You are on page 1of 9

850

Development of glass fiber reinforced plastic


poles for transmission and distribution lines
Sherif Ibrahim, Dimos Polyzois, and Sherif K. Hassan

Abstract: An extensive research project is currently being carried out at The University of Manitoba, Canada, to de-
velop lightweight glass fiber reinforced plastic (GFRP) poles for use in transmission and distribution networks. In this
paper, results from tests involving full-scale tapered GFRP poles with a hollow circular cross section subjected to canti-
lever bending are presented. The filament winding process was employed to produce the poles using polyester resin re-
inforced with E-glass fibers. Cantilever bending tests were conducted on twelve full-scale poles up to failure. Test
parameters included fiber orientation and number of layers. Extensive theoretical work preceded the test program and a
theoretical model was developed for evaluating the failure load. The results to date indicate that the developed theoreti-
cal model can predict quite well the ultimate capacity and behavior performance of GFRP poles. This theoretical model
was used in this investigation to determine the optimum cross-sectional dimensions for 6.1 m (20 ft) and 18.3 m
(60 ft) Class 1 GFRP poles.

Key words: transmission and distribution poles, filament winding, fiber-reinforced plastics.

Résumé : Un projet de recherche approfondie est présentement dirigé à l’Université du Manitoba, Canada, pour déve-
lopper des poteaux en plastique léger renforcé de fibres de verre (PRFV) pour une utilisation dans des réseaux de
transmission et de distribution. Dans cet article, des résultats de tests impliquant des poteaux PRFV en fuseau à échelle
grandeur nature avec une coupe transversale à ouverture circulaire et soumis à un fléchissement en porte-à-faux sont
présentés. Le processus d’enroulement des filaments fut employé pour produire les poteaux en utilisant de la résine en
polyester renforcée avec des fibres de verre de type E. Des tests de fléchissement en porte-à-faux furent conduits sur
douze poteaux à échelle grandeur nature jusqu’à la rupture. Les paramètres des tests ont inclus l’orientation des fibres
et le nombre de couches. Un large travail théorique a précédé le programme de tests et un modèle théorique fut déve-
loppé pour évaluer la charge de rupture. Les résultats à ce jour indiquent que le modèle théorique développé peut bien
prédire la capacité ultime et la performance du comportement des poteaux (PRFV). Ce modèle théorique a été utilisé
dans cette étude pour déterminer les dimensions optimales de la coupe transversale pour un poteau PRFV Classe 1 de
6,1 m (20 pd) et 18,3 m (60 pd).

Mots clés : poteaux de transmission et de distribution, enroulement des filaments, plastiques renforcés de fibres.
[Traduit par la Rédaction] Ibrahim et al. 858

Introduction The main disadvantage of concrete poles is their weight,


which drastically increases transportation and erection costs.
Traditional materials such as wood, steel, and concrete are Chemical influences on the concrete surfaces due to envi-
commonly used as poles in electrical transmission and distri- ronmental impact can also affect their long-term perfor-
bution aerial networks. However, the shortage of wooden mance. As in the case of other concrete structures, concrete
poles, their short life expectancy, and various environmental poles are subject to corrosion of the steel reinforcement, re-
concerns related to their treatment have promoted hydroelec- sulting in further strength deterioration and expensive main-
tric utility companies to search for a cost-effective alterna- tenance. Steel is the most common material for the
tive. Wooden poles are continuously exposed to weather, construction of transmission poles in North America. These
fungi, woodpeckers, etc., which result in a very significant poles, however, are very expensive (Escher 1982). Corrosion
deterioration of their load bearing capacity with time. The protection is of primary concern in steel poles, which must
service life of wooden poles is approximately 20 years be painted or galvanized, a process that does not always
(Vanderbuilt and Criswell 1988). Any extension of this ser- guarantee long-term protection. Generally, traditional poles
vice life requires continuous inspection and follow-up care. made of wood, concrete, or steel are subject to deterioration
In a number of European countries, concrete poles are used. under environmental attacks. Regular maintenance is essen-
tial to prolonging serviceability of these poles.
Received July 28, 1999.
Revised manuscript accepted November 25, 1999. On the other hand, glass fiber reinforced plastic (GFRP)
poles are lightweight and corrosion resistant. Being light-
S. Ibrahim, D. Polyzois, and S.K. Hassan. Department of weight is a major advantage of GFRP poles, making them
Civil and Geological Engineering, The University of suitable for transportation and installation in mountain ter-
Manitoba, Winnipeg, MB R3T 5V6, Canada. rains, marshes, and remote or inaccessible areas. Although
Written discussion of this article is welcomed and will be the initial cost of GFRP poles may be higher than that of tra-
received by the Editor until February 28, 2001. ditional poles, the long-term benefits and the savings in
Can. J. Civ. Eng. 27: 850–858 (2000) © 2000 NRC Canada
Ibrahim et al. 851

