J Mater Sci (2020) 55:829–892 REVIEW Review A review of natural fiber composites: properties, modification and processing techniques, characterization, applications Aliakbar Gholampour1 1 2 and Togay Ozbakkaloglu2,* Department of Infrastructure Engineering, The University of Melbourne, Melbourne, VIC, Australia Ingram School of Engineering, Texas State University, San Marcos, TX, USA Received: 17 May 2019 ABSTRACT Accepted: 4 September 2019 There has been much effort to provide eco-friendly and biodegradable materials for the next generation of composite products owing to global environmental concerns and increased awareness of renewable green resources. An increase in the use of natural materials in composites has led to a reduction in greenhouse gas emissions and carbon footprint of composites. In addition to the benefits obtained from green materials, there are some challenges in working with them, such as poor compatibility between the reinforcing natural fiber and matrix and the relatively high moisture absorption of natural fibers. Green composites can be a suitable alternative for petroleum-based materials. However, before this can be accomplished, there are a number of issues that need to be addressed, including poor interfacial adhesion between the matrix and natural fibers, moisture absorption, poor fire resistance, low impact strength, and low durability. Several researchers have studied the properties of natural fiber composites. These investigations have resulted in the development of several procedures for modifying natural fibers and resins. To address the increasing demand to use eco-friendly materials in different applications, an up-do-date review on natural fiber and resin types and sources, modification and processing techniques, physical and mechanical behaviors, applications, life-cycle assessment, and other properties of green composites is required to provide a better understanding of the behavior of green composites. This paper presents such a review based on 322 studies published since 1978. Published online: 16 September 2019 Ó Springer Science+Business Media, LLC, part of Springer Nature 2019 Introduction The production of petroleum-based materials and glass fibers results in the release of a significant amount of greenhouse gases into the atmosphere [1]. Address correspondence to E-mail: togay.oz@txstate.edu https://doi.org/10.1007/s10853-019-03990-y The use of environmentally friendly green materials, which are recyclable, biodegradable, and renewable, has been recently considered to decrease the environmental impact of petroleum-based materials [2]. Fully green composites are a type of biocomposite 830 produced by a combination of biofibers and resins from renewable agricultural and forestry feedstock [1]. This type of composite is disposed of and composted at the end of their life without harming the environment [3]. Biofibers have been extensively used as reinforcers and fillers in composites in recent years [1]. They come from biological origins and are classified into non-wood fibers (natural fibers) and wood fibers [4]. Natural fibers have superior physical and mechanical properties than wood fibers, and they are high in cellulose content and crystallinity, and lighter in weight than wood fibers [1]. Therefore, this type of fiber could attract more attention in industry [5]. Some desirable properties of natural fibers include high specific strength and modulus, flexibility during processing, low self-weight, low cost, and substantial resistance to corrosion and fatigue [6]. However, natural fibers have drawbacks, such as high moisture absorption, high anisotropy, low compatibility with conventional resins, and less homogeneous than glass and carbon fibers [3, 7]. Biocomposites are classified into fully and partially green composites [4]. Petrochemical- and bio-based resins are used as matrices in the partially and fully green composites, respectively, and natural fibers are used in both [4]. Although all biopolymers are generally compostable, not all biodegradable and compostable plastics are produced from biopolymers; some are produced entirely from non-bio-based materials [4]. Owing to the environmentally friendly characteristic of green composites, they have being increasingly used in industry [1]. Industrial applications include aerospace, automobile, military, construction, packaging, medical, sporting equipment, and railway [3]. Table 1 summarizes the advantages and disadvantages of fully green composites over traditional composites [8]. The physical and mechanical properties of natural fiber composites have been investigated in a few previous review studies [9], which are summarized in Table 2. However, detailed information for treatment and modification techniques, fire resistance, thermal conductivity and analysis, durability properties, and time-dependent properties was not provided in these studies. Additionally, none of the existing review articles published in journals provide any information about creep, shrinkage, wear resistance, air content, compressive behavior, viscoelastic properties, and life-cycle assessment of the natural J Mater Sci (2020) 55:829–892 fiber composites. Therefore, to address the increasing demand to use eco-friendly materials in different applications, an up-do-date review is required to allow a better understanding of the behavior of natural fiber composites manufactured using discontinuous natural fibers. This paper presents such a review based on 322 studies published since 1978. Natural fibers In recent years, more renewable plant resources have been discovered and used because non-renewable resources are becoming scarce [1]. Figure 1 shows the structural organization of a natural fiber cell wall [24]. This type of fiber is strong, light, inexpensive, and renewable [24]. These inexpensive natural fibers can become a viable alternative for expensive and non-renewable synthetic fibers (e.g., e-glass, which is alumino-borosilicate glass, and carbon fibers) when high elastic modulus is not required [25]. However, the hydrophilic nature of natural fibers results in swelling of fibers and the production of voids at the interface between the matrix and fiber, leading to poor mechanical properties of composites prepared using these fibers [26, 27]. Therefore, suitable techniques must be utilized to produce natural fiber composites with high-quality natural fibers. Sources and types of fibers Natural fibers are produced from plant-, animal-, and mineral-based sources [1]. Fibrous plants are abundantly available in agricultural crops and tropical areas [1]. Plant fibers are mainly composed of cellulose, whereas protein is the major component of animal fibers [4]. Plant fibers are classified into primary sources, which are fibers produced as byproducts of other principal products (e.g., food, feedstock, and fuel) for industrial usage, and secondary plants, which are produced as by-products derived from manufacturing processes [4]. Table 3 shows the major natural fiber sources and their main properties. Table 4 shows the main producers and annual production of these fibers worldwide. There are eight major types of plant fibers: bast fibers (jute, ramie, flax, rattan, soybean, hemp, vine, banana, and kenaf), collected from the skin and bast around the plants stem; leaf fibers (abaca, banana, sisal, and pineapple), collected from leaves; seed 831 J Mater Sci (2020) 55:829–892 Table 1 Summary of advantages and disadvantages of fully green composites over traditional petrochemicalbased composites Advantages Disadvantages Less expensive Lower weight Higher flexibility Renewable Biodegradable Good thermal and sound insulation Eco-friendly Nontoxic Lower energy consumption No residues when incinerated No skin irritations Lower mechanical properties (especially impact strength) Higher moisture absorption Lower durability Poor fire resistance Variation in quality Restricted maximum processing temperature Poor microbial resistance Low thermal resistance Demand and supply cycles Table 2 Previous review studies on natural fiber composites Year Reference 2001 George et al. [10] 2008 John and Thomas [11] 2011 Zini and Scandola [12] 2011 La Mantia and Morreale [13] 2011 Ku et al. [14] 2012 Faruk et al. [4] 2012 Dittenber and GangaRao [15] 2012 Abdul Khalil et al. [16] 2013 Koronis et al. [17] 2013 Nguong et al. [18] 2014 Faruk et al. [19] 2015 Satyanarayana [20] 2015 Mohammed et al. [21] 2016 Pickering et al. [22] 2018 Sanjay et al. [23] Chemical Physical Processing modification modification techniques H Water Fire Elastic absorption resistance modulus Tensile strength Flexural strength H H H H H H H H H H H H H H H H H H Impact strength H H H H H H H H H H H H H H H fibers (cotton, coir, and kapok), collected from seeds and seed cases; grass fibers (corn, wheat, bamboo, barley, and rice); core fibers (corn and wheat stalk), collected from the stalks of the plants; wood pulp H H H H fibers; root fibers (luffa, swede, and cassava); and fruit fibers (borassus, tamarind, banana, and coir) [1, 4, 28]. 