Open access peer-reviewed chapter

Strength Analysis and Variation of Elastic Properties in Plantain Fiber/Polyester Composites for Structural Applications

Written By

Christian Emeka Okafor and Christopher Chukwutoo Ihueze

Submitted: 22 November 2019 Reviewed: 18 December 2019 Published: 15 July 2020

DOI: 10.5772/intechopen.90890

From the Edited Volume

Composite and Nanocomposite Materials - From Knowledge to Industrial Applications

Edited by Tri-Dung Ngo

Chapter metrics overview

684 Chapter Downloads

View Full Metrics

Abstract

Plantain fiber-reinforced composite materials have demonstrated significant properties that are applicable in structural design and development. However, two major concerns arise in relation to the obvious material anisotropy and challenges imposed by structural discontinuity encountered as need for use of fasteners arises. The study assesses the extent of variation of elastic properties ( E x , E y , G xy , v xy , v yx , m x , m y ) with fiber orientation using MATLAB functions while considering the extent of variation of the tangential stresses around an idealized functional hole edge. The tensile strength of 410.15 and 288.1 MPa was recorded at 0° fiber orientation angle, while 37.3397 and 33.133 MPa were obtained at fiber orientation angle of 90° for Plantain Empty Fruit Bunch Fiber Composite (PEFBFC) and Plantain Pseudo Stem Fiber Composite (PPSFC), respectively. The tangential stress distribution at hole edge indicated maximum stress value of 119.15 and 100.587 MPa at angular position ø = 90° for PEFBFC and PPSFC, respectively. Judging from various failure indices considered, failure will be initiated at ø = 70° for PEFBFC with stress concentration factor of 2.53 and ø = 65° for PPSFC with stress concentration factor of 2.13, which are less than the stress concentration around the peak stress when angular position is 90°. Both PEFBFC and PPSFC showed similar trends in response to the design scenario considered.

Keywords

  • elastic properties
  • structural application
  • plantain fiber
  • composites
  • matrix

1. Introduction

These days several shortcomings have been observed with respect to the utilization of synthetic fiber-reinforced polymers, hence empowering the drive for more utilization of plant fiber composite in structural designs. The primary weaknesses of synthetic fiber-reinforced polymers which includes issue in afterlife disposal and non-biodegradability are completely settled by utilizing plant fibers in polymer reinforcements. Reinforced composites are prime choice for light weight structural designs and automotive body parts assembly.

Extensive literature on recent advancements in reinforced composites and its reliability are reported in classical reports of Dehmous et al. [1], Okafor et al. [2], Xie and Wang [3], Beaumont et al. [4], Pei et al. [5], Bittrich et al. [6], Prasad et al. [7], Wang et al. [8]. The hygrothermal efficiencies has been reported by Foulc et al. [9], Shettar et al. [10]. In addition, the utilization of reinforced composites predicates the reuse of domestic and agricultural residues. For example, vehicles made with fiber-reinforced composites are lighter and run on smaller engines which produce fewer emissions to the environment. Most items produced using natural fiber composites is a win-win for manufacturers. Most composite material ventures can utilize their genius green item data to build deals because customers comprehend the ecological dangers of synthetic assembling.

Due to scarcely available information regarding some new material response to structural discontinuity, superior properties of those composites are seriously compromised by the utilization of bizarrely enormous factor of safety in design. Accordingly, the quick fate of composite materials as a class of innovative materials may depend more on clear assessment of its performance in various structural design scenario. All inclusive acknowledgment of composites as eco-friendly materials will therefore depend especially on the certainty of the designer and client about the variation of its elastic properties. In a typical fiber-reinforced composites, the polymer matrix serves as a binder and deforming most times for stress distribution purposes. There are different options in the choice of matrix/fibers and the general composition of reinforced composites is shown in Figure 1. The figure identified the three major categories of polymers to include biopolymers, thermoplastics and thermosets. Biopolymers are chain like atoms created by organic biomass. Exceptional nontoxicity and biodegradable properties of biopolymers boosts their applications in composites formulation, hardware and restorative gadgets. Fuse of nano-sized support in the biopolymers to improve the properties contributes to the upgrade functional applications of the matrix.

Figure 1.

Composition of reinforced composites.

Though thermosets and thermoplastics sound similar, they have very different properties and applications [11]. Thermosets typically changes from fluid to solid state after curing chemical reaction initiated by addition of a catalyst, cross-linker, and curing agent. In the course of the chemical reaction, the material solidifies as a result of cross-linking and formation of longer molecular networks. Subsequently, any further exposure to high heat will cause the material to degrade unlike thermoplastic parts that melts and softens whenever exposed to elevated temperature, thermoset simply become set in their physical and mechanical properties after an initial treatment and therefore are no longer affected by additional heat exposure.

Again, thermoplastics are dissolvable plastics. At temperature above liquefying point, the thermoplastic condenses. The thermoplastic sets once again into solid state when the temperature is reduced and the handling temperature dips under its melting point. This inherent characteristic of thermoplastics enables its softening when heated above its melting point and re-forming as the temperature decreases below the melting point. Most of the times, the expenses of materials for creating thermoset are lower when contrasted with thermoplastic. Also thermoset is regularly simple for wetting the reinforcements and shaping last composites items. Thermoplastics will in general be harder than thermosets and require no refrigeration as uncured thermosets as often as possible do, and can be more effectively be reused and fixed. Elastomers are typically thermosets (requiring vulcanization).

Obviously literature has indicated several approximate relationships between some reinforced composites elastic constants and the homogenized modules of elasticity [12, 13, 14]. Also recent research have reported the possibility of measuring variation elastic constants of materials using ultrasonic methods [15, 16, 17, 18]. However, scanty research is available on strength analysis and variation of elastic properties in plantain fiber/polyester composites, a gap that the present study seeks to fill.

Advertisement

2. Background to plantain cultivation and utilization as reinforcement in polymer composites

An expanded enthusiasm for the utilization of agricultural wastes in development of reinforced composites has been on the increase. Natural fibers extracted from bio wastes offer a few points of interest over woody biomass, since they are accessible in huge amounts as leftovers and agricultural wastes [19, 20]. The plantain pseudo stem (PPS) and empty fruit bunch (EFB) strands presented in this chapter are agricultural by-products that are biodegradable and locally available from renewable agricultural sources with potentials to contribute to reduction in environmental pollution when utilized in large scale as polymeric reinforcements.

Plantain fruit is one of the staples in Nigeria and it is mainly cultivated in the tropics and ethnic enclaves [21]. It is evaluated that 70 million individuals in West and Central Africa derive most of their nourishment and vitality requirements from plantain fruit plant [22, 23]. Plantain fruit has a fare potential in the light of its huge cultivation and consumption in Nigeria and many other African countries. Akinyemi et al. [24] reported that plantain plant is the third most important plant grown after cassava and yam in Nigeria; collaborating, Kaine and Okoje [25] showed that plantain production is a very profitable enterprise as every ₦1 naira invested in plantain production yields a return on investment of about ₦12.60 kobo. In a study about economics of plantain production Kainga et al. [26] found that the associated high return on investment and short maturity period for plantain contributes to its massive cultivation in Nigeria.