transportation and erection that these poles provide make Table 1. Properties of E-glass and vinylester resin.
their selection attractive. The use of GFRP poles is not new.
Vinylester resin
A number of companies are already involved in the produc-
Properties E-glass (DERAKANE 470-300)
tion of such poles. Research in this area, however, is limited.
In 1988, Bell initiated its own investigation of GFRP poles Tensile modulus (GPa) 72.4 3.58
and conducted an experimental program at the Centre de Re- Poisson’s ratio 0.2 0.3
cherche du Reseau Exterieur (CERRE) (McClure et al. Tensile strength (MPa) 2400 85
1992). The specimens were tapered with a hollow cross sec- Shear modulus (GPa) 30 1.38
tion and were manufactured by centrifugal casting. The test Density (gm/cm3) 2.54 1.08
results indicated that GFRP poles could safely resist loads
comparable to those of wooden poles. The behavior of these
GFRP poles was found to be truly elastic even for large de- matched the outer diameter of the specimens. The two seg-
flections. ments of the base were clamped together using four dwyidag
An experimental investigation was also conducted by bars, as shown in Fig. 1. To ensure the stability of the por-
Shakespeare Inc. on filament wound GFRP poles in 1993 tion of the specimen within the base, a GFRP tapered sleeve,
(Derrick 1996). Class 4 GFRP poles were tested at the na- 1000 mm long and 25 mm thick, was inserted in the speci-
tional test labs of Engineering Data Management (EDM) in men. The concrete base was clamped to a rigid structural
Colorado, U.S.A. The tests were successful, as each pole wall using four hollow square steel (HSS) sections (250 ×
met or exceeded the strength requirement specified for 250 × 6 mm), two on each side of the base, as shown sche-
wooden poles. matically in Fig. 1. Four high strength dwyidag bars were
The objectives of the research program at The University used to clamp the HSS section and the concrete base to the
of Manitoba, Canada, were to (i) develop a theoretical model structural wall. To ensure full fixity of the specimen within
for determining the ultimate strength and performance of the concrete base, the gap between each specimen and the
GFRP poles, (ii) evaluate the developed theoretical model concrete base was filled with plaster of paris.
through a series of small- and full-scale testing, and (iii) de- The load was applied horizontally, 600 mm below the top,
velop design guidelines for the use of GFRP. according to the ANSI Standard 05.1 (1992). An electronic
To date, twelve small-scale specimens (2.5 m long) and load cell was used to monitor the applied load, as shown in
twelve full-scale specimens (6.25 m long) have been tested Fig. 1. Loading was applied at a rate of approximately
under cantilever bending load up to failure. This paper pres- 0.25 mm/s.
ents the results from the full-scale tests. The lateral deflection of the poles at the loading position
was monitored through an electronic linear measurement
transducer (LMT). The LMT was mounted on a fixed steel
Experimental program column 10 000 mm away from the specimen. The stroke
range for each LMT was 2500 mm.
Specimens and test setup Two LMTs attached on two opposite sides of the poles,
The poles were fabricated through the filament winding approximately 1000 mm above the base, were used to moni-
wet process at the ISIS–Faroex Filament Winding Research tor the change in the specimen diameter, as shown in Fig. 1.
Facility in Gimli, Manitoba. E-glass fibers (1100 TEX) and These LMTs had a range of 350 mm. The difference in the
vinylester resin (DERAKANE 470-300) were used for man- readings of the two LMTs represented the change in the
ufacturing these poles. The material properties and strength specimen diameter at that height. This method allowed con-
for both the fiber and the resin were provided by the manu- tinuous monitoring of the change in the diameter with load-
facturer (Dow Plastics Company 1997) and are presented in ing.
Table 1. Strain along the specimen near the base was monitored
The specimens were tapered hollow sections, 6250 mm in through 24 electrical resistance strain gauges. The strain
length. The inner diameters at the base and at the top were gauges had a 5 mm gauge length and a 120 W electrical re-
416 and 305 mm, respectively. The wall thickness for each sistance. Three strain gauges were mounted at eight different
specimen varied depending on the number of layers, which locations to measure strains in the longitudinal and cir-
ranged from 4 to 8, in two layer increments, giving a total cumferential directions as well as at 45° off-axis. Four strain
thickness of between 2.75 and 5.5 mm at the base. Three fi- gauges were located on the tension side and the other four
ber angles with respect to the longitudinal axis of the pole on the compression side at heights 30, 400, 800, and
were used: ±5, ±10, and ±20. Hoop winding was also em- 1200 mm above the concrete base.
ployed in ten of the specimens with different ratios, while The load cell, the LMTs, and the strain gauges were con-
only two of the specimens were fabricated without hoop nected to a computer-controlled 32-channel data acquisition
winding. The fiber volume fraction was measured during the system to monitor and record all data. Visual observations
manufacturing process by determining the weight of fiber were also made during the test. The load was applied until
and resin used and transforming those weights into volume complete failure of the specimen occurred.
fractions. The configurations as well as the fiber volume
fraction for the tested specimens are listed in Table 2. Test results
The fixed support for the specimens consisted of a square A summary of all test results is shown in Table 3. The
reinforced concrete base measuring 800 mm wide by equivalent wooden pole class is also shown in this table. The
1000 mm high. It was made up of two segments with a ta- classification method of distribution and transmission poles
pered circular hole in the middle with dimensions that is used to compare the strength of poles made from different
© 2000 NRC Canada
852 Can. J. Civ. Eng. Vol. 27, 2000