832 J Mater Sci (2020) 55:829–892 Figure 1 Structural organization of a natural fiber cell wall [24]. There are mainly three types of animal fibers: animal hair, avian fiber, and silk fiber. The fibers of animal hair (wool or hair) are taken from hairy mammals and animals (e.g., sheep, goats, alpacas, and horses) [1, 28]. Silk fibers are taken from the dried saliva of bugs or insects during the preparation of their cocoons. Avian fibers are taken from the feathers of birds [1, 28]. Properties of fibers Properties of a single fiber are dependent on the shape, size, crystallite content, orientation, and thickness of the cell walls [1]. The main characteristics of natural fibers are low energy consumption, low density, non-abrasive nature, low cost, renewability, biodegradability, easy availability, and worldwide abundance [42]. Although plant fibers are normally rigid, unlike brittle synthetic fibers, they are not fractured during processing [43]. Plant fibers exhibit comparable specific strength and stiffness properties to glass fibers [43]. Bast fibers are extracted from the stem ribbon using a retting process. This type of fiber has moderately high tensile strength and stiffness; it is inexpensive, exhibits high performance, and is easily available [1]. This fiber is more applicable when strength, lightweight, and noise absorption are important, such as in the automotive and building industries. Most of the plant fibers are categorized as eco-friendly fibers because they are biodegradable and have no negative effect on the environment, except those grown in temperate zone with extensive use of agrochemicals (e.g., flax) [1]. On the other hand, animal fibers have low stiffness balanced by their elastic recovery and high elongation [1]. They are also less hydrophilic than plant fibers, durable with moderate resistance, poor conductors of heat, very sensitive to some alkalis, and able to provide reinforcement in multi-axial situations [1]. The cell walls of dried plant fibers are mainly composed of lignin and sugar-based polymers (i.e., cellulose and hemicellulose) and small amount of other materials such as starch, extractives, protein, and inorganics [1]. Table 5 presents the chemical components of the most used dried plant fibers in industry with the average amount (in wt%) of the compositions. Cellulosic fibers have a hydrophilic nature (moisture absorbent) under natural conditions [1]. The moisture content in fibers can negatively influence the mechanical behavior of natural fiber composites. Table 6 illustrates the equilibrium moisture amount of natural fibers at the relative humidity (RH) of 65% and temperature of 21 °C. Grass (plant fibers) Seed (plant fibers) Leaf (plant fibers) Up to 3000 Up to 200 Up to 1000 Up to 1500 Up to 350 100–650 20–32 Up to 90 Flowering plant in nettle family (Boehmeria nivea) Abaca plant (Musa textilis) Pineapple leaf (Ananas magdalenae) Agave (Agave sisalana) Raffia palm (Raphia ruffia) Coconut (Cocos nucifera) Shrub (Gossypium) Pentandra tree (Ceiba pentandra) Grass pulp (Bambusoideae) Ramie Pina Sisal Raffia Coir Cotton Kapok Bamboo Abaca Up to 1900 6 Up to 4000 Hibiscus (Hibiscus cannabinus L) Cannabis (Cannabis Sativa L) Kenaf Hemp 20a 10–20 11–22 – 12–25 – 200–400b – 25–30 20–30 16–50a 17–20 Jute Up to 4000 Fiber diameter (lm) 12–16 Fiber length (mm) Up to 900 Flax Bast (plant fibers) Source Herbaceous plant (Linum Usitatissimum) Vegetable plant in linden family [Corchorus capsularis (white jute), Corchorus olitorius (tossa jute)] Fiber Group Table 3 Sources of major natural fibers and their main properties [1, 28–30] Rapid absorption and desorption of water Low thermal conductivity, moderate moisture regains, high insulation, high anti-static properties, growing in tropical areas with a humidity of 60–90% Excellent durability Heat conducting, good dying, good ultraviolet-light blocking, natural antibacterial properties Rapid absorption and desorption of water, low elasticity, easy dying High mechanical strength, buoyancy, resistance to saltwater damage Resistant to salt water, wear resistant Coarse, hard, durable, strong, and stretchable, not easily absorb moisture, resistant to saltwater deterioration, with a fine surface that accepts a wide range of dyes Rough High concentration of lignin, high strength, less flexibility than cotton, unsuitability for dyeing, good resistance to salt water damage and microbial action Rapid moisture absorption, high tensile strength Fluffy Excellent durability, high stability, good tenacity, good flexibility, good ultraviolet radiation resistance, excellent permeability Main properties J Mater Sci (2020) 55:829–892 833 Chicken, birds Lambs Indian cashmere goat North African angora goat Arabian dromedary and Northeast Asian Bactrian camels South America camels Angora rabbit Asbestos Mixed silicates Feather Lambswool Cashmere wool Mohair wool Camel hair Apparent diameter when agrochemicals are not used Apparent diameter b a Mineral-based fibers Saltwater clam Dog hair Muskoxen Yak Rabbits Sheep Byssus Chiengora Qiviut Yak Rabbit Wool Alpaca Angora wool Asbestos cloth Glass Maize (Maı́z) Beech tree (Fagus) Chinese mulberry silkworm Cornstalk Modal Silk Core (plant fibers) Wood pulp (plant fiber) Animal fibers Source Fiber Group Table 3 continued to to to to 50 390 115 125 Up to 150 – 12–300 – Up Up Up Up 3–13 – 25 50–80 16 – Up to 152 Up to 3000 – Up to 1500 Fiber length (mm) 12–29 12–16 0.03–0.035 5–24 15–20 14–19 25–45 15–23 5–50 10–50 – 15–20 16–90 14–16 16–40 40 200a 10–13 Fiber diameter (lm) Soft, warm Soft, good blending with other fibers Fire resistant, lightweight Fire resistant Lightweight, strong Lightweight, soft, wear resistant Good absorbency, low conductivity, easy dying finish Lightweight Lightweight, fluffy Soft, does not shrink Heavy, warm Soft Good thermal and acoustic insulation, high deformability, high durability Lightweight, good thermal and acoustic insulation Soft, warm, elastic Soft Durable, resilient, holding dyes well Warm, lightweight Main properties 834 J Mater Sci (2020) 55:829–892 835 J Mater Sci (2020) 55:829–892 Table 4 Producers and production amount of most widely used natural fibers [1, 4, 31–41] Production amount (9 103 ton) Fiber Producer Abaca Bagasse Coir Cotton Flax Jute Kapok Kenaf Bamboo Philippines (85%), Ecuador 70 Brazil, China, India, Thailand, Australia, USA 75,000 India, Sri Lanka, Thailand, Vietnam, Philippines, Indonesia, Brazil 1200 China, Brazil, India, Pakistan, USA, Uzbekistan, Turkey 25,000 France, Belgium, Netherland, Poland, Russian Federation, China 830 India (60%), Bangladesh, Myanmar, Nepal 3450 Philippine, Malaysia, China, South America, Indonesia, Thailand 101 India (45%), China, Malaysia, USA, Mexico, Thailand, Vietnam 970 China, Japan, India, Chile, Ecuador, Indonesia, Myanmar, Nigeria, Sri Lanka, Philippines, 30,000 Pakistan China (80%), Chile, France, Germany, UK 214 China, Brazil, Lao PDR, Philippines, India 280 Brazil (40%), Kenya, Tanzania, China, Cuba, Haiti, Madagascar, Mexico, Sri Lanka, India 378 China (70%), Brazil, Bulgaria, Egypt, Madagascar, India, Thailand, Vietnam, Uzbekistan, 150 Turkmenistan Australia, Argentina, China, Iran, New Zealand, Russia, UK, Uruguay 2100 China, Mongolia, Australia, India, Iran, Pakistan, New Zealand, Turkey, USA 20 Hemp Ramie Sisal Silk Wool Cashmere wool Mohair wool Camel hair Alpaca Angora wool South Africa, USA China, Mongolia, Afghanistan, Iran Peru, Bolivia, Chile China, Argentina, Chile, Czech Republic, Hungary, France Table 5 Chemical composition of the most widely used plant fibers [44–47] 5 2 7 3 Fiber Cellulose (wt%) Hemicellulose (wt%) Lignin (wt%) Waxes (wt%) Abaca Coir Cotton Flax Jute Kapok Kenaf Bamboo Hemp Pina Ramie Sisal Bagasse Oil palm Curaua Wheat straw Rice straw Rice husk 56–63 32–43 85–90 71 61–71 64 72 26–43 68 81 68.6–76.2 65 55.2 65 73.6 38–45 41–57 35–45 20–25 0.15–0.25 5.7 18.6–20.6 14–20 23 20.3 30 15 – 13–16 12 16.8 – 9.9 15–31 33 19–25 7–9 40–45 – 2.2 12–13 – 9 21–31 10 12.7 0.6–0.7 9.9 25.3 29 7.5 12–20 8–19 20 3 – 0.6 1.5 0.5 – – – 0.8 – 0.3 2 – – – – 8–38 14–17 836 J Mater Sci (2020) 55:829–892 Table 6 Equilibrium moisture content of major natural fibers at RH of 65% and temperature of 21 °C [48] Fiber Moisture content (wt%) Abaca Coir Flax Jute Bamboo Hemp Pina Ramie Sisal Bagasse 15 10 7 12 8.9 9 13 9 11 8.8 degradability. In the fully and partially green composites, natural fibers are used with bio-based (as a fully degradable matrix) and petrochemical-based (as a partly degradable matrix) resins, respectively. Petrochemical-based resins A petrochemical-based matrix is a chemical product derived from petroleum, which is obtained from fossil fuels like coal and natural gas [4]. The two