You are on page 1of 15

International Journal of Fatigue 59 (2014) 114–128

Contents lists available at ScienceDirect

International Journal of Fatigue


journal homepage: www.elsevier.com/locate/ijfatigue

Finite element analysis of the fatigue strength of copper power


conductors exposed to tension and bending loads
Fachri P. Nasution a,⇑, Svein Sævik a, Janne K.Ø. Gjøsteen b,1
a
Department of Marine Technology, NTNU, Otto Nielsens vei 10, 7491 Trondheim, Norway
b
Norwegian Marine Technology Research Institute, MARINTEK, Otto Nielsens vei 10, 7052 Trondheim, Norway

a r t i c l e i n f o a b s t r a c t

Article history: The paper presents FE analyses for predicting longitudinal stresses from tension–tension and tension–
Received 29 April 2013 bending fatigue tests of a 95 mm2 stranded copper power conductor. As fatigue test results indicated that
Received in revised form 27 August 2013 the fatigue performance was dominated by longitudinal stresses, the models were formulated by a com-
Accepted 18 September 2013
bination of elastic beam and elastic–plastic beam-contact elements that included the friction. Two con-
Available online 28 September 2013
tact conditions were investigated: the point (trellis) contact between adjacent layers and the inline
contact within each layer and between centre wire and inner layer. Due to the plastic deformations of
Keywords:
the wires obtained from the manufacturing procedure, a simplified description of the contact behaviour
Copper conductor
Point (trellis) contact
was adopted and calibrated by axial tension testing. The FE models were further validated by calibration
Inline contact testing and mesh sensitivity checks. The simulated stresses were applied to attempt bridging the gap
Plastic deformations between the SN data obtained from full cross-section tension–tension and tension–bending testing
Mesh sensitivity and SN data obtained from individual wires testing.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction (Db) (from pitch and roll motion) acting on the cables. The most
heavily loaded section is close to the attachment point (at section
Floating vessels like ships and semisubmersibles are exten- A–A) to the vessel (Fig. 2). Mean and dynamic tension will be trans-
sively used for production of oil and gas from offshore fields. Also ferred to the wires as tension and shear forces while the dynamic
offshore wind turbine concepts based on floating units have been bending moment will induce local bending as well as axial friction
developed. Electrical power cables are used in conjunction with forces into each individual wire.
floating units for provision of energy to installations on the sea As shown in Fig. 1b, the mean axial force (F x ) in each wire is a
bed, power from land to the floater, or export of power from a wind function of the mean global tension (T) and the mean global torque
turbine to land. Power cables that are linked to a floating unit are moment (M T ). The dynamic axial force (DFx) in each wire is a func-
subjected to fatigue loading from the waves and due to the move- tion of these, the corresponding dynamic quantities DT and DMT,
ment of the vessel in the waves. Therefore, fatigue strength needs the dynamic curvature (Db) and the coefficient of friction (l) be-
to be verified for design. tween the contact surfaces. The dynamic curvature results in local
A power cable consists in general of multiple conductors each bending in each wire where DMx is the dynamic torque moment
representing an assembly of individual wires usually made of cop- about the helix tangential x-direction, DMy is the dynamic bending
per or aluminium. Fig. 1a shows a typical offshore power cable for moment about the helix bi-normal y-direction and DMz is the dy-
an alternating current (AC) three-phase system. One cable usually namic bending moment about principal normal vector of the helix
has three conductors, one for each phase. Each conductor consists curve.
of copper wires helically wound in layers around a centre wire. The wires in a stranded conductor are stranded helically in lay-
During operation, a cable will be exposed to gravity, environ- ers which leads to contact longitudinally both within and between
mental loading from the sea, and to forces due to movements of each layer, which is illustrated in Fig. 3b. Contact within a layer
the vessel. The gravity will induce a mean global tension (T) and and between centre and first layer is denoted inline contact. The
mean global torque moment (M T ). The forces due to movements global axial force will result in both longitudinal and transverse
of the vessel induce a dynamic tensile load (DT) and torque forces within the layer where the transverse forces will cause
(DMT) (due to heave and surge motion) and dynamic curvatures diameter reduction in the inline (hoop) direction of each wire. If
cylindrical bodies with diameter D1 and D2 are pressed together
⇑ Corresponding author. Tel.: +47 735 955 64. with a certain load per unit length, p, a small continuous contact
E-mail address: fachri.nasution@ntnu.no (F.P. Nasution). surface area will occur as shown in Fig. 4.
1
Presently employed at Reinertsen AS, 7010 Trondheim, Norway.

0142-1123/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijfatigue.2013.09.009
F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128 115

Fig. 1. (a) Offshore power cable for three-phase AC system using helically stranded copper conductor inside. (b) Helically stranded copper conductor with lay angle a exposed
to dynamic and static loads.

