You are on page 1of 27

Article

Mathematics and Mechanics of Solids


1–27
Effect of friction on the mechanical Ó The Author(s) 2018
Article reuse guidelines:

behavior of wire rope with hierarchical sagepub.com/journals-permissions


DOI: 10.1177/1081286518816519
journals.sagepub.com/home/mms
helical structures

Dabiao Liu , Shimin Zheng and Yuming He


Department of Mechanics, Huazhong University of Science and Technology, Wuhan, China;
Hubei Key Laboratory of Engineering Structural Analysis and Safety Assessment, Wuhan,
China
Received 1 September 2018; accepted 8 November 2018

Abstract
The effect of friction on the mechanical behavior of wire rope with hierarchical helical structures under tensile and bend-
ing loads is studied. A typical 7 × 7 wire rope with an independent wire rope core is considered. It is assumed that fric-
tion only occurs between adjacent helical wires in the core strand, and between the core wire and the double-helix
wires in the outer strands. Based on Love’s thin rod theory, the mechanical responses of the rope under tensile and
bending loads are analyzed, as well as the contact force. The effects of the chirality of rope and the initial helix angle of
the wire on the distribution of the contact force are analyzed. It is shown that the effect of the chirality on the contact
force of wires and stresses in the rope is insignificant. The global stiffness of the rope can be enhanced by the friction
effect. The contact force of wires changes periodically with the winding angle of the centerline around the axis of the
strand. The current model provides an effective method to assess the local deformation and stresses of the wire rope
with consideration of the inter-wire friction.

Keywords
Wire rope, thin rod theory, friction effect, contact force, hierarchical helical structure

1. Introduction
Wire rope with hierarchical helical structures is widely used in various engineering fields due to its high
strength, exceptional softness and high stability [1]. In practice, wire rope is generally in a complex stress
state, for example, under the combined action of tension, torsion, bending or dynamic loads simultane-
ously. During deformation, the contact between inter wires inevitably leads to friction, which may lead
to wear of the rope. Thus, the study of the friction effect on the mechanical behavior of the wire rope
with hierarchical helical structure is of great significance for ensuring the safety of the rope in the speci-
fied conditions.
Many theoretical models, generally based on the elastic thin rod theory, have been developed for
describing the mechanical performance of the wire rope under tension, torsion and bending loads [2–11],
providing closed-form expressions for their constitutive behavior if the geometric and material para-
meters are given [12]. Among them, the fiber model is the simplest one in which the bending and torsion
rigidity of the wire are neglected. Hruska [13] first proposed a discrete model in which the wires were
assumed to be subjected to pure tensile forces. Also, Knapp [14] proposed a new stiffness matrix for

Corresponding author:
Dabiao Liu Department of Mechanics, Huazhong University of Science and Technology, Wuhan 430074, China.
Email: dbliu@hust.edu.cn or liudb@hust.edu.cn
2 Mathematics and Mechanics of Solids 00(0)

straight cables subjected to tension and torsion, in which the cross-section was considered as a single
composite element. In 1976, Costello and Phillips [15] treated the ropes as groups of separate curved rods
based on the curved rod theory [16]. They provided a rigorous derivation with consideration of both the
bending and torsion stiffness of the wires. The model was then extended by Kumar and Cochran [17] to
a linear and closed-form expression for the global stiffness of the rope. Jolicoeur and Cardou [18] com-
pared the mechanical response of a strand rope predicted by the thin rod model that was estimated by
Hruska’s fiber model [13]. Moreover, the theoretical predictions were also compared with the measure-
ments on a single strand [19]. They found that the fiber model is inappropriate if the twisting moment
and angle have to be considered, while the thin rod model gives a good prediction of the torsion stiffness
of the single strand. In addition to the thin rod models, a semi-continuous model called the orthotropic
sheet model has been developed by Hobbs and his workers [20,21]. This model has been used to predict
the mechanical behavior of cables. In this model, each layer of twisted wire was treated as an orthotropic
complete cylinder. Based on the same homogenization, Blouin and Cardou [22] developed another semi-
continuous model in which each wire layer was considered as a transversely isotropic continuous layer.
The model was later extended by Jolicoeur and Cardou [23] and by Cardou and Jolicoeur [2]. Semi-con-
tinuous models can be used to predict the mechanical behavior of wire rope subjected to torsion and
bending loads. Comparing semi-continuous models with thin rod models for wire rope, Raoof and
Kraincanic [3] found that the thin rod theory is more reliable for small-diameter wire strands, while the
semi-continuous models are much more appropriate for a rope with a large number of individual wires.
Interestingly, Karathanasopoulos and Angelikopoulos [24] recently proposed a quantitative methodol-
ogy to optimally design the structural arrangement of multi-layer helical assemblies. They found that
certain structural arrangements reduce the effective torsional moments developed in helically braided
structures, finally leading to an equilibrated macroscopic mechanical performance.
The aforementioned models mainly focus on rope consisting of single-helix wires. However, most
ropes in practice have a complex multi-strand or hierarchical structure in which most inter wires are
double helices [4]. Lee [25] proposed a mathematical model on the basis of the vector differential geome-
try for investigating the geometrical properties of single-helix and double-helix wires. Many differences
between single-helix and the double-helix wires have been indicated. Stanova et al. [26] developed a con-
cise mathematical model for describing single-helix and the double-helix configurations in the form of
parametric equations. With the strand geometric model, they performed a finite element (FE) simula-
tion on the multi-layered strand under tension tests [27], which gave results in good agreement with
experiments. Recently, Argatov [4] developed an asymptotic simulation method for evaluating the
mechanical behavior of a stranded rope subjected to tensile and torsional loads with considering the
effect of wire flattening. The double-helix configuration was fully analyzed by Elata et al. [5] for model-
ing the mechanical behavior of a wire rope with an independent wire rope core (IWRC) under axial
loads. Usabiaga and Pagalday [6] established another model, based on the work of Ashkenazi et al. [28]
and the general thin rod theory [16], for double-helix cables without taking Poisson’s effect into consid-
eration. In their model, the wires were assumed to be un-lubricated to avoid the relative movement
between adjacent wires. However, instead of attaching the material points of each wire section to the
centerline of the rope [28], Usabiaga and Pagalday [6] attached those points to the centerline of the
wound strand. More recently, Zhao et al. [7] proposed a bottom-up theoretical model to study the
mechanical performance of ropes with hierarchical helical structures. The analytical solutions of wire
ropes subjected to axial tension, torsion and bending loads were obtained. Based on the thin rod theory
of Love [16] and the frictionless assumption, Xiang et al. [8] established a new model to simulate the
mechanical response of the double-helix wire rope (IWRC) under tensile and torsional loads. Both the
global stiffness of the rope and the local deformation and stresses of individual wires were predicted by
the model.
Most of the above models neglect the effect of friction on the mechanical behavior of wire ropes. Le
Clair and Costello [29] studied the axial response of a strand under bending and torsion loads with con-
sidering the friction effect. They treated the inter-wire friction, governed by the equilibrium of the outer
wire, as an internal action in the strand. In this way, both the sliding friction force and the static friction
force in the strand were obtained. Costello [1] analyzed the effect of friction on the mechanical proper-
ties of a simple strand subjected to axial and bending loads. In Costello’s model, the contact
Liu et al. 3