major types of petrochemical-based matrices used for green composites are thermoplastics and thermosets. Polyethylene, polystyrene, polypropylene, and polyvinyl chloride (PVC) are utilized as thermoplastic matrices, whereas epoxy, polyester, vinylester, and phenolic (phenol formaldehyde) are used for thermoset matrices [1, 4]. Matrices for green composites A matrix is used in composites to hold the reinforcing materials together by surface connection. The main responsibilities of the matrix are the environmental tolerance, surface appearance, and durability of the composite [1]. As the matrix is stressed, it transfers the external load uniformly to the fibers, and it is applied to resist the propagation of cracks and damage [1]. In recent years, many studies have been carried out in an attempt to find an alternative for conventional petroleum-based matrices because of limited fossil fuel resources and the environmental impact of using petroleum-based matrices [4]. Figure 2 shows the classification of polymeric matrices used in green composites based on their Thermoplastics Thermoplastic resins are based on polymers that can be shaped easily in the viscous state and solidified by cooling (physical change) [49]. In the melted condition, the viscosity of thermoplastic resins is approximately 500–1000 times higher than that of uncured thermoset resins [4]. They are solid at room temperature and can be reformed and reshaped when heated without chemical reactions [4]. Thermoplastic resins have higher impact resistance (approximately 10 times), more reform-ability, higher damage tolerance, and higher processing temperatures and Matrices for Natural Fiber Composites Fully degradable Natural based • Polylactic acid • Thermoplastic starch • Cellulose • Polyhydroxy alkanoate Partly degradable Oil based • Aliphatic polyester • Aliphatic-aromatic polyester • Poly(ester amide) • Poly(alkyene succinate)s • Poly(vinyl alcohol) Figure 2 Classification of polymeric matrices for natural fiber composites [1]. • • • • Polypropylene Polyester Polyethylene Polyvinyl alcohol 837 J Mater Sci (2020) 55:829–892 Table 7 Advantages and disadvantages of petrochemical thermoplastic resins [50] Resin Advantage Disadvantage Polyethylene High ductility and impact strength Good fatigue resistance Lightweight Low moisture absorption Low cost High temperature resistance High dielectric resistance Excellent chemical resistance Good fatigue resistance Good chemical resistance Resistance to stress cracking Very low moisture absorption Weather resistance High impact resistance Versatility Good chemical resistance Flame retardant Low cost Good dimensional stability Poor weathering resistance Flammable High thermal expansion Polypropylene Polystyrene Polyvinyl chloride Difficult to process Comparatively expensive Limited availability Flammable Low impact resistance Brittle Poor resistance to UV Poor resistance at low and high temperatures Table 8 Advantages and disadvantages of petrochemical thermoset resins [53–56] Resin Advantage Disadvantage Epoxy High thermal and mechanical properties High water resistance Low curing shrinkage Long working times ability Easy to use Lowest cost More expensive than vinylester Corrosive amine hardener Difficult to process Polyester Vinylester Very high chemical and environmental resistance Higher mechanical properties than polyester Phenolic High fire resistance pressures than thermoset resins [1]. Table 7 shows the advantages and disadvantages of petrochemicalbased thermoplastic resins. Thermosets Thermoset resins are infusible and insoluble materials that are cured by heat or a catalyst [51]. Thermosets are completely different from thermoplastics; they cannot be melted and reshaped by heating [52]. Owing to three-dimensional covalent bonds between High curing shrinkage Limited range of working times Moderate mechanical properties High styrene emissions in open molds High curing shrinkage More expensive than polyester High styrene content Requires post-curing for good properties Difficult to process the polymer chains (chemical change), this type of resin has a higher modulus, improved creep resistance, higher thermal stability, and higher chemical resistance than thermoplastic resins [4]. They are also brittle at room temperature and show low fracture toughness. Table 8 shows the advantages and disadvantages of petrochemical-based thermoset resins. Bio-based resins Bio-based resins are polymers that are fully or partially obtained from renewable resources [1]. Bio- 838 Table 9 Advantages and disadvantages of bio-based resins [57–60] J Mater Sci (2020) 55:829–892 Resin Advantage Disadvantage Starch Fully biodegradable Low cost PLA High modulus and strength Nontoxic Relative low cost High molecular weight Fully biodegradable Brittle Difficult to process Water-sensitive Brittle Relatively poor impact strength Low thermal degradation temperature Relative low decomposition temperature Low stability Brittle Low deformability More expensive than other bio-based polymers High moisture absorbance Relative low decomposition temperature PHA Cellulose Abundant Relative low cost Ease to modify Moderate impact resistance Moderate heat resistance based polymers can be produced from plants (e.g., starch and cellulose) or through the polymerization of plant-based sugars and oils [e.g., polylactic acid (PLA), polyethylene terephthalate, and polypropylene] [4]. Based on the physical properties, there are three types of bio-based polymers: fully bio-based and biodegradable, e.g., polyhydroxyalkanoates (PHA) and starch, partially bio-based and biodegradable, e.g., cellulose and PLA, and partially bio-based and non-biodegradable, e.g., bio-polyethylene terephthalate, bio-polyethylene, and biopolypropylene [1, 4]. Table 9 shows the advantages and disadvantages of different types of bio-based resins. It was reported that in 2011 approximately 3.5 million tons of bio-based polymers were produced in the world, whereas 235 million tons of traditional petrochemical-based polymers were produced in that year [1]. The growth rate of bio-based polymer production has been high in recent years, and it has been estimated that the production will reach approximately 12 million tons per year by 2020 [1]. However, this amount is significantly lower than the production of petrochemical-based polymers, and it needs to be increased further to decrease the negative environmental impact of petroleum-based materials [1]. It has been shown that, for the same applications, some types of PLA made from maize starch utilize up to 50% less oil and release 60% fewer greenhouse gases into the atmosphere compared to petrochemicalbased polymers [1]. Fully bio-based and biodegradable polymers contain renewable carbon, which is taken from the atmosphere and then returned to the atmosphere as the polymer decomposes [4]. This renewable carbon can be obtained by collecting the polymers, which are currently used for disposable cutlery and flexible food packaging, in which the plastics are disposed of alongside the food waste [1]. Drop-in bio-based polymers like bio-based polyethylene terephthalate, which are chemically identical to petrochemicalbased polymers, can be traditionally recycled [1]. This type of polymer can be combusted to create renewable energy when it is not possible to recycle it. Bio-based resins offer several advantages over petrochemical-based resins: they are energy efficient in production (65% less energy required to produce), safe (remain nontoxic as they degrade), recyclable (break down faster with less energy), completely renewable (as they are made from biomass), and environmentally friendly (generate 68% less greenhouse gases) [61]. However, their production cost is approximately 10% higher than that of petrochemical-based resins [62]. Modification of natural fiber composites The interfacial adhesion across the phase boundary (i.e., at the interface between two materials) is vital for the mechanical behavior of composites [4]. Poor bonding at the phase boundary leads to a composite with weak mechanical properties. The main concern 839 J Mater Sci (2020) 55:829–892 of the utilization of natural fiber composites is the relative high moisture absorption of natural fibers, which results in weak compatibility between fibers and the matrix [4]. Although moisture absorption mostly affects natural fibers, there are some resins that absorb a large amount of moisture [1]. Therefore, both the fiber surface and the matrix need to be modified to enhance the adhesion of natural fibers to the matrix for improving the strength and stiffness of the natural fiber composite. Fiber modification Owing to the high sensitivity of natural fibers to moisture, moisture absorption results in delamination between the matrix and fiber, severely reducing the mechanical properties of the composite [4]. This is attributed to the fact that because