Table 2. Configuration and fiber volume fraction of full-scale specimens.

Specimen Longitudinal Number of Total number Base thickness Fiber Total weight
No. fiber orientation hoop layers of layers (mm) volume (%) (kg)
1 (10/–10) 0 8 5.5 49 99.40
2 (10/–10) 2 8 5.0 60 80.25
3 (10/–10) 4 8 4.5 60 68.15
4 (10/–10) 2 6 4.0 54 60.85
5 (10/–10) 2 4 2.4 60 39.00
6 (20/–20) 0 8 5.5 53 94.30
7 (20/–20) 2 8 5.0 60 81.00
8 (20/–20) 4 8 4.9 58 72.05
9 (20/–20) 2 6 4.0 57 59.40
10 (20/–20) 2 4 2.75 51 38.20
11 (5/–5) 2 6 4.6 50 62.50
12 (5/–5) 2 4 2.8 52 38.00

Fig. 1. Schematic drawing of the full-scale test setup. tudinal layers and two circumferential layers), had the high-
est performance ratio of approximately 0.47 kN/kg. Speci-
mens 3 and 8, both consisting of eight layers (four
longitudinal layers and four circumferential layers), had a
load-to-weight ratio of 0.4 kN/kg. Specimens 1 and 6, both
consisting of eight longitudinal layers without any cir-
cumferential layers, did not perform as well as the other
eight-layer pole group and had a load-to-weight ratio of 0.32
kN/kg.
A significant drop in the performance ratio was observed
when the number of layers was reduced to six or four layers.
This was due to the high reduction in the wall thickness,
which in turn reduced the buckling load significantly. As
shown in Table 3, the weight ratio for specimen 4 was ap-
proximately 0.34 kN/kg. This is almost the same value as
that of specimen 1, which had eight longitudinal layers with-
out circumferential layers. This result indicates the impor-
tance of including circumferential layers to improve the
performance of poles.
materials under the effect of lateral loads and is employed
The load capacity-to-weight ratio of a class 1 6.1 m (20 ft)
for preliminary design purpose only. It should be noted that
if poles made from different materials share the same class, long wooden pole is 0.095 kN/kg. The load-to-weight ratio
this does not mean that they would have the same strength for an equivalent class 1 GFRP pole (specimen 4 or 11) is
under different loading conditions such as combination of 0.3 kN/kg. This means that GFRP poles are three times more
axial and lateral loading. Specimen 2 failed prematurely in- efficient than wooden poles and are approximately only one
side the concrete base because of lack of internal stiffener in third the weight of equivalent wooden poles.
this region. Two specimens (4 and 11) satisfied the require- The number of circumferential layers has substantial in-
ment of class 1 poles, according to ANSI Standard 05.1 fluence on the local buckling failure load. Circumferential
(1992), by carrying an ultimate load equal to or exceeding layers provide confinement for the longitudinal fibers. For
20 kN (4500 lb). The top displacement of those two speci- GFRP poles with the same number of longitudinal layers,
mens was less than 10% of the free height of the pole. Spec- higher local buckling failure load can be achieved by incor-
imens 4 and 11 consisted of only 6 layers and weighed about porating more number of circumferential layers. This was