deformation between wires was neglected; only the friction between the core wire and the outer wires
was considered. Papailiou [30] considered the interlayer friction and the interlayer slip of the wires in
the cables under simultaneous tensile and bending loads. The author found that the bending stiffness
changes with the bending displacement and the tension applied on the rope. However, Papailiou’s geo-
metrical model is a simple strand consisting of a straight core wire surrounded by helical wires in a sin-
gle layer. Later, to describe the bending behavior of helically wrapped cables under tension, Hong et al.
[11] extended Papailiou’s analysis to a strand with a core wire and multi-layer single-helix wires. Hong
et al.’s model accounts for the nonlinear dissipative behavior of the cable resulting from the slippage of
wires under friction forces. The authors found that the bending stiffness decreases rapidly with increas-
ing the curvature of the cable. Inagaki et al. [31] also made an extension of Papailiou’s model to study
the mechanical response of cables with hierarchical helical structures subjected to a bending load. They
found that the internal stress and the number of slipping wires play important roles in the service life of
the cable. Gnanavel et al. [32] studied the effect of friction on the response of a simple straight strand
subjected to tension and torsion loads under three conditions, i.e. the core–wire contact, the wire–wire
contact in the layer without contacting the core and the coupled contact among the core and all the
wires, respectively. A FE model was also developed to analyze statically indeterminate contacts in a sim-
ple strand under an axial load [33]. Using some bending symmetric boundary conditions, Jiang [34]
developed a concise FE method to calculate the deformation and stress of a simple strand under pure
bending with consideration of the friction effect. The case of coupled core–wire and wire–wire contact
leads to a statically indeterminate contact problem requiring the consideration of local contact deforma-
tion. Recently, Karathanasopoulos et al. [35] established a planar elastoplastic FE model for the simula-
tion of the mechanical response of helical structures to traction loads. Then, they highlighted the effect
of hardening, contact and loading history on the mechanical response of helical structures. More FE
analyses on the wire rope accounting for the friction effect have been conducted by many other authors
[36–38], in which the simulation considering the inter-wire friction agrees better with measurements,
compared with the frictionless case.
Recently, Chen et al. [39–41] and Meng et al. [9] developed a new semi-analytical method for the wire
rope accounting for different contact statuses based on contact mechanics and the thin rod theory, in
which the frictional contact between wires and the Poisson’s ratio of the wires were considered. The
method has been successfully used to perform the contact analysis of wire ropes subjected to various
load conditions [9,39–41]. Spak et al. [42] indicated that it is necessary to include the damping effect
caused by the friction effect for the study of the mechanical behavior of wire rope. However, a more
accurate method is still required for the study of the friction effect on the mechanical behavior of the
rope with hierarchical structures due to the complexity of the problem.
In this paper, we study the effect of friction on the mechanical behavior of the wire rope with a multi-
strand cross-section consisting of both single-helix and double-helix wires. We extend the contact model
of Gnanavel et al. [32] and Costello [1] to a multi-stranded rope – a typical 7 × 7 wire rope (IWRC). As
an example, the structure parameters of the rope are adjusted to ensure that the friction only occurs
between adjacent helical wires in the core strand and between the core wire and helical wires in the outer
strands. Based on the general thin rod theory, a method to calculate the local curvatures and twist of
the wires is developed. The global mechanical response of the rope and the stresses of the inter wires in
the rope are analyzed under conditions with and without the friction effect.

2. Geometrical description of the helical wire


We focus on wire rope with a double-helix configuration and take a typical 7 × 7 wire rope (IWRC, see
Figures 1(a) and (b)) into account. The parameters ds and dw are introduced to denote the chirality of
strands and wires, respectively. If ds = 1 or dw = 1, the chirality of strands or wires is right-handed; while
if ds = 1 or dw = 1, the chirality is left-handed. We define the parameter d = ds dw to characterize the
chirality of the two-level wire rope. If the handedness of the inter wires in the strand is different from
the twist direction of the strand, then d = 1. Otherwise, d = 1.
Mathematically, a two-level helical structure can be realized by creating a helix on the original helix.
For the two-level spiral structure rope shown in Figure 1(a), the strands can be divided into two
4 Mathematics and Mechanics of Solids 00(0)

Figure 1. Geometric construction of the wire rope: (a) the hierarchical helical model of a 7 × 7 rope; (b) the cross-section of the
wire rope; (c) the Frenet–Serret local right-hand axes.

categories, and the wires can be divided into four categories, as shown in Figures 1(a) and (b). In the
straight core strand, the center wire is denoted as wire I, while the other n  1 single-helix wires are
denoted as wires II. In each outer strand, the core wire is referred to as wire III, while the other double-
helix wires are referred to as wires IV.
In order to describe the geometric configuration of the rope, a Cartesian coordinate system fx, y, zg is
introduced, as illustrated in Figure 1(c). Let fi, j, kg be the unit-basis vectors, and the axis of the rope be
along the direction of k. So, the first-level helical curve can be given by
py a
RðS Þ = ry cos aw i + ry sin aw j + k ð1Þ
2p
where S is the arc length of the curve, ry is the helical radius, py = 2pry =tan u is the helical pitch and u is
the spiral angle of the helical wire; aw = a + að0Þ , where a is the winding angle of the rope and að0Þ is the
phase angle related to the arrangement position of the core strand in the rope. The relation between S
and a is given by
dS 1 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

= 4p2 ry2 + p2y ð2Þ
da 2p
At an arbitrary point of the helix, we can establish a local right-hand Frenet–Serret coordinate system
fT, N, Bg. The corresponding tangent, normal and binormal unit vectors are
1  
TðS Þ = qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2pry sin aw i + 2pry cos aw j + py k
4p2 ry2 + p2y
NðS Þ =  cos aw i  sin aw j ð3Þ
1  
BðS Þ = qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi py sin aw i  py cos aw j + 2pry k
4p2 ry2 + p2y

Each of these unit vectors is a function of S or a. The Darboux vector [43] is introduced to describe the
rotation of the Frenet–Serret coordinate system, that is
vF ðS Þ = kBðS Þ + tTðS Þ ð4Þ

Here, k and t are the curvature and torsion of the curve, respectively. In order to obtain a hierarchical
helical curve, one can move the vector vF along a helical curve and let the direction of vF be the same
as the tangent direction of the curve. Therefore, a double-helix curve can be obtained
rðsÞ = RðS Þ + rp cos bw NðS Þ + rp sin bw BðS Þ ð5Þ
Liu et al. 5

where s is the arc length along the second-level helix, also a function of S (or a); rp is the radius of the
second-level helix; bw = b + bð0Þ , where b is the helix angle of the second-level helix, which is a function
of a, and bð0Þ is the phase angle depending on the arrangement position of the wire in the rope. So, one
can use a or b to describe the hierarchical helix.
For the rope in Figure 1(a), the centerline of the wire jðj = 1, 2, . . . nÞ in the strand iði = 1, 2, . . . N Þ can
be derived from Equation (5). By Equations (1) and (5), the function for the core wire (wire I) of the core
strand (i.e. strand I) can be expressed as
py a3
rðs : 1, 1Þ = k ð6Þ
2p
The function for the outer wires (wires II) in the core strand is given by
rðs : 1, jÞ = rp1 cos a2wj i + rp1 sin a2wj j + pp1 a2 =ð2pÞk
= rp1 cos a2wj i + rp1 sin a2wj j + py a3 =ð2pÞk, ð7Þ
ð0Þ
a2wj = a2 + aj = a2 + 2pðj  2Þ=ðn  1Þ, ðj = 2, 3, . . . , nÞ

The function for the single-helix wires (i.e. wires III) in the outer strands (i.e. strand II) is
rðs : i, 1Þ = ry cos a3wi i + ry sin a3wi j + py a3 =ð2pÞk,
ð0Þ
ð8Þ
a3wi = a3 + ai = a3 + 2pði  2Þ=ðN  1Þ, ði = 2, 3, . . . , N Þ

The function for the double-helix wires (i.e. wires IV) in the outer strands is
  1=2 
2 2 2
rðs : i, jÞ = ry cos a3wi  rp2 cos bwj cos a3wi + rp2 py sin bwj sin a3wi 4p ry + py i
  1=2 
2 2 2
+ ry sin a3wi  rp2 cos bwj sin a3wi rp2 py sin bwj cos a3wi 4p ry + py j
ð9Þ
 1=2 
+ py a3 =ð2pÞ + 2pry rp2 sin bwj 4p2 ry2 + p2y k,
ð0Þ
bwj = b + bj = b + 2pðj  2Þ=ðn  1Þ, ði = 2, 3, . . . , N; j = 2, 3, . . . , nÞ

Here, a2 , a3 , b are the winding angles of wires II, III and IV, respectively; a2w , a3w , bw also corre-
spond to wires II, III and IV, respectively; pp1 and pp2 are the helical pitch of wires II and III, respec-
tively; rp1 and rp2 are the helix radius of the core strand and out strands, respectively.
We consider the formation process of the wire rope with a hierarchical helical structure. The wire seg-
ment of length ds is firstly wound into a strand segment dS, which corresponds to an increment db of
the winding angle b. In turn, such a segment of the strand is wound into a rope segment, corresponding
to an increment of winding angle da. Let uf be the helix angle of the wire, defined from the centerline of
the strand to the centerline of the wire; and let up denote the helix angle of the strand, measured from
the centerline of the rope to the centerline of the strand. From Figure 1(a), we can obtain the following
relations
 
uf = arccosðdS=dsÞ = arctan rp db=dS
 
up = arcsin ry da=dS ð10Þ
py = 2pry = tan up

Zhao et al. [7] found that the mechanical behavior of the hierarchical helical structure is sensitive to the
parameters uf and up . If we assume that the wire rope with regular hierarchical helical structures is uni-
form, then db=da = l. Here, l is a structural constant associated with the geometric characteristic of the
rope. If the height h is that at which the point is provided, one can obtain the winding angle as
6 Mathematics and Mechanics of Solids 00(0)

a = 2ph=py ð11Þ

The value of b can be also obtained. In this case, both uf and up are constant, and
 
