Skip to main content

The exchange of Musa spp. fibre in composite fabrication: a systematic review

Abstract

Background

The areas of application of natural fibres have gained popularity in recent times due to their attractive advantages when compared with other materials of engineering. These advantages include lightness, cost-effectiveness, and ease of processing, ecological friendliness, and durability. Previously, farmers only harvest Musa spp. fruits for their food values and packaging purposes.

Main body of the abstract

Several research works have been undertaken which accentuate the applications of the assumed waste portions of Musa spp. (banana and plantain) specifically Musa spp. fibre as a reinforcement material in composite manufacture. As a material for reinforcement in composites, the characterization, treatment, and fabrication techniques; elemental, chemical, and mechanical properties of Musa spp. fibre have been analysed. The mechanical properties of banana fibre reinforcement in polyester, epoxy, cement, and plastics composites were evaluated with those of other biodegradable fibres to explicate their relationships.

Short conclusion

This review aims to explore the current state of knowledge on the interaction of Musa spp. fibre in composite manufacture, to aid intending researchers with ample knowledge on the choice of material in bio-based composite design.

Background

Musa spp. (banana and plantain) are one of the oldest crops grown by farmers the world over. Banana in Arabic means ‘finger’. It belongs to the family of crops known as Musaceae and exists in about 300 species with only 20 varieties consumed as food. About 70 million metric tons of Musa spp. is produced annually in the temperate regions of the world (Muller and Krobjilowski 2003; Joshi et al. 2004; Umair 2006). As a result of identified ecological reasons, the application of natural material in the manufacturing and industrial sector has gained prominence (Xu et al. 2015; Monteiro et al. 2014). Two key strategies on which natural fibres composite (NFC) production is based are preventing the depletion of forest resources and ensuring good economic returns (Sathiyamurthy et al. 2013). Because of its enhanced modulus and lightweight, compared to other traditional engineering materials like wood, metal, and steel, natural fibre-reinforced polymer composites have been greatly studied by several researchers (Adekomaya and Adama 2017; Singha et al. 2010).

‘Due to their unique strength properties, availability, lightweight, easy separation, improved energy recovery, high resilience, non-corrosive design, low density, low cost, renewability, and biodegradability, composites strengthened using natural fibres have advantages over conventional reinforcement materials such as glass and carbon fibre’ (Mahaboob 2011; Arpitha and Yogesha 2017). Composites are structural materials mixed at a macroscopic level consisting of two or more constituents (not soluble in each other) (Arpitha and Yogesha 2017). In the continuous phase or matrix or resin (thermosets and thermoplastics), composites consist of the reinforcing or discontinuous phase (metal, ceramic, or fibre) embedded (Aminudin et al. 2011). For example, naturally occurring lignin and wood composites contain cellulose fibre (Arun et al. 2016). Fibre-reinforced composites are made up of fibres as a matrix and polymer as reinforcement.

Most of the shortcomings in the properties of plastics in manufacturing are improved upon by the introduction of natural fibres into waste plastics (polymer) (Li et al. 2007). It has been discovered that the application of natural fibre to composites decreases their overall weight by 10% with ease of processing improved significantly by 80% also, the cost of components is lowered by 5% in comparison with fibreglass-reinforced components (Sakthivei and Ramesh 2013). Most engineering components adopted in the production of vehicles, aircraft, domestic appliances, and packaging industries are protected from environmental attack and corrosion, using waterproof, relatively good strength, and corrosion-resistant materials (Mohammed et al. 2015). Because of their lightweight and high strength, polymeric composites play a very important role in such applications (Sathiyamurthy et al. 2013). There has been a major change between natural and synthetic (manmade) materials in the past couple of years. We have seen it in the growth of fresh foods and the market for eco-friendly vehicles for almost any product imaginable with natural materials. Natural fibre’s global market size was estimated at $4.46 billion as of 2016 and have been forecasted to have a compound annual growth rate of 11.8% from 2016 to 2024 (https: et al., 10.1155, 2015, 243947. xxxx).

For automotive applications, the lightweight and super-improved mechanical properties of NFC materials made them an excellent choice. ‘In North America, light vehicles are one of the most important markets for plastics and polymer composites and had evolved dramatically over the last five decades. There are more than 1000 plastic components in the typical light vehicle, which make up about 8.6% of the overall weight and about 50% of the vehicle's volume’ (GVR 2018). This helped to make vehicles lighter, increased passenger safety, and also improved fuel efficiency, and reduced emissions of greenhouse gases. Fibre materials' susceptibility to moisture and poor bonding with polymer matrices, however, appear to hamper the growth prospects of NFCs. The application of NFCs to car interiors in the automotive industry has been limited by the swelling reaction when exposed to humidity (Swift et al. 2015). Among the different synthetic materials that had been investigated for automotive use as an alternative to iron and steel, plastics have a large share. The studies of filled plastic composites have stimulated tremendous interest in meeting the scarcity of plastic materials over the last decade (Lightsey 1983). From everyday papers to complex structures, computer parts, etc., plastics are used for almost all (Rai and Jai Singh 1986), as a result of lightweight, low water absorption, elevated stiffness, and strength. Synthetic fibres such as nylon, rayon, aramid, glass, and carbon are widely used as plastic reinforcements. There is currently a need to look for their alternate, which is nothing but normal, due to unpredictable conditions of scarcity and the cost of petroleum and its by-products. Vegetable/plant fibres have proved to be an alternative fibre to their synthetic counterpart in recent years.

Natural fibres are less expensive, biodegradable, and do not present a health risk. The natural fibres are readily accessible and abundant in nature. They serve as a viable alternative in polymer reinforcement because of their low cost and renewability compared to synthetic fibres which are expensive and non-renewable. A multitude of natural fibres including wood sawdust, Borassus fruit, kenaf, okra, bamboo, betel nut, cannabis, abaca, coconut, rice husks, walnut, sisal, Arenga pinnata, roselle, cotton stalk, wheat straw, pineapple leaves, jute, and many others, have been studied as composite fillers (Rao and K. and MohanaRao et al. 2007). In West Africa, Musa spp. is extensively cultivated. Nigeria is among the principal producers of Musa spp. (Venkateshwaran and Elayaperumal 2010; Mukhopadhyay et al. 2008). In general, the plant is cut after harvesting the fruit, leaving the lower part of the plant for the fresh stalk to emerge from Pothan et al. (2007). Growers harvest only Musa spp. fruit for food and fruit wrapping leaves. The other parts of the plant are regarded as waste and a source of natural fibres used in composites for reinforcement (Becker et al. 2013). Banana wastes include rotting berries, peels, a bunch of empty fruit, leaves, rhizome, and pseudostem (Ihueze and Okafor 2014). When properly pretreated and prepared, fruit bunches, pseudo-stems (Abdullah et al. 2014; Imoisili et al. 2017; Srinivasan et al. 2014), and peels (Brindha et al. 2017; Naidu et al. 2013) are very good sources of fibres for the development of reinforced composites and other applications such as pyrolysis feedstock, textiles, and paper manufacture and in the preparation of adsorbent (Srinivasan et al. 2014).

As a result of their excellent mechanical properties, ease of processing, and comparatively low cost, polymer composite reinforced with fibre materials is adapted in various manufacturing processes (Yusuf et al. 2019; Santhosh et al. 2014). A plethora of research work on Musa spp. fibres as fillers in composites manufacture have been performed over the years. The goal of this study is to discuss the existing state of awareness of the interplay of Musa spp. fibre in composite manufacturing to assist planning researchers with comprehensive knowledge of bio-based composite design and choice of material.

Main text

Musa spp. fibre extraction

The extraction of fibre from the body tissue of Musa spp. is known as retting (Amir et al. 2017). Retting can be accomplished through Biological, chemical, physical and mechanical processes. Biological retting can be achieved artificially or naturally with the use of microorganisms or enzymes. Chemical agents, for example, Sodium hydroxide, surfactants, are applied in chemical retting. Steam explosion and ultrasound retting are categorized under physical retting. The production of green fibres is achieved through mechanical means and it is referred to as mechanical retting (Mohanty et al. 2005).

Musa spp. fibre treatment

There are disadvantages associated with using cellulosic fibres in composites as reinforcement materials. These include the incompatibility of hydrophobic polymer matrices with hydrophilic fibres (poor interfacial adhesion), aggregate-form propensity during manufacturing, and low moisture resistance. These properties play important roles in determining the physical, mechanical, structural, and thermal properties of the composites (Sisti et al. 2018). Researchers have shown that chemical modifications of fibres like acetylation, acrylation, and treatment with alkali, polystyrene, styrene–maleic, and anhydride increased the composite's tensile properties by growing the dispersion of the fibres in the composites, reducing the hydrophobicity of the fibre, and improving the compatibility of fibre/matrix through mechanical anchoring, physical, and chemical bonding (Poletto and Zattera 2015; Haneefa et al. 2008).

