Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Composites: Part A 40 (2009) 644–652

Contents lists available at ScienceDirect

Composites: Part A
journal homepage: www.elsevier.com/locate/compositesa

Stress and failure analysis of crimped metal–composite joints used


in electrical insulators subjected to bending
Alain Prenleloup a, Thomas Gmür a,*, John Botsis a, Konstantin O. Papailiou b, Kurt Obrist b
a
Ecole polytechnique fédérale de Lausanne (EPFL), Laboratoire de mécanique appliquée et d’analyse de fiabilité, Bâtiment ME, Station No. 9, 1015 Lausanne, Switzerland
b
Pfisterer-Sefag AG, Elektrotechnische Spezialartikel, Werkstraße 7, 6102 Malters, Switzerland

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

Article history: Experimental and numerical investigations are carried out on metal/fibreglass-reinforced-plastic joints
Received 17 July 2008 integrated in electrical insulators subject to bending. Numerical stress and strain distributions through
Received in revised form 20 February 2009 the bond are calculated with a solid 3D finite element model and the damage initiation in the composite
Accepted 24 February 2009
is highlighted. The simulations are compared to experimental data obtained from several joint specimens
tested under bending on an experimental setup equipped with strain gauges and a six-channel acoustic
emission system. Good correlation between the finite element predictions and the test results is found.
Keywords:
The investigations have identified the stress concentrations in the rod, the onset of damage when the
B. Stress concentrations
C. Finite element analysis (FEA)
load–displacement curve characterizing the bending test deviates from linearity, and the different failure
D. Acoustic emission mechanisms.
E. Joints/joining Ó 2009 Elsevier Ltd. All rights reserved.

1. Introduction In the last decade, several experimental and/or numerical stud-


ies have been reported on the stress, fatigue and failure analysis of
Metal–composite joints are increasingly used in industrial metal–composite single, double or tubular lap joints under trac-
equipments such as automotive, aircraft, aerospace or electrical tion, torsion, bending or multi-axial loading. However, all these
components. As a specific application of mixed joints encountered works are devoted to adhesively bonded joints which do not expe-
in high-voltage networks, silicon composite insulators with rience crimping prior to external mechanical loading. Only a few
crimped metal end-fittings manufactured for high-voltage trans- papers [1–9] have been published on the stress and failure analysis
mission lines or substations are now replacing conventional porce- of composite insulators consisting of a glass-reinforced polymer
lain elements due to the hydrophobic nature of their silicon rod with two metal end-fittings radially crimped onto the compos-
surface, their insensitiveness to dynamic loads such as short-cir- ite during assembly. In these works, numerical and experimental
cuits or earthquakes, and their light weight compared to porcelain investigations are thoroughly conducted on non-adhesively
which facilitates their handling, transport and installation. These bonded joints, but they are restricted to insulators subjected to ax-
structures, however, require two metal end-fittings in order to ial tensile loads. So far, only two studies [10–12] have been carried
transfer loads from the high-voltage conductor to the tower or out on the bending behaviour of crimped mixed metal–composite
the substation equipment. Two different designs are adopted insulators. According to the limited current research and develop-
according to whether the insulator is loaded primarily in bending ment on this type of joints, a systematic and integrated approach
or in traction. In the first case, the core of the insulator is chosen that would account for different types of configurations and lead
in the form of a large cylinder with aluminium fittings bonded to to realistic design methodology for crimped metal–composite
the composite tube with an epoxy adhesive layer. In the second sit- structural joints subject to bending is desirable.
uation, a thin rod-type insulator is preferred where steel end-fit- In comparison to single-material components, mixed structural
tings are crimped to the composite rod with a radial compression elements exhibit differences in the constitutive law, thermal or
by means of a hydraulic press until a sufficient plastic deformation hydrothermal response and failure characteristics between the
is initiated in the metal. Due to the lower assembly cost of the lat- metal, the composite and the possible bonding layer. This intro-
ter solution, the crimping strategy is nowadays also applied to duces zones of high stresses that induce damage initiation and
thick rod-type insulators (with generally a diameter above growth, leading to fracture onset and ultimately failure. Conse-
50 mm) subject to bending. quently, before such composite–metal joints can be used with con-
fidence in structural applications, a thorough characterization of
their mechanical and failure behaviour under static mechanical
* Corresponding author. Tel.: +41 21 6932924; fax: +41 21 6937340.
E-mail address: thomas.gmuer@epfl.ch (T. Gmür).
loads is required. The present paper is aimed at a critical evaluation