Contact between wires in different layers is called point contact Zhou et al. [9] studied the plastic flow, local wear and fretting
(Figs. 3a and 5). failure for overhead electrical conductors imposed by bending fati-
A comprehensive literature study has been conducted with re- gue loads. Papailiou [10] presented a new model of conductors un-
spect to models for describing fatigue and contact stresses in ca- der simultaneous tensile and bending loads. During the bending
bled structures. Mindlin [1] described the contact surfaces of two loads, the model takes into account the interlayer friction and slip
isotropic bodies subjected to normal and tangential loading with- in the conductor. A mesoscale model has been developed by Hong
out friction. Dong and Steidel [2] studied the contact stress condi- et al. [11] in order to describe the bending behaviour of helically
tions between layers of strands using a photoelastic technique. wrapped cable under tension. The model accounts for the nonlin-
Hobbs and Raoof [3] have reported fatigue test results for socketed ear behaviour of the cable due to friction forces. Karlsen [12] sim-
structural strands. They concluded that the fatigue behaviour was ulated the fatigue mechanism in dynamic power cables by
governed by several failure mechanisms related to the contact con- dynamic testing in tension and bending. He concluded that the ef-
ditions close to the socket. Johnson [4] described the Hertzian con- fect of fretting on the fatigue properties was less dominant for cop-
tact stresses between two solid bodies under a certain normal load. per conductors than for steel wires and ropes. However, there is no
A theory of contact was developed for predicting the shape of the detailed information regarding the failure positions on the speci-
contact area and the growth in size with increasing load; the mag- mens. Lévesque et al. [13] presented investigations related to fret-
nitude and distribution of surface tractions, normal and possibly ting fatigue of an overhead electrical conductor focusing on the
tangential forces, transmitted across the interface. Raoof and contact conditions at the trellis contact points.
Hobbs [5] proposed an analytical model for relevant multi-layered The copper power conductor investigated here further pro-
stranded structures for determining the point and inline contact duced by a compacting procedure where the following features
forces and associated relative displacements with its ends fixed were noted:
against rotation. Raoof [6] developed from first principles a theo-
retical model using axial single wire data for predicting the axial 1. Introduction of irregularities in wire geometries, specially for
fatigue of the full cross-section at constant load amplitude, and the outer layer due to irregular supporting point contact
was able to correlate the theoretical predictions to observations conditions.
from experimental testing. Raoof [7] concluded that his theoretical 2. An increase in the contact area, thus reducing the contact
model provides useful upper bounds to the fatigue life of cables stresses.
failing at the end termination and that the termination type signif-
icantly affects the observed fatigue life. Raoof and Huang [8] pro- The effect of irregularities on the fatigue performance of indi-
posed simple methods for estimating strand plane section vidual wires were investigated by Nasution et al. [14–16] where
bending stiffness when imposed to cyclic bending and external a FE model was presented that allowed transforming the nominal
hydrostatic pressures. SN data into SN data of actual longitudinal stresses that included
As opposed to the aforementioned authors, the present study the effect of irregularities. Full cross-section SN data obtained from
focus on copper conductors applied in dynamic cables where the fatigue tension–bending tests results were also presented using a
design is based on using protective steel armours and bending simple analytical stress model. Significant deviations were noted
stiffeners to resist the external forces, thus limiting the tensile between the full cross-section SN data and the individual wire data
forces and curvatures in the copper conductor. which were believed to be due to:
116 F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128

Fig. 2. Flexible power cable attached to a floating offshore structure.

Fig. 3. Point and inline contact in a stranded copper conductor. (a) Point contacts between inner and outer layer. (b) Plan view of conductor cross section. The black arrows
show point contact between wires in adjacent layers (outer and inner layer), the red arrows show inline contact between wires within one layer and between inner layer and
core wire. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128 117

Fig. 4. Inline contact between two wires within layers and between inner and centre wires.

1. The effect of localised bending between layers in the ten-


sion–bending mode.
2. The effect of friction between layers.
3. The effect of the irregularities in terms of stress
concentration.

The specimens tested in [16] were inspected by Scanning Elec-


tron Microscopy (SEM). Although, it was not possible to see any
signs of fretting, neither from the SEM results nor by visual inspec-
tion, it was not possible to completely rule out that fretting con-
tributed to the crack initiation process. Compared to steel rope
and aluminium conductors the contact stresses in copper conduc-
tors are lower, and the displacements larger and it is at least rea-
sonable to assume that fretting plays a less dominant role in the
fatigue performance of compacted copper conductions exposed
to the loading conditions investigated here. This is also inline with
the observations by Karlsen [12]. One reason for this may be the in-
creased contact area and reduced contact stresses from the com-
pacting procedure. If the assumption of negligible influence from
fretting is correct it should be possible to bridge the gap noted be-
tween the full cross-section SN data and the individual wire SN
data obtained from tension testing based on beam and beam con-
tact theory. Fig. 5. Point contact between two cylinder bodies with elliptical contact surface
The purpose of the present paper is therefore to perform beam and angle in the plane.