Coupling agents helped in improving the interfacial adhesion between matrices and fibres, however, most researchers preferred using chemical treatments in resolving these disadvantages (Li et al. 2007; Zin et al. 2018; Lyn 2018; Cai et al. 2015). Physically, techniques such as calendar, heat treatment, electric discharge (corona method), and mercerization, were used in treatment fibre. These methods change the fibres surface and structural properties with a significant impact on polymer adhesion, but they do not influence the fibre’s chemical constituents (Fiore et al. 2016). Mercerization, or alkaline therapy, is a widely used physical process. It is a chemical procedure in which natural fibres are immersed in a relatively concentrated aqueous solution with a solid base to achieve adequate swelling by removing hemicelluloses, lignin, waxy materials, and impurities from the fibre cell wall surface. As suggested by ASTM-D1965, mercerization is a concept method of subjecting vegetable fibres to an interaction with a reasonably concentrated strong base aqueous solution to produce great swelling without interfering with the fine structure, dimension, morphology, and mechanical properties of the fibre (Adekunle 2015). In improving the resistance to water absorption and disrupt the network structure of hydrogen bond, mercerization is typically used to eliminate or change the surface OH groups of fibres made of cellulose, thereby raising the roughness of the fibre’s surface (Osoka and Onukwuli 2015). The strength of the fibre and absorbency resulting from the swelling of its cell walls are improved by the alkali solution (Kim 2015).

Alkaline fibre treatment depolymerizes cellulose and improves the fibre's surface roughness, which in turn allows the fibre–matrix interface to acquire greater hydrogen bond extension. The application of natural fibre alkali treatment also increases the tensile strength, fibre wetting through fibrillation, adhesion to the fibre–matrix caused by the elimination of artificial and natural impurities, and oils that cover the outer surface of the fibre’s cell wall and depolymerizes the architecture of the original fibre to reveal short-length crystallites [41, 39, 47]. The alkali concentration, fibre soaking temperature, and fibre soaking length are the parameters or factors used in mercerization. Okafor et al. (Komal et al. 2019) treated plantain fibres for 4 h with a 5% aqueous NaOH.

Osoka and Onukwuli (Adekunle 2015) treated fibres recovered from the empty fruit bunch of plantain for 30–150 min with 2 to 10 wt % aqueous NaOH. Orekoet al. (Okafor et al. 2012) treated plantain fibres by soaking them for 60 min in 5% aqueous NaOH. Chemically, fibre treatment requires the use of chemical compounds to make two naturally incompatible materials compatible and thus strengthen their mechanical characteristics (Oreko et al. 2018). These fibres are treated in different ways, including enzymatic and alkaline treatment; acetylation, and benzoylation (Xu et al. 2015; Debnath et al. 2019). The widely used binding agents include silane and maleic anhydride as in maleic anhydride-grafted polypropylene (MAPP), and the commonly used maleic anhydride coupling agents.

Maleic anhydride interacts with the OH groups on the fibre surface, replacing the groups of MAPP (Li et al. 2007). This leads to fibre coating with long hydrophobic polymer chains, enhancing hydrophobic matrix compatibility by forming carbon–carbon bonds that enhanced the fibre wettability and interfacial adhesion (Zin et al. 2018; Pannu et al. 2018). Isocyanates, triazine, peroxide, and permanganate (Li et al. 2007; Fiore et al. 2016) are other chemical reagents for natural fibres’ treatment.

Musa spp. composite fabrication

Manual technique

To build Musa spp. composites, the manufacturer lays the fibres in the resin according to the application's reinforcement needs. This has been accomplished by hand historically and it is referred to as the technique of hand lay-up. The most popular method of composite preparation is hand lay-up technique (Brígida et al. 2010), and it is still the most popular for banana composites because it is cheap and does not require much technical expertise. However, concerted efforts are now being made towards the application of automated methods in fibre placement to ensure quality and repeatability (Kulkarni et al. 2019).

Automated techniques

Automated processes for fibre placement fall into two categories: Tape laying and winding of filaments. The tape-laying method required the use of equipment that controlled fibre positioning over resins of narrow fibre tapes. The ‘sharpness’ of the turns needed to position the fibres in the specified direction is determined by the width of the tape. This means that larger tapes were used for incremental turns whereas narrow tapes were used for the sharp turns that are associated with more complex shapes. The smaller units of fibre are used for filament winding. The fibre yarns are woven in given directions in the shape of the part over a rotating mandrel in this system. Successive layers are added until the appropriate thickness is achieved.

While filament winding was initially restricted to simple and direct paths, with the use of robots, the method is now able to produce complex shapes. A second manipulation of the outcome of tape laying or filament winding must be achieved when thermosets are used as polymer resins. In an autoclave or oven, this is normally done by heating the completed structure (Kulkarni et al. 2019). The benefit of online consolidation is provided by thermoplastic systems to remove the high energy and capital costs associated with the curing phase. For manufacturing composites, extrusion is the most common continuous process (Kulkarni et al. 2019).

On the extrusion line at one end, fibres and the resin are pushed through a heated die or shaping tool and then cooled and pulled out at the end. This process can be applied to both thermoplastic and thermoset polymers. RTM (Resin transfer moulding) is a type of composite processing that offers a high flexibility potential, but is limited to low viscosity thermosetting polymers (easily flowing). In the process, a braided, woven, or knitted fibre is placed in a mould, which is further closed and injected with resin (Santhosh et al. 2014; Adeniyi et al. 2019). Following curing, the object is removed by opening the mould. The fibres are generated in a wide range of designs, and for better mechanical properties, many can be joined together in varying orientations. Together with extrusion and RTM, researchers have used injection moulding and compression moulding, which showed that both methods are better suited for thermoplastics than for thermosets (Brígida et al. 2010).

Characterization of Musa spp. fibre

X-Ray fluorescence of Musa spp. fibre

Using X-ray fluorescence (XRF) (e.g. Nitron 3000), raw plantain fibres are analysed for elemental and oxide compositions. After initialization, the system is turned on and allowed to stabilize for 5 min. The tool commonly used was the Cu–Zn system, which, due to its strength, usually detects a large number of elements and sesquioxides. While the ray point was placed over it, the sample was placed on the sample holder and the ray button was pressed to observe log data. The data are obtained in triplicates and the average is immediately taken (Alavudeen et al. 2015).

The elemental composition of raw plantain fibres is provided in Table 1 (Adeniyi et al. 2020). The data revealed that the five highest inorganic constituents of plantain fibre in the descending order were indium, silicon, potassium, calcium, and aluminium. These elements and others present in smaller amounts are underpinned by the use of harvested plantain fibres in various engineering applications. For instance, indium and silicon are useful in the manufacture of electronic devices such as transistors, insulators, rectifiers, and chips, given that they are present in the quantities extracted. For the production of thin conductive transparent films for photovoltaic use, indium oxide can also be doped with tin oxide on its own. The comparable use of bio-extracted silicon and indium has been verified by previous works (Ju´ stiz-Smith et al. 2008; Vu et al. 2010). Plantain fibre can also be reprocessed due to the potassium and silicon content of the fibre for use in fertilizer production (Zhang et al. 2015). Potassium for industrial and health applications such as blood pressure control can also be derived from plantain fibre (Mohiuddin et al. 2014). In the manufacture of yarn, thread, high-quality textiles, and auto-body works, plantain fibres can also be used as a composite material (Becker et al. 2013; Okafor et al. 2012; Ramesh et al. 2014).

Table 1 Musa spp. fibre elemental compositions (Alavudeen et al. 2015; Adeniyi et al. 2020)

Fourier transform infrared spectroscopy (FTIR) of plantain fibres

The study of fibres functional group using FTIR is an important indicator of the possibility of the integration of fibres with polymer matrices. To qualitatively classify the constituents of the fibre, FTIR Spectroscopy was carried out on the treated and untreated plantain fibre samples. The strength of light absorbed or released by the fibre at a given wavelength corresponding to a given functional group in the fibre is obtained by FTIR spectroscopy. The FTIR spectra for samples of plantain fibres, NT, 2 T, 4 T, and 6 T are shown in Fig. 1. The characterization spectra of untreated and treated plantain fibres as derived from Fig. 1 are summarized in Table 2 (Alavudeen et al. 2015).