1359-835X/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compositesa.2009.02.020
A. Prenleloup et al. / Composites: Part A 40 (2009) 644–652 645

of the bending behaviour of specific steel–glass/epoxy crimped is measured as 56%. Geometrically, the pultruded rod considered
joints used in thick-rod composite insulators connecting trans- has a nominal diameter between 51 and 88 mm and a length,
formers and substations to overhead lines up to the highest voltage adjustable to customer’s need, ranging from 0.4 to 2.0 m. Tem-
levels. Numerical simulations of the stress distributions through- pered unalloyed carbon steel is chosen for the metal end-fittings
out the joints are performed by means of full 3D finite element fabricated directly by the manufacturer of the insulators. Achieved
models. The different damage mechanisms observed experimen- by the insulator manufacturer, the crimping of the end-fittings on
tally are identified and the main parameters influencing the tough- the composite rod is performed by means of an eight-jaw hydraulic
ness of the joints and the distributions of the internal stresses are press. A high internal pressure, corresponding to a crimping load of
investigated. An experimental setup based upon acoustic emission up to 600 kN and inducing a radial shortening of the steel adher-
and strain gauges is also developed in order to validate the numer- ends, is required in order to reach a prescribed value, depending
ical models and to follow, up to failure, the damage progress in the on the diameter of the rod, after relaxing the crimping pressure.
joints. This work should serve as a basic step towards the optimi- It should however be stressed that the crimping phase is not cru-
zation of the mechanical bearing capacity of thick-rod insulators cial for rods subjected to bending in contrast with insulators under
subject to bending. tensile loading where crimping has to prevent the rod from sliding
out of the end-fitting.
2. Description of the steel–glass/epoxy crimped joint under The values of the constitutive parameters for the epoxy resin
investigation and the ECR-glass fibres have been chosen in the literature. By
using the classical rule of mixtures applied to the whole homoge-
Manufactured by Pfisterer-SefagÒ (Switzerland), the thick rod- nized GRP core and knowing the volume fraction of the fibres, the
type composite insulators used in this study are made of an epoxy elastic properties of the unidirectional composite rod could be eas-
rod reinforced with ECR-glass fibres and of two steel fittings ily estimated. Some of these evaluated properties have been cor-
strongly crimped to both ends of the rod (Fig. 1). For preventing roborated with the values provided by the supplier of the
degradation in corrosive environments and electrical leakage cur- composite rod and have been confirmed with a normalized tensile
rents, the composite is covered with a silicone rubber housing or test performed on an in-house axial/torsion testing system. The
shed which is not considered in the analysis. latter has also been used for determining the elastic and plastic
The composite solid core, obtained by a pultrusion technique, is constitutive parameters of the steel end-fittings. No heat or surface
a unidirectional glass-reinforced polymer (GRP) rod. The matrix is treatment is applied to the fittings. The properties found for the
made of epoxy resin and the volume fraction of the ECR-glass fibres different materials are listed in Table 1 and the stress–strain graph
derived from the load–displacement curve measured for the steel
is shown in Fig. 2.