of the presence of non-cellulosic components (i.e., pectin, lignin, and hemicelluloses), natural fibers in nature are polar and hydrophilic, and thus, they create active conditions (i.e., accessibility to hydroxyl (OH) and carboxylic acid groups) for water absorption [63]. Furthermore, differences in the environmental conditions, such as the amount of sun, rain, soil conditions, and the amount of water the plant receives during the growing period, as well as the processing and production conditions, may also affect the natural fibers performance [1]. Therefore, performance and properties of natural fibers may be different in every harvesting season and even in the same cultivation population. The other limiting factor for the utilization of the natural fibers in composites is their low thermal stability [4]. However, the problems regarding the utilization of natural fibers in the composite can be solved by physical and chemical modifications. Physical modification Physical modification on natural fibers enhances the mechanical adhesion between the natural fiber and matrix by enhancing the interface without changing the chemical properties of the fibers [4]. Extensive researches have been conducted to understand the influence of physical treatments on natural fibers. Physical treatment methods include corona, plasma, ultraviolet (UV), fiber beating, and heat treatment. These techniques are applied only to change the surface properties of natural fibers. The surface energy of cellulose fibers with corona treatment is changed to improve the compatibility between the hydrophilic fiber and matrix [64]. In this approach, using a high voltage at low temperature an atmospheric pressure plasma is generated [65, 66]. The surface energy of cellulose fibers is changed in the plasma treatment method by a surface modification technique similar to that of the corona treatment [67]. However, in plasma treatment, the gas type, flow, pressure, and concentration are controlled; in corona treatment they are not [67, 68]. The UV treatment method increases the polarity of the fiber surface, which leads to better wettability of the fibers and increased strength of the composite [64, 69]. In the fiber-beating approach, an increase in the surface area, defibrillation of the fibers, and mechanical interlocking can result in 10% increase in the strength of natural fibers [70]. In heat treatment, fibers are heated to a temperature close to the fiber degradation temperature. This condition affects the physical, mechanical, and chemical properties of the fibers, including their water content, chemistry, strength, cellulose crystallinity, and degree of polymerization [71, 72]. Chemical modification Chemical modification on natural fibers improves the adhesion between the matrix and natural fibers via chemical reactions. Extensive studies have been performed to understand the effect of chemical treatment on natural fibers. The hydrophilic nature of natural fibers and the hydrophobic nature of matrices are considered two different phases, resulting in weak bonding at the interfaces of natural fiber composites [4]. The chemical treatment of natural fibers decreases the inherent hydrophilic behavior of the fibers and improves the adhesion properties of the matrix and fiber [1]. Chemical treatment methods include alkaline, silane, acetylation, benzoylation, peroxide, maleated coupling agents, sodium chlorite, acrylation and acrylonitrile grafting, isocyanate, stearic acid, permanganate, triazine, oleoyl chloride, and fungal treatments [19, 73]. Alkaline treatment The alkaline treatment approach is one of the simplest and most economic and effective methods for improving the adhesion properties of natural fibers to the matrix [74]. In this method, the cellulosic molecular structure of the natural fibers is 840 modified using sodium hydroxide (NaOH) [74]. Alkaline treatment increases the speed of fiber fragmentation and disaggregation [75]. The orientation of the highly packed crystalline cellulose order is changed by creating amorphous regions in which cellulose micro-molecules are separated and the spaces are filled with water molecules [1]. Alkalisensitive OH groups are broken down and moved out from the fiber structure, and a fiber cell–O–Na group is created between cellulose molecular chains by the remaining reactive molecules. Therefore, the number of hydrophilic OH groups decreases, the resistance of the fiber to moisture increases, and a certain amount of hemicelluloses, lignin, pectin, wax, and oil are taken out [1]. The surfaces of the fibers become clean and uniform, which improves the stress transfer capacity between the cells. An optimum alkali concentration should be obtained to prevent extra delignification of the fibers, but higher concentrations can weaken and damage the fibers. With an optimum alkali concentration, the diameter of the fibers is decreased, resulting in better adhesion due to the increased effective fiber surface area and aspect ratio (length/diameter) [76]. Based on the report by Brigida et al. [77], alkaline treatment is the most efficient approach for cellulose exposition of natural fibers. They also found that alkaline treatment can maintain the native hydrophilic characteristic of green coconut fibers and increase their thermal stability. Figure 3 shows scanning electron microscopy Figure 3 SEM micrographs of longitudinal views of a untreated and b 6% alkalinetreated kenaf fiber [78]. J Mater Sci (2020) 55:829–892 (SEM) micrographs of longitudinal views of untreated and 6% alkaline-treated kenaf fibers [78]. Silane treatment In this approach, silane coupling agents coat micro-pores on the fiber surface. Silane forms a chemical link between the fiber surface and the matrix via a siloxane bridge [4]. At the initial stage of this treatment, silanols are created using existing moisture and hydrolyzable alkoxy groups [4]. One end of the created silanol reacts with a cellulose OH group, and the other end reacts with the matrix functional group by the condensation process. As a result, molecular continuity is formed across the interface of the composite and a hydrocarbon chain is obtained to restrain the fiber swelling into the matrix [1]. It has been shown that fibers subjected to silane treatment exhibit better tensile strength than those subjected to alkali treatment [24]. Figure 4 shows SEM micrographs of surface views of untreated and 2%, 4%, and 6% silane-treated hemp fibers [79]. Acetylation treatment In acetylation treatment, acid catalysts are used to graft acetyl groups to the cellular structure of the fibers [80]. The fibers are firstly soaked in the acetic acid and then treated with the acetic anhydride for 1–3 h at high temperature to accelerate the reaction because acetic acid and acetic anhydride cannot separately react with the fibers [4]. In this approach, an esterification reaction happens between OH and carboxylic/anhydride groups of the natural fibers [81]. Figure 5 shows SEM micrographs of surface views of 841 J Mater Sci (2020) 55:829–892 Figure 4 SEM micrographs of surfaces of a untreated and b 2%, c 4%, and d 6% silanetreated hemp fiber [79]. 100μm 100μm 100μm 100μm untreated and 18% acetylation-treated flax fibers [80]. In the acetylation treatment, the wax and cuticle on the surface of the fiber are removed by the interaction with acetyl, resulting in a smoother surface. Based on the report by Tserki et al. [82], acetylation treatment of flax and hemp fibers resulted in their reduced moisture absorption, removal of the non-crystalline constituents of the fibers, production of a smooth fiber surface, and improvement of the stress transfer efficiency. In addition, based on the study reported by Zafeiropoulos et al. [83], acetylation changed the bulk properties of the flax fibers, as well as their surface properties. Benzoylation treatment To decrease the hydrophilic nature of natural fibers, benzoyl chloride is used to Figure 5 SEM micrographs of surfaces of a untreated and b 18% acetylation-treated flax fiber [80]. improve the interfacial adhesion and thermal stability of fibers [84]. First, in a pretreatment step, lignin, waxes, and covering oil are extracted and reactive OH groups are exposed on the fiber surface. The fibers are then treated with the benzoyl chloride, where the OH groups of the fibers are replaced with benzoyl groups. Last, OH groups are attached to the cellulose backbone [81]. Figure 6 shows SEM micrographs of surface views of untreated and benzoylation-treated flax fibers [85]. Peroxide treatment In peroxide treatment, peroxideinduced grafting of polyethylene is cohered to the fiber surface and free radicals of peroxide react with the OH groups of the fiber and matrix [86]. This 842 Figure 6 SEM micrographs of surfaces of a untreated and b benzoylation-treated flax fiber [85]. J Mater Sci (2020) 55:829–892 10μm 10μm (a) method improves the adhesion of the fibers to the matrix at the interface, decreases the moisture absorption ability of the fibers, and increases the thermal stability of the fibers [81]. Figure 7 shows SEM micrographs of surface views of untreated and peroxide-treated flax fibers [85]. Maleated coupling agents Maleated coupling agents provide sufficient interaction between the functional surface of the fibers and the matrix, and they can form carbon–carbon bonds to the polymer chains of the matrix [87]. The fiber surfaces are coated with long-chain polymers due to the reaction of maleic anhydride with the OH groups in the amorphous region of the fiber cellulose structure [88, 89]. The coating excludes the OH groups from the fiber cells, reducing the hydrophilic tendency of the fibers, which leads to the formation of a bridge interface and efficient interlocking between the fiber and the matrix because of the covalent bonding between OH groups of the fibers and the anhydride groups of the maleic anhydride [81, 90]. Figure 8 shows SEM micrographs Figure 7 SEM micrographs of surfaces of a untreated and b peroxide-treated flax fiber [85]. 