Fig. 1
figure 1

IR spectra of treated and non-treated plantain fibres (Alavudeen et al. 2015)

Table 2 IR spectral analysis of treated and untreated plantain fibre (Alavudeen et al. 2015)

For the raw fibre, the broadband at 3344 cm−1 corresponds to cellulose vibrational stretching of hydrogen bonded-OH (Xu et al. 2015). This stretching is correlated with the involvement of OH in cellulose, lignin, hemicellulose, and a carboxylic acid, according to Monteiro et al. (Monteiro et al. 2014). The absorption observed at 2917.15 and 2850.66 cm−1 corresponds to the alkyl cellulose group's asymmetric and symmetric stretching vibrations, which are typical of any natural fibre. The absorption band value of 1739.13 cm−1 and about 1650.66 cm−1 can be resulting from the stretching of hemicellulose and probably fatty acids from lignin of the carbonyl (C = 0) ester and carboxyl groups (Ortega et al. 2016; Bakri and Jayamani 2016).

The absorption at 1425.86 cm−1 corresponds to aromatic ring vibrations, while the small peaks between 1372.03 and 1242.22 cm−1 describe the C–H bending bond structure of the cellulose, hemicellulose, and lignin functional group, while the peak at 1039.58 cm−1 corresponds to cellulose, hemicellulose, and minor lignin material C–O-symmetric stretching vibration (Ortega et al. 2016; Sefadi and Luyt 2012). The β-glucoside linkages between the sugar units in cellulose and hemicellulose are due to the small absorption band at 878.10 cm−1 (Xu et al. 2015).

After treatment of the fibres, the strength of the peak about 3344 cm−1 earlier noted for the raw fibre increased to around 3354 cm−1, this increase may be due to the OH group's breakdown of lignin and part hydrogen bond (Reddy et al. 2013). This hydrogen-bonded (−OH) vibration characteristic peak indicates the presence of intermolecular hydrogen bonding that appears to change in improved treated fibres to higher absorbance value (Reddy et al. 2013). From the data presented, it can be observed that only 2% of the treated fibres had higher absorbance compared to NT fibres, which means that 2 T fibres are stronger than untreated fibres, whereas 4 T and 6 T fibres have decreased intensity.

If the amount of the alkali increases, the OH band becomes wider and this implies the presence of a high cellulose concentration. Also, the band 1739.31 cm−1 which is absent in the treated fibres due to the CH3COO− and COOH functional group of hemicellulose found in the raw fibre spectra. The absence of this characteristic vibrational stretching indicates that the hemicellulose material has been substantially eliminated by alkaline therapy (Sefadi and Luyt 2012). Fibres are further demonstrated in literature in the functional group study for further fibre treatment as shown in FTIR spectra [41, 5, 62, 64]. The results of the FTIR study showed that NaOH treatment resulted in the gradual elimination of hemicellulose and lignin, which act as fibre binding components. As demonstrated by the widening of the OH band, the therapy was also successful in exposing lignin. Free hydroxyl groups that can react with polymer groups in composite production are shown by large concentrations of cellulose on fibre surfaces (Alavudeen et al. 2015).

Chemical composition of Musa spp. fibre

As stated by Adeniyiet al. (Kulkarni et al. 2019) and Geethama et al. (Ihueze et al. 2015), Table 3 presents the characteristic chemical composition of the banana and plantain.

Table 3 Chemical composition of extracted Musa spp. fibres (Kulkarni et al. 2019; Ihueze et al. 2015)

Physicomechanical properties of natural Musa spp. fibre

Fibre samples subjected to standardized characterization tests such as ash and charcoal contents, water absorption, moisture content, tensile strength, elemental analysis, and chemical analysis, and the presence of metal elements as tabulated (Tables 1, 2, 3, 4). Also, ions in natural fibres and the different chemical properties of cellulosic fibres had been similarly reported (Rai and Jai Singh 1986; Rao and K. and MohanaRao et al. 2007; Alavudeen et al. 2015; Adeniyi et al. 2020).

Table 4 Physicomechanical properties of isolated Musa spp. fibres (Rao and K. and MohanaRao et al. 2007; Alavudeen et al. 2015)

Also, the bananas cross-sectional area was examined by Murali and Mohan (Reddy and Yang 2003) using optical laser beam equipment, demonstrating that the banana fibre’s cross-sectional area taken for the study is 0.3596mm2. At various degrees of orientation of the fibre cross section, the cross section measurements of the fibres are determined. This investigation projected that the cross sections of the fibre were roughly circular. The magnitude of laser beam diffraction was also found to be dependent on the curvature of the cross section and the variety of the fibre. The test, therefore, gives relative fibre shapes rather than the real area of the cross section.

The optical microscopy study by Kulkarni et al. (Geethama et al. 1998) has revealed the banana fibre cell structure. It has four cell types, namely phloem, sclerenchyma, xylem, and parenchyma, which made up the banana fibre. Besides, in this analysis, they discovered the variations of mechanical properties with the fibre diameter (Rao and K. and MohanaRao et al. 2007). Using banana fibre combined with epoxy resin and hardener, Al-Qureshi (Kulkarni et al. 1982) designed and tested a truck ‘Manaca’ model. However, some rare and necessary fibreglass/banana chopped fibre/epoxy hybrid composite panels were laminated. It has gone through several years of research for road quality and has shown excellent results in conclusion. The percentage moisture content and density of individual fibres are shown in Table 2 (Alavudeen et al. 2015).

Geethama et al. (1998) measured the tenacity and elastic modulus of banana fibre in the range of 529–759 MPa and 8–20 GPa, respectively, with a percentage elongation of banana fibre breakage varying from 1.0 to 3.5 (Alavudeen et al. 2015).

Musa spp. composites characterization

According to the ASTMD 3039 (Standard Test Method), universal testing machine (UTM) was used for the analysis of composites’ Tensile Properties (Al-Qureshi 1999). The samples selected were tested to study the morphological and functional responses of the composite, prepared entirely at room temperature during cold processing, with the highest levels of factor combinations. To qualitatively classify the constituents of the fibre using a Thermo-Scientific Nicolet iS5-iD1 unit, Fourier transform infrared (FTIR) spectroscopy is carried out on the treated and untreated composite samples (with the highest degree of factor combination). Using a scanning electron microscope (SEMPhenomProX) with a microscope acceleration voltage set to 15 kV, the surface properties of composites are also carried out (Alavudeen et al. 2015).

Tensile characteristics of Musa spp. composites

Figure 2 shows the tensile strength, as applied to the identification of the samples, where the treated fibres are represented with samples 1 to 30. In brief, composites produced with treated and untreated plantain fibres were considered and subjected to fibre-length effects and fibre concentration at different levels of their combinations. It was verified, on a general basis, that composites in samples 1 to 9 have a tensile strength higher than composites in samples 10 to 30. This may be explained by the adhesive effect of the handled polystyrene matrix as previously described by Abdulkareem et al. (ASTM et al. 2017; Abdulkareem and Adeniyi 2017). Figure 2 displays the graphical presentation of the tensile strength obtained from the study of composite samples. The graph showed a high tensile strength yielded by samples 6, 9, and 5 in descending order, with a maximum tensile strength of 401 MPa for sample 5 (Alavudeen et al. 2015).

Fig. 2
figure 2

Tensile strengths response of plantain fibre polystyrene composites (Alavudeen et al. 2015)

This research indicated that untreated plantain fibre had the highest tensile strength when combined with polystyrene resin. As shown in the result tensile strength improved as fibre length and fibre concentration increased. This is distinct from similar works with different polymer matrices (Arun et al. 2016; Abdulkareem and Adeniyi 2017); the treated fibres were believed to have the highest fibrous composite tensile strength. However, in this research, interfacial adhesion was not an obstacle, previous works in the literature reported that polystyrene contributes to the adhesive matrix Abdulkareem et al., (ASTM et al. 2017; Abdulkareem and Adeniyi 2017), which makes untreated fibre stronger throughout curing. The whole composite processing was also performed at room temperature in cold pressing and there was no polymer melting or hot pressing involved, it was a non-thermal mixing process.

FTIR Characteristics of Musa spp. composites

The FTIR spectra for composites of treated and untreated plantain fibres are shown in Fig. 3. The composites with treated and untreated samples were picked from the high-level samples because the fibre length and the fibre concentration did not influence the composition of the final composite (Kulkarni et al. 2019). The possible functional groups found are O–H (hydroxyl), C=C (alkenes), CH2, C–O, CO–OH (carboxyl), and aromatic carbon groups (Abdulkareem et al. 2020; Herman et al. 2015). A contrast of the fibre composite spectra of not treated and treated showed that the amplitude of the PFRPC peaks of not treated was greater than that of the treated composites. This indicates that the absorbance ratios of the processed fibre composites are greater than those of the untreated fibres. The mean light intensities are both 88.18 and 88.46 for untreated and treated PFRPC composites, which revealed the influence of fibre alteration on the absorption of light. This means that, while composites can have comparable properties, treated fibre composites may be more brittle than untreated fibre composites (Reddy et al. 2013). This indicates that the treated composites have higher fibre–matrix adhesion relative to untreated fibre composites, as already seen in the SEM findings (Alavudeen et al. 2015).