Africa cultivates over 50% of worldwide production of plantain and Nigeria is one of the biggest plantain producing nations in the planet. Therefore the interest in plantain plant fiber for polymer reinforcement was as a result of its abundance and accessibility as it is evaluated that over 15.07 million tons of plantain fruit is produced each year in Nigeria with about 2.4 million metric tons produced from southern Nigeria [27, 28]. Plantain fiber also satisfied over 50% conditions for eco-friendly materials as shown in Figure 2 and can make strong reinforcement in composites. A composite which can be characterized as a physical blend of at least two unique materials, has properties that are commonly superior to those of any of the establishing materials. It is important to utilize blends of materials to tackle issues in light of the fact that any one material alone cannot suffix effectively in eco-friendly materials technology at an acceptable performance [29].

Figure 2.

Properties of eco-friendly building materials.

Cadena Ch et al. [30] and Adeniyi et al. [31] orchestrated the potentials of natural fibers from plantain pseudo stem for use in fiber-reinforced composites. It is therefore important to assess the extent of variation of elastic properties in plantain fiber-reinforced polyester composites to guard against out of plane failure during structural applications. Unfortunately most studies involving plantain fiber-reinforced composites has dwelt on assessment of tensile, flexural and hardness properties [32], optimization of hardness strengths [33], effect of water and organic extractives removal [34], effects of fiber extraction techniques [35], optimization of flexural strength [36], compressive and impact strength evaluation [37, 38, 39], effect of high-frequency microwave radiation [40], effect of chemical treatment on the morphology [41], implications of interfacial energetics on mechanical strength [42]. Although Ihueze, Okafor and Okoye [43] has reported the longitudinal (1) and transverse (2) properties of plantain fiber-reinforced composites in Figure 1, there is still need to establish the essential elastic constants at directions other than the material axis directions 1–2.

The present research efforts will further drive the interests of structural designers in the use of plantain fiber-reinforced composites because the superior strength of materials are rarely utilized to full as a result of incomplete knowledge of elastic properties which are related to various fundamental solid-state characteristics of the composites. In essence, the elastic constants of plantain fiber-reinforced composites is expected to describe the material response to external stressor and provide useful information about bonding characteristics and structural stability. Kenedi et al. [44] assessed the orthotropic elastic properties in a sandwiched composites laminates and proposed models for estimating the orthotropic elastic properties of composite materials. Hwang and Liu [45] reported that elastic modulus and Poisson’s ratio vary significantly with different braid angles in carbon fabric/polyurethane composites. Ren et al. [46] reported that elastic modulus and tensile strengths are overly dependent on the angles of fiber orientation. Kumar et al. [47] studied the influence of ±0°, ±10°, ±30°, ±40°, ±45°, ±55°, ±65°, ± 75°, and ± 90° angle ply on mechanical properties of glass-polyester composite laminate and found that that glass/polyester with 0° fiber orientation angle yields’ high strength. Cordin et al. [48] experimentally examined the effect of 0°, ±22.5°, ±45°, ±67.5° and 90° fiber orientation angles on the mechanical properties of polypropylene-lyocell composites. Ihueze et al. [49] optimally determined the tensile strengths of plantain fiber-reinforced composites considering 30°, 45° and 90° fiber orientation angles. The application of these previous studies are limited to fiber orientation angles studied, however failure may be initiated from angles other than those considered hence the need to verify the variation of important elastic constants within a wide range of fiber orientation coverage are necessary.

Additionally, researchers have provided various theoretical strategies for determination of elastic constants in reinforced composites using software codes to cover the wide range of fabricating conditions. Jules et al. [50] ascertained the effect of fibers orientation on the predicted elastic properties of long fiber composites using Monte-Carlo simulation to assign the in plane and out of plane orientation values. Venetis and Sideridis [51] developed a model to find the approximate elastic constants in unidirectional fiber-reinforced composite materials in terms of the constituent material properties. Cuartas [52] theoretically determined the elastic properties in CFRP composites which compared favorably with other methods based on tensile tests and ultrasonic characterization. Rahmani et al. [12] found that MATLAB codes are capable of predicting the elastic constants of composites with reasonable confidence.

Advertisement

3. Mathematical framework for assessment of extent of variation of elastic properties in plantain fiber-reinforced polyester composites

One significant property of composite materials is their plainly visible macroscopic anisotropy, which means that the properties estimated in the longitudinal direction are by far not the same as those measured in transverse direction. There are no material planes of symmetry, and normal loads create both normal strains and shear strains. This anisotropic characteristic of reinforced composites results in low mechanical properties in the out-of-plane orientation where the matrix carries the primary load. Consequently the application of reinforced composites is limited in scenarios prone to complex load paths such as lugs and fittings [53].

By implication any endeavor to comprehend the structural application of plantain fiber-reinforced polyester composite must assess the inborn anisotropy. Composites are a subclass of anisotropic materials that are delegated orthotropic. Orthotropic materials have properties that are unique in three directions with perpendicular axes of symmetry. In this way, orthotropic mechanical properties depend heavily on fiber orientation. An orthotropic ply is thus defined as that having two different material properties in two mutually perpendicular directions at a point and the two mutually perpendicular directions also form the planes of material properties symmetry at the point.

3.1 Determination of reduced stiffness matrix and compliance matrix

Considering two possible loading conditions of longitudinal (direction 1) and transverse (direction 2) in the matrix as shown in Figure 3, the resulting direct strains from Hooks law are respectively e 1 = v 21 σ 1 E 1 and e 2 = v 12 σ 1 E 1 where v 12 = major Poisson’s ratio and v 21 = minor Poisson’s ratio.

Figure 3.

Stressed single thin composite lamina. From Ref. [43].