stress analysis to verify whether this is the case. The following pro-
cedure is adopted:
stresses obtained in the full cross-section tension–bending
tests presented in [16] and the tension–tension tests pre-
1. Build a FE model based on beam and beam-contact
sented here.
elements.
7. Use the simulated stresses and plot the results into the
2. Formulate a simplified contact stiffness and friction model
individual wire actual longitudinal stress range SN-curve.
representative for large contact areas (plasticised contact
surfaces).
3. Calibrate the normal contact stiffness model by full cross- 2. Experimental tests
section axial stiffness testing.
4. Calibrate the friction model by measured friction 2.1. General
coefficients.
5. Present SN data from full cross-section tension–tension The specimens used in this work were taken from a 95 mm2
tests. copper power conductor (ETP copper), designated by the UNS
6. Perform FE stress analyses, include the stress concentration C11000 series. The definition of ETP copper is related to copper al-
factors (SCFs) obtained from measured geometric irregular- loy purity of at least 99.95% and characterised by a very high elec-
ities obtained [15,16] and estimate the actual longitudinal trical conductivity and ductility. The conductor cross section
118 F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128

consisted of 19 wires, each with a nominal diameter of 2.5 mm. A was more severe for the outer layer, as noted for the individual
centre wire is followed by 6 and 12 helically wound wires in two wires fatigue data reported in [15,16]. The detailed listing of the
layers. The pitch lengths were measured to be 220 mm for both tension–tension test results is found in Table 1.
layers. When using the right hand lay rule the obtained lay angles Fig. 8 displays the full cross-section fatigue data in tension–ten-
were 4.1° and +7.3° for the inner and outer, respectively. The sion mode compared to the individual outer wires fatigue data
modulus of elasticity (E) for copper is 115 GPa [12]. based on the nominal stress ranges Dr. Noting that the black line
shows average of fatigue data taken from individual outer wires in
2.2. Previous fatigue tests tension–tension mode. The upper and lower bounds of the SN
curve are based on nominal stress range with the plus/minus
As previously reported elsewhere [15,16], SN curves based on two standard deviations scatter band.
nominal stress range (Dr) and actual longitudinal stress range The results show that by using the same level of longitudinal
(Drxx) have been constructed based on individual outer wires test- stress ranges as the individual wires, the fatigue strength obtained
ing in tension–tension mode. The actual longitudinal stresses (rxx) from full cross-section testing is very close to the fatigue strength
of individual wires were assessed based on the measured SCFs. obtained from individual wire testing in tension–tension mode for
The maximum SCF in the axial direction, SCF(a), occurred at thin- the outer layer.
nest section as the observed irregularities and was calculated as:
rpeak 2.5. Friction tests
SCF ðaÞ ¼ ð1Þ
rnominal
A series of tests on friction performance of copper wire against
where rpeak represents the maximum stress occurring and rnominal is copper wire were conducted. The purpose of these tests was to
the nominal applied axial stress. The mean SCFs based on Eq. (1) investigate the coefficient of friction between the copper wires.
were found to be 1.238 and 1.058, for the outer and inner wires, Several constant normal force conditions were investigated in
respectively. the tests. The tangential and normal loads were monitored through-
The results showed that the fatigue lifes for the outer wires were out the tests and the associated coefficient of friction l calculated.
consistently below the fatigue lifes for the inner layer and centre Small variations were noted between the tests and the coeffi-
wire. All the observed fatigue failures occurred at the thinnest sec- cient of friction was found to be 0.2.
tion. This indicates that in these tests the fatigue failures were gov-
erned by the irregularities resulting from additional bending stresses.
Fatigue bending tests of full cross-section conductors were also 3. Layer contact formulations
conducted. All fatigue failures occurred in the inner layer, in the
section that had been bent over a bellmouth. The observed fatigue 3.1. General
data obtained from full cross-section testing and the stress range
estimated by only considering wire bending showed that the fati- In the present study, the layer contact formulation in terms of
gue strength of the full cross-section conductors was significantly copper wire interaction between adjacent or within layers are split
lower than the fatigue strength found for individual wires. This
may be explained by the fatigue strength being governed by local-
ised bending and friction effects between layers, which is the main
topic of this work.

2.3. Calibration test

The calibration test used to tune the FE model of the 95 mm2


copper power conductor was performed by investigating the mea-
sured axial stiffness vs. FE and analytical results.
A copper power conductor with free span length 1000 mm was
mounted in a servo-hydraulic test machine. On each individual
wire of the outer layer strain gages were attached for measuring
the axial strain. In addition, the specimen was equipped with a
extensometer for monitoring the axial strain of conductor.

2.4. The conductor tension–tension fatigue test

Six specimens with free span length 1000 mm and cross section
area 95 mm2 were subjected to nominal stress ranges (Dr) of
160 MPa and 190 MPa with stress ratio, R = 0.1. All the specimens were
terminated at the ends using standard sockets filled with high strength
epoxy resin. The intention of the tests was to provide full scale bench-
mark data on fatigue performance in tension–tension mode.
The tension–tension fatigue testing was carried out in a stan-
dard two-column servohydraulic test machine with hydraulic
clamp at both ends (Fig. 6). The loading was sinusoidal with fre-
quency 2 Hz, in load control.
All the fatigue failures consistently developed at the outer layer
in the thinnest section of the wire (at point contact area), as shown
in Fig. 7. This complies with the observation that the irregularities
resulting from plastic deformation from the compacting process Fig. 6. Standard fatigue test rig of a full cross-section by mounting the both ends.
F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128 119

Fig. 9. Inline contact between two wires within a layer and between inner and
central wires by introducing total equivalent width beq.

into two distinct conditions based on the area of contact. The con-
dition includes inline contact governing the contact between wires
within layers and between inner layer and centre wire. The second
condition is point contact between outer and inner layers.
The Hertzian contact theory governs when the contact zone is
limited compared to the surrounding material volume (elastic
range and small strains). In the present case, the contact condition
between wires (inline and point contact) have developed from the
Fig. 7. A failed full cross-section copper power conductor at outer layer in tension–
manufacturing processes with associated large plastic deformation
tension mode.
between wires.