Fig. 3
figure 3

FTIR spectra of treated and untreated composites (Alavudeen et al. 2015)

SEM characteristics of Musa spp. composites

The internal cracks and internal structure composite samples were observed using SEM micrographs (Mohiuddin et al. 2014). Among the high-level samples of treated and untreated specimens, the composites used were chosen. This was to show the effects of fibre treatment on the polystyrene matrix. The micrograph showed clearly the effect of reinforcement with plantain fibres, the dispersion of fibres, and the effect of the treated fibres (Fig. 4).

Fig. 4
figure 4

SEM micrographs of a untreated and b treated composites (Alavudeen et al. 2015)

Concerning the untreated PFRPC, the surface of the treated PFRPC had a heterogeneous surface with a relatively homogeneous surface. As alkaline treatment of fibres increased their surface roughness was exposed. A rougher surface relative to that of the untreated composite was observed. Li et al. (2007) also found out that proper bonding of microfibrils in untreated PFRPC was found in contrast with that of treated PFRPC whose microfibrils were dispersed and not properly bonded. This phenomenon was attributed to the removal by alkaline treatment of hemicellulose and pectin as binding agents between fibre-forming microfibrils (Li et al. 2007; Ramesh et al. 2014). Figure 4 displays fibre pull-outs, as seen by the white circle, thus creating voids. Higher interfacial bonding was attributed to the presence of hackles (shown by the red circle). Low compatibility with polystyrene of untreated plantain fibre (Song et al. 2018) was observed in SEM micrographs of untreated PFRPC. The treated PFRPC micrograph reveals that the fibres on the composite surfaces were twisted, bent, and almost broken; hence, there were a negligible number of voids around the fibres and a few pull-outs of fibre. These findings indicate the high compatibility of fibre and resin (Bakri and Jayamani 2016; Bennet et al. 2014). This means that processed plantain fibres have better adhesion to the polystyrene matrix than untreated fibres. Since composites need to have high tensile strength and good interfacial adhesion to be used successfully in structural applications (Singh and Palsule 2014), more research is required to obtain solid and robust PFRPC for optimum treatment concentration and structure of the fibre–matrix.

Physicomechanical properties of Musa spp. fibre-reinforced composites

Several studies have been performed to predict the various mechanical properties of polymer-reinforced banana fibre composites, such as tensile strength and flexural strength. The outstanding mechanical properties exhibited by the banana fibres are shown in Table 5. Tensile checks were conducted under ASTM-D 3379-75 (Reddy and Yang 2003). The physicochemical properties of natural Musa spp. fibre are shown in Table 4.

Table 5 Summary of the physicomechanical properties of Musa spp. fibre composites (Kulkarni et al. 2019)

Banana–polyester composites

Several studies have been performed with the base matrix of commercially sourced unsaturated polyester resin (HSR 8131), ECMALON 4411, C-451, and P-9509 (Adeniyi et al. 2019; Geethama et al. 1995; Idicula et al. 2010; Shesan et al. 2019; Cadena et al. 2017; Rao et al. 2010). Alavudeen et al. (2015) investigated the relationship of other polyester composites that made use of other forms of natural fibres, banana fibre–polyester composites were used. It was observed in the study (Fig. 3a, columns B and F) that banana fibres had greater tensile strength compared to sisal fibres (Geethama et al. 1995). Rao et al. (Cadena et al. 2017) (Fig. 3a, columns D and G) also confirm the result. In the same research, however, Rao et al. (Cadena et al. 2017) found that both bamboo and vakka fibres used in reinforced composites had higher tensile strength compared to banana fibres (especially bamboo fibres).

Compared to others, studies by Samivel and Babu (Rao et al. 2010) at a slightly lower fibre loading showed that banana fibre-reinforced composites also have higher tensile strength than those of kenaf. We may also infer that the natural fibre-reinforced composites have tensile strength which decreases significantly at low fibre loading. It can be summarized that banana fibre-reinforced polystyrene composites have 55–75 MPa as their tensile strength for fibre loading within the range of 30 to 40%. The perceived tensile strength decreasing order in comparison with other fibres is bamboo, vakka > banana > sisal, kenaf. Bananas can be an option of choice for material developers interested in obtaining maximum tensile strength from their composites based on the position and availability of various natural fibre sources.

Figure 5b shows the flexural strength comparison of other fibre materials to banana fibre. It was revealed in Idicula et al. (Geethama et al. 1995) study (columns C and I) that composites of banana fibre have lower flexural strength compared to sisal. This was also reiterated (columns G and J) by Rao et al. (Cadena et al. 2017). Further analysis by Rao et al. (Cadena et al. 2017) also found that banana fibre-reinforced composites' flexural strength is lower than that of both bamboo and vakka composites. However, Samivel and Babu (Rao et al. 2010) explained that, compared to kenaf fibres, banana fibre-reinforced composites have greater flexural strength. It is difficult to draw some inferences considering the banana fibre’s loading with flexural power, as there tend to be large variations between studies. The perceived decreasing order of flexural strength is bamboo, vakka, sisal > banana > kenaf in contrast to other fibres. Based on this, it follows that banana fibre-reinforced composites in the flexural strength domain do not perform favourably with those of other natural fibre-reinforced composites.

Fig. 5
figure 5

Tensile (a) and flexural (b) strength of natural fibre-reinforced polyester composites (Kulkarni et al. 2019)

In comparison with those of other materials, Fig. 6 shows the impact power of banana fibre-reinforced composites. Samvel and Babu (Rao et al. 2010) was the only published comparative research on the impact strength of banana fibre-reinforced polystyrene composites. Thus, banana has a much stronger impact of power than kenaf fibres.

Fig. 6
figure 6

Impact strength of natural fibre-reinforced polyester (Kulkarni et al. 2019)

Banana–epoxy composites

Adeniyi et al. (Alavudeen et al. 2015) compared the role of Musa spp. fibre in epoxy composites with other types of fibres from natural origin. The tensile properties of these epoxy composites were studied. Both comparative and non-comparative studies had also been undertaken by several authors (Arun et al. 2016; Sakthivei and Ramesh 2013; Yusuf et al. 2019; Samivel and Babu 2013; Mangalaraja et al. 2019). The study carried out by Arun et al. (2016), Sakthivei and Ramesh (2013) used commercially sourced epoxy resin LY556 and DGEBP-A. It was discovered that composites of banana fibre-reinforced epoxy have lower tensile strength relative to coir (Fig. 7, columns A and F). Banana fibre-reinforced composites had greater tensile strength (Mangalaraja et al. 2019) than hemp fibres (columns B and G). Banana fibre-reinforced composites also have improved tensile strength (Sakthivei and Ramesh 2013) than as in columns C and H. However, for sisal and banana fibres, as in columns E and I, the tensile strength of sisal fibres is greater (Alavudeen et al. 2015).

Fig. 7
figure 7

Tensile strength of natural fibre-reinforced epoxy composites (Kulkarni et al. 2019)

The perceived decreasing order of tensile strength is coir, sisal > banana > hemp, jute, as opposed to other fibres. When compared to other fibres that have been studied comparatively, banana fibres displayed an intermediate tensile strength. Therefore, the preference of banana fibres over other natural fibres will be subjectively dependent on the availability of these fibres at the site when considering tensile strength as the property of interest. Compared with coir, flexural strength for banana fibre-reinforced epoxy composites is greater in Fig. 8a (columns A and E) (Arun et al. 2016). The flexural intensity in columns B and F is only slightly higher compared to that of jute (Sakthivei and Ramesh 2013). Compared to banana fibres, sisal fibre-reinforced epoxy composites possess greater flexural strength in columns D and G (Bhoopathi et al. 2017). The perceived flexural strength decreasing order in comparison with other fibres is sisal > banana > coir, jute. It can be inferred that when compared with those of other natural fibres, banana fibre-reinforced composites have an intermediate flexural strength (Kulkarni et al. 2019).

Fig. 8 a
figure 8

, b Flexural and Impact strength of natural fibre-reinforced epoxy composites (Kulkarni et al. 2019)

Banana fibre-reinforced epoxy composites have higher impact strength relative to jute fibres in Fig. 8b (columns A and D) (Sakthivei and Ramesh 2013). Sisal fibres, however, demonstrate higher impact strength relative to banana fibres, as seen in columns C and E (Bhoopathi et al. 2017). The perceived declining order of tensile strength for other fibres is sisal > banana > jute. As for the other mechanical properties of banana fibre-reinforced epoxy composites; it can be concluded that, compared to other recorded natural fibres, flexural strength for the banana fibre-reinforced composites were intermediate (Kulkarni et al. 2019).