10μm (b) of untreated and maleic anhydride (MA)-treated jute fiber surfaces [91]. The MA-treated fiber shows improved fiber–resin adhesion at the interface, which leads to failure of the fiber by tearing instead of interfacial failure. Sodium chlorite treatment In sodium chlorite treatment, fibers are bleached in an acid solution using sodium chlorite (NaClO2) [92]. This approach removes moisture from the fiber and increases their hydrophobic nature, which increases the flexibility of the fibers [81]. Figure 9 shows SEM micrographs of surface views of untreated and sodium chloritetreated flax fibers [93]. Acrylation and acrylonitrile grafting In the acrylation and acrylonitrile grafting method, more access is provided for the reactive cellulose macro-radicals to the polymerization medium by the reaction of acrylic acid (CH2=CHCOOH) with the cellulosic OH groups of the fibers [94]. Ester linkages with the cellulose OH groups are produced by the carboxylic acid of 10μm (a) (b) J Mater Sci (2020) 55:829–892 843 Figure 8 SEM micrographs of surfaces of a untreated and b maleic anhydride-treated jute fiber [91]. Figure 9 SEM micrographs of surfaces of a untreated and b sodium chlorite-treated flax fiber [93]. coupling agents, and fiber moisture is eliminated by a reduction of the hydrophilic OH groups of the fiber structure [94]. Peroxide radicals start to graft with acrylic acid on the matrix, producing oxygen–oxygen bonds by extracting the hydrogen atoms of the polymer chains [1]. Therefore, the bonding capacity and stress transfer of the interface increase by the coupling mechanism of the fibers and the matrix [81]. Figure 10 shows SEM micrographs of surface views of untreated, acrylated, and acrylonitrile-graftedtreated oil palm fiber [95]. Isocyanate treatment Isocyanate treatment increases the bonding properties between the fibers and the matrix by providing strong covalent bonds between them [96]. It also increases the moisture resistance properties of the fibers [96]. The strong covalent bonds (chemical bondage) and moisture resistance are obtained by the reaction of the functional groups (–N=C=O) of isocyanate with the OH groups of the cellulose and lignin constituents of the fibers [81]. Figure 11 shows SEM micrographs of the surface views of untreated and isocyanate-treated oil palm fiber [95]. Stearic acid treatment In stearic acid treatment, the water resistance of the fibers increases due to the reaction of the carboxyl groups of stearic acid, obtained from an ethyl alcohol solution, with the hydrophilic OH groups of the fibers, thus eliminating the non-crystalline constituents of the fiber structure [97, 98]. Therefore, the fibers are dispersed better in the matrix by breaking down the fiber bundles with more fibrillation [81]. Figure 12 shows SEM micrographs of surface views of untreated and stearic acidtreated flax fibers [98]. Oleoyl chloride treatment Oleoyl chloride (fatty acid derivate) reacts with the OH groups of the fibers and improves their wettability and adhesion properties [99]. This improvement is obtained by making the fibers more hydrophobic by eliminating the hydrophilic OH groups from the external surfaces of the fibers [81]. Figure 13 shows SEM micrographs of 844 J Mater Sci (2020) 55:829–892 Figure 10 SEM micrographs of surfaces of a untreated, b acrylated-treated, and c acrylonitrile-grafted-treated oil palm fiber [95]. 200μm 1μm (a) (b) 200μm (c) Figure 11 SEM micrographs of surfaces of a untreated and b isocyanate-treated oil palm fiber [95]. 1μm (a) Figure 12 SEM micrographs of surfaces of a untreated and b stearic acid-treated flax fiber [98]. 200μm (b) 845 J Mater Sci (2020) 55:829–892 Figure 13 SEM micrographs of surfaces of a untreated and b 24-h oleoyl chloride-treated jute fiber in pyridine [99]. surface views of untreated and 24-h-oleoyl chloridetreated jute fibers in pyridine [99]. triazine with the hydrophilic OH groups of the cellulose and lignin constituents [81]. Permanganate treatment Chemical interlocking at the interface between natural fibers and the matrix is improved by treating the fibers with permanganate [100]. In this approach, better adhesion between the fibers and the matrix is provided by the reaction of potassium permanganate (KMnO4) with the cellulose OH groups and lignin constituents of the fibers, thus improving the thermal stability of the fibers [81]. Figure 14 shows SEM micrographs of the surface views of untreated and permanganate-treated flax fibers [100]. Fungal treatment Fungal treatment is a new biological treatment in which specific enzymes are applied to eliminate the non-cellulosic components of the fiber surface [102]. Extracellular oxidase enzymes are produced from white-rot fungi, which react with the lignin peroxidase to remove lignin from the fibers. Better interlocking between the fibers and matrix creates fine holes in the fiber surface via producing hyphae [81]. Figure 15 shows SEM micrographs of surface views of untreated and fungal-treated hemp fibers [102]. Triazine treatment Triazine (C3H3N3) treatment increases the adhesion properties of the fibers and matrix by providing covalent bonds between them, and it enhances the moisture resistance of the fibers [101]. The covalent bond and moisture resistance are obtained by the reaction of the functional groups of Figure 14 SEM micrographs of surfaces of a untreated and b permanganate-treated flax fiber [100]. Matrix modifications In general, the concern of moisture absorption in composites stems mainly from the natural fibers. However, some matrices (especially bio-based resins) such as soy protein resins absorb a large amount of 846 J Mater Sci (2020) 55:829–892 Figure 15 SEM micrographs of surfaces of a untreated and b fungal-treated hemp fiber [102]. moisture [1]. Soy protein concentrate resin was modified by Chabba and Netravali [103] through adding glutaraldehyde in one case and vinyl alcohol in the other. Both methods resulted in approximately 2% less moisture absorption as compared to that of the unmodified resin. Kumar and Zhang [104, 105] improved the moisture absorption of soy protein resin made with different superplasticizers using benzilic acid. They found that immersing the soy protein resin in 0.5% benzilic acid for 26 h resulted in 61% (with thiodiglycol superplasticizer) to 76% (with formamide superplasticizer) decrease in the moisture absorption of the resin. Newill et al. [106] found that a 0.254-mm-thick polymeric coating resulted in an approximately 47% reduction in the moisture absorption of the matrix. Doherty et al. [107] illustrated that the moisture absorption of a matrix could be reduced by 61% by applying a natural ligninbased coating with a 0.204-mm thickness. Processing techniques The main parameters influencing the processing of natural fiber composites are moisture, fiber type, fiber volume fraction, and temperature of the composite. The moisture of both the fiber and the matrix must be controlled before processing, and the required modifications must be performed if moisture is present [1]. The length, aspect ratio, and chemical composition of the fibers also have great effects on the processing, sustainability, and performance of composites [1]. An increase in the fiber volume fraction of the composite increases the composite’s stiffness, strength, and water uptake and decreases its deformability [1]. The processing temperature is another important factor. The maximum temperature that can be used in the processing period to avoid degradation of the most of natural fibers is 200 °C within 20 min [108, 109]. Any temperature above this limit can cause degradation, shrinkage, and low performance of the natural fiber composite because the chemical, physical, and mechanical properties of the natural fibers