60 kg. Specimen 9, which also consisted of 6 layers but had observed on specimens 3 and 4. Both specimens have four
a fiber orientation of ±20°, failed at an ultimate load of 18.9 longitudinal layers; however, specimen 3 has four circumfer-
kN. This is less by 5% than that required for class 1, so this ential while specimen 4 has two circumferential layers. Con-
specimen was classified as class 2 pole. Five specimens (1, sequently, the local buckling failure load of specimen 3 is
3, 6, 7, and 8) exceeded the class 1 requirement and were about 45% higher than that of specimen 4. The same phe-
classified in the heavier categories (H2 and H3). The maxi- nomenon was observed for specimens 8 and 9.
mum lateral displacement of the poles tested was less than The failure mechanism of specimen 6 is shown in Fig. 2.
12.7%. A diagonal crack extending to a height between 200 and
Since each specimen had a different weight and a different 1200 mm above the base developed on the compression side.
load carrying capacity, the load capacity-to-weight ratio Specimen 6 was made from eight longitudinal layers at ±20°
(kN/kg) was used to compare the performance of the various with respect to the vertical axis. In this case, the longitudinal
specimens. Specimen 7, consisting of eight layers (six longi- bending stresses would have substantial components in the

© 2000 NRC Canada


Ibrahim et al. 853

Table 3. Performance of full-scale GFRP poles.

Specimen Weight Ultimate Top displacement Equivalent Displacement as Load/weight


No. (kg) capacity (kN) (mm) wooden pole class % of free height (kN/kg)
1 99.40 34.2 486 H3 9.7 0.34
2 80.25 * * * * *
3 68.15 30.1 608 H2 12 0.44
4 60.85 20.8 430 1 8.6 0.34
5 39.00 7.55 240 6 4.8 0.19
6 94.30 30.4 431 H2 8.6 0.32
7 81.00 36.4 634 H3 12.7 0.47
8 72.05 28.1 593 H2 11.9 0.39
9 59.40 18.9 416 2 8.3 0.32
10 38.20 6.9 252 6 5.0 0.18
11 62.50 20 403 1 8.1 0.32
12 38.00 8.8 278 5 5.6 0.23
*This specimen failed prematurely as a result of the lack of internal stiffener within the concrete base.

transverse to fiber direction as well as in-plane shear stress. Fig. 2. Diagonal fracture failure mode.
Since GFRP has very low strength in the transverse direc-
tion as well as low in-plane shear, failure in the transverse to
fiber direction would occur before local buckling load is
reached. On the other hand, specimen 1 was made from
eight longitudinal layers at ±10°. Because of that small fiber
angle, the component of the longitudinal bending stress in
the transverse to fiber direction was not critical and failure
due to local buckling was observed 200 mm above the base.
Moreover, local buckling failure was combined with exces-
sive splitting and delamination of the fibers because of the
lack of any confinement.
All specimens with circumferential layers (with the excep-
tion of specimen 2) failed by local buckling on the compres-
sion side of the specimens at a height that varied from 200
to 800 mm above the concrete base, as shown in Fig. 3. In
those specimens, circumferential layers resisted all the trans-
verse and in-plane shear stresses and prevented any failure in
the transverse direction.