up = arctan 2pry =py ð12Þ

3. Friction states for the wire rope


We use the friction model for the wire rope developed by Gnanavel et al. [32]. For the rope with a
straight core wire, p there are three types of friction
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi between wires [1]: (i) if the radius of the helix wire
1 + cot2 ½p=ðn  1Þsec2 uf 0ffi , there only exists contact between the outer wires; (ii) if
rp0 = rc0 + rw0 . rw0pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rp0 = rc0 + rw0 \rw0 1 + cot2 ½p=ðn  1p Þsec 2 u , there only exists contact between the core wire and the
f0
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

outer wires; (iii) if rp0 = rc0 + rw0 = rw0 1 + cot2 ½p=ðn  1Þsec2 uf 0 , the outer wires would simply touch
each other and touch the core wire as well. Here, rc0 is the radius of the straight core wire, rw0 is the
radius of the helical wires at the initial helix angle uf 0 and n  1 is the number of the outer helical wires.
Obviously, the contact condition between wires strongly depends on the geometrical parameters of the
rope. So, the frictional effect depends on the geometrical parameters as well. In what follows, as an
example, we adjust the geometrical parameters of the rope to ensure that friction only occurs between
adjacent helical wires in the core strand, and between the core wire and the double-helix wires in the
outer strands. Other types of inter-wire friction mentioned above may happen, but the analysis method
is similar. In addition, the contact between wires II and IV is generally point contact. As indicated by
Costello [1], the effect due to friction in this case is very small, which is neglected in the current work.
Therefore, the stress concentrations at the points of contact cannot be accounted for with the present
method.
The contact line between the wires is a helix. Along an infinitesimal distance ds of the helical wire,
the directions of the loads acting on the wires are along the contact line. If we only consider the friction
between the adjacent helical wires and neglect the contact deformation, the distributed forces per unit
length in the normal, binormal and axial directions of contact line are S1 , P1 and Q1 (see Figure 2(a)),
respectively. If we only consider the friction between the core wire and outer helical wires, the distribu-
ted forces per unit length in the normal, binormal and axial directions of the contact line are denoted as
S2 , P2 and Q2 , respectively, as shown in Figure 2(b).

Figure 2. Loads on a helical wire due to friction: (a) the friction between outer helical wires; (b) the friction between an outer wire
and the core wire.
Liu et al. 7

4. Internal forces and strain of the wires


We assume that the radius of curvature of the rope is known since it can be measured directly in practice.
We now analyze the equilibrium of forces of wires with and without the friction effect.
For the frictionless case, when the wire rope is subjected to an axial load, it usually undergoes elonga-
tion and twist simultaneously. The core strand of the rope undergoes both tensile force and the torque,
while the outer strands undergoes the bending moment, tensile force and torque. In the bottom-up anal-
ysis method we used here, the double helices (wires IV) can be simplified as a series of single-helix wires
in the xoy plane under pure bending. We firstly consider the local deformation of the wires in the strands
that undergo axial tension and torsion. Then, we analyze the mechanical response of the wires in the
strands under bending. When the friction effect is considered in the rope, the curvature of the rope plays
an important role in the analysis. If the rope is under uniaxial loading, both the core strand and the outer
strands undergo the bending moment in addition to the tensile force and the torque. One can directly
calculate the local deformation of the wires in the strands under axial tension, torsion and bending.

4.1 Equilibrium equations


Consider an infinitesimal segment of wire subjected to the most general case of loading, as illustrated in
Figure 3. Let s be the arc length along the wire. Let n, b and t represent the normal, binormal and tan-
gential vectors of an arbitrary point in the centerline, respectively. N and N 0 are the shear forces of the
wire in the directions of n and b, respectively; T is the tensile force in the direction of t; G, G0 and H
denote three internal moments in the directions of n, b and t, respectively; X, Y and Z are the external
line loads per unit length of the centerline of the wire in the directions of n, b and t, respectively; K, K 0
and Y are the external moments per unit length of the wire in the directions of n, b and t, respectively.
The equilibrium equations of a thin rod are given by [16]
dN
 N 0 t + T kB + X = 0 ð13Þ
ds
dN 0
 T kN + Nt + Y = 0 ð14Þ
ds
dT
 N kB + N 0 kN + Z = 0 ð15Þ
ds
dG
 G0 t + HkB  N 0 + K = 0 ð16Þ
ds

Figure 3. Local coordinate systems and loads acting on a thin slender rod.
8 Mathematics and Mechanics of Solids 00(0)

dG0
 HkN + Gt + N + K 0 = 0 ð17Þ
ds
dH
 GkB + G0 kN + Y = 0 ð18Þ
ds
where t, kN and kB denote the components of the Darboux vector in the directions t, n and b,
respectively.

4.2 Local deformation of the wires in the frictionless rope


We use the superscript (0) to indicate the parameters in the initial (un-deformed) configuration of the
hierarchical helical structure. For a helical wire, we have the following geometric relations [7], that is
df sin uf 0 cos uf 0 sin2 uf 0
t0 = , kN 0 = 0, kB0 = ð19Þ
rp0 rp0

df sin uf cos uf sin2 uf


t= , kN = 0, kB = ð20Þ
rp rp

The expressions relating the changes in curvature and twist per unit length to the internal forces are
given by [1]
prf4 prf4 prf4 E
G= EðkN  kN 0 Þ = 0, G0 = EðkB  kB0 Þ, H = ðt  t 0 Þ ð21Þ
4 4 4ð1 + nÞ
where E is the Young’s modulus of the material and rf is the radius of the deformed wire. The tensile
force in the wire is
T = prf2 Eef ð22Þ

The relation between the radius rf and the normal strain ef of the wire is given by
 
rf = rf 0 1  nef ð23Þ

where n is the Poisson’s ratio. Considering that the wire rope under tension and torsion deforms uni-
formly, we obtain that both uf and ef are not affected by the arc length s. Therefore, all internal forces
in the wire do not change with s, that is
dT dN dN 0 dH dG dG0
= = = = = =0 ð24Þ
dS dS dS dS dS dS
Substituting Equations (21), (22), and (24) into Equations (13)–(18), we obtain
X = N 0 t  TkB ð25Þ
N 0 = HkB  G0 t ð26Þ
N = 0, Z = 0, Y = 0 ð27Þ
If uf and ef are known, all the internal forces of wires can be determined from Equations (21), (22),
(25) and (26).
Since the normal strain ey and the maximum shear strain g y of the rope can be measured readily, we
now try to derive the helix angle uf and the axial strain ef of the wire from ey and g y , as shown in
Figure 4. We denote the twist per unit length of the rope as uy . From Figure 4, we obtain the following
relations
Liu et al. 9

Figure 4. A rope under tensile deformation.

a3  a30
uy = dp = tp , g y = dp ry0 uy ð28Þ
ly0

where ly0 is the initial length of the rope, a30 is the winding angle of the outer strand at the initial state,
t p is the torsion of the outer strand and ry0 is the radius of the un-deformed rope. Neglecting the contact
deformation of wires, we have
ry0 = rf 10 + 2rf 20 + rf 30 + 2rf 40 ð29Þ

rp10 = rf 10 + rf 20 ð30Þ

rp20 = rf 30 + rf 40 ð31Þ

where rf 10 , rf 20 , rf 30 , rf 40 are the radius of wires I, II, III and IV at the initial state, respectively; rp10 , rp20
are the spiral radius of the core strand and the outer stand at the initial state, respectively.
Figure 5 shows the planar view of a single-helix centerline and a double-helix centerline at the initial
state and at the deformed state. From Figure 5(a), the normal strain ep1 and the maximum shear strain
g p1 of the core strand are [7]
lp1  lp10   cos uf 2
ep1 = = 1 + ef 2 1 ð32Þ
lp10 cos uf 20

rp10 1 + ep1 1
gp1 =  ð33Þ
rp1 cot uf 2 cot uf 20

In view of the geometric relations of the wire rope, we have


ep1 = ey = ef 1 , gp1 = g y = g f 1 ð34Þ

By Equations (23) and (30), we obtain the radius of the core strand at the deformed state
10 Mathematics and Mechanics of Solids 00(0)

Figure 5. Geometric relations at the un-deformed state and the deformed state: (a) single-helix centerline; (b) double-helix
centerline.

   
rp1 = rf 10 1  nef 1 + rf 20 1  nef 2 ð35Þ

Similar to relation (28), one can derive the relation between the twist per unit length up1 and the max-
imum shear strain g p1 of the core strand, that is
a2  a20
up1 = df 2 , g p1 = df 2 rp10 up1 ð36Þ
lp10

By Equations (30) and (32)–(36), we have the following conclusions: if g y + tan uf 20 = 0, then uf 2 = 0;
if g y + tan uf 20 6¼ 0, then
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n
2 2
 