are changed by depolymerization, oxidation, hydrolysis, decarboxylation, dehydration, and recrystallization [110]. The major manufacturing methods for natural fiber composites are compression molding, extrusion molding, injection molding, and resin transfer molding techniques. Table 10 presents the advantages and disadvantages of the processing techniques for manufacturing natural fiber composites. Processing techniques of thermoplastic composites Compression molding Compression molding has been used since 1990s for thermoplastic composites as the demand for lightweight and high-performance materials increased [1]. In this technique, preheated materials are initially placed in the molding cavity. Then, they are compressed and deformed by the core side of the mold while subjecting the cavity to high pressure [1]. Before opening the mold and removing the composite, the high pressure is maintained until the composite solidifies. The important parameters that must be considered with this technique are the amount of material, heating time, pressure applied to the mold, and cooling time [109]. Sheet molding, as a type of compression molding technique, is one of the main processing approaches for composite production. Figure 16 shows the process diagram of sheet molding technique [123]. In this technique, a measured amount of a specific resin is 847 J Mater Sci (2020) 55:829–892 Table 10 Advantages and disadvantages of different processing techniques [109, 111–122] Technique Advantage Disadvantage Compression molding Fast setup time Low wasted material Low cost for large and complex composites Good surface finish Even pressure distribution on the composites Excellent part reproducibility Very high volume production ability Low labor requirements on a production level Fast setup time Low initial setup costs Low production costs Low production speed Suitable only for flat or moderately curved composite shapes Sheet molding Extrusion molding Injection molding Hand lay-up Resin transfer molding Resin infusion molding Low operational cost Low cost in mass production High throughput Flexibility to make parts with complex shapes High precision Simple principles to teach Low tooling cost Wide choice of material types and suppliers Flexibility in material design Better product consistency than that of compression molding Tighter tolerance and more intricate parts than injection molding Fast setup time and low setup costs Low maintenance costs Excellent surface quality on both sides Short process time Highly even quality and material thickness dispensed with a paste reservoir into the plastic carrier film, which is passed under a chopper to cut the fibers onto the surface [123]. Another sheet is added on top of the layer after placing the fibers in the resin paste. The sheets are then compacted and sent through to the take-up roll for storing the product. The carrier film is removed, and the material is cut into charges and molded under heat and pressure to prepare the composite based on the requested shape. Finally, the product is removed from the mold after full curation [1]. Suitable only for the preparation of composites with a low fiber volume fraction Moderate production speed Suitable only for the preparation of composites with a uniform cross section Mediocre precision High initial setup costs Labor intensive Styrene emission from unsaturated polyester and vinylester resins Dependency of the quality of the laminate to the skill of laminators Resins need to be low viscosity to be workable by hand More material is wasted than with compression molding Production speed is lower than that of injection molding Slow cycle times and high consumable costs Extrusion molding Extrusion molding is one of the most widely used procedures in the manufacturing process of natural fiber composites. This technique is preferred because of the high stiffness and strength of the composites and because the formation of the composites with this technique is very easy [1]. This technique begins with storing the thermoplastic material in the form of pellets or granules in a hopper. Then, they are delivered to a heated barrel to be molten. The molten plastics are subsequently used for the required shape of the composite. The final stage is cooling the product [1]. 848 J Mater Sci (2020) 55:829–892 Figure 16 Process diagram of sheet molding technique for composite production [123]. Figure 17 Process diagram of resin transfer molding technique for composite production [123]. Injection molding Injection molding is the most common approach utilized in the manufacturing of mass-produced composites. This method is initiated by putting the polymers in the form of pellets or granules into the hopper and heating them to be molten. The molten materials are then injected into a chamber formed by a split die mold and kept in the mold. The melt is chilled down to solidify before opening the mold [120]. Processing techniques of thermoset composites Hand lay-up Hand lay-up is the most common, simplest, and cheapest technique for production of composites [15]. In this technique, a release agent as an anti-adhesive agent is initially applied to the open mold, and the fibers are then placed in the mold. Resins are applied to the fibers by pouring and brushing with a roller or brush. Lay-up is made by building layer upon layer till the desired thickness. Entrapped air in the laminate is removed manually with squeegees or rollers. 849 J Mater Sci (2020) 55:829–892 Table 11 Density of major natural and synthetic fibers and polymers [15, 22, 129] Fiber and polymer type Density (kg/m3) Cotton Jute Flax Hemp Kenaf Ramie Sisal Coir Pina Abaca Bamboo Bagasse Banana Oil Palm Curaua Pulp Silk Feather Wool Carbon Glass Polypropylene Low-density polyethylene High-density polyethylene Polystyrene Polyester Nylon 6 Nylon 6,6 Vinylester Epoxy Phenolic Starch PLA PHA 1520–1560 1440–1520 1420–1520 1470–1520 1435–1500 1450–1550 1400–1450 1150–1220 1440 1500 600–1100 1250 1350 700–1550 1400 1500 1300 900 1300 1800–1840 2550–2600 890–910 910–925 940–960 960–1040 1040–1400 1120–1140 1130–1150 1200–1400 1110–1400 1160–1210 1000–1390 1210–1250 1180–1260 Vacuum is often used to avoid air bubbles and to help draw the resin into the cavity [116]. Resin infusion molding Resin transfer molding In resin transfer molding (RTM), the resin is preheated and loaded into the holding chamber instead of loading the polymer into an open mold [124]. This technique is most suitable for medium-volume production of large components. Figure 17 shows the process diagram of RTM technique [123]. Layers of the textile are prearranged in the solid mold, and the resin is then injected to impregnate the preforms. Physical properties Unit weight and density Table 11 shows the densities of most widely used natural and synthetic fibers and polymers, showing that the natural fibers have a lower density than synthetic fibers. Air content Over several years of producing synthetic composite materials, it is now possible to control and optimize the void content to be lower than 2% by volume of the composites [130]. Baley et al. [131] reported different techniques to determine the void content of 60 Flexural Strength (MPa) Laminates are then left to cure under standard atmospheric conditions [113]. Resin infusion molding (RIM) is a double-mold process with medium flow as one type of vacuum-assisted RTM, in which the second mold face is replaced with a flexible membrane [125]. The process takes place on timescales where the fibers swell during infusion [126, 127]. It is initially started by laying dry fibers in the metal face, and the flexible tool face is then pressed down over the composite object. The liquid resin is injected under moderate pressure into the mold and saturates the fibers. The mold can be subjected to a vacuum to minimize air pockets to enhance the quality of the composite [128]. 50 40 30 20 10 0 0 10 20 30 40 50 60 70 Void Content (%) Figure 18 The effect of void content on the flexural strength of a vacuum-treated random-oriented wood pulp fiber/starch composite manufactured using hot pressing [135]. 