You are on page 1of 8

AN ALTERNATIVE FIBER-REINFORCED PLASTIC POLE DESIGN

Gustavo Abreu Birchal, Estevam Barbosa Las Casas, Structural Engineering Department, Federal University of Minas Gerais, Belo Horizonte, Brazil Carlos Alberto Cimini Jr., Mechanical Engineering Department, Federal University of Minas Gerais, Belo Horizonte, Brazil Stephen W. Tsai Department of Aeronautics and Astronautics, Stanford University, Stanford, USA

SUMMARY: This paper presents the concepts used in an alternative pole design, applying
fiber-reinforced plastics (FRP) to semi-monocoque structures towards the production of lightweight poles. Based on the ultimate strength capacity and maximum allowed displacements, an user-friendly computer code was implemented as tool for designers. Six proposed poles were compared to two existing monocoque designs made of steel and E-Glass epoxy, highlighting the advantages offered by the presented concept.

KEYWORDS: Poles, composites, semi-monocoque structures, cost effective design INTRODUCTION


The necessity of investments in data communication and energy transmission infrastructure, and the growing use of mobile phones are providing an extremely favorable setting for searching new technologies in the area of civil materials and construction techniques. Great part of the existing structures is made of self-supported metal truss towers, with heights varying from 30 to 100 m. In high urban density places, however, this height rarely exceeds 30 m. This fact, together with the lack of space in cities and the concern of the organized civil society related to visual pollution, is leading the service companies to choose poles, instead of truss towers. Poles used today are, in their vast majority, made of wood, reinforced concrete or steel. Nevertheless, these poles have particular features that raise considerably costs of design, manufacture, installation and maintenance. Wooden poles have short life expectancy with significant deterioration of their load bearing capacity due to continuous exposure to w eather, fungi and woodpeckers, bringing serious environmental concerns. The main disadvantages of reinforced concrete poles are their heavy weight, which drastically increases transportation and erection costs, and the subjection of the steel reinforcement to corrosion and to other chemical influences arising from environmental impacts. The use of steel also results in additional costs as a consequence of corrosion and fatigue preventive maintenance that, jointly to its density, make its use very expensive. A possibility that may improve substantially the above deficiencies rely on the use of new materials and new structural concepts. One viable proposal is to use semi-monocoque structures, a design concept largely used in aeronautics industries, together with composites that come up as natural candidates among new materials, since they not only present

outstanding strength and stiffness features, but also show noticeable weight gain and desirable resistance to corrosion and fatigue added to poor electric al conducting. The use of composite materials in the civil construction, initiated in the sixties and experienced an accelerated growth since then due to some composite inherent properties of extreme importance to this field. Approximately 6.6% of the use of composites occurs today in the field of civil construction, as may be seen in the pie chart on Figure 1 [2], with prospects of increasing even more.

Transportation 18%

Industry 14%

Clothing 14%

Geotextile 9%

Leisure 9%

Construction 7%

Medicine 7%

Agriculture 4%

Others 19%

Figure 1 Composite use on industry [2] An informal research project has been jointly carried out at the Federal University of Minas Gerais, Brazil, and Stanford University, USA, to develop an alternative pole design, applying fiber-reinforced plastics (FRP) to semi-monocoque structures towards the production of lightweight poles. The current research program contemplates: A. The development of a computer code allowing the automated design of the proposed pole, taking into account aspects such as ultimate strength capacity and maximum allowed displacements; B. The evaluation of the developed theoretical models through a series of reduced-scale testing; C. The optimization of the design and manufacturing process for the pole. To date, just the stage A has been concluded, and stage B is in development. This paper will present the design procedures adopted, as well as six theoretical models with their respective analysis.

POLE MODELING
The pole was modeled as an isostatic cantilever beam clamped at one end, made of two concentric modules (skins) reinforced by equally spaced longitudinal bars (stringers), as depicted in Figure 2.