Theoretical analysis
Material properties and strength
Glass fiber-reinforced plastic was used as the construction
material for the poles investigated. The lamina properties
were derived from the material properties of the E-glass fi-
ber and the vinylester resin given in Table 1.
These properties were used to derive effective moduli of the fact that the stresses in fibers and matrix are not equal
the composite material based on micromechanical models. under the corresponding loading conditions. The Tsai–Hahn
The rule of mixture, verified by Adams (1987), was used to equations can be expressed in terms of the fiber and matrix
evaluate the modulus of elasticity in the fiber direction (E1) properties as follows:
and the major Poisson’s ratio (m12) as follows:
(uf + hum )
[1] E1 = Ef uf + Em um [3] E2 =
é uf hum ù
[2] m12 = m f uf + m m um ê + ú
ë Ef Em û
where uf and um are the fiber and the matrix volume ratio, (uf + hum )
respectively; Ef and Em are the fiber and the matrix modulus [4] G12 =
of elasticity, respectively; mf and mm are the fiber and the ma- é uf hum ù
ê + ú
trix Poisson’s ratio, respectively. ëGf Gm û
For calculating the effective Young’s modulus in the trans-
verse direction (E2) and the shear modulus (G12), the Tsai– where h is the stress partitioning parameter. It was found that
Hahn approach (1980) was used. This approach is based on a value of h = 0.5 yields accurate predictions based on com-

© 2000 NRC Canada


854 Can. J. Civ. Eng. Vol. 27, 2000

Fig. 3. Local buckling failure mode.

parison of theoretical results and experimental data for glass as well as the large deflection at the top of the poles. The
epoxy (Tsai and Hahn 1980). Gf and Gm are the fiber and specimens were discretized into 784 elements as shown in
the matrix shear modulus, respectively. Fig. 5. The circumference was divided into 16 elements. In
Equations [1]–[4] were used to determine the effective the longitudinal direction, the element length was taken
material properties for the composite lamina. The fiber vol- equal to 100 mm at the base and gradually increased to
ume ratio, uf , and the matrix volume ratio, um, were mea- 185 mm near the top. This variation in element size was
sured during the manufacturing process. used to provide a fine mesh near the area of maximum stress
The strength of the composite can be determined in a sim- and anticipated failure zone. The model developed was used
ilar manner based on a micromechanical model and ad- to predict the failure load, whether this was due to instability
vanced elasticity methods. The strength of the composite (local buckling) or material failure. To determine the ulti-
depends on the relative properties of fibers and matrix and mate load of the modeled GFRP poles, a load step was spec-
their corresponding volume fractions. Detailed calculations ified. This load step was automatically divided into unequal
of strength are given in Adams and Doner (1967) and substeps whose range depended on the behavior of the pole
Agarwal and Broutman (1980). in the previous substep. At the end of each substep, the
ANSYS program adjusted the stiffness matrix to reflect any
Finite element model nonlinear changes in the pole stiffness. Ultimate loads were
The ANSYS program (1995) was used to develop a non- determined on the basis of the Tsai and Wu (1971) failure
linear finite element model of GFRP poles. The geometric criterion unless divergence was detected due to structural in-
boundaries of the poles were defined as the first step in the stability (i.e., local buckling) before reaching this failure cri-
modeling process. Eight-node quadrilateral layered shell ele- terion.
ments were used to model the poles as shown in Fig. 4. For A comparison between the experimental and the predicted
each layer within the element, orthotropic material proper- ultimate loads for the tested specimens is shown in Fig. 6.
ties in the fiber and transverse to the fiber direction were de- The average ratio of the experimental-to-theoretical ultimate
fined. Fiber orientation for each layer was specified by load was approximately 0.99 with a standard deviation of
defining the fiber angle with respect to the element axes as 12%. As evident from this figure, there is a strong correla-
shown in Fig. 4. Detailed formulation of this element is tion between the results obtained from the proposed model
given in Yunus et al. (1989). The portion of the specimen and the experimental results.
that was embedded in the concrete base was also included in The load–deflection curves obtained from both the experi-
the model. In the current study, a geometric nonlinear analy- mental and the finite element results are shown in Fig. 7 for
sis was used, taking into account the cross-section distortion specimens with ±10° longitudinal fiber angle. It can be seen
© 2000 NRC Canada
Ibrahim et al. 855