2 o
A + B rf 20 n rf 1 + rf 20 6 A2 B2 A2  B2 rf220 n2 + rf 1 + ð1 + nÞrf 20
ef 2 =  ð37Þ
A2  B2 rf220 n2

and

B
uf 2 = arccos ð38Þ
1 + ef 2
     
where A = rp10 1 + ey = g y + tan uf 20 , B = 1 + ey cos uf 20 and uf 2 2 ½0, p=2.
From Figure 5, one can obtain both the tensile strain ey and the maximum shear strain g y of the rope
ly  ly0   cos up2
ey = = 1 + ep2 1 ð39Þ
ly0 cos up20
ry0 1 + ey 1
gy =  ð40Þ
ry cot up2 cot up20

By Equations (23) and (29), one can derive the radius of the rope at the deformed state
Liu et al. 11

       
ry = rf 10 1  nef 1 + 2rf 20 1  nef 2 + rf 30 1  nep2 + 2rf 40 1  nef 4 ð41Þ

The relation between up2 and up2 for each outer strand is

sin up2 cos up2 sin up20 cos up20
up2 = dp2  ð42Þ
ry ry0

Similarly, the maximum shear strain of the outer strand is



sin up2 cos up2 sin up20 cos up20
gp2 = df 4 rp20 up2 = d  rp20 ð43Þ
ry ry0

According to the geometric relations in Figure 5(b), we can obtain the tensile strain ep2 and the maxi-
mum shear strain g p2 of the outer strand
  cos uf 4
ep2 = 1 + ef 4 1 ð44Þ
cos uf 40
rp20 1 + ep2 1
gp2 =  ð45Þ
rp2 cot uf 4 cot uf 40

Since ef 3 = ep2 , the radius of the outer strand at the deformed state is given by
   
rp2 = rf 30 1  nep2 + rf 40 1  nef 4 ð46Þ

Combining Equations (39)–(41) and Equations (43)–(46), one can obtain the parameters rp2 , ry ,ep2 , ef 4 ,
up2 , uf 4 , g p2 . Here, up2 and uf 4 2 ð0, p=2Þ.

4.3 Mechanical response of the wires in the strand under bending


We now analyze the mechanical response of the wires in the outer strand under pure bending, as illu-
strated in Figure 6. A Cartesian coordinate system fx, y, zg is introduced with the standard basis vectors
fi, j, kg. The core strand is considered to be deformed into an arc of radius r and the centerline of the
strand can be described by
RðS Þ = r cos a3w i + r sin a3w j ð47Þ
where S is the arc length of the deformed strand centerline and r = dS=da. By Equation (20), we obtain
ry
r= ð48Þ
sin2 up2

The tangential, normal and binormal unit vectors of the arc are given by
TðS Þ =  sin a3w i + cos a3w j
NðS Þ =  cos a3w i  sin a3w j ð49Þ
BðS Þ = k

Let the Darboux vector vF move along the arc and its direction coincide with the direction TðS Þ. We
then obtain a helix wound about the arc.
The centerline of wires IV in the outer strand is given by
rðsÞ = RðS Þ + rp2 cos bw NðS Þ + rp2 sin bw BðS Þ
    ð50Þ
= r  rp2 cos bw cos a3w i + r  rp2 cos bw sin a3w j + rp2 sin bw k

where s is the arc length of the wire, and


12 Mathematics and Mechanics of Solids 00(0)

Figure 6. Deformation of a strand under pure bending.

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
ds 1  2 db
= r  rp2 cos bw + rp2 ð51Þ
dS r da

The deformation of wires IV can be described by either the winding angle a or b. As shown in
Figure 6, we consider the following decomposition. A segment of wire with an infinitesimal length
ds is wound into a segment of strand AC with an infinitesimal length dS; the winding angle of the
c with a radius of curvature r.
wire is uf 4 . In turn, the strand segment AC is bent into segment AB
During this process, the increment of the winding angle is da. According to the geometry of the
deformed wire, one has
db r
= l, l = tan uf 4 : ð52Þ
da3 rp2

For an infinitesimal wire segment, the bending moment is given by M = EIf k. Here, k is the local cur-
vature of the wire and If = prf4 =4 is the moment of inertia of the cross-section. Accounting for the rela-
tive position of the wire in the strand, we know that k is a function of the winding angle a3w or bw . So,
the bending moment in the wire j along the z direction is
 
MZj = EIf kj bj  k ð53Þ

where kj and bj are the curvature and the binormal unit vectors of wire j, respectively. The expression of
kj bj  k has been derived by Zheng et al. [44] for a helical wire, which is
h   2   i
kj bj  k = s3 a r 2
p2 1  l 2
cos b w  rrp2 2 + l 2
cos b w + r 2
+ 2r 2 2
p2 l ð54Þ
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2  
ds 2 tan2 u , s dsa 1
where sa = da 3
= r  r p2 cos b w + r f 4 aa = da3 = s a r p2 l r  r p2 cos b w sin bw .

4.4 Local deformation of the wires in the rope with friction effect
We now analyze the equilibrium of forces of the wires in the core strand and outer stands, respectively,
because of the different friction conditions in them.

4.4.1 Equilibrium equations of the core strand. Along any contact line between the adjacent helical wires, there
are the normal distributed force S1 and the tangential distributed forces P1 (see Figure 2(a)). The interfa-
cial forces can be related to the forces and moments of the wires by
X =  2S1 cos f1 ð55Þ
Y = 2P1 cos f1 ð56Þ
Liu et al. 13

Figure 7. The centerline of an outer wire in a deformed position where the outer wire is wrapped around the core wire.

Z = 2Q1 ð57Þ
K =0 ð58Þ

K 0 = 2Q1 rf 2 cos f1 ð59Þ

Y =  2rf 2 P1 cosð2f1 Þ ð60Þ

where f1 is the contact angle and rf 2 is the wire radius in the core strand. Considering the slipping
between helical wires in a strand under tension, we express the relation between the distributed forces
P1 , Q1 and the contact force S1 as
P1 = mS1 , Q1 = mS1 ð61Þ

where m is the coefficient of friction.


Following Costello [1], we analyze the friction between wires in the core strand under bending (see
Figure 7). Here, nc , bc and tc denote the normal, binormal and tangential unit vectors of the core wire,
respectively; nw , bw and tw denote the normal, binormal
 and tangential unit vectors of the outer helical
wires, respectively. Assuming cot uf 2 = ry uw = rp1 cy , one can derive the curvature and the twist angle
per unit length of wires II
cos uf 2
kN = cos cy ð62Þ
ry

sin2 uf 2 cos2 uf 2
kB =  sin cy ð63Þ
rp1 ry

df sin uf 2 cos uf 2 df sin uf 2 cos uf 2


t= + sin cy ð64Þ
rp1 ry

If the bending deformation of the rope is very small, the length of the rope can be expressed as h’ry uw .
So, cy = h tan uf 2 =rp1 , that is, cy = a2w is the helix angle of wires II. The bending moments and the torque
of wires II are given by
!
pErf42 cos uf 2
G= cos cy  kN 0 ð65Þ
4 ry
14 Mathematics and Mechanics of Solids 00(0)

!
0
pErf42 sin2 uf 2 cos2 uf 2
G =  sin cy  kB0 ð66Þ
4 rp1 ry
!
pErf42
df sin uf 2 cos uf 2 df sin uf 2 cos uf 2
H= + sin cy  t 0 ð67Þ
4ð1 + nÞ rp1 ry

The tensile force is


T = prf22 Eef 2 ð68Þ

The relation between the arc length s and the angle cy of wires II is s = rp1 cy =sin uf 2 . By Equation (16),
we have the shear force
pErf42 cos uf 2 df sin uf 2
N0 =  sin cy  G0 t + HkB ð69Þ
4 ry rp1

By Equation (18), we obtain


pErf42 sin uf 2 cos uf 2 sin uf 2
Y=  cos cy + GkB  G0 kN ð70Þ
4ð1 + nÞ ry rp1

Using Equation (69), one may derive


dN 0 dN 0 dcy
=
ds dcy ds
2 3
pErf42 cos uf 2 df sin uf 2 pErf42 cos2 uf 2
6 cos cy + cos cy t 7 ð71Þ
6 4 ry rp1 4 ry 7 df sin uf 2
=66 4
7
7 rp1
pEr 2
4 0 d f sin u f 2 cos uf 2 f 2 d f sin u f 2 cos uf 2 cos u f 2 5
G cos cy + cos cy kB  H cos cy
ry 4ð1 + nÞ ry ry

Combining Equations (14), (17), (56), (61), (59) and (71), we obtain
w1  rf 2 w2
N= ð72Þ
1  rf 2 t
pErf42 cos2 uf 2 df sin uf 2 0
where w1 = 4 ry cos cy rp1 + HkN  Gt, and w2 =  dN
ds + T kN . Therefore

Y = w2  Nt ð73Þ

Z = NkB  N 0 kN ð74Þ

K 0 = w1  N ð75Þ
dN
X=  + N 0 t  TkB ð76Þ
ds
Combining Equations (55) and (61), we obtain the contact force and the friction force of the wire
X
S1 =  ð77Þ
2 cos f1
mX
P1 = mS1 =  ð78Þ
2 cos f1
Liu et al. 15

 
where cos f1 = K 0 = Zrf 2 . Obviously, if m = 0, then P1 = 0.
Now we consider the internal forces of wire I. The tensile force is
T = prf21 Eef 1 ð79Þ

The torque is
Eprf41 gf 1
H= ð80Þ
4ð1 + nÞ rf 10

The bending moment is


1
M= EIf 1 ð81Þ
ry

where If 1 = prf41 =4 is the moment of inertia of the cross-section of wire I.