Table 1
Fatigue test data for full cross-section due to tension–tension loads.

No Dr MPa No. cycles to failures [cycles] Irregularities [mm] Failure position (Layer)
1 160 776,070 0.24 Outer
2 160 797,203 0.235 Outer
3 160 898,909 0.20 Outer
4 190 308,412 0.165 Outer
5 190 409,001 0.125 Outer
6 190 421,446 0.115 Outer

Fig. 8. SN fatigue data for full cross-section compared to nominal SN curve based on individual wires testing in tension–tension mode.
120 F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128

Fig. 10. Contact geometry area between inner and outer layer.

bodies. Since the strategy adopted here is based on beam elements,


believing that the fatigue performance is primarily governed by
longitudinal stresses, the contact conditions need only to be repre-
sented with sufficient accuracy with respect to the load sharing be-
tween layers, local bending at the trellis points and friction due to
the relative displacements between wires. Therefore, a simplified
procedure for representing the contact conditions has been
adopted applying the concept of elastic springs in the normal
direction and elastic–plastic springs in the tangential direction to
deal with friction.

3.2. Contact in normal direction

3.2.1. Inline contact


The total equivalent width, beq, of the contact zone between
wires within a layer and also between inner and central wire is
shown in Fig. 9, where A denotes the cross section area of the
wires, taken as 5 mm2 and r is the wire nominal radius of 1.25 mm.
The inline contact force finline due to the elastic deformation is
shown in:

Ebeq
finline ¼ C in dn ð2Þ
R

where Cin is a calibration factor for inline contact and dn denotes the
A
normal contact displacement. By substituting R = 2r and beq = 2r , Eq.
(2) can be rearranged as

EA
finline ¼ C in dn
4r2 ð3Þ
EA
Fig. 11. Stick–slip displacement on the wire interfaces. (a) Curvilinear beam kinline ¼ C in 2
deformations. (b) Stick–slip displacement quantities between helix wire and centre. 4r

The inline contact stiffness kinline represents the normal stiffness to


The surface areas are therefore enlarged with associated in- be applied for the inline contacts within inner and outer layers and
creased resistance against relative motion between contacting between inner and central wires.
F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128 121

3.2.2. Point contact


Fig. 10 represents the contact geometry area, Ac, as a parallelo-
gram between outer and inner wires indicated by the shaded area.
The contact area Ac is the product of Leqbeq, given as
D
Ac ¼ D ð4Þ
jsinða2  a1 Þj
where D is the wire diameter, a2 and a1 are the lay angles of outer
and inner wires, respectively.
The point contact force, Fpoint, acting on the inner wires at point
O is calculated as,
Ac E
F point ¼ C po dn ð5Þ
2r

EAc
kpoint ¼ C po ð6Þ
2r
where Cpo is a calibration factor for point contact.

3.3. Stick–slip model


Fig. 13. Individual wire model. (a) Outer wire model [15,16]; (b) Ovalised wire
The contact elements applied are based on an ideal elastic–plas- cross section at section B–B (at the thinnest section).
tic relation where the stick stiffness is given by
kstick ¼ lfinline =D ð7Þ
where D represent the relative displacement at which slip occurs,
after which the stiffness is zero (Coulomb friction).
The D-parameter needs to be selected such that adequate stiff-
ness is obtained, i.e., Navier hypothesis of plane surfaces remain
plane applies before slip occurs. In order to ensure this, sandwich
beam theory is applied looking at the internal work performed
along the helix path using the approach proposed by Sævik
[17,18]. Noting that the free relative displacement in one layer is
proportional to the helix radius squared, a layer by layer approach
is assumed. This means that the stick stiffness setting at the inter-
faces belonging to one layer is based on the helix radius in this
layer. The stick stiffness can be studied by considering the internal
work done (Wi) along a quarter pitch between the wire and the
surrounding interfaces which for inline contacts can be written as:
Z p=2    
@ v @up @ v @up
Wi ¼ EA þ d þ
0 @x @x @x @x
X
n
R
þ kstick v dv dW ð8Þ
i¼1
sin a

where EA is axial stiffness in the layer, v = uw  up, is the relative


displacement at the interface between helix wire and the centre Fig. 14. A conductor model subjected to tension–tension loads (Dr) with stress
ratio R = 0.1.
along the helical path (see Fig. 11), uw denotes the longitudinal helix
wire displacement, n is the number of interfaces sharing the inter-
action process and up represents the wire displacement that would is the helix radius of wire, j represents the global curvature and W
occur if plane surfaces remained plane during the stick condition. R is the angular position, see Fig. 11a. The plane surfaces remain plane
strain is taken as:
dup
¼ R cos2 a cos Wj ð9Þ
dx
Furthermore, by assuming the relative displacement (v) as:
v ¼ v 0 sin W ð10Þ
The following expression is obtained for v0:
2a
R2 cos j
v0 ¼  sin a
kstick R2
 ð11Þ
1 þ n EA sin2 a

where it has been assumed that each interface carries equal load,
i.e., shared among n interfaces.
The upper term represents the free displacement which
Fig. 12. Complete 3D model of full cross-section power copper conductor. approaches zero when the shear interaction stiffness approaches
122 F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128