3. Numerical stress and failure analysis

3.1. Finite element models

Full 3D solid finite element models of the different joints have


been elaborated using the commercially available ABAQUSÒ soft-
ware. Non-linear simulations are performed successively on the
51- to 88-mm-thick composite insulators for the three following
steps: the crimping of the end-fitting onto the composite rod, the
relaxation of the crimping pressure and the bending of the entire
joint.
The numerical models are based upon hexahedral quadratic 20-
node 3D solid elements with a reduced integration scheme for the
computation of the structural components (C3D20R in the ABA-
QUSÒ element nomenclature). By symmetry, only a half of the
complete joint is modelled. The mesh, represented in Fig. 3 for a
63-mm-thick insulator, is composed of 112 finite elements per
hardened steel jaw or blade compressing the end-fitting on the
composite rod, 4915 elements for the steel half-fitting and 6912
elements for the composite half-cylinder. As illustrated in Fig. 3,
the grid for the composite rod is refined at the outlet of the end-fit-
ting since this zone is subject to stress concentrations during the
bending phase as shown in the next subsections.
For the crimping process, a Coulomb contact law is chosen for
symbolizing the contact at the interfaces between the press and
the end-fitting and between the latter and the composite rod. A
friction coefficient of 0.35, corresponding to the mean value of data
measured on several specimens, is adopted for both interfaces. This
value has been measured through a pull-out test applied on vari-
ous slices of a rod constrained by an end-fitting and also by means
of a tensile test on flat GRP specimens cut out of a rod and com-
pressed between two steel blocks with the same surface roughness
as on the inner surface of the end-fitting. For the bending step, a
Fig. 1. Schematic configuration of a silicon composite insulator. full damage tensor criterion (similar to the Tsai–Wu criterion) is
646 A. Prenleloup et al. / Composites: Part A 40 (2009) 644–652

Table 1
Material properties (engineering constants).

Component Young’s modulus Shear modulus Poisson’s ratio


E1a (GPa) E2, E3 (GPa) G23 (GPa) G31, G12 (GPa) m23 ( ) m31, m12 ( )
End-fitting 200 – – – 0.29 –
Rod 44 10.3 4.3 5.1 0.5 0.32
a
For the composite, subscripts 1, 2 and 3 refer to the fibre and both transverse directions.

Table 2
Ultimate strengths for the composite rod.

Direction Value (MPa)


Longitudinal (tension) 1038
Longitudinal (compression) 794
Transverse (tension) 32
Transverse (compression) 141
Longitudinal–transverse 54

crimping process is, as aforementioned, not critical for thick insu-


lators. Since these structures are primarily designed for resisting to
bending moments caused by loads mainly orthogonal to the longi-
tudinal or glass fibre axis, the crimping pressure is chosen suffi-
ciently low to avoid any damage of the composite rod. The stress
analysis is thus focused here on the behaviour of the structures
during the bending phase. Furthermore, the results presented are
Fig. 2. Stress–strain curve for the carbon steel used for manufacturing the end-
limited to the 51- and 63-mm-thick insulators since they are in es-
fittings. sence similar for all the range of diameters.
A mechanical loading of 9 and 10 kN is applied at the free end of
the 51- and 63-mm-thick models, respectively. Located in Fig. 4,
applied to the elastic transversely isotropic composite. Derived the distributions of the longitudinal and radial normal stresses
from in-house performed axial tensile and compression, in-plane rzz and rrr along the outer compressive fibre of the rod are shown
shear and torsion, and biaxial compression strength tests on spec- in Fig. 5. Situated at the outlet of the end-fitting, a maximum lon-
imens cut in rods of various diameters, the ultimate strengths mea- gitudinal stress rzz of 910 MPa and a highest radial stress rrr of
sured are shown in Table 2. These values have been obtained with 280 MPa are then reported in the composite for the thinner insu-
cylindrical longitudinal specimens (axial tension and compres- lator, while these values are reduced to 810 and 260 MPa for
sion), flat transverse coupons cut in thick rods (transverse tension the thicker composite rod. The distributions of the radial–axial
and compression), cylindrical samples with two half-section shear stresses srz on the axis of the rod and on the quasi outer com-
notches (longitudinal–transverse shear) and cubic specimens pressive fibre of the composite (Fig. 4) are illustrated in Fig. 6. It
(biaxial compression). should be noted that according to the hertzian contact law the
highest shear stresses within the built-in length appear in the com-
3.2. Numerical stress analysis posite at a distance from the interface between the rod and the
end-fitting corresponding to 3% of the radius of the rod. Maximum
It can be shown [11] that, contrary to the situation inherent in shear stresses srz of 70 MPa and 100 MPa are predicted on the
thin composite insulators subject to tensile loading [8,9], the axis of the 51-mm-thick rod in the middle of the end-fitting and

Fig. 3. Modelling of the composite joint with 3D finite elements (63-mm-thick composite insulator).
A. Prenleloup et al. / Composites: Part A 40 (2009) 644–652 647

Fig. 4. Loading description and location of reported stress distributions.