Fig. 4. Eight-node layered shell element.

that the load–deflection relationship obtained from the finite Fig. 5. Finite element mesh for analysis of test specimen 4 with
element is stiffer than the experimental results. In addition, cantilever height of 5 m and fiber volume of 54%.
the finite element results are linear because progressive fail-
ure was not accounted for in the analysis.
When a GFRP pole is subjected to bending moment, the
tensile and compressive stresses resulting from the applied
moment tend to flatten or ovalize the cross section. This re-
duces the flexural stiffness of the GFRP pole as the curva-
ture increases. Moreover, the deformation of the cross
section increases the axial bending stresses and lowers the
local buckling load (Ibrahim and Polyzois 1999). The load
versus the ratio of the change in cross-section radius
(ovalization), at 1 m above the concrete base for the ±10°
specimens, obtained from the finite element model as well as
the experimental investigation is shown in Fig. 8. It is evi-
dent in this figure that the load–ovalization curve obtained
from the finite element model has the same nonlinear trend
as the experimental results. This figure indicates the ability
of the finite element model to account for the nonlinear dis-
tortion of the cross section. The ovalization predictions of
the finite element model are less than the experimental mea-
surements because progressive failure was not accounted for
in the analysis.

Optimization study
An extensive theoretical study involving 6.1 m (20 ft) and
18.3 m (60 ft) GFRP poles was performed to determine the
optimum cross-section diameter of these poles that would
meet the same strength criteria as class 1 wooden poles. Ac-
cording to ANSI Standard 05.1 (1992), class 1 wooden poles
are required to have a load carrying capacity of 20 kN
(4500 lb) applied horizontally at 600 mm (2 ft) from the top
of the poles. In the optimization procedure, the fiber angles
were assumed to be ±10° with the two external layers placed
in the circumferential directions, one inner and one outer
layer. The fiber volume ratio of 0.6 was used in calculating
the material properties. The taper ratio of the pole consid-
ered was 0.83% (1 in./10 ft). The embedded portion of the tively. For the 6.1 m (20 ft) specimen, the bottom diameter
GFRP poles under the ground line was 1 m (3.28 ft) and was varied from 152.4 mm (6 in.) to 508 mm (20 in.), while
2.44 m (8 ft) for the 6.1 m and 18.3 m specimens, respec- for the 18.3 m (60 ft) poles, it was varied from 457 mm
© 2000 NRC Canada
856 Can. J. Civ. Eng. Vol. 27, 2000

Fig. 6. Comparison between theoretical and experimental ultimate load.

Fig. 7. Comparison between experimental and finite element load–deflection relationship for 10° full-scale specimens.

(18 in.) to 1016 mm (40 in.). In each case, the wall thickness amined in this program was 305 mm (12 in.). A similar opti-
of the specimen was determined by trying different number mization study was conducted for the 18.3 m (60 ft) GFRP
of layers and the least number of layers that met class 1 poles, shown in Fig. 10. The optimum bottom diameter in
strength requirements was considered. The thickness of each that case was 559 mm (22 in.).
layer was assumed to be 0.44 mm (0.0173 in.). Using the de-
signed wall thickness, the total weight of the poles was de- Conclusions
termined. The optimum bottom diameter can be obtained by
plotting the total weight versus the bottom diameter. The re- An experimental study was used to investigate the ulti-
lationship between the bottom diameter and the weight for mate capacity and performance of tapered filament wound
the 6.1 m (20 ft) GFRP poles is shown in Fig. 9. As shown GFRP poles subjected to cantilever bending. The study
in the figure, the optimum bottom diameter for the case ex- showed that the ultimate load capacity of GFRP poles is

© 2000 NRC Canada


Ibrahim et al. 857

Fig. 8. Comparison between experimental and finite element load–ovalization relationship for 10° full-scale specimens.