Hence the application of both direct stresses σ 1 and σ 2 will yield corresponding strains as follows:

e 1 = σ 1 E 1 v 21 σ 2 E 2 E1
e 2 = σ 2 E 2 v 12 σ 1 E 1 E2

Putting Eqs. (1) and (2) in a matrix form, yields

e 1 e 2 = 1 E 1 v 21 E 2 v 12 E 1 1 E 2 σ 1 σ 2 E3
σ 1 σ 2 = E 1 1 v 12 v 21 v 21 E 1 1 v 12 v 21 v 12 E 2 1 v 12 v 21 E 2 1 v 12 v 21 e 1 e 2 E4

Eq. (3) is symetric about the loading diagonal such that

v 21 E 2 = v 12 E 1 E5

A combined effect of shear and direct stresses gives the reduced stiffness matrix as in Eq. (6) and reduced compliance matrix as in Eq. (7)

σ 1 σ 2 τ 12 = E 1 1 v 12 v 21 v 21 E 1 1 v 12 v 21 0 v 12 E 2 1 v 12 v 21 E 2 1 v 12 v 21 0 0 0 G 12 e 1 e 2 e 12 E6
e 1 e 2 e 12 = 1 E 1 v 21 E 2 0 v 21 E 1 1 E 2 0 0 0 1 G 12 σ 1 σ 2 τ 12 E7

Minor Poisson’s ratio is the strain resulting from a stress in the axial direction, Ihueze et al. (2013) calculated the major Poisson’s ratio v 12 for plantain fiber/polyester composites. However, there is need to further assess the minor Poisson’s ratio v 21 using Eq. (5) and Table 1 as follows

Composites Properties
S u 1 (MPa) S u 2 (MPa) S y (MPa) E 1 (MPa) E 2 (MPa) E (MPa) v 12 τ max (MPa) G 12 (MPa)
PEFBFC 410.15 37.3397 33.69 14,922 7030.962 9990.10 0.38 19.3100 3622.99
PPSFC 288.10 33.1330 29.24 13027.5 6817.175 9146.305 0.29 15.5700 3332.835

Table 1.

Evaluated mechanical properties of plantain fibers and plantain fibers reinforced polyester composites. From Ref. [43].

S u 1 , S u 2 are tensile strengths in the longitudinal and transverse directions respectively.

v 21 ,   PEFBFRP = E 2 v 12 E 1 = 7030.962 0.38 14,922 = 0.179
1 v 12 v 21 = 1 0.38 0.179 = 1 0.068 = 0.93
v 21 ,   PPSFC = E 2 v 12 E 1 = 6817.175 0.29 13027.5 = 0.152
1 v 12 v 21 = 1 0.29 0.152 = 1 0.068 = 0.956

The reduced stiffness matrix ( ) for PEFBFC and PPSFC is obtained from Eq. (6).

For PEFBFC

E 1 1 v 12 v 21 = 14922 1 0.38 0.179 = 16011.07
v 21 E 1 1 v 12 v 21 = 0.179 16011.07 = 2865.98
E 2 1 v 12 v 21 = 7030.962 1 0.38 0.179 = 7544.113
PEFBFC = 16011.07 2865.98 0 2865.98 7544.113 0 0 0 3622.99 MPa

For PPSFC

E 1 1 v 12 v 21 = 13027.5 1 0.29 0.152 = 13628.23
v 21 E 1 1 v 12 v 21 = 0.152 13628.23 = 2071.49
E 2 1 v 12 v 21 = 6817.175 1 0.29 0.152 = 7131.53
PPSFC = 13628.23 2071.49 0 2071.49 7131.53 0 0 0 3332.835 MPa

The reduced compliance matrix ( β ) is obtained from Eq. (7).

For PEFBFC

1 E 1 = 1 14922 = 6.7 × 10 5 1 / MPa
v 21 E 2 = 0.179 7030.962 = 2.5 × 10 5 1 / MPa
1 E 2 = 1 7030.962 = 1.4 × 10 4 1 / MPa
1 G 12 = 1 3622.99 = 2.8 × 10 4 1 / MPa
β PEFBFC = 6.7 × 10 5 2.5 × 10 5 0 2.5 × 10 5 1.4 × 10 4 0 0 0 2.8 × 10 4 1 / MPa

For PPSFC

1 E 1 = 1 13027.5 = 7.7 × 10 5 1 / MPa
v 21 E 2 = 0.152 6817.175 = 2.2 × 10 5 1 / MPa
1 E 2 = 1 6817.175 = 1.5 × 10 4 1 / MPa
1 G 12 = 1 3332.835 = 3.0 × 10 4 1 / MPa
β PPSFC = 7.7 × 10 5 2.2 × 10 5 0 2.2 × 10 5 1.5 × 10 4 0 0 0 3.0 × 10 4 1 / MPa

3.2 Transformation of elastic constants

The know of the stress-strain relationship in the plantain/polyester composite is completely comprehended by knowing the associated independent engineering elastic constants (E 1, E 2, G 12 and v 12) as previously determined by Ihueze et al. (2013) as in Tables 1 and 2. However, there is need to further establish these elastic properties at different directions of fibers other than directions 1 and 2. Datoo [54] derived various expressions for determination of the elastic properties in the reference axes x-y for any fiber orientation as expressed in Eqs. (8)(14) where c = cosθ , s = sinθ , E x , E y , G xy , v xy , m x and m y are the elastic properties at any fiber orientation θ relative to a reference direction x-y.

Property Polyester resin
Density (g/cm3) 1.2–1.5 (1400 kg/m3)
Young modulus (MPa) 2000–4500
Tensile strength (MPa) 40–90
Compressive strength (MPa) 90–250
Tensile elongation at break (%) 2
Water absorption 24 h at 20°C 0.1–0.3
Flexural modulus (GPa) 11.0
Poisson’s ratio 0.37–0.38
Plantain pseudo stem fibers
Young modulus (MPa) 23,555
UTS (MPa) 536.2
Strain (%) 2.37
Density (kg/m3) 381.966
Plantain empty fruit bunch fibers
Young modulus (MPa) 27,344
UTS (MPa) 780.3
Strain (%) 2.68
Density (kg/m3) 354.151

Table 2.

Mechanical properties of plantain fibers and polyester resin. From Ref. [43].

1 E x = c 4 E 1 + s 4 E 2 + c 2 s 2 1 G 12 2 v 12 E 1 E8
1 E y = s 4 E 1 + c 4 E 2 + c 2 s 2 1 G 12 2 v 12 E 1 E9
1 G xy = c 2 s 2 4 E 1 + 4 E 2 + 8 v 12 E 1 + c 2 s 2 1 G 12 E10
v xy = E x c 4 + s 4 v 12 E 1 c 2 s 2 1 E 1 + 1 E 2 1 G 12 E11

The minor Poissons ratio with respect to the material reference axes is obtained from Eq. (5) such that

v xy E x = v yx E y E12
m x = E x c 3 s 1 G 12 2 v 12 E 1 2 E 1 c s 3 1 G 12 2 v 12 E 1 2 E 2 E13
m y = E y c s 3 1 G 12 2 v 12 E 1 2 E 1 c 3 s 1 G 12 2 v 12 E 1 2 E 2 E14

3.3 Variation of engineering elastic constants with fiber orientation θ

Considering fiber orientation θ° ranging from 0° to 90° in increments of 5° the variation of E x , E y , G xy , v xy , m x and m y has been assessed Plantain Empty Fruit Bunch Fiber Composite (PEFBFC) and Plantain Pseudo Stem Fiber Composite (PPSFC) using Eqs. (8)(14) and presented in Tables 3 and 4, respectively.