dictions given by the strength-of-materials theory for a cantilever


beam subject to bending with shearing. It should be observed that
within the bond all the normal and shear peak stresses overrun the
corresponding ultimate strengths (rlim or slim) given in Table 2 for
the longitudinal, transverse and longitudinal–transverse direc-
tions. Since at the loading level chosen the integrity of the insula-
tors is still preserved as it will be seen later, a hidden damage has
probably been initiated at the current stress state in the composite
rod, which will be confirmed in the section devoted to the experi-
mental analysis of the joints. As an additional information, the
intensity of the normal and shear stresses in the mid-plane of
the rod for the 63-mm-thick insulator is illustrated in Fig. 7, where
the stress concentrations are highlighted with a circle.

Fig. 5. Numerical distribution of the longitudinal and radial normal stresses for the
51- and 63-mm-thick composite rods loaded at 9 and 10 kN, respectively.

Fig. 6. Numerical distribution of the radial–axial shear stresses on the neutral axis
and on the nearly outer compressive fibre for the 51- and 63-mm-thick composite
rods loaded at 9 and 10 kN, respectively.

at the outlet of the fitting on the nearly outer compressive fibre,


respectively. For the 63-mm-thick insulators, the corresponding
stresses are equal to 65 and 90 MPa.
It can be seen in Figs. 5 and 6 that the stress distributions are
very similar for both insulators and that the peak stresses appear
at the same relative locations. Outside the bond, the normal and
Fig. 7. Numerical stress distribution in the joint for the 63-mm-thick composite rod
shear stresses follow exactly the linear and constant analytical pre- loaded at 10 kN (a circle indicates maximum stress location).
648 A. Prenleloup et al. / Composites: Part A 40 (2009) 644–652

3.3. Numerical damage analysis bond, play an important role in the optimization of the joints [8],
only the last two characteristics are crucial for the bending phase,
In order to verify the rod integrity, a Tsai–Wu-type damage ten- so that the sensitivity analysis is limited here to these parameters.
sor criterion is applied to the unidirectional composite. The results The sensitivity of the stress–strain state captured during the
are expressed in the form of a normalized index (stress ratio) cor- bending step to the radial shortening applied to the composite rod
responding to the inverse of the factor by which the applied bend- during the crimping phase is analysed for the 51-mm-thick insulator
ing load can be linearly magnified before failure [13], an index of 1 loaded at 9 kN when ranging the parameter between 70% and 140%
indicating the onset of damage. The predictions found for the dam- of its nominal value. The resulting maximum normal and shear
age criterion when subjecting the 51- and 63-mm-thick insulators stresses, referred to the corresponding ultimate strengths (Table
to a bending load of 9 and 10 kN, respectively, are shown in Fig. 8. 2), are illustrated in Fig. 9a, where it can be observed that the longi-
It can be observed that the unit limit value for the damage index is tudinal stress rzz on the outer fibre of the composite is only slightly
exceeded slightly but very locally. The critical zones, marked with affected by a change in radial shortening, but the shear stress srz on
circles in the figure, are the middle of the 63-mm-thick composite the neutral axis is drastically reduced by almost 40% with a 40% in-
rod, where the shear stresses are dominant, and for both insulators crease of this parameter. For the same increase, the radial normal
the vicinity of the two contact points with the end-fitting, sub- stress rrr on the outer fibre and the shear stress srz on the nearly out-
jected to high radial stresses and located at the beginning of the er fibre are however increased by up to 40% since the crimping effect
rod for the part of the composite under tension and at the bond is more pronounced with a larger radial shortening.
end for the part in compression. At this loading intensity, only local The sensitivity of the stresses to the length of the joint is next
damage is however initiated, which explains that the mechanical studied, again for the 51-mm-thick insulator under a bending load
bearing capacity of these insulators is actually much higher, as it of 9 kN. The parameter is varied from 75% to 150% of the nominal
will be shown later. It should also be added that the zone appar- value chosen as reference. Presented in Fig. 9b, the results show
ently critical on the end face of the composite is not due to the con- that all the maximum normal and shear stresses, referred to the
tact of the rod with the bottom of the end-fitting (a sufficiently corresponding ultimate strengths (Table 2), are decreasing with
large gap is present between the rod and the bottom of the fitting), an enlargement of that parameter, especially the shear stress srz
but is in fact to be attributed to the warping of the end face, induc- predicted on the neutral axis which is reduced by more than 40%
ing normal radial and hoop tensile stresses which overrun in this for an increase by 50% of the built-in distance. It is thus seen that
region the relatively low ultimate transverse strength in tension
(Table 2); the resulting damage index reported on the end face of
the rod is however not critical for the bending phase of the insula-
tors, since the corresponding local stress–strain state does not af-
fect the overall bending stiffness of the structure.