For the point contacts this is modified into:

3R2 F point l
Dpoint ¼  
2
ð14Þ
1
F tol
 1 Lp EA sin a

where Lp is the length between contact points (see Fig. 3).

4. Finite Element (FE) modelling

4.1. General

Uflex3D is a finite element tool for 3D stress analysis of umbili-


cals [19], however, only including beam stresses due to internal
pressure, external pressure, tension, torsion and bending loads.
The copper power conductor model was made using elastic
beam elements for all wires with a constant elastic bending, axial
and torsion stiffness.
The full cross-section copper conductor model was 1000 mm
long including one centre wire, six wires for inner layer and twelve
wires for outer layer, see Fig. 12. The model was meshed with 500
elements per wire in the longitudinal direction.
The contact elements include four parts, namely:

1. Point contact elements. Positioned between inner and outer


layers.
2. Inline contact elements, within layers.
Fig. 15. Tension–bending model at initial condition. 3. Inline contact between centre wire and inner wires.
4. Contact between the outer layer of the conductor and the bend
surface (i.e., bellmouth).
infinite. However, the shear interaction stiffness must be selected
based on a compromise between numerical stability and accuracy.
The bellmouth was modelled as a geometrical surface, where a
In this study, the shear interaction stiffness is selected by allowing
penalty stiffness is applied during wire-surface contact. No friction
<10% relative displacement in the stick phase. This gives the fol-
is included in this contact element.
lowing requirement:
The wire to wire contact elements were given an elastic stiff-
  2
1 1 EA sin a ness in the normal direction whereas friction was treated by an
kstick P 1 2
ð12Þ elastic–plastic model using the stick displacement parameter
n F tol R
established from the above.
where Ftol denotes the allowable tolerance for stick relative dis- In order to perform the fatigue calculation by FE analysis, this
placement. The stick–slip parameter for the inline contact is then has to be performed in two steps and based on the actual longitu-
established using Eq. (7) and selecting n = 3 as, dinal stresses (rxx) given as:
3R2 finline l rxx ¼ rxa SCF ðaÞ þ rxby SCF ðbyÞ þ rxbz SCF ðbzÞ ð15Þ
Dinline ¼  
2
ð13Þ
1
F tol
 1 EA sin a
where rxa represents the nominal axial stresses, rxby is the nominal
bending stresses in y-direction (weak axis), and rxbz is the nominal

Fig. 16. Comparison results of axial force vs. strain obtained from calibration tests, FE and theoretical model.
F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128 123

Table 2
Detailed contact parameters of full cross-section copper conductor.

Interface Contact forces Nominal stiffness per interface Stick–slip per interface (Eq. (7)) Slip parameters [mm] (Eqs. (13) and (14))
2 2
Centre-inner layer 0.65 N/mm 200 N/mm 1411.5 N/mm 9.21  105
Inner layer–inner layer 0.65 N/mm 200 N/mm2 1411.5 N/mm2 9.21  105
Outer layer–outer layer 1.8 N/mm 200 N/mm2 1411.5 N/mm2 2.55  104
Outer layer–inner layer 12 N 930 N/mm 2.16  104 N/mm 1.11  104

Fig. 19. Axial force distributions of individual wires from different layer due to
Fig. 17. Point contact distribution between outer and inner layers due to tensile tensile load 9500 N.
load 9500 N.

Fig. 18. Inline contact distribution due to tensile load 9500 N (Red line is hidden
behind the blue line). (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)

bending stresses in z-direction (strong axis) and according to the


following procedure:

1. Calculate the fatigue representative for the sections away from


the point contacts (away from the thinnest section) by using 1.0
for all SCFs in all layers.
2. Calculate the fatigue at the point contacts having the thinnest
sections by using the obtained data for the stress concentration
factors due to axial loads, SCF(a), as reported previously [15,16].

For the tension–bending model, the SCFs in the thinnest sec-


tions also included the weak (y-direction), SCF(by), and strong (z-
direction) axes, SCF(bz). Fig. 13 illustrates the associated wire cross Fig. 20. Longitudinal stresses due to tension–tension loads. (a) Dr = 160 MPa; (b)
Dr = 190 MPa.
section at the thinnest section (section B–B). The compacting pro-
cess lead to plastic deformation such as ovalising the wire cross
section at the thinnest section due to point contact. The SCF in Dnominal  t reduced
SCF ðbyÞ ¼ ð16Þ
bending about weak axis, SCF(by), at section B–B is given by Dnominal
124 F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128

Fig. 21. Fatigue damage distribution of full cross-section conductor model by


Dr = 160 MPa. (a) Outer layer; (b) Inner layer.