Fig. 9. Optimization of 6.1 m (20 ft) GFRP poles.

equivalent to that of wooden poles. However, a significant mens tested because of the high radius-to-thickness ratio of
saving in the weight is achieved by utilizing GFRP. The the specimens. Only one specimen suffered material failure
load capacity-to-weight ratio of GFRP poles is almost three on the compression side because of the absence of cir-
times higher than that of wooden poles. Maximum load ca- cumferential layers in that specimen. The finite element
pacity-to-weight ratio was achieved by incorporating cir- method used in this investigation provided an excellent pre-
cumferential winding along with longitudinal fibers. Of the diction of the critical buckling and material failure loads, as
eight-layer GFRP poles tested, the maximum load capacity- well as the corresponding modes of failure for thin-walled
to-weight ratio was attained when two circumferential layers GFRP poles. Optimum cross-section dimensions for both
were combined with six longitudinal layers. Local buckling 6.1 m (20 ft) and 18.3 m (60 ft) class 1 GFRP poles were
was the most dominant mode of failure in most of the speci- obtained theoretically using the finite element model.

© 2000 NRC Canada


858 Can. J. Civ. Eng. Vol. 27, 2000

Fig. 10. Optimization of 18.3 m (60 ft) GFRP poles.

Acknowledgements Structures Workshop Proceedings, Electric Power Research In-


stitute, July 25–26, Denver, Colo., pp. 55–61.
The research project in this study was sponsored by ISIS- Dow Plastics Company. 1997. Technical product information. Mid-
Canada, Faroex Ltd. of Manitoba, and the Province of Mani- land, Mich.
toba. The specimens were fabricated at the ISIS–Faroex Fil- Escher, G.A. 1982. Design and testing of an optimum fiberglass re-
ament Winding Research Facility, and the testing was inforced pole and market study on the demand for utility poles
conducted at the Structural Engineering and Construction in Canada. Report prepared for the Canadian Electrical Associa-
R&D Facility at The University of Manitoba. tion, Montreal, Que.
Ibrahim, S., and Polyzois, D. 1999. Ovalization analysis of fiber-
reinforced plastic poles. Composite Structures, 45: 7–12.
References
McClure, G., Boire, L., and Carriere, J.C. 1992. Applications of
Adams, R.D. 1987. Damping properties analysis of composites. In advanced composite materials in overhead power lines and tele-
Engineering materials handbook. Edited by T.J. Reinhart et al. communications structures. In Advanced composite in bridges
ASM International, Materials Park, Ohio, Vol. 1, Composites, and structures. Edited by K.W. Neale and P. Labossière. Cana-
pp. 206–217. dian Society for Civil Engineering, Montreal, Que., pp. 543–549.
Adams, F.D., and Doner, D.R. 1967. Transverse normal loading of a Tsai, S.W., and Hahn, H.T. 1980. Introduction to composite materi-
unidirectional composite. Journal of Composite Material, 1: 152. als. Technomic Publishing Co., Lancaster, Pa.
Agarwal, B., and Broutman, L. 1980. Analysis and performance of Tsai, S.W., and Wu, E.M. 1971. A general theory of strength for
fiber composites. Wiley Interscience publication, New York. anisotropic materials. Journal of Composites Materials, 5: 58–80.
ANSI. 1992. American national standard for wood poles 05.1, Vanderbuilt, M.D., and Criswell, M.E. 1988. Reliability analysis of
specifications and dimensions. American National Standards In- pole-type transmission structures. Computer and Structures,
stitute, New York, N.Y. 6(2): 335–343.
ANSYS, Inc. 1995. ANSYS user’s manual for revision 5.2 — Yunus, S., Kohnke, P.C., and Saigal, S. 1989. An efficient through-
Volumes I to IV. Houston, Pa. thickness integration scheme in unlimited layered doubly curved
Derrick, G.L. 1996. Fiberglass composite distribution and trans- isoparametric composite shell element. International Journal for
mission poles. Manufactured Distribution and Transmission Pole Numerical Methods in Engineering, 28: 2777–2793.

© 2000 NRC Canada

You might also like