S/N θ° E x E y G xy v xy v yx m x m y E x E 2 G xy E 2
1 0 14922.000 7030.962 3622.990 0.380 0.179 0.000 0.000 2.122 0.515
2 5 14769.770 7053.373 3629.270 0.378 0.180 0.116 0.037 2.101 0.516
3 10 14337.670 7121.323 3647.477 0.372 0.185 0.221 0.073 2.039 0.519
4 15 13690.480 7236.950 3675.727 0.362 0.191 0.304 0.111 1.947 0.523
5 20 12911.620 7403.786 3710.985 0.350 0.201 0.362 0.150 1.836 0.528
6 25 12081.300 7626.692 3749.256 0.336 0.212 0.395 0.190 1.718 0.533
7 30 11262.730 7911.712 3785.946 0.320 0.225 0.405 0.231 1.602 0.538
8 35 10497.990 8265.787 3816.391 0.304 0.239 0.398 0.271 1.493 0.543
9 40 9810.462 8696.213 3836.528 0.288 0.255 0.377 0.310 1.395 0.546
10 45 9209.655 9209.655 3843.571 0.271 0.271 0.346 0.346 1.310 0.547
11 50 8696.213 9810.462 3836.528 0.255 0.288 0.310 0.377 1.237 0.546
12 55 8265.787 10497.990 3816.391 0.239 0.304 0.271 0.398 1.176 0.543
13 60 7911.712 11262.730 3785.946 0.225 0.320 0.231 0.405 1.125 0.538
14 65 7626.692 12081.300 3749.256 0.212 0.336 0.190 0.395 1.085 0.533
15 70 7403.786 12911.620 3710.985 0.201 0.350 0.150 0.362 1.053 0.528
16 75 7236.950 13690.480 3675.727 0.191 0.362 0.111 0.304 1.029 0.523
17 80 7121.323 14337.670 3647.477 0.185 0.372 0.073 0.221 1.013 0.519
18 85 7053.373 14769.770 3629.270 0.180 0.378 0.037 0.116 1.003 0.516
19 90 7030.962 14922.000 3622.990 0.179 0.380 0.000 0.000 1.000 0.515

Table 3.

Variation of engineering elastic constants with fiber orientation θ in PEFBFC.

S/N θ° E x E y G xy v xy v yx m x m y E x E 2 G xy E 2
1 0 13027.500 6817.175 3332.835 0.290 0.152 0.000 0.000 1.911 0.489
2 5 12897.620 6830.651 3343.613 0.290 0.154 0.114 0.023 1.892 0.490
3 10 12530.190 6872.014 3375.039 0.291 0.159 0.214 0.047 1.838 0.495
4 15 11983.280 6944.024 3424.350 0.291 0.168 0.293 0.073 1.758 0.502
5 20 11330.960 7051.119 3486.843 0.290 0.180 0.344 0.103 1.662 0.511
6 25 10643.310 7199.198 3555.901 0.287 0.194 0.369 0.136 1.561 0.522
7 30 9974.406 7395.333 3623.334 0.282 0.209 0.371 0.173 1.463 0.532
8 35 9359.192 7647.394 3680.229 0.275 0.224 0.356 0.212 1.373 0.540
9 40 8816.012 7963.509 3718.337 0.265 0.239 0.328 0.252 1.293 0.545
10 45 8351.207 8351.207 3731.758 0.253 0.253 0.292 0.292 1.225 0.547
11 50 7963.509 8816.012 3718.337 0.239 0.265 0.252 0.328 1.168 0.545
12 55 7647.394 9359.192 3680.229 0.224 0.275 0.212 0.356 1.122 0.540
13 60 7395.333 9974.406 3623.334 0.209 0.282 0.173 0.371 1.085 0.532
14 65 7199.198 10643.310 3555.901 0.194 0.287 0.136 0.369 1.056 0.522
15 70 7051.119 11330.960 3486.843 0.180 0.290 0.103 0.344 1.034 0.511
16 75 6944.024 11983.280 3424.350 0.168 0.291 0.073 0.293 1.019 0.502
17 80 6872.014 12530.190 3375.039 0.159 0.291 0.047 0.214 1.008 0.495
18 85 6830.651 12897.620 3343.613 0.154 0.290 0.023 0.114 1.002 0.490
19 90 6817.175 13027.500 3332.835 0.152 0.290 0.000 0.000 1.000 0.489

Table 4.

Variation of engineering elastic constants with fiber orientation θ in PPSFC.

Figure 4 shows the variation of elastic modulus with fiber orientation, it can be seen that the highest value of 14,922 and 13027.5 MPa in the reference x-direction ( E x ) is attained in the fiber orientation angle 0° for PEFBFC and PPSFC respectively. However as fiber orientation angle changes, there is a sharp drop in the value of elastic modulus in the reference x-direction to a respective lowest value of 7030.96 and 6817.18 MPa as the fiber orientation angle increased to 90°. On the contrary, the lowest value of 7030.96 and 6817.18 MPa were recorded for elastic modulus in the reference y-direction ( E y ) as the fiber orientation angle increased from 0° to 90° reaching a peak value of 14,922 and 13027.5 MPa for PEFBFC and PPSFC respectively. The implication is that reinforcements are required to be aligned in the direction of applied load [55]. Although Jones [56] intuitively suggested that highest value of material properties may not necessarily occur along the principal material directions, rather it is essential that transverse reinforcement is needed in unidirectional fiber composites which are subjected to multi axial loading [57].

Figure 4.

Variation of elastic modulus with fiber orientation in PEFBFC and PPSFC.

As can be seen in Figure 5 that shear modulus peaked at 45° fiber orientation and shear modulus was symmetric at about 45° fiber orientation angle for both PEFBFC and PPSFC considered. This implies that the higher in-plane shear resistance is achievable when fiber orientation is 45°. Also the respective minimum value of 3622.99 and 3332.83 MPa at fiber orientation 0° for PEFBFC and PPSFC can be seen to gradually increase to maximum values of 3843.57 and 3731.758 MPa at fiber orientation 45° and then reversed parabolically at 90° where it again reaches to 3622.99 and 3332.84 MPa. Similar trend was obtained by Farooq and Myler [58] who developed efficient procedures for determination of mechanical properties of carbon fiber-reinforced laminated composite panels. This trend in which the value of G xy peaks at 45° fiber orientation angle and lowers at 0° and 90° fiber orientation angle indicates that off-axis reinforcement is very necessary for robust shear stiffness in unidirectional composites [57].

Figure 5.

Variation of shear modulus with fiber orientation in PEFBFC and PPSFC.

Figure 6 shows variation of Poisson’s ratio with fiber orientation, the graph depicts a gradual drop of major Poisson’s ratio ( v xy ) for PEFBFC and PPSFC respectively from 0.38 and 0.29 when fibers are aligned at 0° orientation angle to a lowest value of 0.18 and 0.15 value when fibers were aligned at 90° orientation angle. Additionally, the minor Poissons ratio ( v yx ) for PEFBFC and PPSFC increased respectively from 0.18 and 0.15 when fibers are aligned at 0° orientation angle to a highest value of 0.38 and 0.29 value when fibers were aligned at 90° orientation angle.