Outer skin

Inner skin

Stringer

Figure 2 Pole design concept In this module, stringers are responsible for carrying the axial loading due to structural weight and bending loading. Skins are modeled to mainly support shear and torsion, although they are also responsible for axial tensile stresses and contribute to the effectiveness of stringers on axial compression. Skins also provide an outer surface for wind force distribution along the structure. Fibers were oriented on angle -ply direction (+/ ) at the skins and on axial direction at the stringers, taking advantage of the best mechanical properties that unidirectional composites have to offer, which are the stiffness and the strength on the fibers direction. This concept fits exactly with the semi-monocoque configuration, recommended for attainment of lightweight structures, that differs from the monocoque concept, which relies entirely on the skin to carry all the applied loads. One important function of the outer skin is to form an impermeable surface for supporting and distributing the aerodynamic wind pressure. These aerodynamic forces are transmitted to the stringers through plate and membrane action. Resistance to shear and torsional loads is supplied by shear stresses developed in the skins while axial and bending loads are supported by the combined action of skins and longitudinal stiffeners. Although the skin is efficient for resisting such loads, it buckles under low compressive stresses. Rather than increase the skin thickness and suffer a consequent weight penalty, stiffening stringers are attached to the skin, thereby dividing it into small panels and increasing the buckling and failure stresses [3]. The manufacturing process of the pole can be completely automated, coupling pulltrusion for processing the stringers and filament winding for wrapping the skins, which results in high economical efficiency and significant cost reduction on a large scale production.

FAILURE CRITERION
A computer code was developed as an user-friendly tool for designers in order to allow the complete automation of the pole project. The code flow chart is shown in Figure 3. Data input is done in the first procedure. Regarding the loading, there are three options available: standard loading for telecommunication poles, standard loading for transmission poles and random loading. For the first type, the loading represents the dead weight of the structure and accessories (such as aerials and their supports, platforms, lightning conductors and ladders), accidental loads and wind forces acting on the structure and on the accessories [4] [5]. The transmission poles loading follows the same guidelines, and additionally includes the forces transmitted by the transmission cables [6]. There is an optional procedure for those who prefer providing the properties of fibers and matrix individually, instead of feeding the software with information related to the composite. In this case, the properties of the material will be evaluated through the micromechanics theory of composite materials [7] [8].

Data Entry Composite Properties Evaluation

Loading Evaluation and Stress Analysis

Ultimate Strength Capacity Checking

Maximum Allowed Displacements Checking

Stresses Redistribution

Initial Design Changes

Results Presentation Figure 3- Computer code flow chart The next step is the evaluation of the stress resultants based on the Euler beam theory. The distribution of axial strength at the stringers and skins is done, initially, with respect to the areas, moments of inertia and tensile moduli for each piece. On the other hand, the shear stresses must be entirely resisted by the skins. The proposed analysis procedure makes use of the shear flow concept and the theory of shear stresses along thin walled tubes reinforced with booms, where the booms may cause discontinuities at the tube walls, interrupting the shear flow. The failure verification is done through the Tsai-Wu criterion, based on invariants derived from stress tensor components and accounting for the difference between tensile and compressive strength [9]. The reduced form of the Tsai-Wu criterion implemented at the computer code is showed by Equation 1 f 1 1 + f 2 2 + f 11 1 2 + f 22 2 2 + f 66 6 2 + 2 f 12 1 2 = 1 (1)

where f i and f ij are respectively the second and fourth order strength tensors components, i are the axial stresses and is the shear stress [8]. The verification procedure then checks whether the pole displacements do not exceed the limits established in the Brazilian Standards [10] for transmission lines, for each kind of pole, also checking the overall instability and stringer crippling. It is important to emphasize that, for accurate results, both the static and dynamic loading must be taken into account. For static loading, basic assumptions of structural stability were assumed, using the beam differential equation and its natural and essential boundary conditions, obtained through the first variation of the energy functional [11]. Displacement in each point of the structure and

ultimate compressive force are then calculated through Equation 2, which incorporates second order bending effects, where E is the Youngs modulus, I is the cross section moment of inertia, v is the beam transversal displacement , N is the axial load and P is the shear distributed load. (EI v) (N v) P = 0 (2)