Fig. 23. Longitudinal stresses for outer layer in tension–tension mode. (a)
Dr = 160 MPa; (b) Dr = 190 MPa.

Fig. 24. Curvature along the model at maximum deformation.

2.5 mm, resulting in SCFby equal to 0.92 and 0.96 for the outer
and inner wires, respectively. For the SCF in bending about strong
axis, SCF(bz), was taken to be 1.0 because it will not be important
for the maximum stress.
Fig. 22. Predicted fatigue damage distribution of full cross-section conductor model
by Dr = 190 MPa. (a) Outer layer; (b) Inner layer.
4.2. Tension–tension model

where the average treduced was found to be 0.21 and 0.1 mm for out- The principal model of full cross-section copper conductor ex-
er and inner wires [15,16], respectively. Dnominal was taken to be posed to the tension–tension loads with stress ratio, R = 0.1 is
F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128 125

Fig. 25. Axial force histories acting at end beams due to reversed bending loads.

Fig. 27. Predicted fatigue damage distribution of copper conductor model due to
reversed bending motion with radius of curvature q1 = 1500 mm. (a) Inner layer;
(b) Outer layer.

bellmouth where both ends were axially pretensioned by end


beams with a constant tensile force of 9500 N. The test setup in-
cluded 3000 mm model length. However, the full detail FE model
length was truncated to 1000 mm (five pitches) due to computing
limitations.
Two radii of bellmouth curvature were investigated
q1 = 1500 mm and q2 = 3000 mm at a constant tension of 9500 N
according to the test parameters.
Fig. 15 presents the complete FE model at the initial condition.
The end beams were connected to the outer external tubular
frames 1 and 2, where the cyclic bending is induced by prescribed
displacement of the beam actuator. This motion is then transferred
to the external frames by means of the transverse beam and the
attachment arms at 1 and 2.

5. FE results

5.1. Calibration

Fig. 26. Actual longitudinal stresses vs. curvature hysteresis loops due to reversed
Fig. 16 shows the correlation results obtained from the proce-
bending motions at the location of maximum fatigue stress. (a) q1 = 1500 mm; (b)
q2 = 3000 mm. dure described in Section 3. The results show good agreement be-
tween the calibration test, FE and theoretical results. Similar
correlation was also obtained by Papailiou [10]. The obtained con-
shown in Fig. 14. The applied nominal stress ranges, Dr, were 160
tact calibration factors for inline (Cin) and point contact (Cpo) were
and 190 MPa, for the fatigue load cases in accordance with the test
2.2  106 and 2.1  102, respectively. The associated stick–slip
procedure.
parameters are given in Table 2.
The point contact force distribution due to maximum tensile
4.3. Tension–bending model load 9500 N is shown in Fig. 17. The point contact forces show that
the quantities are nearly similar over the model length unless close
FE analyses were also carried out for the full cross-section con- to the end fittings, the average point contact forces indicating
ductor model subjected to reversed bending loads over the 12 N.
126 F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128

Fig. 28. Predicted fatigue damage distribution of copper conductor model due to
reversed bending motion with radius of curvature q2 = 3000 mm. (a) Inner layer; (b)
Outer layer.

The inline contact forces distribution within inner layer, be-


tween centre and inner layer and within outer layer are compared
in Fig. 18. The results demonstrate the inline contact force distribu-
Fig. 29. Actual longitudinal stresses vs. curvature hysteresis loops due to reversed
tion within inner layer is similar to the inline contact force be- bending motions for different number of elements at maximum deformation. (a)
tween centre and inner layer where the maximum inline contact q1 = 1500 mm; (b) q2 = 3000 mm.
force is approximately 0.65 N/mm. For the outer layer the maxi-
mum inline contact force is close to 1.8 N/mm. The axial force
distributions of individual wires for the different layers is pre- layers, respectively. Similar to above results, the first predicted fa-
sented in Fig. 19. The highest axial force occurred in the centre wire tigue failure occurred at outer layer as observed in the test.
with 518 N. For the inner and outer wires, the axial forces are In order to ensure sufficient accuracy, a mesh sensitivity check
505 N and 496 N, respectively, which demonstrates almost equal was carried out with respect to actual stress range and predicted
load sharing between layers. fatigue life using 800 elements.
Fig. 23 shows the actual longitudinal stress results for outer
5.2. Tension–tension fatigue wires by varying number of elements. The outer layer longitudinal
stresses converged to 203 MPa and 235 MPa for applied nominal
The obtained actual longitudinal stress history of outer and in- stress ranges Dr = 160 and 190 MPa, respectively.
ner wires from full cross-section and individual wire models are The first predicted fatigue failures took place at outer layer
presented in Fig. 20. It is seen that the results compare well with where the predicted fatigue life converged to 1.17  106 and
longitudinal stresses of individual wires for both load cases. 3.33  105 cycles for applied nominal stress ranges Dr = 160 and
The fatigue damage distributions of the models for one load cy- 190 MPa, respectively.
cle have been calculated. Thus the expected fatigue life can be as-
sessed as the inversion of the fatigue damage. The red spots
indicates the occurred maximum fatigue damage. 5.3. Tension–bending fatigue
Fig. 21 shows the fatigue damage distributions for outer and in-
ner layers for nominal stress range Dr = 160 MPa. The predicted The curvature along the model at maximum deformation is
number of cycles to failure are 1.17  106 and 2.13  106 cycles shown in Fig. 24 for q1 = 1500 and q2 = 3000 mm.
for outer and inner layers, respectively. The first predicted fatigue The end beams were applied to simulate the die springs by
failure is located in the outer layer in accordance with the test. associating the same axial stiffness and then introducing a preten-
For nominal stress range Dr = 190 MPa, the fatigue damage dis- sion of 9500 N as indicated at point A in Fig. 25. Next, the reversed
tributions are displayed in Fig. 22. The predicted number of cycles bending loads were exposed to the midspan of the model by the
to failure are 3.33  105 and 5.26  105 cycles for outer and inner bellmouth contact surface.
F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128 127