Figure 6.

Variation of Poisson’s ratio with fiber orientation in PEFBFC and PPSFC.

Figure 7 depicts the variation of shear coupling coefficient with fiber orientation, equal magnitude of shear coupling effect was obtained at 45° fiber orientation angle for both PEFBFC and PPSFC considered. Gibson [57] reported that shear coupling coefficient is a measure of the amount of shear strain developed in the xy plane per unit normal strain along the direction of the applied normal stress σ x . Figure 7 clearly indicate that the maximum value of the shear coupling coefficient in the reference x-direction for PEFBFC and PPSFC was attained at 30° fiber orientation angle while the coefficient in the reference y-direction for PEFBFC and PPSFC was attained at 60° fiber orientation angle. This is an indication that as the shear-coupling ratio increases, the amount of shear coupling increases.

Figure 7.

Variation of shear coupling coefficient with fiber orientation in PEFBFC and PPSFC.

Advertisement

4. Tsai-Hill failure criteria assessment of longitudinal tensile strength

Failure theory is essential in determining whether the composite has failed. Literature review has shown that results of failure prediction depend on failure criterion applied and one major failure criteria used in the industries is Tsai-Hill and failure criteria. Additionally, since composites ultimate tensile strength and strain depend on the fiber orientation, a failure criterion must be used in which the applied stress system is also in material axis [54]. Tsai-Hill theory considers an interaction of the stresses in the fiber direction. It postulates that failure can only occur in reinforced composites when the failure index exceeds 1, hence Eq. (19) must be satisfied to avoid failure.

By considering an arbitrary positive angle θ with reference to the x-axis in Figure 3, Ihueze et al. [43] transformed the stresses within the global axes (x-y) into material axes 1–2 as given in Eq. (15)

σ 1 σ 2 τ 12 = c 2 s 2 2 sc s 2 c 2 2 sc sc sc c 2 s 2 σ x σ y τ xy E15

where c = cosθ and s = sinθ . Taking longitudinal direction stresses as σ y = τ xy = 0 and thus

σ 1 = σ x cos 2 θ E16
σ 2 = σ x sin 2 θ E17
τ 12 = σ x cosθsinθ E18

Considering Tsai-Hill failure criterion and setting the failure index as 1 for the composite failure to occur:

σ 1 S u 1 2 + σ 2 S u 2 2 + τ 12 τ max 2 σ 1 S u 1 σ 2 S u 1 = 1 E19

Substituting the appropriate value in Eq. (19) we have for PEFBFC

σ x cos 2 θ 410.15 2 + σ x sin 2 θ 37.3397 2 + σ x cosθsinθ 19.3100 2 σ x cos 2 θ 410.15 σ x sin 2 θ 410.15 = 1
σ x ,   PEFBFC = 1 cos 4 θ 410.15 2 + sin 4 θ 37.3397 2 + cos 2 θ sin 2 θ 19.3100 2 cos 2 θ sin 2 θ 410.15 2

And for PPSFC

σ x cos 2 θ 288.10 2 + σ x sin 2 θ 33.1330 2 + σ x cosθsinθ 15.5700 2 σ x cos 2 θ 288.10 σ x sin 2 θ 288.10 = 1
σ x ,   PPSFC = 1 cos 4 θ 288.10 2 + sin 4 θ 33.1330 2 + cos 2 θ sin 2 θ 15.5700 2 cos 2 θ sin 2 θ 288.10 2

Hence the value for σ x is then calculated for orientation ranging from 0° to 90° as shown in Table 5.

S/N Orientation θ PEFBFC PPSFC
1 0 410.15 288.1
2 5 195.86 152.635
3 10 108.786 86.8937
4 15 75.4556 60.6747
5 20 58.6356 47.3142
6 25 48.8262 39.5106
7 30 42.6439 34.6088
8 35 38.6038 31.4352
9 40 35.9607 29.4005
10 45 34.3043 28.1833
11 50 33.3926 27.5977
12 55 33.0716 27.5303
13 60 33.2311 27.9038
14 65 33.774 28.6492
15 70 34.5913 29.679
16 75 35.5418 30.8595
17 80 36.4447 31.9912
18 85 37.0995 32.8242
19 90 37.3397 33.133

Table 5.

Longitudinal tensile strength variation with fiber orientation angle.

The variation of longitudinal tensile strength with fiber orientation for PEFBFC and PPSFC has been presented in Figure 8, it can be seen that the tensile strength equals 410.15 and 288.1 MPa which are the longitudinal tensile strength for PEFBFC and PPSFC respectively when fiber orientation angle is 0°; on the other hand, the tensile strength equals 37.3397 and 33.133 MPa which are the transverse tensile strength for PEFBFC and PPSFC respectively when fiber orientation angle is 90°.

Figure 8.

Variation of longitudinal tensile strength with fiber orientation for PEFBFC and PPSFC.

Advertisement

5. Variation of tangential stress and modulus around a structural discontinuity

Structural discontinuity arising from holes in reinforced composites created for joining or access purposes causes stress concentration at the point of discontinuity [59]. Adequate comprehension of stress redistribution pattern and concentrations is helpful for proficient and safe structural designs. Unlike in ductile materials where stress concentration is of no much ado, plantain fiber-reinforced composites may be sufficiently brittle, hence every form of stress concentration and structural discontinuity has to be properly designed. In a typical scenario where a circular hole is created in the composite as shown in Figure 9, assuming no interlaminar stresses exist around the free edge of the hole, the ply is nominally stressed by σ 1 , σ 2 , σ 12 some distance away from the hole as indicated. Lekhnitskii [60] derived various useful expressions for stress distribution around holes in a composite plate, the tangential elastic modulus E at an angular position is determined using Eq. (20).

Figure 9.

Depiction of hole in the plantain fiber-reinforced composites sample.

E = 1 sin 4 E 1 + 1 G 12 2 v 12 E 1 sin 2 cos 2 + cos 4 E 2 E20

Hence the tangential stress σ at the perifery of the hole with an angle is found from Eq. (21).