Banana fibre–cement composites

Several scholars documented the inclusion of banana fibre in cement as reinforcement (Mukhopadhyay et al. 2008; Venkateshwaran et al. 2011; Bledzki and Gassan 1999; Savastano et al. 2005). The air-cured banana fibre-reinforced cement composites were investigated by Zhu et al. (Savastano et al. 2005) and found that flexural strength is about 25 MPa at a fibre load of 14% by mass and the fracture toughness value was 1.74 kJm2. They indicated that these mechanical properties were ideal for the use of building materials with water absorption of less than 25%. The Brazilian fibres (sisal, banana waste, and Eucalyptus Grandis residue from the pulp mill) were examined by Savastano et al. (Bledzki and Gassan 1999) as reinforcements for cement-based composites. The study observed that the inclusion of 12% by weight of Brazilian waste fibre pulps into cement produced composites in the range of 1.0–1.5 kJ/m2 with a flexural strength of about 20 MPa and fracture toughness. Only for low-cost housing construction can these values of composites be used. Besides, banana kraft pulp with a high aspect ratio in ordinary Portland cement offers greater fibre–matrix bonding with increased fibre fracturing and reduced fibre pull-out, and lower toughness (Rao and K. and MohanaRao et al. 2007).

Banana fibre–thermoset plastic composites

The banana fibres use in thermoset–plastics composites manufacture had been earlier investigated (Poletto and Zattera 2015; Geethama et al. 1995; Jannah et al. 2009; Zhu et al. 1994; Mishra et al. 2000; Pothan et al. 2003; Pothan et al. 2003). It was discovered that the use of banana fibre in the scanning electron microscope analysis of cotton fabric that the composites unique module appears to be in similar order compared to glass-fibre-reinforced plastic (UdayKiran et al. 2007). The swelling actions and mechanical properties of fibre/novolac composites were researched by Mishra et al. (Zhu et al. 1994), using a compatible agent (maleic anhydride). It was discovered during the study that the maleic anhydride, when applied, significantly led to a reduction in the steam and water absorption rates associated with an improvement in the module, stiffness, and impact strength of Young, thus mechanical properties’ improvement (Rao and K. and MohanaRao et al. 2007). The research also established the treatment of composites with maleic anhydride, ensuring a higher consistency of composite materials reinforced with plant fibre. Joseph et al. (Jannah et al. 2009) investigated the relationship between glass fibres and banana fibres, and phenol–formaldehyde-reinforced composites’ mechanical properties. They determined that glass fibres and banana fibres, at optimum fibre length have maximum tensile stress of 40 and 30 mm, respectively. They discovered also that banana/phenol–formaldehyde composites increased Young's modulus and tensile strength values up to 320 and 400%, respectively, with fibre loading application of 48% when a neat resin was used. The composite’s Flexural modulus was increased to 25% for banana/phenol–formaldehyde composites by increasing the fibre loading up to 45% (Bhoopathi et al. 2017). The polymeric materials’ structure and viscoelastic behaviour for different applications to assess their relevant stiffness and damping characteristics have been investigated using dynamic mechanical test methods over a calculated range of temperatures and frequencies (Bhoopathi et al. 2017). Pothan et al. (Mishra et al. 2000; Pothan et al. 2003) discovered that better fibre/matrix bonding occurred at 40% fibre loading in their study on dynamic mechanical analysis of banana fibre strengthened with polyester composites, with debonding occurring at 10 and 20%, respectively, when analysed through SEM. Idicula et al. (Nystrom et al. 2007) found at random that using a short banana and polyester-reinforced sisal fibres, the composites provided the highest value in their work with a fraction of 0.40 by volume. Using composites with relative sisal and banana volume fraction, high tensile strength, flexural modulus, and lowest impact strength, improved fibre/matrix adhesion and stress transfer, composites with relative sisal volume, and a 3:1 fraction of banana volume are obtained.

When the matrix of the polyester was mixed with banana/sisal fibre it resulted in a decline in the thermal conductivity efficiency of the composite. This demonstrated that polymeric matrix thermal conductivity was more significant. The treated fibre exhibited a major improvement in banana/sisal fibre density and thermal conductivity value, while the opposite effect was seen for composites of pineapple leaf fibre (PALF)/glass fibre. Pothan et al. (Mukhopadhyay et al. 2008) studied the fibre/matrix interaction feature with chemical surface adjustment on complex mechanical properties and found 80 °C and 130 °C failure modulus curves and damping curve, respectively. The peaks around 130 °C are correlated with the matrix ‘Tg’ glass transition temperature and that is due to the interlayer transition around 80 °C.

Uday et al. (Pothan et al. 2006) found in their work that polyester-reinforced banana fibre has a tensile strength of 59 MPa when the fibre length was 30 mm and 51% fibre weight. Haneefa et al. (Poletto and Zattera 2015) in their study on the hybridization of banana fibre and glass fibre identified improved Young's modulus \and tensile strength with an increased glass fibre's volume fraction due to greater compatibility polystyrene matrix with the glass fibre.

Banana fibre–thermoplastic composites

Banana fibre applied as reinforcement in thermoplastic composites has been documented in several studies (Idicula et al. 2006; Sapuan et al. 2006; Maleque et al. 2006; Habibi et al. 2008; Paul et al. 2008). Sapuan et al. (Idicula et al. 2006) and Maleque et al. (Sapuan et al. 2006) studied the mechanical properties of woven banana fibres reinforced with epoxy composites. The x direction maximal stress value was estimated to be 14.14MN/m2 and 3.398MN/m2 in the y direction. Analysis of the traction, flexural effects, and fracture surface of woven pseudo stem banana fibre reinforced with epoxy composite by hand lay-up technique was also studied. Tensile test results reveal that, relative to unreinforced epoxy, the ultimate intensity of banana-reinforced epoxy was improved from 25 to 47 MPa. The Young's module has also been enhanced from 1300 to 1850 MPa. Similarly, flexure power increased from 45 MPa to 75, and Young's flexural module increased from 1525 to 1825 MPa. In their work using agro-industrial residues like cotton stalk, rice straw, bagasse, banana plant waste, etc., the fracture surface analysis by scanning electron microscope showed that the banana fibre composite was characteristic of a ductile type of failure with limited plastic deformation. Habibi et al. (Maleque et al. 2006) reported that except for low-density polyethylene (LDPE), maleated low-density polyethylene (MLDPE) in the 0.75:0.25 blend-based composite ratios, Young's modulus exhibited a gradual increase in value with the inclusion of MLDPE in the blend, while elongation at break remained unevenly constant. Paul et al. (Merlini et al. 2011) discovered in their work on the effects of fibre loading on banana/polypropylene (PP) composite, that thermal diffusivity as well as thermal conductivity, both, respectively, decreased with increased fibre loading (from 0.24 W/mK for tidy polypropylene matrix to 0.217 and 0.157 W/mK for 0.10 and 0.50 of volume fractions, respectively). Chemical treatments (alkali, silane, permanganate, and benzoyl chloride) (Idicula et al. 2010) have also been found to cause increased thermal diffusivity in addition to conductivity (Bhoopathi et al. 2017; Paul et al. 2008).

Banana fibre-biodegradable composites

The preparation or surface modification of strengthening fillers like natural fibres to make them more compatible with protein matrices is the other well-studied way to increase the interface and compatibility of protein composites. For example, Kumar et al. (Zainuddin et al. 2008) treated banana fibres with alkali solution and the treated fibres resulted in improved tensile strength and composite modulus of protein. The tensile strength and modulus of the protein composite made of treated fibres showed an 82% and 963% increase in tensile strength and modulus, respectively, relative to soy protein film without fibres. The biodegradability test demonstrated that composites are 100% biodegradable (Zainuddin et al. 2008).

Conclusions

Considering the desire for materials that are cost-effective, eco-friendly, lightweighted, and durable in the manufacturing industry, Musa spp. fibre used as reinforcement material in composites manufacture comes in handy. In composite manufacture, banana fibre applied as a reinforcing component positively modifies the physical, mechanical, and structural characteristics of the composites. Composites dependent on these fibres have very strong potential in the manufacture of various sections’ components in the manufacturing, automobile, machinery industries, as a result of low density, improved tensile strength, high tensile modulus, and low elongation split. Most countries like Nigeria, India, etc., harvest Musa spp. in very large quantities. The application of its fibre perceived as waste by the majority of their population in composites manufacture for producing useful industrial components would be very attractive for their economies. Musa spp. fibre and its composites could further be made attractive if a suitable cost-effective design method for the separation of its fibre and composite production is developed and engineered. This would enhance its research and application to a greater extent. This work seeks to provide a collective and articulate source of information for systematic and persistent research in Musa spp. fibre composites progress and application.

Availability of data and material

Not applicable.