850 composites. The void content in natural fiber composites is a concern because of the voids inherent in natural fibers, which affect the transverse failure of the composites [132]. Air or other volatile substances can be trapped inside the composites during insertion of the fibers into the resin, and micro-voids can form after curing of the composites. Voids can form during compounding or melt-flow processing, or because of uneven shrinkage owing to thermal gradients during the solidification process (cooling). An increase in the cooling rate results in an increase in the void content of composites. Micro-voids create poor mechanical properties, which result in sudden failure of the composites [87]. The fiber volume fraction and fiber length in the composites are the main factors for void or bubble formation. Higher fiber content and length in the composite increase the probability of void formation [133]. It was reported that a high void content in graphite fiber/polyimide and glass fiber/polystyrene composites (i.e., over 20% by volume) results in low fatigue resistance [133, 134]. The shapes of the voids can be changed from spherical to elongated, and the void content can be decreased by a post-extrusion hot-drawing approach [133]. Dong and Takagi [135] assessed the effect of void content on the flexural properties of a vacuum-treated random-oriented wood pulp fiber/ starch composite manufactured using hot pressing (see Fig. 18). Judd and Wright [136] studied quantitatively the effect of voids on the mechanical properties of fiber/resin composites and reported that, regardless of resin, fiber type, and fiber surface treatment, the interlaminar shear strength of a composite decreases about 7% for each 1% void. Ghiorse [137] showed that laminate lay-up type affects the void distribution, and void length, area, and aspect ratio increase with an increase in the void content. J Mater Sci (2020) 55:829–892 Tensile properties The tensile properties of natural fiber composites are mainly affected by the interfacial adhesion between the resin and fibers [1]. Physical and chemical modification of the fiber and resin enhances the tensile properties of composites [14]. Tables 13 and 14 present the tensile properties of natural fibers and polymers (thermoplastics and thermosets), respectively. The tensile properties of natural fiber composites are highly dependent on the fiber volume fraction in the matrix resin. Although a number of researchers showed irregular trends for tensile properties of composites as a function of fiber volume fraction, it is generally true that by an increase in the fiber volume fraction below an optimum value, the load is distributed to more fibers and the matrix can carry the applied load after fibers fracture. This can lead to a higher tensile strength for the composite [14]. With further increases in the fiber volume fraction after the optimum amount, brittle fracture occurs in the fibers and the matrix cannot support the additional load from the fibers. Under these conditions, the low tensile strength eventually leads to failure of the whole composite [160]. The contrary and irregular trends for the tensile properties can be because of many factors, including incompatibility between the fibers and matrix, fiber degradation, and incorrect manufacturing processes [14]. Figures 19 and 20 illustrate the influence of fiber content on the tensile strength and elastic modulus of flax (manufactured using extrusion and injection molding) and hemp fiber/high-density polyethylene (HDPE) composites (manufactured by compression molding) [14, 93, 161]. Figure 21 shows the influence of the fiber content by weight on the elongation at break of cross-plied flax fiber/HDPE, cotton fiber/corn starch (CS), and jute fiber/polybutylene succinate (PBS) composites [14, 93, 145, 162]. Mechanical properties Flexural properties In general, the mechanical properties of natural fiber composites are lower compared to those of synthetic fiber composites. However, these properties can be enhanced by proper modifications of the natural fibers and matrices through the application of techniques summarized in previous sections. Table 12 presents a summary of the existing studies on the mechanical properties of natural fiber composites. Flexural stiffness is one of the key parameters for measuring the resistance of a composite against bending deformation. Flexural properties are mainly dependent on two parameters: the modulus of the composite material and the moment of inertia [4]. Table 15 presents the flexural properties of polymers used in natural fiber composites. An increase in the fiber content up to the optimum amount increases the 851 J Mater Sci (2020) 55:829–892 Table 12 Summary of existing studies on the mechanical properties of natural fiber composites Reference Composite Apparent fiber diameter (lm) Manufacturing method Observation Zampaloni et al. Random-oriented hemp, flax, kenaf, sisal, – [138] and coir fiber/polypropylene composite Compression molding Joseph et al. [139] Cross-plied abaca fiber/phenolic composite – Compression molding Biswas et al. [140] Random-oriented coir fiber/epoxy composite – Hand lay-up Gao and Mader [141] Random-oriented jute fiber/ polypropylene composite 22 Injection molding Karmaker and Schneider [142] Shibata et al. [143] Cross-plied jute and kenaf fiber/ polypropylene composite – Injection molding Random-oriented kenaf and bagasse fiber/corn-starch composite Kenaf: 140a Bagasse: 394a Compression molding Sharma and Cross-plied banana fiber/polyurethane Kumar [144] composite Liu and Hughes Woven flax fiber/epoxy composite [145] 80a Compression molding Resin infusion Zhang et al. [146] Unidirectional flax fiber/phenolic composite – Li et al. [147] Unidirectional flax fiber/epoxy composite 5–35a 40% by weight hemp fiber composite exhibited a higher tensile and flexural strength than 40% by weight kenaf, flax, coir, or sisal fiber composites Fiber length of 30 mm was optimum for achieving maximum tensile, flexural, and impact strength Fiber length of 30 mm was optimum to achieve maximum tensile and flexural strength. An increase in fiber length resulted in an increase in the impact resistance of the composite Fiber length is an important parameter that can affect the tensile behavior of the composite Superior mechanical properties for the jute fiber composite than kenaf fiber composite 60% and 66% increase in flexural modulus of kenaf and bagasse fiber composite with increasing fiber volume fraction, respectively 15% fiber content by weight resulted in the maximum flexural strength Fracture toughness is mostly dependent on the fiber volume fraction rather than reinforcement architecture Crack propagation was accompanied by extensive fiber bridging in the analysis of interlaminar fracture energy Approximately 30% increase in the interlaminar fracture energy of the composite using 1% by weight of multi-walled carbon nanotubes on the surface of flax fibers Improved resistance to crack propagation when fiber treated by styrene and composite cured by an out-of-autoclave curing process Maximum increase in fracture toughness of the composite by adding both micro-rubber and nanosilica particles – Compression molding Resin infusion Kafi et al. [148] Woven jute fiber/polyester composite – Compression molding Kinloch et al. [149] 17 Resin infusion Woven flax fiber/epoxy composite 852 J Mater Sci (2020) 55:829–892 Table 12 continued Reference Composite Apparent fiber diameter (lm) Manufacturing method Observation Hsieh et al. [150] Random-oriented rubber fiber/epoxy composite 1.24 Resin infusion Agunsoye and Aigbodion [151] Wong et al. [152] Random-oriented bagasse fiber/ polyethylene composite 13 Compression molding 175% increase in the fracture toughness with the addition of 20% by weight silica nanoparticles Fracture toughness steadily decreased with increasing fiber content Random-oriented bamboo fiber/polyester composite 300–380a Hand lay-up Alamri and Low [26] Random-oriented recycled cellulose/ epoxy composite 5–10 Compression molding Muralidhar [153] Unidirectional flax fiber/epoxy composite – Hand lay-up Bledzki et al. [154] Woven jute and flax fiber/epoxy composite Jute: 37 Flax: 20 Compression molding Cao et al. [155] Bagasse fiber/polyester composite 390a Compression molding Biswas et al. [34] Sisal fiber/polyester composite – Compression molding Dhakal et al. [156] Random-oriented hemp fiber/polyester composite 20 Hand lay-up Jandas et al. [157] Random-oriented banana fiber/polylactic acid composite – Compression molding Liang et al. [158] Unidirectional flax fiber/epoxy composite – Hand lay-up a The highest fracture toughness was achieved using 50% volume fraction and a fiber length of 10 mm Fracture toughness increased with an increase in the fiber content up to 46% by weight Compressive strength and stiffness of the composite was because of reinforcing the matrix rather than the fiber/matrix synergy An increase in the fiber content enhanced the impact strength of the composite. Flax fiber composite had higher impact resistance than jute fiber composite, and an increase in the number of voids decreased the impact strength of composite Impact strength of the composite 30% increased with an increase in the fiber content up to an optimum amount of 65% At 0.5% fiber volume fraction, sisal fiber/polyester composite had comparable impact strength to glass fiber/polyester composite Increasing the hemp fiber volume fraction up to 10% led to a 228% increase in the impact strength An increase in the fiber volume fraction up to 25% resulted in the maximum impact strength Compressive strength of the composite 15–30% decreased