, improvements—including multilayered fibrous structures. Internat J Solids Structures 2007;44:2924–42.
[6] Liu F. Static response model of synthetic fiber lines with internal friction.
the consideration of the loss of symmetry due to the failure of Report TM 44-89-14. Port (Hueneme, CA): Naval Civil Engineering Laboratory;
individual rope components—are still needed before it can be 1989.
reliably used for detailed design calculations. Due to the generality [7] Liu F. Computer modeling of high-strength synthetic lines under static and
cyclical tensile loads. Technical report TR-2052-OCN. Port (Hueneme, CA):
of its formulation, the proposed model can be extended to consider Naval Facilities Engineering Services; 1995.
multiple failure regions along the length of a given rope, it [8] Lo KH, Xu H, Skogsberg LA. Polyester rope mooring design considerations.
can incorporate several axial-torsional elements throughout the In: International offshore and polar engineering conference. 1999.
[9] Karayaka M, Srinivasan S, Wang S. Advanced designed methodology for
recovery length considering the appropriate stiffness of the rope
synthetic mooring. In: Offshore technology conference. OTC 10912. 1999.
components, and it can represent damaged ropes comprised of [10] Seo MH, Backer S, Mandell JF. Modelling of synthetic fiber ropes deterioration
materials other than PET. caused by internal abrasion and tensile fatigue. Washington (DC, USA): MTS
Science and Technology for a New Oceans Decade; 1990.
[11] Beltran JF, Williamson EB. Degradation of rope properties under increasing
Appendix monotonic load. Ocean Eng 2005;32:823–44.
[12] Beltran JF, Williamson EB. Numerical simulation of damage localization in
polyester mooring ropes. J Eng Mech 2010;136:945–59.
Rope response is computed by summing up the contribution [13] Costello GA. Theory of wire rope. New York (USA): Springer-Verlag; 1990.
of each component that forms the rope that is being analyzed. [14] Leech CM. Theory and numerical methods for the modeling of synthetic ropes.
Based on experimental results [38], PET fibers experience damage Commun Appl Numer Methods 1987;3:407–713.
[15] Leech CM. The modeling of friction in polymer fibre ropes. Int J Mech Sci 2002;
when subjected to stretching. Beltran and Williamson [11]
44:621–43.
proposed a model to compute rope response considering the [16] Li D, Miyase A, Williams JG, Wang SS. Damage tolerance of synthetic-fiber
degradation of its mechanical properties. Thus, the damage- mooring ropes: small-scale experiments and analytical evaluation of damaged
dependent constitutive relationship that governs the normal stress subropes and elements. Technical report CEAC-TR-03-0101. University of
Houston; 2002.
for each rope component that belongs to the lowest hierarchical [17] Beltran JF, Williamson EB. Investigation of damage-dependent response of
structure of the rope cross-section modeled, σd (ε, ε 2 , . . .), is mooring ropes. J Eng Mech 2009;135:1237–47.
J.F. Beltrán, E.B. Williamson / Engineering Structures 33 (2011) 1698–1709 1709

[18] Ward EG, Ayres RR, Banfield S, O’Hear N, Smith CE. Experimental investigation [28] Raoof M, Kraincanic I. Determination of wire recovery length in steel cables
of damage tolerance of polyester ropes. In: Proceedings of the 4th international and its practical applications. Comput Struct 1998;68:445–59.
conference on composite materials for offshore operations. 2005. [29] Winkler MM, McKenna HA. The polyester rope taut leg mooring concept:
[19] Hankus J. Safety factor for hosting rope weakened by fatigue cracks in wires. a feasible means for reducing deepwater mooring cost and improving
In: Organisation internationale pour l’Etude de l’Endurance des cables. OIPEEC. stationkeeping performance. In: 27th Annual offshore technology conference.
Round table conference. 1981. 1995.
[20] Chaplin CR, Tantrum N. The influence of wire break distribution on strength. [30] Gjelsvik A. Development length for single wire in suspension bridge cable.
In: Organisation internationale pour l’Etude de l’Endurance des cables. OIPEEC. J Struct Eng 1991;117(4):1189–201.
Round table conference. 1985. [31] LeClair R. Axial response of multilayered strands with compliant layers. J Eng
[21] Cholewa W, Hansel J. The influence of the distribution of wire rope faults Mech 1991;117(12):2884–903.
on the actual breaking load. In: Organisation internationale pour l’Etude de [32] Phillips JW, Miller RE, Costello GA. Contact stresses in a straight cross-lay
l’Endurance des cables. OIPEEC. Round table conference. 1981. wire rope. In: Proc., 1st annual wire rope conf. Engineering extension service.
[22] Cholewa W. Wire fracture and weakening of wire ropes. Wire rope discard Washington State University. 1980.
criteria. Round table conference. Swiss federal institute of technology. [33] Beltran JF. Computational modeling of synthetic–fiber ropes. Ph.D. disserta-
ETH. Institute of lightweight structures and ropeways. September. Zurich tion. Austin (TX): The University of Texas at Austin; 2006.
(Switzerland); 1989. [34] Banfield S. Personal communication.
[23] Evans JJ, Ridge IML, Chaplin CR. Wire failures in ropes and their influence [35] Ayres R. Determining interim criteria for replacing damaged polyester
on local wire strain behaviour in tension–tension fatigue. J Strain Anal 2001; mooring rope. Final report to BP exploration. PN 6941-RRA. 2001.
36(2):231–44. [36] Leech CM, Hearle JWS, Overington MS, Banfield SJ. Modelling tension and
[24] Chien CH, Costello GA. Effective length of a fractured wire in wire rope. J Eng torque properties of fibre ropes and splices. In: Proceedings of the third
Mech 1985;111(7):952–61. international offshore and polar engineering conference. 1993.
[25] Raoof M. Wire recovery length in a helical strand under axial-fatigue loading. [37] Banfield SJ, Hearle JWS, Leech CM, Tebay R, Lawrence CA. Fibre rope modeller
Int J Fatigue 1991;13(2):127–32. (FRM): a cad program for the performance and prediction of advanced cords
[26] Raoof M, Huang YP. Wire recovery length in suspension bridge cable. J Struct and ropes under complex loading environments. Germany: Techtextil; 2001.
Eng 1992;118(12):3255–67. [38] Van den Heuvel CJM, Heuvel HM, Faasen WA, Veurink J, Lucas LJ. Molecular
[27] Raoof M, Kraincanic I. Recovery length in multilayered spiral strands. J Eng changes of PET yarns during stretching measured with rheo-optical infrared
Mech 1995;121(7):795–800. spectroscopy and other techniques. J Appl Polym Sci 1993;49:925–34.

Das könnte Ihnen auch gefallen