σ = E E 1 A σ 1 + B σ 2 + C σ 12 E21

where

A = cos 2 + 1 + p sin 2
B = q q + p cos 2 sin 2
C = 1 + q + p p sin 2
p = 2 q v 12 + E 1 G 12
q = E 1 E 2

Using the stress transformation matrix and replacing the axes system 1–2 by radial (r)-tangential (∅), we can resolve the tangential stress σ back into the material axesin Eq. (22) for proper strength evaluation

σ r σ τ r = c 2 s 2 2 sc s 2 c 2 2 sc sc sc c 2 s 2 σ x σ y τ xy E22

At the edge of the hole, only the tangential stress σ > 0, thus σ r = σ r = 0 in Eq. (22), therefore

σ 1 = σ x = σ sin 2 E23
σ 2 = σ y = σ cos 2 E24
σ 12 = σ xy = σ cos sin E25

Using the maximum stress criterion, the material will fail when any stress value in the material axes exceeds their respective ultimate strenght. Such that

σ 1 S u 1 < 1 E26
σ 2 S u 2 < 1 E27
σ 12 τ max < 1 E28

The left hand side of Eqs. (26)(28) represents the failure indices (FI). The maximum failure index (FI) for the applied stress is factored in to obtain the load factor. Due to the inherent material orthotropy, the failure zone of the plantain fiber-reinforced composite as a result of structural discontinuity may not necessarily occur at the point of maximum stress concentration, therefore it is important to assess the extent of variation of the tangential stress around the hole edge and the failure index using maximum stress theory at other points aside the point of maximum stress concentration. Also in the present consideration we take a simplified scenario where σ 2 = σ 12 = 0 such that the ply of dimensions 150 × 19.05 × 3.2 mm with a circular hole at the center is subjected to only nominal axial stress σ PEFBFC = 34 MPa and σ PPSFC = 29 MPa . Tables 6 and 7 depict the values of tangential stress, material axis stress and tangential modulus as computed using Eq. (20)(25).

S/N Eø σø σ1 σ2 σ12 F.I.1 F.I.2 F.I.12 Stress conc.
0 7030.962 −23.338 0.000 −23.338 0.000 0.000 0.625 0.000 −0.69
5 7053.373 −22.807 −0.173 −22.634 1.980 0.000 0.606 0.103 −0.67
10 7121.323 −21.211 −0.640 −20.571 3.627 0.002 0.551 0.188 −0.62
15 7236.950 −18.542 −1.242 −17.300 4.636 0.003 0.463 0.240 −0.55
20 7403.786 −14.786 −1.730 −13.056 4.752 0.004 0.350 0.246 −0.43
25 7626.692 −9.917 −1.771 −8.146 3.799 0.004 0.218 0.197 −0.29
30 7911.712 −3.903 −0.976 −2.927 1.690 0.002 0.078 0.088 −0.11
35 8265.787 3.303 1.087 2.217 −1.552 0.003 0.059 0.080 0.10
40 8696.213 11.751 4.855 6.896 −5.786 0.012 0.185 0.300 0.35
45 9209.655 21.484 10.742 10.742 −10.742 0.026 0.288 0.556 0.63
50 9810.462 32.515 19.080 13.434 −16.010 0.047 0.360 0.829 0.96
55 10497.994 44.784 30.050 14.733 −21.042 0.073 0.395 1.090 1.32
60 11262.726 58.103 43.577 14.526 −25.159 0.106 0.389 1.303 1.71
65 12081.304 72.076 59.203 12.873 −27.607 0.144 0.345 1.430 2.12
70 12911.622 86.025 75.962 10.063 −27.648 0.185 0.269 1.432 2.53
75 13690.476 98.951 92.323 6.628 −24.738 0.225 0.178 1.281 2.91
80 14337.671 109.599 106.294 3.305 −18.742 0.259 0.089 0.971 3.22
85 14769.765 116.668 115.782 0.886 −10.130 0.282 0.024 0.525 3.43
90 14922.000 119.152 119.152 0.000 0.000 0.291 0.000 0.000 3.50
95 14769.765 116.668 115.782 0.886 10.130 0.282 0.024 0.525 3.43
100 14337.671 109.599 106.294 3.305 18.742 0.259 0.089 0.971 3.22
105 13690.476 98.951 92.323 6.628 24.738 0.225 0.178 1.281 2.91

Table 6.

Variation of tangential stress, material axis stress and tangential modulus at the edge of material discontinuity in PEFBFC.

S/N Eø σø σ1 σ2 σ12 F.I.1 F.I.2 F.I.12 Stress conc.
0 6817.175 −20.978 0.000 −20.978 0.000 0.000 0.633 0.000 −0.72
5 6830.651 −20.459 −0.155 −20.304 1.776 0.001 0.613 0.114 −0.71
10 6872.014 −18.909 −0.570 −18.339 3.234 0.002 0.554 0.208 −0.65
15 6944.024 −16.346 −1.095 −15.251 4.086 0.004 0.460 0.262 −0.56
20 7051.119 −12.791 −1.496 −11.295 4.111 0.005 0.341 0.264 −0.44
25 7199.198 −8.269 −1.477 −6.792 3.167 0.005 0.205 0.203 −0.29
30 7395.333 −2.793 −0.698 −2.095 1.209 0.002 0.063 0.078 −0.10
35 7647.394 3.635 1.196 2.439 −1.708 0.004 0.074 0.110 0.13
40 7963.509 11.024 4.555 6.469 −5.428 0.016 0.195 0.349 0.38
45 8351.207 19.391 9.695 9.695 −9.695 0.034 0.293 0.623 0.67
50 8816.012 28.736 16.863 11.873 −14.150 0.059 0.358 0.909 0.99
55 9359.192 39.014 26.179 12.835 −18.331 0.091 0.387 1.177 1.35
60 9974.406 50.087 37.565 12.522 −21.688 0.130 0.378 1.393 1.73
65 10643.308 61.651 50.640 11.011 −23.614 0.176 0.332 1.517 2.13
70 11330.958 73.175 64.615 8.560 −23.518 0.224 0.258 1.510 2.52
75 11983.277 83.856 78.239 5.617 −20.964 0.272 0.170 1.346 2.89
80 12530.185 92.667 89.873 2.794 −15.847 0.312 0.084 1.018 3.20
85 12897.622 98.526 97.778 0.748 −8.554 0.339 0.023 0.549 3.40
90 13027.500 100.587 100.587 0.000 0.000 0.349 0.000 0.000 3.47

Table 7.

Variation of tangential stress, material axis stress and tangential modulus at the edge of material discontinuity in PPSFC.

Tangential stress distribution at hole edge for PEFBFC and PPSFC are shown in Figure 10, the maximum stress value of 119.15 and 100.587 MPa was attained at angular position ø = 90° for PEFBFC and PPSFC respectively. However, considering various failure indices in Tables 5 and 6, failure will be initiated at ø = 70° for PEFBFC with stress concentration factor of 2.53 and ø = 65° for PPSFC with stress concentration factor of 2.13 which are less than the stress concentration around the peak stress when angular position is 90°.

Figure 10.

Tangential stress distribution at a hole edge for PEFBFC and PPSFC.

Advertisement

6. Conclusions

The utilization of plantain fiber-reinforced composites in structural applications empowers architects to acquire huge accomplishments in the usefulness, security and economy of development. These materials have high proportion of strength-to-density ratio, can be tailored to posses certain mechanical properties. The elastic constants of plantain fiber-reinforced composites depend greatly on fiber orientation with notable anisotropic characteristics which makes it less attractive for applications involving lugs and fittings. The present report amplified some notable design procedures in handling such limitations in plantain fiber-reinforced composites using relevant failure theories. Both plantain EFBFRC and PSFRC showed similar trends in response to the design scenario considered. Be that as it may be, a proficient utilization of plantain fiber-reinforced composites in structural applications requires a cautious assessment of all influential factors.