3.4. Numerical sensitivity analysis

The numerical stress and damage study is here completed by a


brief sensitivity analysis of the main parameters influencing the
load bearing capacity of the joints when subjected to bending.
Although for the crimping step several parameters such as the fric-
tion coefficient between the GRP rod and the end-fitting, the size of
the initial gap between the rod and the fitting, the thickness of the
end-fitting, the radial shortening of the rod and the length of the

Fig. 8. Normalized damage index for the (a) 51- and (b) 63-mm-thick composite Fig. 9. Sensitivity of the maximum stresses to the radial shortening and the bond
rods loaded at 9 and 10 kN, respectively (a circle indicates highest index location). length in the 51-mm-thick composite rod.
A. Prenleloup et al. / Composites: Part A 40 (2009) 644–652 649

a longer end-fitting contributes to reduce the failure index and,


consequently, to delay the onset of damage.

4. Experimental stress and failure analysis

4.1. Description of the experimental setup

In order to validate the numerical stress state and damage index


generated during the bending phase of the insulators, a specific
experimental setup has been developed for the analysis of the
composite–metal joint under flexure.
Applying a transverse load at the free end of a built-in insulator,
a servo-hydraulic test frame, provided by the manufacturer of the
insulators, has been constructed for subjecting joint specimens to
bending loads up to 50 kN (Fig. 10). The experimental system is
equipped with a hydraulic jack and a load cell provided by Vibrom-
eterÒ (Switzerland), and instrumented with an LVDT (linear vari-
able differential transformer) displacement transducer. The
regulation loop for the control of the hydraulic group and the data
transfer are performed with a multi-channel acquisition card
implemented on a PC. While the end-fitting is firmly bolted to Fig. 11. Location of the strain gauges (a) and the acoustic emission microphones (b)
(51-mm-thick composite insulator).
the rigid frame in order to obtain perfect clamping conditions,
the free end is coupled to the hydraulic jack through a ring placed
in a circular groove manufactured on the composite and by means 4.2. Experimental stress analysis
of a universal joint for avoiding any parasitic axial loading.
For the stress analysis, six HBMÒ 3/350LY13 resistive strain Several 51-, 63- and 88-mm-thick insulators have been assem-
gauges are pasted at different locations on the composite rod and bled by the manufacturer of the joints by using the nominal pres-
the outer surface of the fixed end-fitting (Fig. 11a). For the failure sure [8] for crimping the end-fittings on the composite rod. These
study, a six-channel acoustic emission (AE) system (Mistras joints are next mounted in the aforementioned test frame and a
2001) provided by Euro Physical AcousticsÒ is added to the test transverse load is progressively applied with a controlled loading
equipment to shed light in the process of damage activity on the velocity (protocol fixed by IEEE standards) at the free end of the
joint as a function of the bending moment (Fig. 11b). specimens clamped at the other side.

Fig. 10. Experimental setup used for the joint specimens subject to bending (63-mm-thick composite insulator).
650 A. Prenleloup et al. / Composites: Part A 40 (2009) 644–652