Abbreviations

SPP:

Species

NFC:

Natural fibre composites

ASTM:

American Standard Testing

NaOH:

Sodium hydroxide

OH:

Hydroxide

MAPP:

Maleic anhydride-grafted polypropylene

RTM:

Resin transfer moulding

XRF:

X-ray fluorescence

Cu–Zn:

Copper–zinc

FTIR:

Fourier transform infrared spectroscopy

T:

Transmittance

IR:

Infrared

MP:

Megapascal

GP:

Gigapascal

kg/m3 :

Kilogram per metre cubed

UTM:

Universal testing machine

SEM:

Scanning electron microscope

PFRPC:

Plantain fibre polystyrene composites

wt%:

Weight percentage

KJ/m2 :

Kilojoule/metre square

MLDPE:

Maleated low-density polyethylene

LDPE:

Low-density polyethylene

References

  • Abdulkareem SA, Adeniyi AG (2017) Tensile and water-absorbing properties of natural fibre reinforced plastic composites from waste polystyrene and rice husk. Journal of Technological Development Nigerian 14(1):18–22

    Article  Google Scholar 

  • Abdulkareem SA, Adeniyi A, Amosa M, Raji S (2020) Development of plastic composite using waste sawdust, rice husk, and bamboo in the polystyrene-based resin (PBR) matrix at ambient conditions: valorization of biomass to value-added commodities. Springer, Berlin, pp 423–438

    Google Scholar 

  • Abdullah N, Sulaiman F, Miskam MA, Taib RM (2014) Characterization of banana (Musa spp.) pseudo-stem and fruit-bunch-stem as potential renewable energy resources. Int J Biol Vet Agric Food Eng 8:712

    Google Scholar 

  • Adekomaya O, Adama K (2017) Banana and plantain fibre-reinforced polymer composites. J Technol 36:782–787

    Google Scholar 

  • Adekunle KF (2015) Surface treatments of natural fibres—a review: part 1. Open J Polym Chem 5:41

    Article  CAS  Google Scholar 

  • Adeniyi AG, Ighalo JO, Onifade DV (2019) Banana and plantain fibre-reinforced polymer composites. J Polym Eng 39(7):597

    Article  CAS  Google Scholar 

  • Adeniyi, A. G., Onifade, D. V., Ighalo, J. O., Abdulkareem, S. A., and Amosa, M. K. (2020).Extraction and Characterization of Natural Fibres from Plantain (Musa paradisiaca) Stalk Wastes. Iranian (Iranica) Journal of Energy and Environment 11(2): 116–121, 2020

  • Alavudeen A, Rajini N, Karthikeyan S, Thiruchitrambalam M, Venkateshwaren N (2015) Mechanical properties of banana/kenaf fibre-reinforced hybrid polyester composites: Effect of woven fabric and random orientation. Material Design 1980–2015(66):246–257

    Article  CAS  Google Scholar 

  • Al-Qureshi HA (1999) The Use of Banana Fiber Reinforced Composites for the Development of a Truck Body. 2nd International Wood and Natural Fiber Composite Symposium, June. Kassel, Germany, pp 28–29

    Google Scholar 

  • Aminudin E, Din MFM, Mohamad Z, Noor ZZ, Iwao K (2011) International conference on environment and industrial innovation IPCBEE

  • Amir N, Abidin KAZ, Shiri FBM (2017) Effects of fibre configuration on mechanical properties of banana fibre/PP/MAPP natural fibre reinforced polymer composite. Proc Eng 184:573–580

    Article  CAS  Google Scholar 

  • Arpitha GR, Yogesha B (2017) An overview of mechanical property evaluation of natural fibre-reinforced polymers. Mater Today Proc 4:2755–2760

    Article  Google Scholar 

  • Arun A, Sathyaseelan R, Tamilselvan M, Gowtham M, Karthikeyan A (2016) Influence of weight fractions on mechanical, water absorption and corrosion resistance behaviors of untreated hybrid (coir/banana) fiber reinforced epoxy composites. Int J ChemTech Res 9:932–940

    CAS  Google Scholar 

  • ASTM, A.S.f.T.a M. (2017). ASTM D3039-standard test method for tensile properties of polymer matrix composite materials. In: Book ASTM D3039-standard test method for tensile properties of polymer matrix composite materials (ASTM International, 2017)

  • Bakri, M. K. B., and Jayamani, E. (2016). Comparative study of functional groups in natural fibres: Fourier transforms infrared analysis (FTIR). International Conference on Futuristic Trends in Engineering, Science, Humanities, and Technology (FTESHT-16), pp.167–174.

  • Becker H, Matos RF, Souza JA, Lima DA, Souza FTC, Longhinotti E (2013) Pseudo-stem banana fibers: characterization and chromium removal. Orbital Electron J Chem 5:164–170

    Google Scholar 

  • Bennet, C., Rajini, N., WinowlinJappes, J.T., Venkatesh, A., Harinarayanan, S., and Vinothkumar, G. (2014). “Effect of lamina fibre orientation on tensile and free vibration (by impulse hammer technique) properties of coconut sheath/sansevieria cylindrical hybrid composites” Advanced material research. Vol.984–985, pp. 172–177

  • Bhoopathi, R., Ramesh, M., Rajaprasanna, R., Sasikala, G. and Deepa, C. (2017).Physical properties of glass-hemp-banana hybrid fibre-reinforced polymer composites. Indian Journal of Science & Technology, 10: 1–7.

  • Bledzki AK, Gassan J (1999) Composites Reinforced with Cellulose Based Fibers. Prog Polym Sci 24(2):221–274

    Article  CAS  Google Scholar 

  • Brígida AIS, Calado VMA, Gonçalves LRB, Coelho MAZ (2010) Effect of chemical treatments on properties of green coconut fibre. Carbohyd Polym 79:832–838

    Article  CAS  Google Scholar 

  • Brindha R, Narayana CK, Vijayalakshmi V, Nachane RP (2017) Effect of different retting processes on yield and quality of banana pseudostem fibre. J Nat Fibers 16:1–10

    Google Scholar 

  • Cadena EM, Vélez RJM, Santa JF, Otálvaro GV (2017) Natural Fibers from Plantain Pseudostem (Musa Paradisiaca ) for Use in Fiber-Reinforced Composites. Journal of Natural Fibers 14:678–690

    Article  CAS  Google Scholar 

  • Cai M, Takagi H, Nakagaito AN, Katoh M, Ueki T, Waterhouse GI, Li Y (2015) Influence of alkali treatment on internal microstructure and tensile properties of abaca fibres. Ind Crops Prod 65:27–35

    Article  CAS  Google Scholar 

  • Debnath, K., Roy, S., Bose, T., Adhikari, S., Sinha, V., and Rabha, V. A. (2019). Processing of Green Composites, Springer: Switzerland, 2019: 65–80.

  • Fiore V, Scalici T, Nicoletti F, Vitale G, Prestipino M, Valenza A (2016) A new eco-friendly chemical treatment of natural fibres: effect of sodium bicarbonate on properties of sisal fibre and its epoxy composites. Compos B Eng 85:150–160

    Article  CAS  Google Scholar 

  • Geethama VG, Reethama J, Thomas S (1995) Short Coir Fiber Reinforced Natural Rubber Composites: Effect of Fiber Length, Orientation, and Alkali Treatment. J Appl Polym Sci 55:583–594

    Article  Google Scholar 

  • Geethama VG, Mathew KT, Lakshminarayana R, Thomas S (1998) Composite of Short Coir Fibers and Natural Rubber: Effect of Chemical Modification, Loading and Orientation of Fiber. J Polym Sci 65:1483–1491

    Google Scholar 

  • GVR (2018) Natural fiber composites (NFC) market size. Share Trends Anal Rep 2018:139

    Google Scholar 

  • Habibi Y, Ibrahim MM, Dufresne A (2008) Processing and Characterization of Reinforced Polyethylene Composite made with Lignocellulosic Fibers from Egyptian Agroindustrial Residues. Journal of Composites Science and Technology 20:1877–1885

    Article  CAS  Google Scholar 

  • Haneefa A, Bindu P, Arvind I, Thomas S (2008) Studies of tensile and flexural properties of short banana/glass hybrid fiber reinforced polystyrene composites. J Compos Mater 42:1471–1489

    Article  CAS  Google Scholar 

  • Herman V, Takacs H, Duclairoir F, Renault O, Tortai J, Viala, B.(2015). Core double-shell cobalt/graphene/polystyrene magnetic nanocomposites synthesized by in situ sonochemical polymerization. RSC Advancements, 5(63):51371–51381

  • Idicula M, Malhotra SK, Joseph K, Thomas S (2005) Dynamic mechanical analysis of randomly oriented intimately mixed short banana/sisal hybrid fibre reinforced polyester composites. Composite Science & Technology 65:1077–1087