For dynamic loading, the principle of vortex formation, critical whenever the Von Karman vortex trail appears [12], are applied to the fundamental equation in structural dynamics and linear vibration theory for a single degree-of-freedom system [13], allowing the evaluation of the displacements according to Equation 3. m u + c u + k u = p(t) (3)

where m is the total mass of the system, c is the viscous damping coefficient, k is the spring constant and the excitation p(t) is a function of the force due to the vorticity of the viscous damping factor and of the frequency ratio, defined as the induced frequency divided by the undamped circular natural frequency and through which the resonance effects may be observed by appearance of frequency ratios similar to the unity. Since there is no analytical method for prediction of crippling strength of the stringers, empirical techniques [14] are used whenever the stringers are under analysis. On the other hand, there is no need of checking whether the skins will buckle or not since the structure is semi-monocoque, allowing skin local instability. Therefore all tensile and compressive loading is modeled to be resisted by the stringers. This is done through the stress redistribution procedure. Skins, however, contribute with an effective width to local buckling/crippling of stringers. The last two stages of the software are used, respectively, for changes on the entered data and for presentation of the resulting designed pole.

THEORETICAL MODELS
In order to verify the efficiency of the proposed design, six designs made of E-Glass fibers embedded in epoxy matrix were studied so that they could be compared to existing designs. Properties are listed in Table 1, which also shows geometric features of each structure. A number of comparison parameters among the six models and the two existing designs, such as dead weight, ultimate compressive force, displacement at the top of the structure and the shear force / dead weight ratio, are also listed. For the latter parameter, the shear force is the summation of all shear forces acting on the critical pole section. Poles 1 to 3 were designed according to the proposed model for comparison with pole 4, an existing steel 30 m high pole for telecommunication purposes, with variable cross section and thickness. Since they all are cell phone poles, their loading includes, besides the dead weight of the structure and wind forces, the weight of the following accessories: aerials (2.4 kN), aerials support (2.9 kN), platforms (1.2 kN), lightning conductors (0.4 kN), ladders (1.5 kN) and vertical accidental loads (4.3 kN). Analysis of pole 1 shows that the appropriate use of the unidirectional composite properties at the fibers direction allows the design of a structure over 5 times lighter than the steel existing design, although it was not tapered. However, this pole does not pass the displacement design criterion, presenting excessive displacements of 13.42 m at the top compared to the allowed 2.10 m, falling beyond the linear small displacement theory used. This pole also showed symptoms of stability loss.

Table 1 Comparison among designs


Items Design Material P ole 1 E-Glass / Epoxy 17.66 38.60 8.26 4.14 0.26 1,060.00 610.00 31.00 117.00 71.90 30.00 0.0007 0.0007 0.0007 0.0007 0.6347 0.6347 0.5647 0.5647 3 9o 0
o

Comparison 1 Pole 2 E-Glass / Epoxy 17.66 38.60 8.26 4.14 0.26 1,060.00 610.00 31.00 117.00 71.90 30.00 0.0015 0.0015 0.0015 0.0015 1.1015 1.1015 0.9015 0.9015 7 1.1 o 80 / -80 90 o 8.50 21.27 40.60 16.67 4.78 195.21 1.14 2.10 90.04% 56.41% Crippling of stringers
o o

Comparison 2 Pole 4 Steel COR-500 77.01 210 210 150 0.3 N/A N/A N/A N/A N/A 30.00 0.0048 0.0030 N/A N/A 0.7940 0.4740 N/A N/A N/A N/A
o

Pole 3 E-Glass / Epoxy 17.66 38.60 8.26 4.14 0.26 1,060.00 610.00 31.00 117.00 71.90 30.00 0.0010 0.0010 0.0010 0.0010 1.0010 1.0010 0.8601 0.8601 7 1.1 o 80 / -80 90 o 5.37 18.15 40.60 16.67 7.56 106.15 1.89 2.10 4.20% 27.32%
Rupture of skins due to shear
o

Pole 5 E-Glass / Epoxy 17.66 38.60 8.26 4.14 0.26 1,060.00 610.00 31.00 117.00 71.90 6.25 0.0009 0.0009 0.0009 0.0009 0.3705 0.3705 0.3500 0.3500 9 9o 3 / -3 90 o 0.50 0.50 36.40 6.25 72.80 109.69 1.71 0.44 1.20% 3234.80%
Rupture of skins due to shear
o o