References

  1. 1. Dehmous H, Welemane H, Karama M, Tahar KA. Reliability approach for fibre-reinforced composites design. International Journal for Simulation and Multidisciplinary Design Optimization. 2008;2(1):1-9
  2. 2. Okafor CE, Okafor EJ, Obodoeze JJ, Ihueze CC. Characteristics and reliability of polyurethane wood ash composites for packaging and containerisation applications. Journal of Materials Science Research and Reviews. 2018;1(3):1-10
  3. 3. Xie HB, Wang YF. Reliability analysis of CFRP-strengthened RC bridges considering size effect of CFRP. Materials. 2019;12(14):2247
  4. 4. Beaumont PW, Soutis C, Hodzic A, editors. The Structural Integrity of Carbon Fiber Composites: Fifty Years of Progress and Achievement of the Science, Development, and Applications. Switzerland: Springer; 2016
  5. 5. Pei X, Han W, Ding G, Wang M, Tang Y. Temperature effects on structural integrity of fiber-reinforced polymer matrix composites: A review. Journal of Applied Polymer Science. 2019;136(45):48206
  6. 6. Bittrich L, Spickenheuer A, Almeida JHS, Müller S, Kroll L, Heinrich G. Optimizing variable-axial fiber-reinforced composite laminates: The direct fiber path optimization concept. Mathematical Problems in Engineering. 2019;2019:1-11
  7. 7. Prasad SV, Kumar GA, Sai KP, Nagarjuna B. Design and optimization of natural fibre reinforced epoxy composites for automobile application. In: AIP Conference Proceedings. Vol. 2128(1). Melville, NY, USA: AIP Publishing; 2019. p. 020016
  8. 8. Wang Y, Cui G, Shao Z, Bao Y, Gao H. Optimization of the hot pressing process for preparing flax fiber/PE thermoplastic composite. Mechanical Engineering Science. 2019;1(1):41-45
  9. 9. Foulc MP, Bergeret A, Ferry L, Ienny P, Crespy A. Study of hygrothermal ageing of glass fibre reinforced PET composites. Polymer Degradation and Stability. 2005;89(3):461-470
  10. 10. Shettar M, Chaudhary A, Hussain Z, Kini UA, Sharma S. Hygrothermal studies on GFRP composites: A review. In: MATEC Web of Conferences. Vol. 144. Karnataka, India: EDP Sciences; 2018. p. 02026
  11. 11. Ngo TD. Natural fibers for sustainable bio-composites. In: Natural and Artificial Fiber-Reinforced Composites as Renewable Sources. Rijeka: IntechOpen; 2018. p. 107
  12. 12. Rahmani H, Najaf SHM, Ashori A, Golriz M. Elastic properties of carbon fibre-reinforced epoxy composites. Polymers and Polymer Composites. 2015;23(7):475-482
  13. 13. Schürmann H. Konstruieren mit Faser-Kunststoff-Verbunden. Berlin Heidelberg: Springer-Verlag; 2007
  14. 14. Younes R, Hallal A, Fardoun F, Chehade FH. Comparative review study on elastic properties modeling for unidirectional composite materials. In: Hu N, editor. Composites and their Properties. London: IntechOpen; 2012
  15. 15. Goncalves R, Trinca AJ, Pellis BP. Elastic constants of wood determined by ultrasound using three geometries of specimens. Wood Science and Technology. 2014;48:269-287
  16. 16. Longo R, Delaunay T, Laux D, El Mouridi M, Arnould O, Le Clézio E. Wood elastic characterization from a single sample by resonant ultrasound spectroscopy. Ultrasonics. 2012;52:971-974
  17. 17. Majano-Majano A, Fernandez-Cabo JL, Hoheisel S, Klein MA. Test method for characterizing clear wood using a single specimen. Experimental Mechanics. 2012;52:1079-1096
  18. 18. Vazquez C, Gonçalves R, Bertoldo C, Bano V, Vega A, Crespo J, et al. Determination of the mechanical properties of Castanea sativa mill. using ultrasonic wave propagation and comparison with static compression and bending methods. Wood Science and Technology. 2015;49:607-622
  19. 19. Adewole S. Plantain (Musa acuminata) value chain analysis in Ondo State, Nigeria. Scientific Papers: Management, Economic Engineering in Agriculture & Rural Development. 2017;17(3):25-36
  20. 20. Widsten P, Kandelbauer A. Adhesion improvement of lignocellulosic products by enzymatic pre-treatment. Biotechnology Advances. 2008;26:379
  21. 21. Ojediran EO, Ibrahim HK, Adebisi LO, Belewu KY, Owolawi S. Analysis of profitability and determinants of plantain production in IFE agricultural development project (ADP) zone of Osun state, Nigeria. Fuoye Journal of Agriculture and Human Ecology. 2017;1(1):77-86
  22. 22. HTA. Constraints to Banana Production. Ibadan, Nigeria: International Institute of Tropical Agriculture; 1990. pp. 142-144
  23. 23. Norgrove L, Hauser S. Improving plantain (Musa spp. AAB) yields on smallholder farms in West and Central Africa. Food Security. 2014;6(4):501-514
  24. 24. Akinyemi SOS, Aiyelaagbe IOO, Akyeampong E. Plantain (Musa spp.) cultivation in Nigeria: A review of its production, marketing and research in the last two decades. In: Dubois T, Hauser S, Staver C, Coyne D, editors. Proceedings of an International Conference on Banana & Plantain in Africa Harnessing International Partnerships to Increase Research Impact. Acta Horticulturae. Vol. 879. 2010. pp. 211-218
  25. 25. Kaine AIN, Okoje LJD. Estimation of cost and return of plantain production in Orhionwon Local Government Area, Edo State, Nigeria. Asian Journal of Agriculture and Rural Development. 2014;4(2):162-168
  26. 26. Kainga PE, Nnadi CD, Jimmy SP, Ugorji KS. Economics of plantain production in oil producing communities of Bayelsa State, Nigeria. Nigerian Journal of Agricultural Economics. 2015;5(1):54-60
  27. 27. Ahmed S. The Survey of Nigeria Agriculture by Raw Materials Research and Development Council (RMRDC). 2004:123
  28. 28. Olumba CC. Productivity of improved plantain technologies in Anambra State, Nigeria. African Journal of Agricultural Research. 2014;9(29):2196-2204
  29. 29. Samuel OD, Agbo S, Adekanye TA. Assessing mechanical properties of natural fiber reinforced composites for engineering applications. Journal of Minerals and Materials Characterization and Engineering. 2012;11(1):1-5
  30. 30. Cadena Ch EM, Vélez R, M J, Santa JF, Otálvaro G. Natural fibers from plantain pseudostem (Musa paradisiaca) for use in fiber-reinforced composites. Journal of Natural Fibers. 2017;14(5):678-690
  31. 31. Adeniyi AG, Ighalo JO, Onifade DV. Banana and plantain fiber-reinforced polymer composites. Journal of Polymer Engineering. 2019;39(7):597-611