Fig. 30. Comparison between the predicted fatigue life using individual wires SN data and full cross-section fatigue testing in tensiontension mode. The individual wires SN
curve is based on the actual longitudinal stresses.

Fig. 31. Comparison between the predicted fatigue life using individual wires SN data and full cross-section fatigue testing in tensionbending mode. The individual wires SN
curve is based on the actual longitudinal stresses.

Fig. 26 presents the longitudinal stresses vs. curvature hystere- fatigue lifes are 1.88  106 and 5.0  106 cycles for the inner and
sis loops during the reversed bending motion. The hysteresis loops outer layers, respectively.
corresponds to the frictional energy dissipated at the contact inter- A mesh sensitivity check was performed using 800 elements.
faces. The results exhibit that frictional energy dissipated in the Fig. 29 presents the hysteresis loops of actual longitudinal stresses
cyclic bending regime is larger for radius of curvature vs. curvature in the inner layer comparing to different number of
q1 = 1500 mm than q2 = 3000 mm. The larger energy dissipated, elements. For radius of curvature q1 = 1500 mm and
the larger longitudinal stress range. Therefore, the accumulated en- q2 = 3000 mm, the longitudinal stress ranges converged to 241
ergy dissipated at the contact wire surfaces play an important role and 191 MPa, respectively.
in assessing fatigue damage of copper power conductors. The ac-
tual longitudinal stress ranges (Drxx) are 241 and 191 MPa for ra-
dius of curvature q1 = 1500 and q2 = 3000 mm, respectively, as 6. Discussion of FE analyses and test results
compared to 190 MPa and 95 MPa only considering pure bending.
The predicted fatigue damage distribution results for different FE analyses of a full cross-section copper power conductor ex-
layers are displayed in Fig. 27 for radius of curvature 1500 mm. posed to tension–tension and tension–bending loads have been
The results show that the predicted first fatigue failure occurred carried out including mesh sensitivity checks.
at inner layer within the bellmouth. This is in accordance with The predicted first fatigue failure of copper conductor model in
the observed fatigue failures seen in the tests [16]. The predicted tension–tension fatigue mode is located at outer layer as shown in
number of cycles to failure are 2.33  105 and 4.0  105 cycles Figs. 21 and 22. This is in accordance with the observed fatigue fail-
for inner and outer layers, respectively. ure as shown in Fig. 7. It indicates that the fatigue failures were
For radius of curvature of 3000 mm, the predicted fatigue dam- governed by the local bending effects due to surface irregularities
age distribution is shown in Fig. 28, the predicted first fatigue fail- on the wires where the most severe irregularities were found in
ure located at inner layer inside the bellmouth. The estimated the outer layer [16].
128 F.P. Nasution et al. / International Journal of Fatigue 59 (2014) 114–128