Pole 6 E-Glass / Epoxy 17.66 38.60 8.26 4.14 0.26 1,060.00 610.00 31.00 117.00 71.90 6.25 0.0010 0.0010 0.0010 0.0010 0.4210 0.4210 0.4010 0.4010 9 16 o 7 / -7 90 o 0.80 0.80 36.40 6.25 45.79 266.02 0.65 0.44 1.31% 116.74%
Rupture of skins due to shear
o o

Pole 7 E-Glass / Epoxy 17.6 6 38.60 8.26 4.14 0.26 1,060.00 610.00 31.00 117.00 71.90 6.25 0.0015 0.0015 0.0015 0.0015 0.4210 0.4210 0.3915 0.3915 7 16 o 80 / -80 90 o 1.01 1.01 36.40 6.25 36.04 426.30 0.44 0.44 33.96% 7649.70%
Rupture of skins due to shear
o o

Pole 8 E-Glass / Epoxy 17.66 38.60 8.26 4.14 0.26 1,060.00 610.00 31.00 117.00 71.90 6.25 0.0050 0.0050 N/A N/A 0.4260 0.3150 N/A N/A N/A N/A 80 o / -80 o N/A 0.80 0.80 36.40 6.25 45.79 130.82 0.64 0.44 N/A N/A N/A

Density (kN/m3) Longitudinal modulus (GPa) Transverse modulus (GPa) In-plane shear modulus (GPa) Major Poissons ratio Ultimate longitudinal tensile strength (MPa) Ultimate longitudinal compressive strength (MPa) Ultimate transverse tensile strength (MPa) Ultimate transverse compressive strength (MPa) Ultimate in-plane shear strength (MPa) Height (m) Outer skin thickness. at the base (m) Outer skin thickness. at the top (m) Inner skin thickness. at the base (m) Inner skin thickness. at the top (m) Outer diameter of the outer skin. at the base (m) Outer diameter of the outer skin. at the top (m) Outer diame ter of the inner skin. at the base (m) Outer diameter of the inner skin. at the top (m) Number of stringers Inner angle o f the stringers Longitudinal fiber orientation at the skins Longitudinal fiber orientation at the stringers Pole weight (kN) Pole weight + accessories (kN) Applied shear force (kN) Applied shear force height (m) Ratio applied shear force / pole weight (kN/kN) Ultimate compressive force (kN) Displacement at the top (m) Maximum allowed displacement (m) Safety factor for skins Safety factor for stringers Failure probable cause N/A not applicable

N/A N/A 19.78 32.56 40.60 16.67 2.05 130.51 1.18 2.10 N/A N/A N/A

90 o 3.86 16.63 40.60 16.6 7 10.52 25.60 13.42 2.10 3.73% 7.54% Stability loss

Pole 3 was designed to overcome pole 1 deficiency, resulting in a top displacement of 1.89 m for a maximum allowable of 2.1 m, and without any sign of stability loss. In spite of having its cross-section stiffened, the new pole, which initial failure may be observed through the rupture of skins due to shear, is almost 4 times lighter than the steel design. Pole 2 was designed to obtain similar displacements to the steel design and even then was shown to be over 2 times lighter, and presenting a shear force/dead weight ratio almost 2.5 times higher. To make this possible, an increase at both skins diameters was introduced, clearly demonstrating the tendency for crippling of stringers. A second set of poles, numbered from 5 to 7, is also described in Table 1, this time for comparison with the pole 8, a 6.25 m E-Glass/epoxy monocoque pole presented at ICCM12 [1]. The considered load case consisted of a 36.40 kN shear force applied at the top as point load and the dead weight of the structure. The analysis of pole 5 shows that the use of semi-monocoque structures allows weight gains by a factor of 1.6, when considering just the ultimate strength. However, excessive displacements (1.71 m) were observed when compared to the allowable 0.44 m (for rural application). Pole 6 was designed aiming at displacements similar to those presented by pole 8, but did not show any weight gain when compared to pole 8. Displacements, however, again exceeded the allowable in this case (0.65m compared to the allowable 0.44 m). Pole 7 was designed to improve the deficiencies above mentioned, agreeing with the maximum allowed displacements. Since it had its cross-section considerably stiffened thereby suffering a weight penalty around 25%, this pole didnt show any sign of failure for the applied set of loads. It is important to highlight that in all the presented cases almost the totality of displacements were caused by the static loading, with slight contribution of dynamic loading, usually smaller than 1 mm. This is explainable by the induced frequencies of the structures that were always nearly a hundred times over the undamped natural frequencies, avoiding therefore the risk of resonance. It is also interesting to notice that the stringers were extremely effective in resisting the axial stresses. In neither of the six analyzed poles they were close to failure, showing cases where the safety factor was found to be higher than 7.