  32. 32. Chimekwene CP, Fagbemi EA, Ayeke PO. Mechanical properties of plantain empty fruit bunch fiber reinforced epoxy composite. International Journal of Research in Engineering, IT and Social Sciences. 2012;2(6):86-94
  33. 33. Okafor EC, Ihueze CC, Nwigbo SC. Optimization of hardness strengths response of plantain fibers reinforced polyester matrix composites (PFRP) applying Taguchi robust design. International Journal of Science & Emerging Technologies. 2013;5(1):1-11
  34. 34. Alvarez-López C, Rojas OJ, Rojano B, Ganán P. Development of self-bonded fiberboards from fiber of leaf plantain: Effect of water and organic extractives removal. BioResources. 2014;10(1):672-683
  35. 35. Obijiaku JC, Kamalu CIO, Osoka EC, Onyelucheya OE, Uzondu FN, Obibuenyi JI. Effects of extraction techniques on the yield and mechanical properties of empty plantain fruit bunch fibers. International Journal of Engineering and Management Research (IJEMR). 2015;5(6):494-500
  36. 36. Ihueze CC, Okafor CE. Optimal design for flexural strength of plantain fibers reinforced polyester matrix. Journal of Innovative Research in Engineering and Sciences. 2016;4(4):520-537
  37. 37. Ihueze CC, Okafor EC. Response surface optimization of the impact strength of plantain fiber reinforced polyester for application in auto body works. Journal of Innovative Research in Engineering and Science. 2014;4:505-520
  38. 38. Okafor CE, Godwin HC. Evaluation of compressive and energy adsorption characteristics of plantain fiber reinforced composites. World Journal of Engineering and Physical Sciences. 2014;2(3):036-048
  39. 39. Oreko BU, Otanocha OB, Emagbere E, Ihueze CC. Analysis and application of natural fiber reinforced polyester composites to automobile fender. Covenant Journal of Engineering Technology (Special Edition). 2018;1(1):1-12
  40. 40. Imoisili PE, Tonye DI, Victor PA, Elvis OA. Effect of high-frequency microwave radiation on the mechanical properties of plantain (Musa paradisiaca) fibre/epoxy biocomposite. Journal of Physical Science. 2018;29(3):23-35
  41. 41. Imoisili PE, Fadare OB, Popoola AV, Okoronkwo AE. Effect of chemical treatment on the morphology and mechanical properties of plantain (Musa paradisiaca) fibre. IOSR Journal of Applied Chemistry. 2017;10(5):70-73. DOI: 10.9790/5736-1005017073
  42. 42. Sinebe JE, Chukwuneke JL, Omenyi SN. Implications of interfacial energetics on mechanical strength of fiber reinforced polymer matrix. International Journal of Materials Engineering. 2019;9(1):1-7. DOI: 10.5923/j.ijme.20190901.01
  43. 43. Ihueze CC, Okafor CE, Okoye CI. Natural fiber composite design and characterization for limit stress prediction in multiaxial stress state. Journal of King Saud University-Engineering Sciences. 2015;27(2):193-206
  44. 44. Kenedi PP, Vignoli LL, Duarte BT, Matos FCDA, Dias HOT. Orthotropic elastic properties assessment of sandwich laminates. Journal of Aerospace Technology and Management. 2017;9(3):389-396
  45. 45. Hwang SF, Liu HT. Prediction of elastic constants of carbon fabric/polyurethane composites. In: Solid State Phenomena. Vol. 258. Baech, Switzerland: Trans Tech Publications; 2017. pp. 233-236
  46. 46. Ren B, Noda J, Goda K. Effects of fiber orientation angles and fluctuation on the stiffness and strength of sliver-based green composites. Journal of the Society of Materials Science, Japan. 2010;59(7):567-574
  47. 47. Kumar KV, Reddy PR, Shankar DR. Influence of angle ply orientation of stacking on mechanical properties of glass-polyester composite laminate. International Journal of Engineering and Advanced Technology (IJEAT). 2013;2249:8958
  48. 48. Cordin M, Bechtold T, Pham T. Effect of fibre orientation on the mechanical properties of polypropylene–lyocell composites. Cellulose. 2018;25(12):7197-7210
  49. 49. Ihueze CC, Okafor EC, Ujam AJ. Optimization of tensile strengths response of plantain fibers reinforced polyester composites (PFRP) applying Taguchi robust design. Innovative Systems Design and Engineering. 2012;3(7):64-76
  50. 50. Jules EJ, Tsujikami T, Lomov SV, Verpoest I. Effect of fibres length and fibres orientation on the predicted elastic properties of long fibre composites. Macromolecular Symposia. 2004;17:1-109
  51. 51. Venetis J, Sideridis E. Elastic constants of fibrous polymer composite materials reinforced with transversely isotropic fibers. AIP Advances. 2015;5(3):037118
  52. 52. Cuartas VM, Perrin M, Pastor ML, Welemane H, Cantarel A, Karama M. Determination of the elastic properties in CFRP composites: Comparison of different approaches based on tensile tests and ultrasonic characterization. Advances in Aircraft and Spacecraft Science. 2015;2(n° 3):249-260
  53. 53. Clyne TW, Hull D. An Introduction to Composite Materials. Cambridge, United Kingdom: Cambridge University Press; 2019
  54. 54. Datoo MH. Mechanics of Fibrous Composites. New York: Elsevier Science Publishing Co. Inc; 2012
  55. 55. Maheshwari P, Misra A, Kumar A. Investigate the variation in elastic constants and failure strength of lamina with FVF and FOA for JFRP lamina. International Journal of Applied Engineering Research. 2018;13(9):157-161
  56. 56. Jones RM. Mechanics of Composite Materials. 2nd ed. Philadelphia, PA: Taylor and Francis, Inc.; 1999
  57. 57. Gibson RF. Principles of Composite Material Mechanics. 4th ed. Boca Raton, FL: CRC Press; 2016
  58. 58. Farooq U, Myler P. Efficient determination of mechanical properties of carbon fibre-reinforced laminated composite panels. ARPN Journal of Engineering and Applied Sciences. 2017;12(5):1375-1392
  59. 59. Koussios S, Beukers A. Lekhnitskii’s formalism for stress concentrations around irregularities in anisotropic plates: Solutions for arbitrary boundary conditions. In: Variational Analysis and Aerospace Engineering. New York, NY: Springer; 2009. pp. 243-265
  60. 60. Lekhnitskii SG. Anisotropic Plates. 2nd ed. London, UK: Gordon and Breach; 1968

Written By

Christian Emeka Okafor and Christopher Chukwutoo Ihueze

Submitted: 22 November 2019 Reviewed: 18 December 2019 Published: 15 July 2020