    Article  CAS  Google Scholar 

  • Idicula M, Devi U, Thomas S (2006) Thermo Physical Properties of Natural Fiber Reinforced Polyester Composite. Journal of Composite Science and Technology 66:2719–2725

    Article  CAS  Google Scholar 

  • Idicula M, Joseph K, Thomas S (2010) Mechanical performance of short banana/sisal hybrid fibre reinforced polyester composites. Journal of Reinforced Plastic Composites 29:12–29

    Article  CAS  Google Scholar 

  • Ihueze CC, Okafor EC (2014) Response surface optimization of the impact strength of plantain fibre reinforced polyester for application in auto body works. J Innov Res Eng Sci 4(4):505–520

    Google Scholar 

  • Ihueze CC, Okafor CE, Okoye CI (2015) Natural fibre composite design and characterization for limit stress prediction in a multiaxial stress state. Journal of the King Saud University of Engineering Science 27(2):193–206

    Article  Google Scholar 

  • Imoisili PE, Fadare O, Popoola A, Okoronkwo A (2017) Effect of chemical treatment on the morphology and mechanical properties of plantain (Musa paradisiaca) fiber. IOSR J Appl Chem 10:70–73

    Article  CAS  Google Scholar 

  • Jannah M, Mariatti M, Abu B, A. and Abdul Khalil, H. P. (2009) Effect of Chemical Surface Modifications on the Properties of Woven Banana-Reinforced Unsaturated Polyester Composites. Journal of Reinforced Plastic Composites 28:1519–1532

    Article  ADS  CAS  Google Scholar 

  • Joseph S, Sreekala MS, Oommen Z, Koshy P, Thomas S (2002) A comparison of the mechanical properties of phenol formaldehyde composites reinforced with banana fibres and glass fibres. Composite Science & Technology 62:1857–1868

    Article  CAS  Google Scholar 

  • Joshi SV, Drza LT, Mohanty AK, Arora S (2004) Are natural fibers composites environmentally superior to glass fiber reinforced composites? Compos A 35:371–376

    Article  CAS  Google Scholar 

  • Ju´ stiz-Smith, N. G., Virgo, G. J. and Buchanan, V. E. (2008) Potential of Jamaican Banana, Coir, Bagasse Fiber as Composite Materials. Journal of Material Characterization 59:1273–1278

    Article  CAS  Google Scholar 

  • Kim H (2015) Hybrid composites with natural fibres. University of Birmingham, Birmingham, pp 97–108

    Google Scholar 

  • Komal UK, Sharma H, Singh I (2019) Processing of green composites. Springer, Switzerland, pp 15–30

    Book  Google Scholar 

  • Kulkarni AG, Satyanarayana KG, Rohatgi PK, Vijayan K (1982) Mechanical Properties of Banana Fibers. Journal of Material Science 18:2290–2296

    Article  ADS  Google Scholar 

  • Kulkarni, M., Radhakrishnan, S., Samarth, N. and Mahanwar, P. A. (2019).Structure, mechanical and thermal properties of polypropylene-based hybrid composites with banana fibre and fly ash. Materials Research Express, 6: 075318.

  • Kumar R, Choudhary V, Mishra S, Verma IK (2008) Banana fibre-reinforced biodegradable soy protein composites. Frontier Chemistry China 3:243–250. https://doi.org/10.1007/s11458-008-0069-1

    Article  Google Scholar 

  • Li X, Tabil LG, Panigrahi S (2007) Chemical treatments of natural fiber for use in natural fiber-reinforced composites: a review. J Polym Environ 15(1):25–33

    Article  CAS  Google Scholar 

  • Lightsey GR (1983) Polymer application of renewable resource materials characteristic of sisal fiber. Plenum Press, New York

    Google Scholar 

  • Lyn N (2018) Effect of the chemical treatment on the inorganic con-tent of Kenaf fibers and the performance of Kenaf-polypropylene composites. University of Waterloo, Canada

    Google Scholar 

  • Mahaboob S (2011) Banana fiber reinforced composite materials. In: Singh DH (ed). IIT Ropar: India

  • Maleque MA, Belal EY, Sapuan SA (2006) Mechanical properties of study of pseudo-stem banana fiber reinforced epoxy composite. Arab J Sci Eng 32:359–363

    Google Scholar 

  • Mangalaraja R, Raj TM, Prabha DR (2019) Statistical features of epoxy resin based hybrid composites reinforced with jute, banana, and flax natural fibers. J Chem Technol Metall 54:309–316

    Google Scholar 

  • Merlini C, Soldi V, Barra GM (2011) Fibre-matrix relationship for composite preparation. Polym Test 30:833–840

    Article  CAS  Google Scholar 

  • Mishra S, Naik JB, Patil YP (2000) The compatibilising effect of maleic anhydride on swelling and mechanical properties of plant fiber reinforced novolac composites. J Compos Sci Technol 60:1729–1735

    Article  CAS  Google Scholar 

  • Mohammed L, Ansari MNM, Pua G, Jawaid M, Islam MS (2015) A review on natural fiber reinforced polymer composite and its applications. Int J Polym Sci 1:15. https://doi.org/10.1155/2015/243947

    Article  Google Scholar 

  • Mohanty AK, Misra M, Drzal LT (2005) Natural fibers, biopolymers, and biocomposites. Taylor & Francis, Texas

    Book  Google Scholar 

  • Mohiuddin AKM, Saha MK, Hossian MS, Ferdoushi A (2014) The usefulness of banana (Musa paradisiaca) wastes in manufacturing of bio-products: a review. Agriculturists 12(1):148–158

    Article  Google Scholar 

  • Monteiro SN, Margem FM, Loiola RL, de Assis FS, Oliveira MP (2014) Flexural mechanical characterization of epoxy composites reinforced with continuous banana fibers. Mater Sci Forum 775:250–254

    Article  Google Scholar 

  • Mukhopadhyay S, Fanguerio R, Arpac Y, Sentrik U (2008) Banana fibers-variability and fracture behavior. J Eng Fibers Fabr 3:39–45

    CAS  Google Scholar 

  • Muller DH, Krobjilowski A (2003) New discovery in the properties of composite reinforced with natural fibers. J Ind Text 33:111–130

    Article  CAS  Google Scholar 

  • Naidu AL, Raghuveer D, Suman P (2013) Experimental study of the mechanical properties of banana fiber and groundnut shell ash reinforced epoxy hybrid composite. Int J Sci Eng Res 4:844–851

    Google Scholar 

  • Nystrom B, Joffe R, Langstrom R (2007) Microstructure and strength of injection molded natural fiber composites. J Reinf Plast Compos 26:579–599

    Article  CAS  Google Scholar 

  • Okafor C, Ihueze C, Ujam A (2012) Optimization of tensile strengths response of plantain fibres reinforced polyester composites (PFRP) applying Taguchi robust design. Innov Syst Des Eng 3:64–76

    Google Scholar 

  • Oreko BU, Otanocha OB, Emagbere E, Ihueze CC (2018) Analysis and application of natural fiber reinforced polyester composites to automobile fender. Covenant J Eng Technol (spec Edn) 1(1):1–12

    Google Scholar 

  • Ortega Z, Morón M, Monzón MD, Badalló P, Paz R (2016) Production of banana fiber yarns for technical textile reinforced composites. Materials 9(370):1–16

    Google Scholar 

  • Osoka E, Onukwuli O (2015) Optimum conditions for mercerization of plantain empty fruit bunch fiber. Int J Adv Mater Res 1:95–101

    CAS  Google Scholar 

  • Pannu AS, Singh S, Dhawan V (2018) Impact and fatigue properties of natural fibre composites. A review paper on biodegradable composites made from banana fibers. Asian J Eng Appl Technol 7:7–15

    Google Scholar 

  • Paul SA, Reussmann T, Mennig G, Lampke T, Pothen LA, Mathew GG, Joseph K, Thomas S (2007) Impact and fatigue properties of natural fibre composites. Compos Interfaces 14:849–867

    Article  CAS  Google Scholar 

  • Paul SA, Boudene A, Joseph K, Thomas S (2008) Effect of fiber loading and chemical treatments on thermo physical properties of banana fiber/polypropylene commingled composite materials. Compos Part A 39:1582–1588

    Article  CAS  Google Scholar 

  • Poletto M, Zattera AJ (2015) Mechanical and dynamic mechanical properties of polystyrene composites reinforced with cellulose fibres: coupling agent effect. J Thermoplast Compos Mater 30:1242–1254

    Article  CAS  Google Scholar 

  • Pothan LA, Oommen Z, Thomas S (2003) Dynamic mechanical analysis of banana a fiber reinforced with polyester composites. J Compos Sci Technol 63:283–293