CONCLUSIONS
A methodology was implemented in order to analyze the behavior of semi-monocoque poles made of composite materials. This research accounted not only the ultimate strength, but also the maximum allowed displacements as prescribed by the Brazilian Standards, with both the static and dynamic loading considered. Six poles made of E-Glass/epoxy were designed for comparison to two existing designs. In spite of presenting constant cross-section and skin thickness along the whole length, these poles turned out to be almost 5 times lighter than the metallic design, and nearly 1.6 lighter than the monocoque composite design, when considering only ultimate strength. When maximum allowed displacements are taken into account, weight gain is lower, although noticeable. This fact is due to the stiffness similarity in all models, and can be minimized using ribs along the pole length, thus reducing buckling stresses as in the aeronautical structures.

The performance of the proposed poles could be improved by taking two measures not included in this first approach. The first one is modeling the structure as a telescope mast. The other one is the use of shape optimization during the design procedures. Finally it should be emphasized that the use of other fiber materials, such as Carbon or Aramid embedded in epoxy matrix, might give results even more satisfactory.

REFERENCES
1. Polysois, D., Ibrahim, S., Burachynsky, V. and Hassan, S.K., Glass fibre-reinforced plastic poles for transmission and distribution lines: An experimental investigation, Proceedings of the 12th International Conference on Composite Materials, ICCM-12, Paris, 1999, CD-ROM. Miravete, A., Los nuevos materiales en la construccin, 2 edicin, Antonio Miravete, INO Reproducciones, S.A., Zaragoza, 1995. (in Spanish)
nd Megson, T.H.G., Aircraft structures for engineering students, 2 edition, Halsted Press (imprint of John Wiley & Sons), New York, 1990.

2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14.

American Society of Civil Engineers Design of Steel Transmission Pole Structures, 2nd edition, American Society of Civil Engineers, New York, 1990. Brazilian Technical Standard NBR-6123/88, Associa o Brasileira de Normas Tcnicas, Rio de Janeiro, 1988. (in Portuguese). Gontijo, C.A., Design of Towers for Transmission, Instituto de Engenharia Aplicada Editora, Belo Horizonte, 1994 (in Portuguese). Agarwal, B.D. and Broutman, L.J., Analysis and performance of fiber composites, John Wiley & Sons, New York, 1990. Daniel, I.M. and Ishai O., Engineering mechanics of composite materials, Oxford University Press, Oxford, 1994. Tsai, S.W. and Hahn, H.T., Introduction to composite materials, Technomic Publishing, Lancaster, 1980. Telebrs Documentation System Engineering Series 240-410-600 (Standard), Sistema de Documentao Telebrs, 1996. (in Portuguese)
nd Timoshenko, S.P. and Gere, J.M., Theory of elastic stability, 2 edition, McGrawHill, New York, 1961.

Simiu, E. and Scanlan, R.H., Wind effect on structures An introduction to wind engineering, Wiley-Interscience, New York, 1978. Craig, R.R.Jr., Structural Dynamics An introduction to computer methods, John Wiley & Sons, New York, 1981. Boeing Commercial Airplane Company, Boeing design manual, section 6, The Boeing Company, Seattle, 1973.

You might also like