4.4.2 Equilibrium equations of the outer strand. Let the strand only be subjected to bending deformation where
the contact is maintained between the outer wires and the core wire. Between any two cross-sections
along an outer wire, the only way in which an external load can be applied to the wire is along the con-
tact line. Therefore, if a force per unit length acts on the boundary of a wire, of radius rf4 , as shown in
Figure 2(b), the following relations can be obtained by transforming the force on the boundary to an
equivalent force and (or) moment acting along the centerline of the wire [1,29]
X = S2 ð82Þ
Y = Q2 sin f2 + P2 cos f2 ð83Þ
Z = Q2 cos f2  P2 sin f2 ð84Þ
K =0 ð85Þ
K 0 =  Zrf 4 ð86Þ
Y = Yrf 4 ð87Þ

where f2 is the contact angle, rf 4 is the radius of wire IV in the outer strand and K, K 0 and Y are the
components of the external moments per unit length along the centerline of the wire. The radius of cur-
vature of the outer strand is rp = ry =sin2 up2 . The curvatures and the twist per unit length of wires IV are
derived as
cos uf 4
kN = cos cp ð88Þ
rp

sin2 uf 4 cos2 uf 4
kB =  sin cp ð89Þ
rp2 rp

df sin uf 4 cos uf 4 df sin uf 4 cos uf 4


t= + sin cp : ð90Þ
rp2 rp

Here, cp = bw is the helix angle of wire IV. The bending moments and the torque of wires IV can be cal-
culated as
!
pErf44 cos uf 4
G= cos cp  kN 0 ð91Þ
4 rp
16 Mathematics and Mechanics of Solids 00(0)

!
pErf44 sin2 uf 4 cos2 uf 4
0
G =  sin cp  kB0 ð92Þ
4 rp2 rp
!
pErf44 df sin uf 4 cos uf 4 df sin uf 4 cos uf 4
H= + sin cp  t0 ð93Þ
4ð1 + nÞ rp2 rp

The tensile force is


T = prf24 Eef 4 ð94Þ

For wire IV, the relation between the arc length s and the helix angle cp is s = cp rp2 =sin uf 4 . Therefore,
by Equation (16), we have
pErf44 cos uf 4 df sin uf 4
N0 =  sin cp  G0 t + HkB ð95Þ
4 rp rp2

By Equation (18), one may have


pErf44 sin uf 4 cos uf 4 sin uf 4
rf 4 Y =  cos cp + GkB  G0 kN ð96Þ
4ð1 + nÞ rp rp2

By Equation (95), we obtain


dN 0 dN 0 dc
=
ds dc ds
2 3
pErf44 cos uf 4 df sin uf 4 pErf44 cos2 uf 4
6 cos cp + cos cp t 7 ð97Þ
df sin uf 4 6
6
4 rp rp1 4 rp 7
7
= 6 4 7
rp2 4 df sin uf 4 cos uf 4 pEr f 4 df sin uf 4 cos uf 4
2
cos uf 4 5
G0 cos cp + cos cp kB  H cos cp
rp 4ð1 + nÞ rp rp

Loads N and Z acting on wire IV can be written as


0
dN
N=   T kN + Y =t ð98Þ
ds

Z = NkB  N 0 kN ð99Þ

Therefore
dN df sin uf 4 ðw3  w4 + w5 Þt  w6 ðdN 0 =ds  TkN + Y Þ
= ð100Þ
ds rp2 t2

Here
2 3
pErf44 cos uf 4 df sin uf 4 pErf44 cos2 uf 4
6 4 sin cp  sin cp t 7
6 rp rp2 4 rp 7
6 7
6 pEr cos2 u4
df sin uf 4 cos uf 4 7
6+ f4 f4
cos cp cos cp 7
6 7
6 4 rp rp 7 df sin uf 4
6 7
w3 = 6
pEr 4 7 rp2 ,
6 0 d f sin u f4 cos uf4 f4 df sin u f4 cos uf4 7
6 +G sin cp  sin cp kB 7
6 rp 4ð1 + nÞ rp 7
6 7
6 4 7
4 pErf 4 df sin uf 4 cos uf 4 2
cos uf 4 2
cos uf 4 5
 cos cp coscp + H sin cp
4ð1 + nÞ rp rp rp
Liu et al. 17

cos uf 4
w4 =  T sin cp ,
rp
1 pErf44 sin uf 4 cos uf 4 sin uf 4
w5 = sin cp
rf 4 4ð1 + nÞ rp rp2
!
1 pErf44 cos uf 4 cos2 uf 4
 sin cp kB + G cos cp
rf 4 4 rp rp
!
1 pErf44 cos2 uf 4 df sin uf 4 cos u f 4
+  cos cp kN + G0 cos cp :
rf 4 4 rp rp
d sin u cos u
and w6 = f f4
rp
f4
cos cp .
From Equations (13) and (82), we obtain the expression for the contact force S2 on wire IV
dN
S2 = X = N 0 t  T kB  ð101Þ
ds
The friction force is
f2 = mS2 ð102Þ

Now we analyze the loads on wires III. A statically equivalent uniformly distributed lateral pressure is
introduced to replace the line force S2 along the contact of wire III, which is
n1
m= S2 ð103Þ
2prf 3

The tensile strain and force of wire III are


!
1 T m  
ef 3 = 2
+ , T = prf23 ef 3 E  m=n ð104Þ
E prf 3 n

Similar to wire II, the radius of curvature of wire III is ry , while the curvature and the twist per unit
length are
cos uf 3
kN = cos c0y ð105Þ
ry

sin2 uf 3 cos2 uf 3
kB =  sin c0y ð106Þ
ry ry

df sin uf 3 cos uf 3 df sin uf 3 cos uf 3


t= + sin c0y ð107Þ
ry ry

where c0y = a3w is the helix angle of wire III. According to Equation (21), the bending moments and the
torque of wire III are given by
!
pErf43 cos uf 3
G= cos c0 y  kN 0 ð108Þ
4 ry
!
0
pErf43 sin2 uf 3 cos2 uf 3
G =  sin c0 y  kB0 ð109Þ
4 ry ry
18 Mathematics and Mechanics of Solids 00(0)

!
pErf43
df sin uf 3 cos uf 3 df sin uf 3 cos uf 3
H= + sin c0 y  t0 ð110Þ
4ð1 + nÞ ry ry

where kN0 , kB0 and t 0 are the curvatures and the twist per unit length of the wire at the initial state.
Transforming the loads on the line of contact to the centerline of wire III, we have
X
n1  
X3 = Y4 cos uf 4 sin bw + Z4 df sin uf 4 sin bw
2
X
n1  
Y3 = Y4 cos uf 4 cos bw  Z4 df sin uf 4 cos bw
2
X
n1  
Z3 = Y4 df sin uf 4  Z4 cos uf 4
2
2 3
K 0 4 cos uf 4 sin bw + Y4 df sin uf 4 sin bw
X
n1 
6 + Y cos u cos b  Z d sin u cos b r sin b d sin u  7
K3 = 4 4 f4 w 4 f f4 w p2 w p p2 5
2
   ð111Þ
+ Y4 df sin uf 4  Z4 cos uf 4 rp2 cos bw sin a2w + rp2 sin bw cos a2w cos up2
2 0 3
K 4 cos uf 4 cos bw  Y4 df sin uf 4 cos bw
X6 
n1  7
K 03 = 4  Y4 df sin uf 4  Z4 cos uf 4 rp2 cos bw cos a2w  rp2 sin bw sin a2w cos up2 5
2
 
rp2 sin bw df sin up2 Y4 cos uf 4 sin bw + Z4 df sin uf 4 sin bw
2 3
K 0 4 df sin uf 4  Y4 cos uf 4
6 Y cos u sin b + Z d sin u sin b  7
X6
n1 6 
4 f4 w 4 f f4 w
7
7
Y3 = 6  r cos b sin a + r sin b cos a cos u p2 7
6 p2  w 2w p2 w 2w
 7
2 6 7
4 + Y4 cos uf 4 cos bw  Z4 df sin uf 4 cos bw 5
 
 rp2 cos bw cos a2w  rp2 sin bw sin a2w cos up2

Considering Equations (85), (108) and (16), we obtain


pErf43 cos uf 3 df sin uf 3
N0 =  sin c0y  G0 t + HkB ð112Þ
4 ry ry

Combining Equations (17) and (109), we have


pErf43 cos2 uf 3 df sin uf 3
N= cos c0y + HkN  Gt  K 0 ð113Þ
4 ry ry

To sum up, if the axial strain and the helix angle of the deformed wire are provided, one then can
obtain the internal forces and the contact force of each wire. The final curvatures and twist per unit
length are assumed and the loads required to keep the wires in equilibrium are determined from the
equations of equilibrium. Neglecting the effect of contact deformation, we use the procedure in Section
4.2 to derive the axial strain and the helix angle of the deformed wire from axial strain ey and shear
strain g y of the rope.