The experimental load–displacement curves (deflection at the


LVDT location) and the corresponding strain–displacement curves
for the six strain gauges pasted on the rod and the end-fitting
(Fig. 11a) are recorded for each specimen. Since the graphs are
nearly coincident for all the joints, only the results obtained with
one of the 51-mm-thick insulators are considered here. The data
are reported in Fig. 12, where it can be observed that up to 9 kN
the load and strain curves are linear with the transverse displace-
ment and that the metal fitting even experiences none deformation
as highlighted by the response of the strain gauge pasted on the
adherend (gauge #1). By increasing the load up to 13.5 kN, all
the curves deviate from linearity. The signals of the strain gauges
on the rod (gauges #3, #4 and #6) show that the composite is sub-
jected to high deformations and the gauge #1 confirms that the
steel end-fitting undergoes a plastic deformation at this load level.
When the transverse load is still increased from 13.5 kN up to the
ultimate load of 16 kN, the signals of the strain gauges are lost one
by one and the joint fails as described in the next subsection. Fig. 13. Numerical and experimental load–displacement curves for three 51- and
The simulated load–displacement curves for both the 51- and 63-mm-thick insulators.
63-mm-thick insulators are compared in Fig. 13 with the corre-
sponding experimental data (three specimens have been tested
for each diameter). For the sake of clarity, the curves are not fully severe AE activity appears at the onset of non-linearity (10 kN) in
plotted after reaching the maximum mechanical load. An excellent the load–displacement curve and that the intensity of the activity
agreement between the two sets of results can be observed in the increases gradually up to the ultimate mechanical loading. A slight
nearly linear behaviour range of the experiments. It could also be AE activity can however already be observed at load levels below
shown that in this range the numerical strain predictions fully the limit of linearity, indicating a possible damage initiation in
agree with the measurements derived from the strain gauges. the composite rod.
However, the onset of non-linearity in the load–displacement The damage mechanism for the 51-mm-thick composite rod is
curves suggests that a local damage is initiated in the composite shown in Fig. 15, where it is seen that failure is caused by a trans-
rod at 50–60% of the maximum mechanical loading, which corre- verse crack propagation due to the normal stresses in the compres-
sponds to a bending load of 9–10 kN. This confirms the suspicion sive part of the rod near the outlet of the end-fitting. This
of a damage initiation as already mentioned in Sections 3.2 and 3.3. phenomenon, due mainly to the bending of the rod and not the
crimping of the end-fitting, is to be related with the damage index
4.3. Experimental failure analysis shown in Fig. 8a which reaches its highest value precisely in the
zone near the initiation of the transverse cracks. The observation
In parallel to the experimental stress analysis described in the of the damage area with a microscope reveals that near the outer
previous subsection, the test frame is also used to highlight the ini- surface of the composite rod the damage corresponds to a fracture
tiation of the local damage in the different insulators instrumented of the polymer matrix and the glass fibres (Fig. 16a), whereas in-
with the AE sensors. While the metal–composite joints are loaded side the composite the damage is governed by a micro-buckling
up to failure, the progress of the damage activity is measured with
the AE equipment. The maximum amplitude (highest peak volt-
age), expressed on a logarithmic scale (1 lV = 0 dB), of the AE
waveforms and the so-called AE energy of the measured acoustic
activity [14] in function of the applied bending load is illustrated
in Fig. 14 for a 51-mm-thick insulator. The LVDT deflection in func-
tion of the applied load is also given. It is clearly apparent that a

Fig. 12. Load–displacement and strain–displacement curves for the 51-mm-thick Fig. 14. Intensity of the acoustic emission activity in function of the bending load
insulator (the location of the strain gauges is shown in Fig. 10a). amplitude for the 51-mm-thick insulator.
A. Prenleloup et al. / Composites: Part A 40 (2009) 644–652 651

Fig. 17. Failure mode of the 63-mm-thick insulator: transverse and longitudinal
Fig. 15. Failure mode of the 51-mm-thick insulator: transverse cracking.
cracking.

end-fitting. They have also highlighted that the damage is initiated


in the composite rod when the load–displacement curve character-
izing the bending test deviates from linearity, the acoustic emis-
sion activity being then drastically increased. Moreover, an
excellent correlation between the numerical simulations and the
experimental results issued from several test specimens have been
observed, at least in the linear range of loading. Micrographs of the
failed joints have allowed identifying the damage mechanisms
which are different according to the thickness of the composite
rod. The analysis has also shown that the some geometrical param-
eters are crucial for the integrity and performance of the joint.