    Article  CAS  Google Scholar 

  • Pothan LA, Thomas S, Habler R (2005) The static and dynamic mechanical properties of banana and glass fiber woven fabric reinforced polyester composites. J Compos Mater 39:1007–1025

    Article  CAS  Google Scholar 

  • Pothan LA, Thomas S, Groeninckx G (2006) The role of fibre/matrix interactions on the dynamic mechanical properties of chemically modified banana fiber/polyester composites. Compos Part A 36:1260–1269

    Article  CAS  Google Scholar 

  • Pothan LA, George CN, Jacob M, Thomas S (2007) Effect of chemical modification on the mechanical and electrical properties of banana fiber polyester composites. Compos Mater 41(9):2371–2386

    Article  CAS  Google Scholar 

  • Rai M, Jai Singh MP (1986) Advances in building and construction materials. Central Building Research Institute, Roorkee

    Google Scholar 

  • Ramesh M, Atreya TSA, Aswin US, Eashwar H, Deepa C (2014) Processing and mechanical property evaluation of banana fibre reinforced polymer composites. Proc Eng 97:563–572

    Article  CAS  Google Scholar 

  • Murali Mohan Rao K, Mohana Rao K (2007) Extraction and tensile properties of natural fibers: vakka, date and bamboo. J Compos Struct 77: 288–295

  • Rao KMM, Rao KM, Prasad AVR (2010) Fabrication and testing of natural fibre composites: vakka, sisal, bamboo, and banana. Material Design 31:508–513. https://doi.org/10.1016/j.matdes.2009.06.023

    Article  CAS  Google Scholar 

  • Reddy N, Yang Y (2003) Bio-fibres from agricultural by-products for industrial application. Trends Biotechnol 23:22–27

    Article  CAS  Google Scholar 

  • Reddy KO, Maheswari CU, Shukla M, Song JI, Rajulu AV (2013) Tensile and structural characterization of alkali-treated Borassus fruit fine fibres. Compos B Eng 44(1):433–438

    Article  CAS  Google Scholar 

  • Sakthivei M, Ramesh S (2013) Mechanical properties of natural fibre (banana, coir, sisal) polymer composites. Sci Park Int Recogni Res J 1:1–6

    Google Scholar 

  • Samivel P, Babu AR (2013) Int J Mech Eng Robot Res 2:348–360

    Google Scholar 

  • Santhosh J, Balanarasimman N, Chandrasekar R, Raja S (2014) Study of properties of banana fiber reinforced composites. Int J Res Eng Technol (IJRET) 3(11):144–150

    Article  Google Scholar 

  • Sapuan SM, Leeni A, Harimi M, Beng YK (2006) Mechanical properties of woven banana fiber reinforced epoxy composites. Journal of Materials and Design 27:689–693

    Article  CAS  Google Scholar 

  • Sathiyamurthy S, Thaheer A, Jayabal S (2013) Optimum drilling parameters of coir fiber-reinforced polyester composites. Indian J Eng Mater Sci 20:59–67

    CAS  Google Scholar 

  • Savastano H Jr, Warden PG, Coutts RSP (2005) Brazilian waste fiber as reinforcement for cement-based composites. J Cement Concr Compos 22:379–384

    Article  Google Scholar 

  • Sefadi JS, Luyt AS (2012) Morphology and properties of EVA/empty fruit bunch composites. J Thermoplast Compos Mater 25(7):895–914

    Article  Google Scholar 

  • Shesan OJ, Stephen AC, Chioma AG, Neerish R, Rotimi SE (2019) Improving the mechanical properties of natural fibre composites for structural and biomedical applications: renewable and sustainable composites (IntechOpen, 2019)

  • Singh AA, Palsule S (2014) Coconut fibre-reinforced chemically functionalized high-density polyethene (CNF/CF-HDPE) composites by Palsule process. J Compos Mater 48(29):3673–3684

    Article  CAS  Google Scholar 

  • Singha AS, Rana Raj K, Rana AK (2010) “Natural fibre-reinforced polystyrene matrix based composites” advanced materials research. Transf Technol Publ Switz 123–125:1175–1178

    Google Scholar 

  • Sisti L, Totaro G, Vannini M, Celli A (2018) Lignocellulosic composite materials. Springer, Switzerland, pp 97–135

    Book  Google Scholar 

  • Song XC, Lin QB, Chen CF, Chen JH, Hu CY (2018) Discrimination between virgin and recycled polystyrene containers by Fourier transform infrared spectroscopy and principal component analysis. Packag Technol Sci 31(8):567–572

    Article  CAS  Google Scholar 

  • Srinivasan V, Rajendra B et al (2014) Effect of fiber orientation and stacking sequence on mechanical and thermal characteristics of Banana–Kenaf hybrid epoxy composite. Mater Des 60:620–627

    Article  CAS  Google Scholar 

  • Swift TK, Moore M, Sanchez E (2015) Plastics and polymer composites in light vehicles. Economics and Statistics Department/American Chemistry Council

  • UdayKiran C, Ramachandra Reddy G, Dabade BM, Rajesham S (2007) Tensile properties of sun hemp, banana and sisal fiber reinforced with polyester composites. J Reinf Plast Compos 26:1043–1050

    Article  CAS  Google Scholar 

  • Umair S (2006) Environmental effect of fiber composite materials-study of life cycle assessment of materials used for ship structure. MS thesis dissertation, Royal Institute of Technology, Stockholm

  • Venkateshwaran N, Elayaperumal A (2010) Banana fiber reinforced polymer composites: a review. J Reinf Plast Compos 29:15

    Article  CAS  Google Scholar 

  • Venkateshwaran N, ElayaPerumal A, Alavudeen A, Thiruchitrambalam M (2011) Mechanical and water absorption behaviour of banana/sisal reinforced hybrid composites. Material Design 32(7):4017–4021

    Article  CAS  Google Scholar 

  • Vu XT, Stockmann R, Wolfrum B, Offenhäusser A, Ingebrandt S (2010) Fabrication and application of a microfluidic-embedded silicon nanowire biosensor chip. Phys Status Solidi (a) 207(4):850–857

    Article  ADS  CAS  Google Scholar 

  • Xu S, Xiong C, Tan W, Zhang Y (2015) Microstructural, thermal, and tensile characterization of banana pseudo-stem fibres obtained with mechanical, chemical, and enzyme extraction. BioResources 10(2):3724–3735

    CAS  Google Scholar 

  • Yusuf NAAN, Abdul Razabb MK, Bakara MBA, Yena KJ, Tunga ChW, Ghanic RSM, Nordin MN (2019) Determination of structural, physical, and thermal properties of biocomposite thin film from waste banana peel. Jurnal Teknologi (sci Eng) 81(1):91–100

    Google Scholar 

  • Zainuddin ES, Sapuan SM, Abdan K, Mohamad MTM (2008) Thermal degradation of banana pseudo-stem filled unplasticized polyvinyl chloride composites. J Mater Des 30(3):557–562

    Article  CAS  Google Scholar 

  • Zhang J, Fleury E, Brook MA (2015) Foamed lignin–silicone bio-composites by extrusion and then compression moulding. Green Chem 17(9):4647–4656

    Article  CAS  Google Scholar 

  • Zhu WH, Tobias BC, Coutts RSP, Langfors G (1994) Air-cured Banana-fiber-reinforced cement composites. J Cement Concrete Compos 16:3–8

    Article  CAS  Google Scholar 

  • Zin M, Abdan K, Norizan MN, Mazlan N (2018) Pertan J Sci Technol 26:161–176

    Google Scholar 

Download references

Acknowledgements

This work received administrative support from the Faculty of Engineering, Nnamdi Azikiwe University, Awka, Nigeria, and the Department of Science Laboratory Technology, OSISATECH Polytechnic Enugu, Nigeria.

Funding

The authors received no funding for this study.

Author information

Authors and Affiliations

Authors

Contributions

OR sourced and supplied the pieces of the literature used for the review; EE conceived and initiated the review process. He also sourced some of the literature. MM authenticated the information retrieved from the existing pieces of the literature supplied, and ODO proofread the article and guided text formatting, APC undertook the compilation and typesetting of the review article. With the submission of this manuscript, I undertake that it has not been published, accepted for publication, or under editorial review elsewhere and that all the authors have read and approved the manuscript for submission.

Corresponding author

Correspondence to Ernest Mbamalu Ezeh.

Ethics declarations

Ethics approval and concept to participate

Not applicable.

Consent for publications

Not applicable.

Competing interest

The authors declare no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Odera, R.S., Okechukwu, O.D., Ezeh, E.M. et al. The exchange of Musa spp. fibre in composite fabrication: a systematic review. Bull Natl Res Cent 45, 145 (2021). https://doi.org/10.1186/s42269-021-00604-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s42269-021-00604-z

Keywords