5. Internal forces and stresses of the rope


5.1 The frictionless case
5.1.1 Internal forces in the frictionless wire rope under tensile loads. In Section 4.2, we have shown that, if the axial
strain ey and the shear strain g y of the rope are known, the axial strain ef and the helix angle uf of the
wire can be derived. By Equations (21), (22), (25) and (26), one may finally determine the internal forces
Liu et al. 19

of the wire. A method for calculating the internal forces in the rope was proposed by Zhao et al. [7], in
which the internal forces of each wire were firstly projected to the centerline of the strand, then the inter-
nal forces of each strand were projected to the rope axis. Using this method, we calculate the total axial
force and the total torque in strand i by
X
n  
Tpi = Tij cos uf j + N 0ij df sin uf j ð114Þ
j=1

X
n  
Hpi = Hij cos uf j + G0 ij df sin uf j + Tij rpi df sin uf j  N 0 ij rpi cos uf j ð115Þ
j=1

The bending moment in strand i is


X
n X
n  
0
Gpi = G0 pij = MZj  MZj0 ð116Þ
j=2 j=2

Similar to Equation (26), the internal force along the binormal direction of strand i is

sin2 up 0 dp sin up cos up


Npi0 = Hpi kStrand
Bi
0 Strand
 Gpi ti = Hpi  Gpi ð117Þ
ry ry

The influence of the shear force Npi0 on the deformation is insignificant. So, the total tensile force and the
total torque of the rope can be respectively calculated as
X
N  
Fy = Tpi cos upi + N 0 pi dp sin upi ð118Þ
i=1

X
N  
My = Hpi cos upi + G0 pi dp sin upi + Tpi ry dp sin upi  N 0 pi ry cos upi : ð119Þ
i=1

5.1.2 Stresses in the frictionless wire rope under tensile load. When the wire rope undergoes the axial tension and
torsion, the internal forces in the wires can be calculated by Equations (21), (22), (25) and (26). The ten-
sile stress in a wire mainly results from tensile force T and bending moments G and G0 , while the shear
stress is mostly attributed to torque H. For wire j in strand i, its internal forces are denoted by Tij , Gij0
and Hij . The maximum stresses in the wire are given by
!  !  !
Tij 4G0  2H 
 ij   ij 
sT = max , sG0 = max  3  , t H = max  3  ð120Þ
2
prfij  prfij  prfij 

For wire jðj = 2, 3, . . . , nÞ in strand ið2, 3, . . . , N Þ, the maximum tensile stress due to bending is
 0   !
G pij rfij  4G0 
 
sMPB = max   = max  pij
3 
ð121Þ
Ifij   prfij

The maximal tensile stress sy and the maximal shear stress t y in a wire of the rope are given by
   !
  Tij 4G0  4G0 
 ij   pij 
sy = max sij = max + +  ð122Þ
prf2ij prf3ij   prf3ij 

ty = tH : ð123Þ
20 Mathematics and Mechanics of Solids 00(0)

5.2 The friction effect


5.2.1 Internal forces of the friction wire rope under tensile and bending loads. For strand i, its internal forces are given
by
X
n  
Tpi = Ti1 + df N 0 ij sin uf j + Tij cos uf j ð124Þ
j=2

X
n  
Npi0  = Ni10 + Nij sin bwj  N 0 ij cos uf j cos bwj + df Tij sin uf j cos bwj ð125Þ
j=2

The bending moments of the cross-section of strand i are


X
n  
Hpi = Hi1 + G0 ij df sin uf j + Hij cos uf j + Tij rpi df sin uf j  N 0 ij rpi cos uf j ð126Þ
j=2

n
X
Gij sin bwj  G0 ij cos uf j cos bwj + Hij sin uf j cos bwj
G0pi = Gi1
0
+ ð127Þ
N 0 ij rpi cos bwj df sin uf j  Tij rpi cos bwj cos uf j
j=2

The tensile force and the torque of the rope are


N 
X 
Fy = Tp1

+ Npi0  dp sin up2 + Tpi cos up2 ð128Þ
i=2

N 
X 
My = Hp1

+ 0
Gpi  dp sin up2 + Hpi cos up2 + Tpi ry dp sin up2  Npi0  ry cos up2 : ð129Þ
i=2

5.2.2 Stresses in the friction wire rope under tensile and bending loads. For wire I, the maximum absolute value of
stress is located at the point farthest from the neutral surface. It is
1
sb11 = Erf 1 ð130Þ
ry

Accounting for the normal stress due to the tensile force, one can obtain the maximum normal stress of
wire I, that is
s11 = Eef 1 + sb11 ð131Þ

The maximum shear stress of wire I is


 
2H 
 11 
t11 =  3  ð132Þ
 prf 1 

For wires II, III and IV, the normal stress due to the bending moments depends on the distance from
the neutral surface. Therefore, the maximum normal stress of wire j in strand i is
!
Tij 4Gij Tij 4G0 ij
sij = max 6 , 6 ð133Þ
prf2ij prf3ij prf2ij prf3ij

The maximum shear stress is


Liu et al. 21

 
 2H 
 ij 
tij =  3  ð134Þ
prf ij 

The maximum normal stress and shear stress are


   
sy = max sij , ty = max tij ð135Þ

respectively.

6. Numerical results and discussion


A basic 7 × 7 stranded rope, as shown in Figure 1(a) and (b), is chosen for evaluating the developed
model. The radius of curvature of the rope is assumed to be ry = 3835mm. We introduce the parameter
d to characterize the handedness of the two-level rope, as indicated in Section 2. The Young’s modulus
of the wire material is E = 188 GPa, the Poisson’s ratio is n = 0:3 and the friction coefficient is m = 0:115
[45], which has been widely used by many other authors [45]. The initial radii of wires I, II, III and IV
are rf 10 = 1:97 mm, rf 20 = 1:865 mm, rf 30 = 1:6 mm and rf 40 = 1:5 mm, respectively. If there is no special
instruction, we assume that the initial helix angles of wires II and III are uf 20 = 18:998 , uf 40 = 18:548 ,
respectively, and the initial helix angle of the outer strand is up20 = 15:558 . Accordingffi to the given struc-
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 2
tural parameters, pwe have rp10 = 3:825 mm\r
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi f 20 1 + cot ½p=ðn  1Þsec uf 20 = 3:892 mm and
2 2
rp20 = 3:1 mm . rf 40 1 + cot ½p=ðn  1Þsec uf 40 = 3:086 mm. This means that friction only exists
between the helical wires (wires II) in the core strand, and between the core wire (wires III) and the
outer helical wires (wires IV) in the outer strand. The friction between wires II and IV is neglected since
it is very small, as indicated by Costello [1]. It should be mentioned that the linear elastic hypothesis is
made throughout this paper, which constrains the range of strains considered in what follows below the
threshold of 1%.

6.1 Internal forces of the rope


The evolution of the tensile force and the torque with the normal strain ey of the rope is given in Figure
8. The results of the internal forces and moments for the 7 × 7 stranded rope without the frictional effect
are also plotted in Figure 8 for comparison purposes. We assume that the cross-section is at height
h = 18mm and the twist per unit length of the rope is uy = 0. Both d = 1 and d =  1 are considered for
studying the influence of handedness on the mechanical response of the rope. It is shown that both the
tensile force Fy and the torque My increase linearly with the strain of the rope ey . For a given value of ey ,
the magnitudes of Fy and My with frictional effect are generally larger than the results in the frictionless
case. Figure 8(a) shows that, for the case d = 1, the friction effect increases the value of Fy by 4.39%;
meanwhile, for the case d =  1, the friction effect increases the value of Fy by 5.05%, as shown in
Figure 8(b). However, the magnitude of Fy for d = 1 is smaller than that for d =  1 by about 2.90%.
This indicates that the handedness plays an insignificant role in the rope under tensile load. From
Figures 8(c) and (d), one can see that, for the case d = 1, the friction effect increases the value of My by
about 16.82%; meanwhile, for the case d =  1, the friction effect increases the value of My by about
9.04%. In addition, in the friction case, the magnitude of My for d = 1 is larger than that for d =  1 by
about 78.12%. Comparing Figure 8(a) with (c) or Figure 8(b) with (d), one can conclude that, for a
given chirality, the friction effect plays a more important role for My than that for Fy . In addition, if we
consider the effect of chirality on the tensile force and torque of the rope by comparing Figure 8(a) with
(b) and Figure 8(c) with (d), we may find that the effect of chirality on the torque My is more significant
than that on the tensile force Fy .
Now we analyze the evolution of the contact force S and the friction force f with the normal strain of
the rope ey . Since f = mS, the trend of f with ey is the same with S. Plots of S versus ey are given in Figure
9. Here, we use Sij to denote the contact force of wire j due to other wires in the strand i. If i = 1, S1j rep-
resents the contact force of wire j resulting from other helical wires; if i = 2, S2j denotes the contact force
of wire j due to the core wire.
22 Mathematics and Mechanics of Solids 00(0)

Figure 8. Plots of the tensile force Fy and the torque My versus the strain of rope ey at the cross-section h = 18mm. Here, uy = 0 is
assumed.