Acknowledgements

The authors wish to thank Philippe Bonhôte (HEIG-VD, Yverdon,


Switzerland) for his careful advices and appreciated help in several
aspects of this study. This work was partially supported by the
Swiss Commission for Technology and Innovation (CTI), Grant
No. 6798.1.
Fig. 16. Micrographs of the fibre and matrix breaking (a) and of the fibre micro-
buckling (b) for the 51-mm-thick insulator.
References

[1] Bansal A, Schubert A, Balakrishnan MV, Kumosa M. Finite element analysis of


of the fibres (Fig. 16b). It should be noted that the delamination substation composite insulators. Compos Sci Technol 1995;55(4):375–89.
visible on the tensile part of the rod (Fig. 15) is a post-failure dam- [2] Lanteigne J, Lalonde S, de Tourreil C. Optimization of stresses in the end-
fittings of composite insulators for distribution and transmission lines. J Reinf
age induced by the relaxation of the bending load.
Plast Comp 1996;15(6):467–78.
For the 63- and 88-mm-thick insulators, a longitudinal crack [3] Bansal A, Kumosa M. Mechanical evaluation of axially loaded composite
propagation due to the shear stresses in the centre of the rod oc- insulators with crimped end-fittings. J Compos Mater 1997;31(20):2074–104.
curs in addition to the transverse crack propagation, as shown in [4] Kumosa M, Han Y, Kumosa L. Analyses of composite insulators with crimped
end-fittings: Part I. Non-linear finite element. Compos Sci Technol
Fig. 17 for the 63-mm-thick insulator. This observation is again 2002;62(9):1191–207.
confirmed by the distribution of the damage index in Fig. 8b, where [5] Kumosa M, Armentrout D, Kumosa L, Han Y, Carpenter SH. Analyses of
it is seen that a second critical failure area appears in the middle of composite insulators with crimped end-fittings: Part II. Suitable crimping
conditions. Compos Sci Technol 2002;62(9):1209–21.
the thicker composites. [6] Mobasher B, Kingsbury D, Montesinos J, Gorur RS. Mechanical aspects of
crimped glass reinforced plastic (GRP) rods. IEEE Trans Power Deliv
2003;18(3):852–8.
5. Conclusions
[7] Duriatti D, Béakou A, Levillain R. Optimisation of the crimping process of a
metal end-fitting onto a composite rod. Compos Struct 2006;73(3):278–89.
The stress behaviour and the failure process under bending [8] Prenleloup A, Gmür T, Botsis J, Papailiou KO. Acoustic emission study and
loading of an electrical silicon composite insulator made of a strength analysis of crimped steel–composite joints under traction. Compos
Struct 2006;74(3):370–8.
glass/epoxy rod with crimped steel end-fittings have been ana- [9] Prenleloup A, Gmür T, Bonhôte P, Botsis J, Papailiou KO. Experimental and
lysed in this paper. The metal/composite joint has been studied numerical strength analysis of mixed metal–composite crimped or adhesively
numerically with a 3D finite element model and experimentally bonded joints. In: Proceedings of CompTest 2006 conference, Porto, April,
2006. p. 25–6.
with a test equipment including an acoustic emission system for [10] Prenleloup A, Gmür T, Botsis J, Papailiou KO, Obrist K, Bonhôte P. Acoustic
monitoring the damage progression during the solicitation. The emission inspection and analysis of crimped metal–composite joints subjected
investigations have shown that stress concentrations appear in to bending. In: Proceedings of fourth ICNDT conference, Chania, October, 2007.
Paper 5-2.
the composite near the interface with the steel fitting at the outlet [11] Prenleloup A. Analyse de l’état de contrainte et de l’endommagement
of the joint and on the mid-surface of the rod in the middle of the d’assemblages sertis en matériau mixte métal–composite sollicités en
652 A. Prenleloup et al. / Composites: Part A 40 (2009) 644–652

traction ou en flexion. Ph.D. Thesis No. 4005. Lausanne: Ecole polytechnique [13] Kollár LP, Springer GS. Mechanics of composite
fédérale de Lausanne; 2008. structures. Cambridge: Cambridge University Press; 2003.
[12] Dumora D, Feldmann D, Gaudry M. Mechanical behavior of flexurally stressed [14] Pollock AA. Acoustic emission inspection. In: Davis JR, editor. Metals
composite insulators. IEEE Trans Power Deliv 1990;5(2): handbook. Nondestructive evaluation and quality control, vol. 17. Metals
1066–73. Park, OH: ASM International; 1989. p. 278–94.

You might also like