In Fig. 30, the fatigue data of full copper cross-section in ten- 4. For the tension–bending mode, all the predicted fatigue
sion–tension mode is compared to the SN fatigue data of individual failures invariably occurred at inner layer inside the bell-
outer wires in tension–tension mode based on the actual longitu- mouth. This is in accordance with the observed fatigue fail-
dinal stress range. The predicted number of cycles to failure show ures in the fatigue bending tests. This indicates that the
good correlation with the fatigue life data based on the actual lon- effect of friction between outer and inner layers at the
gitudinal stresses. All the fatigue life results are within the fatigue point contact area plays an important role in assessing
scatter band. the fatigue life of full cross-section copper conductors.
Moreover, the predicted first fatigue failure due to reversed 5. As long as the longitudinal stresses governs the fatigue per-
bending motions took place at inner layer within the bellmouth. formance FE analysis with beam elements can be used to
This is in accordance with the observed fatigue failures from the predict realistic fatigue lifes.
testing. This indicates that the fatigue bending failures are caused
by friction and localised bending at the point contact between out- Work is on-going in terms of FE analyses and experimental
er and inner layers. The friction effect created enlarged longitudi- works for studying the fatigue strength of a 300 mm2 copper
nal stresses at the point contact section as shown in Fig. 26. The power conductor subjected to tension–bending mode and individ-
fact that friction forces will be largest in the inner layer has also ual wires tested in tension–tension mode applying similar
been noted previously by Papailiou [10] and Hong et al. [11]. methodologies.
The comparison between predicted fatigue life and fatigue life
data in tension–bending mode are presented in Fig. 31, noting that Acknowledgement
the stress values used to present the test data only include the ana-
lytical bending term. The test data are seen to be within the scatter The authors acknowledge ABB, Karlskrona, Sweden for provid-
band of the predicted fatigue lifes using FE analysis and individual ing test materials applied in the test.
wires fatigue data.
References
7. Conclusions and future works
[1] Mindlin RD. Compliance of elastic bodies in contact. J Appl Mech
1949;71:259–68.
2
Fatigue strength investigations of a 95 mm copper power con- [2] Dong RG, Steidel Jr FR. Contact stress in stranded cable. Exp Mech
ductor by FE analyses have been conducted in tension–tension and 1965;5(5):142–7.
[3] Hobbs RE, Raoof M. Interwire slippage and fatigue prediction in stranded
tension–bending modes. FE models were established based on cables for TLP tethers. In: Chryssostomidis C, Connor JJ, editors. Proceedings of
beam and beam contact elements, where a simplified contact mod- third international conference on behaviour of offshore structures, Boston, vol.
elling procedure was adopted. The FE model was calibrated by ax- 2; 1982. p. 77–99.
[4] Johnson KL. Contact mechanics. Cambridge University Press; 1985.
ial stiffness testing and validated by mesh sensitivity checks.
[5] Raoof M, Hobbs RE. Analysis of multilayered structural strands. J Eng Mech
Relevant friction coefficients were obtained by testing. Two alter- 1988;114:1166–82.
native full cross-section fatigue tests of a 95 mm2 copper power [6] Raoof M. Axial fatigue of multilayered strands. J Eng Mech 1989;116:2083–99.
conductor were conducted in tension–tension and tension–bend- [7] Raoof M. Axial fatigue life prediction of structural cables from first principles.
Proc Inst Civil Eng 1991;91:19–38.
ing modes, a total of twelve conductor specimens were tested [8] Raoof M, Huang Yu Ping. Cyclic bending characteristics of sheathed spiral
and the results were compared to FE analysis and applying SN fa- strands in deep water. Int J Offshore Polar Eng 1993;3:189–96.
tigue data obtained from individual wires tests. Several conclu- [9] Zhou ZR, Cardou A, Fiset M, Goudreau S. Fretting fatigue in electrical
transmission lines.. Wear 1994;173:179–88.
sions can be drawn in this paper in terms of fatigue performance [10] Papailiou KO. On the bending stiffness of transmission line conductors. IEEE
of copper power conductors. Trans Power Deliv 1997;12.
[11] Hong K-J, Der Kiureghian A, Sackman JL. Bending behavior of helically
wrapped cables. J Eng Mech 2005;131.
1. In order to perform accurately FE analysis calibration tests [12] Karlsen S. 2010. Fatigue of copper conductors for dynamic subsea power
are required to capture realistic load sharing between lay- cables. In: Proceedings of the OMAE 2010 29th international conference on
ers and in assessing the fatigue life. ocean, offshore and arctic engineering, Shanghai, China.
[13] L évesque F, Goudreau S, Cloutier L. Elastic–plastic microcontact model for
2. For the tension–tension mode, all the predicted first fatigue elliptical contact areas and its application to a treillis point in overhead
failures were initiated at the outer layer. The obtained electrical conductors. J Tribol 2011;133.
actual longitudinal stresses from full cross-section model [14] Nasution FP, Sævik S, Gjøsteen JKØ. Fatigue analysis of copper conductor for
offshore wind turbines by experimental and FE method. Energy Procedia. In:
are close to the actual longitudinal stresses of individual
Proceedings of 9th deep sea offshore wind R&D, Trondheim; 2012.
wires. [15] Nasution FP, Sævik S, Gjøsteen JKØ. Study of fatigue strength of copper
3. For the tension–tension test of full cross-section copper conductor considering irregularities surfaces by experimental testing and FE-
conductors, all the first fatigue failures were initiated at analysis. In: The ASME 2012 31st international conference on ocean, offshore
and arctic engineering, Rio de Janeiro; 2012.
outer layer in the thinnest section. This indicates that the [16] Nasution FP, Sævik S, Gjøsteen JKØ, Berge S. Experimental and finite element
fatigue failures were governed by the local stress concen- analysis of fatigue performance for copper power conductor. Int J Fatigue
tration factors (SCFs) due to the effect of surface irregular- 2013;47:244–58.
[17] Sævik S. Theoretical and experimental studies of stresses in flexible pipes.
ities in the outer wires. Good correlation between tests and Comput Struct 2011;89:2272–91.
FE analysis was obtained. [18] Sævik S. On stresses and fatigue in flexible pipes. Ph.D thesis at Department of
Marine Structures. NTNU, Norway; 1992.
[19] UFLEX3D version 1.0.1 theory and user manual.

You might also like