Figure 9. Plots of the contact force versus the normal strain of the rope: (a) and (b) the contact forces in the core strand (i = 1);
(c) and (d) the contact forces in the outer strand (i = 2).

Here, we focus on the cross-section at height h = 18mm and assume uy = 0. Comparing Figures 9(a)
and (b) or 9(c) and (d), one may find that the effect of chirality on the contact forces of the wires is
Liu et al. 23

Figure 10. The evolution of the contact force with the normal strain of the rope at h = 0 mm: (a) core strand; (b) outer strand.

insignificant. Figures 9(a) and (b) show that, in the core strand, the magnitude of the contact forces
between the helical wires decreases nonlinearly with the normal strain of the rope; and that the value
approaches to zero. Note that the sign ‘‘–’’ simply indicates that the direction of the contact force
changes. For the outer strand, as shown in Figures 9(c) and (d), the contact force between the core wire
and the helical wires decreases linearly with the tensile strain of the rope. Comparing Figures 9(a) and
(c), one may find that, in the rope studied here, the contact force between the core wire and the helical
wires, that is, S2j , is much larger than the contact force between the outer helical wires, that is, S1j .
The evolution of the contact forces with the normal strain of the rope at the cross-section h = 0mm is
given in Figure 10. Here, we assume uy = 0 and d = 1. Comparing Figure 9(a) with 10(a) or 9(c) with
10(b), one may find that the position of the cross-section plays an important role in the contact force. It
means that the contact force of the wire depends on its arrangement position in the rope. In the cross-
section at h = 0mm, the helix angle of the wire is zero and several contact forces are equal to each other,
for example, S13 = S17 and S14 = S16 (see Figure 10(a)) and S22 = S25 , S23 = S24 and S26 = S27 (see Figure
10(b)).
Next, we analyze the relation between the contact force and the helix angle of the wires in the core
strand and the outer strand, as shown in Figure 11. The following parameters are assumed: ey = 0:003,
uy = 0, d = 61. The contact force of the wires changes periodically with the winding angle of the helical
wires, which is in agreement with the previous studies [1,41]. The peak value of the contact force of the
adjacent wires is also adjacent. According to Figure 11(a), at a2 = 08 , h = 0, ey = 0:003, there are two
intersection points between S14 and S16 and between S13 and S17 , respectively. This explains the reason
why there are overlaps between S14 and S16 and between S13 and S17 at ey = 0:003, as shown in Figure
10(a). Similarly, in Figure 11(c), at b = 08 , h = 0, ey = 0:003, there are three intersection points between
S22 and S25 , S23 and S24 and S26 and S27 . This also explains the reason why there are three overlaps
between S22 and S25 , S23 and S24 and S26 and S27 in Figure 10(b) at ey = 0:003. Comparing Figures 11(a)
and (b) or 11(c) and (d), one can conclude that the influence of chirality on the contact force of the wire
along the helix angle is insignificant. Comparing Figures 11(a) and (c), one can see that the friction
between helical wires in the core strand changes with the winding angle in a sinusoidal fashion; mean-
while, the friction between the core wire and helical wires in the outer strand changes with the winding
angle in a ‘‘W’’-shape fashion, which has two crests and two troughs.

6.2 Stresses in the rope


The evolutions of the maximum tensile stress sy and the maximum shear stress t y of the rope with the
normal strain ey are given in Figure 12. The numerical results of the maximum stresses without the fric-
tion effect are also given in Figure 12 for comparison purposes. Here, the cross-section is assumed to be
at h = 18mm, and uy = 0. From Figures 12(a) and (b), one can see that the maximum tensile stress of the
rope sy increases linearly with the normal strain ey , while the increasing rate of sy is slower for the fric-
tion wire rope than that for the frictionless rope. For the friction rope, the influence of the chirality on
the value of sy is minor. If the wire material is frictionless, at the same strain, the value of sy for the case
d =  1 is a little bit larger than that for the case d = 1. As show in Figures 12(c) and (d), the maximum
24 Mathematics and Mechanics of Solids 00(0)

Figure 11. Plots of the contact forces versus the helix angle of the wires: (a) and (b) core strand; (c) and (d) outer strand.

Figure 12. Plots of the maximum normal stress and the maximum shear stress versus the normal strain of the rope: (a) and (b)
correspond to the maximum normal stress for d = 1 and d =  1, respectively; (c) and (d) correspond to the maximum shear stress
for d = 1 and d =  1, respectively.

shear stress of the rope t y increases linearly with the normal strain ey . If the value of ey is given, the value
of t y for the case d = 1 is about 23.5% larger than that for the case d =  1. Comparing Figures 12(c)
Liu et al. 25

Figure 13. Plots of the maximum normal stress (a) and the maximum shear stress (b) versus the winding angle.

and (d), one can see that the influence of the chirality on the value of t y for the friction rope is larger
than that for the frictionless rope.
The evolutions of sy and t y with the helix angle b of wires IV are given in Figure 13. Here, ey = 0:003
and uy = 0 are assumed. From Figure 13(a), one can see that the maximum normal stress sy remains
unchanged with increasing the value of b. Further analysis indicates that the maximum normal stress is
always equal to s11 , that is, the maximum normal stress of wire I in strand I, for both d = 1 and
d = 1. According to Equations (130) and (131), the value of s11 depends on the curvature of rope ry ,
which is a constant here. That is the reason why the value of sy is a constant with helix angle b. Figure
13(b) shows that the effect of chirality on the value of t y is significant. For d = 1, the maximum shear
stress t y changes in a wave-like manner, whereas for d =  1, the maximum shear stress t y nearly
remains constant with increasing b, and the value is smaller than that for d = 1. This phenomenon
seems not rational in physics since a simple change of the chirality may not affect the type of the depen-
dence of the maximum shear stress on the helix angle of the structure. The difference between the maxi-
mum shear stress for d = 1 and d =  1 is mainly due to the definition of the maximum shear stress in
wire rope. According to Equations (134) and (135), the maximum shear stress is defined as the maxi-
mum value of the shear stress among all wires in the rope. Further analysis on the maximum shear
stress of each type of wire in the rope shows the following: if d =  1, the maximum shear stress is in
wires II; if d = 1, the maximum shear stress is sometimes in wires IV and sometimes in wires II. That is
the reason why the value of t y changes in a wave-like manner for d = 1, while nearly remaining constant
for d =  1.

7. Conclusions
The influence of the friction effect on the mechanical behaviors of the wire rope with hierarchical helical
structure (a typical 7 × 7 rope) has been studied. The internal forces, for example, the tensile force and
the torque, of the rope with friction effect are compared with the results without friction effect. It is
found that the internal forces of the rope with friction effect are generally larger than the results without
friction effect. Interestingly, the internal forces of the friction wire rope are less affected by the chirality
than that of the frictionless wire rope. It is also found that the maximum tensile stress of the friction wire
rope is less affected by the chirality than that of the frictionless wire rope, while the maximum shear
stress of the friction wire rope is more affected by the chirality than that without friction.
The evolution of the contact force of wires with the normal strain of the rope or the helix angle of the
wires has been analyzed. It is found that the influence of the chirality of the rope is insignificant, whether
on the contact force between helical wires in the core strand or on the contact force between the core
wire and the helical wires in the outer strand. In the core strand, the contact force between helical wires
tends to zero with increasing the axial normal strain of the rope. In the outer strand, the contact force
between the core wire and the helical wires decreases linearly with the axial normal strain of the rope. In
addition, the contact force of the wires depends on their arrangement position in the strands, and peri-
odically changes with the winding angle.
26 Mathematics and Mechanics of Solids 00(0)

It should be pointed out that the contact analysis carried out above is based on the continuous
assumption. The assumption is only valid for the continuous contact regions, for example, the regions
between adjacent helical wires in the core strand, and between the core wire and the double-helix wires
in the outer stands; it cannot be used for the non-continuous contact regions, for example, the contact
between wires II and wires IV. The friction in the non-continuous contact regions is quite complicated.
An exact approach for describing such a contact condition is not available now. It is clearly a limitation
of the current work. For future study, experiments on the stranded rope are required to verify the pre-
dictions reported here in the future. However, as indicated by Usabiaga and Pagalday [6], experiments
may not provide the local stresses or deformation information of the inner wires. It will be interesting
to perform the FE simulation, which accounts for both the friction effect and the Poisson effect to com-
pare the global and local response of the rope investigated here.

Funding
The author(s) disclosed receipt of the following financial support for the research, authorship, and/or publication of this article:
This work was supported by the National Natural Science Foundation of China (Grant Nos. 11702103, 11472114 and
11772138), the Young Elite Scientist Sponsorship Program by CAST (No. 2016QNRC001), the Fundamental Research Funds
for the Central Universities (HUST: No. 2018KFYYXJJ008).

ORCID iD
Dabiao Liu https://orcid.org/0000-0002-6453-5659

References
[1] Costello, GA. Theory of wire rope. New York: Springer Science & Business Media New York, 1997.
[2] Cardou, A, and Jolicoeur, C. Mechanical models of helical strands. Appl Mech Rev 1997; 50: 1–14.
[3] Raoof, M, and Kraincanic, I. Critical examination of various approaches used for analysing helical cables. J Strain Anal
Eng Des 1994; 29: 43–55.
[4] Argatov, I. Response of a wire rope strand to axial and torsional loads: asymptotic modeling of the effect of interwire
contact deformations. Int J Solids Struct 2011; 48: 1413–1423.
[5] Elata, D, Eshkenazy, R, and Weiss, MP. The mechanical behavior of a wire rope with an independent wire rope core. Int J
Solids Struct 2004; 41: 1157–1172.
[6] Usabiaga, H, and Pagalday, JM. Analytical procedure for modelling recursively and wire by wire stranded ropes subjected
to traction and torsion loads. Int J Solids Struct 2008; 45: 5503–5520.
[7] Zhao, Z-L, Zhao, H-P, Wang, J-S, et al. Mechanical properties of carbon nanotube ropes with hierarchical helical
structures. J Mech Phys Solids 2014; 71: 64–83.
[8] Xiang, L, Wang, HY, Chen, Y, et al. Modeling of multi-strand wire ropes subjected to axial tension and torsion loads. Int
J Solids Struct 2015; 58: 233–246.
[9] Meng, F, Chen, Y, Du, M, et al. Study on effect of inter-wire contact on mechanical performance of wire rope strand
based on semi-analytical method. Int J Mech Sci 2016; 115–116: 416–427.
[10] Karathanasopoulos, N. Torsional stiffness bounds of helical structures under the influence of kinematic constraints.
Structures 2015; 3: 244–249.
[11] Hong, KJ, Kiureghian, AD, and Sackman, JL. Bending behavior of helically wrapped cables. J Eng Mech 2005; 131:
500–511.
[12] Nikolaos, K, and Gerald, K. Mechanical response of a helical body to axial, torsional and radial strain. Int J Mech Sci
2015; 94–95: 260–265.
[13] Hruska, FH. Radial forces in wire ropes. Wire Wire Prod 1952; 27: 459–463.
[14] Knapp, RH. Derivation of a new stiffness matrix for helically armoured cables considering tension and torsion. Int J
Numer Meth Eng 1979; 14: 515–529.
[15] Costello, GA, and Phillips, JW. Effective modulus of twisted wire cables. ASCE J Eng Mech Div 1976; 102: 171–181.
[16] Love, AEH. A Treatise on the Mathematical Theory of Elasticity. New York: Dover Publications. 1944.
[17] Kumar, K, and Cochran, JE Jr. Closed-form analysis for elastic deformations of multilayered strands. J Appl Mech Trans
ASME 1987; 54: 898–903.
[18] Jolicoeur, C, and Cardou, A. A numerical comparison of current mathematical models of twisted wire cables under
axisymmetric loads. J Energ Resour Technol Trans ASME 1991; 113: 241–249.
[19] Utting, WS, and Jones, N. The response of wire rope strands to axial tensile loads-Part I. Experimental results and
theoretical predictions. Int J Mech Sci 1987; 29: 605–619.
Liu et al. 27

[20] Hobbs, RE, and Raoof, M. Interwire slippage and fatigue prediction in standard cables for TLP tethers. In: Proceedings of
the International Conference Behaviour of Off-Shore Structures, 1983, 77–99.
[21] Nabijou, S, and Hobbs, RE. Relative movements within wire ropes bent over sheaves. J Strain Anal Eng Des 1995; 30:
155–165.
[22] Blouin, F, and Cardou, A. A study of helically reinforced cylinders under axially symmetric loads and application to
strand mathematical modelling. Int J Solids Struct 1989; 25: 189–200.
[23] Jolicoeur, C, and Cardou, A. Analytical solution for bending of coaxial orthotropic cylinders. J Eng Mech 1994; 120:
2556–2574.
[24] Karathanasopoulos, N, and Angelikopoulos, P. Optimal structural arrangements of multilayer helical assemblies. Int J
Solids Struct 2016; 78–79: 1–8.
[25] Lee, WK. An insight into wire rope geometry. Int J Solids Struct 1991; 28: 471–490.
[26] Stanova, E, Fedorko, G, Fabian, M, et al. Computer modelling of wire strands and ropes Part I: theory and computer
implementation. Adv Eng Softw 2011; 42: 305–315.
[27] Stanova, E, Fedorko, G, Fabian, M, et al. Computer modelling of wire strands and ropes part II: finite element-based
applications. Adv Eng Softw 2011; 42: 322–331.
[28] Ashkenazi, R, Weiss, MP, and Elata, D. Torsion and bending stresses in wires of non-rotating tower crane ropes.
In: OIPEEC Technical Meeting: Experiences with Ropes 2003, 77–99.
[29] Le Clair, RA, and Costello, GA. Axial, bending and torsional loading of a strand with friction. J Offshore Mech Arctic
Eng 1988; 110: 38–42.
[30] Papailiou, KO. On the bending stiffness of transmission line conductors. IEEE T Power Del 1997; 12: 1576–1588.
[31] Inagaki, K, Ekh, J, and Zahrai, S. Mechanical analysis of second order helical structure in electrical cable. Int J Solids
Struct 2007; 44: 1657–1679.
[32] Gnanavel, BK, Gopinath, D, and Parthasarathy, NS. Effect of friction on coupled contact in a twisted wire cable. J Appl
Mech Trans ASME 2010; 77: 1–6.
[33] Jiang, WG, Warby, MK, and Henshall, JL. Statically indeterminate contacts in axially loaded wire strand. Eur J Mech A
Solid 2008; 27: 69–78.
[34] Jiang, WG. A concise finite element model for pure bending analysis of simple wire strand. Int J Mech Sci 2012; 54: 69–73.
[35] Karathanasopoulos, N, Reda, H, and Ganghoffer, J-f. Finite element modeling of the elastoplastic axial-torsional
response of helical constructions to traction loads. Int J Mech Sci 2017; 133: 368–375.
[36] Sun, JF, Wang, GL, and Zhang, HO. FE analysis of frictional contact effect for laying wire rope. J Mater Process Technol
2008; 202: 170–178.
[37] Kmet, S, Stanova, E, Fedorko, G, et al. Experimental investigation and finite element analysis of a four-layered spiral
strand bent over a curved support. Eng Struct 2013; 57: 475–483.
[38] Zhang, D, and Ostoja-Starzewski, M. Finite element solutions to the bending stiffness of a single-layered helically wound
cable with internal friction. J Appl Mech Trans ASME 2016; 83.
[39] Chen, Y, Meng, F, and Gong, X. Full contact analysis of wire rope strand subjected to varying loads based on semi-
analytical method. Int J Solids Struct 2017; 117: 51–66.
[40] Chen, Y, Meng, F, and Gong, X. Interwire wear and its influence on contact behavior of wire rope strand subjected to
cyclic bending load. Wear 2016; 368: 470–484.
[41] Chen, Y, Meng, F, and Gong, X. Study on performance of bended spiral strand with interwire frictional contact. Int J
Mech Sci 2017; 128–129: 499–511.
[42] Spak, K, Agnes, G, and Inman, D. Cable modeling and internal damping developments. Appl Mech Rev 2013; 65:010801.
[43] Oprea, J. Differential geometry and its applications. Washington, DC: Mathematical Association of America, 2007.
[44] Zheng, S, Liu, D, and He, Y. The influence of fiber migration on the mechanical properties of yarns with hierarchical
helical structures. J Strain Anal Eng 2018; 53: 88–105.
[45] Jiang, WG, and Henshall, JL. The analysis of termination effects in wire strand using the finite element method. J Strain
Anal Eng 1999; 34: 31–38.

You might also like