by applying 10 J impact loading Diameter of a fiber bundle flexural strength of natural fiber composites [4]. However, further increases in the fiber content decrease the flexural strength, owing to defects in the wetting of fibers that can induce stress concentration points in the composites. An increase in the fiber content also increases the flexural modulus of the composites [4]. The fracture toughness of natural fiber composites, which means a composite’s resistance to crack propagation, is mainly influenced by the intrinsic properties of the matrix, the fiber volume fraction, 853 J Mater Sci (2020) 55:829–892 Table 13 Tensile properties of most widely used natural fibers [14, 15, 22, 159] Fiber Tensile strength (MPa) Elastic modulus (GPa) Elongation (%) Cotton Jute Flax Hemp Kenaf Ramie Sisal Coir Pulp Pina Abaca Bamboo Bagasse Banana Oil Palm Curaua Silk Feather Wool 287–800 393–860 345–1500 550–920 195–666 400–938 468–790 135–240 1000 413–1627 430–760 140–800 222–290 500 80–248 87–1150 100–1500 100–203 50–315 5.5–12.6 13–60 27.6–90 55–70 53–66 61.4–128 9.4–25 4–6 40 34.5–82.5 6.2–20 11–32 17–27.1 12 0.5–3.2 11.8–96 5–25 3–10 2.3–5 7–12 1.5–1.8 2.7–3.2 2–4 1.3–5.5 3.6–3.8 2–7 15–40 4.4 1.6 1–10 2.5–3.7 1.1 1.5–9 17–25 1.3–4.9 15–60 6.9 13.2–35 Table 14 Tensile properties of polymers used in natural fiber composites [129, 159] Resin Tensile strength (MPa) Elastic modulus (GPa) Elongation (%) Type of resin Polypropylene Low-density polyethylene High-density polyethylene Polystyrene Nylon 6 Nylon 6,6 Starch PLA PHA Polyester Vinylester Epoxy Phenolic 26–41.4 40–78 14.5–38 25–69 43–79 12.4–94 5–6 21–60 18–24 41.4–89.6 69–83 55–130 50–60 0.95–1.77 0.055–0.38 0.4–1.5 4–5 2.9 2.5–3.9 0.125–0.85 0.35–3.5 0.7–1.8 2.07–4.41 3.1–3.8 3–6 4–7 15–700 90–800 2–130 1–2.5 20–150 35–300 31–44 2.5–6 3–25 2–2.6 4–7 2–10 1 Thermoplastic and the interfacial bonding strength [4]. The energy dissipation is influenced by pull out of fibers in the wake of crack propagation of the composites [149]. Compressive properties The compressive behavior of natural fiber composites is highly dependent on the reinforcement architecture and fiber volume fraction of the composite [4, 163]. With increasing fiber volume fraction below the optimum amount, higher compressive strength is Thermoset obtained for the composite, owing to a decrease in the number of voids [4]. Increased compressive strength is also attributed to good homogeneity and high compaction between the fibers and the matrix. However, as the fiber volume fraction exceeds the optimum value, the compressive strength decreases. Generally, when the fiber volume fraction exceeds 1.5%, a reduction of compressive strength of approximately 8.5% occurs for every 0.5% increase in fiber volume [164, 165]. However, it was shown that the compressive strength of natural fiber composites 854 J Mater Sci (2020) 55:829–892 Tensile Strength (MPa) 50 40 30 20 10 Flax-HDPE Hemp-HDPE 0 0 10 20 30 40 50 Fiber Volume Fraction (%) Figure 19 The effect of fiber content on the tensile strength of flax (manufactured using extrusion and injection molding) and hemp fiber/high-density polyethylene (HDPE) composites (manufactured by compression molding). Elastic Modulus (GPa) 8 6 4 2 Flax-HDPE Hemp-HDPE 0 0 10 20 30 40 Figure 20 The effect of fiber content on the elastic modulus of flax (manufactured using extrusion and injection molding) and hemp fiber/high-density polyethylene (HDPE) composites (manufactured by compression molding). Elongation (%) Flax-HDPE 100 Impact strength The impact behavior of natural fiber composites mainly depends on the bonding level between the matrix and fiber [4]. This property plays a major role during the service life of natural fiber composites [4]. The impact properties of natural fiber composites can be improved by modification methods. The impact loading may be the result of bumps, crashes, and falling objects and debris. Some of the influential parameters on the impact strength of composites are the level of adhesion, favorable bonding, fiber pullout, and energy absorption [4]. 50 Fiber Volume Fraction (%) 120 architecture parameter on the compressive behavior of a hemp fiber/polyester composite tube manufactured by filament winding technique with fiber orientation of 10°, 30°, 45°, 60°, and 90°. Figure 22 shows the macro- and micro-images of the tubes with fiber orientation of 45° and 90° [163]. They found that a fiber orientation of 10° with the loading direction resulted in the highest compressive strength and modulus, and the observed fracture modes were diamond-shaped buckling, micro-buckling, concertina-shape buckling, and progressive crushing. Cotton-CS Effect of fiber/matrix treatment on the mechanical properties Several researchers have assessed the effect of fiber or matrix treatment on the mechanical properties of natural fibers composites. Table 16 presents a summary of the existing studies on the effect of fiber/matrix treatment on the mechanical properties of natural fiber composites. Jute-PBS 80 Effect of crystallinity on the mechanical properties 60 40 20 0 0 5 10 15 20 25 30 35 Fiber Content by Weight (%) Figure 21 The effect of fiber content on the elongation at break of cross-plied flax fiber/HDPE, cotton fiber/corn starch (CS), and jute fiber/polybutylene succinate (PBS) composites. gradually reduced with increasing fiber volume fraction [165, 166]. Wecławski et al. [163] assessed the influence of the fiber orientation as a reinforcement Crystallinity restricts the mobility of the molecular chains of the material and plays a major role in the physical and mechanical behavior of natural fiber thermoplastic matrix composites [207]. The structure of a cellulose material composes amorphous and crystalline regions. The amorphous region absorbs chemicals, such as dyes and resins, whereas the compactness of the crystalline region makes it difficult for chemical penetration [207]. Chemical treatments on the fiber increase its crystallinity, owing to the removal of lignin and hemicellulose after 855 J Mater Sci (2020) 55:829–892 Table 15 Flexural properties of the polymers used in the natural fiber composites [1] Resin Flexural strength (MPa) Flexural modulus (GPa) Polypropylene Low-density polyethylene High-density polyethylene Polystyrene Nylon 6 Nylon 6,6 Polyester Vinylester Epoxy Phenolic Starch PLA PHA 40 9 32 70 85 103 70–110 130–140 110–150 80–135 52 51–70 94 1.5 0.2 1.2 2.5 2.3 3.1 2–4 3 3–4 2–4 2.4 4.2 2.7 Figure 22 a A diagram with positioning of cross section and fiber orientation, and macro- and micro-images of hemp fiber/polyester composite with fiber orientation (h) of b 45° and c 90° [163]. (a) Hemp fibers Section A-A 500μm (b) Hemp fibers Hemp fibers Section A-A 500μm (c) Section B-B 500μm 856 treatment [208, 209]. A limited number of studies assessed the effect of crystallinity on the mechanical properties of natural fiber composites. Rong et al. [72] investigated the effect of different fiber treatments on the crystallinity and found that heat treatment (at 150 °C for 4 h) of unidirectional sisal fibers led to slightly higher crystallinity index (66%) than those under acetylation (65%) and alkaline treatment (61%). They found that alkaline-treated sisal fiber/epoxy composites manufactured by compression molding technique had the highest tensile strength and modulus, acetylated-treated sisal fiber/epoxy composites had the highest flexural strength, and heat-treated sisal fiber/epoxy composites had the highest flexural modulus. Barone [210] investigated the influence of the degree of crystallinity of the polymer on the tensile strength and modulus of random-oriented keratin feather fiber/polyethylene composite manufactured by compression molding technique. They found that an increase in the crystallinity of the polymer obtained from heat treatment (at 170 °C for 15 min) led to an increase in the tensile strength and modulus of the composite because of the increased fiber/polymer interaction. Zafeiropoulos et al. [211] and Joseph et al. [87] reported that cooling rate is one of the main parameters affecting the crystallinity degree of the natural fiber-reinforced composites. They found that lower cooling rates result in smaller fragment lengths of random-oriented flax fiber/ polypropylene and sisal fiber/polypropylene composite, indicating a better stress transfer-ability and stronger interface by slow cooling. Suryanegara et al. [212] reported that random-oriented pulp fiber/PLA composite manufactured by compression molding technique with 3% fiber weight fraction and fully crystallized polymer (heat-treated at 100 °C for 1 h) had 4% and 16% higher tensile strength and modulus compared to those with fully amorphous polymer, respectively. However, they found that the crystallinity of polymer increased the brittleness of the composite. Figure 23 shows the axial tensile stress– strain curves of pulp fiber/PLA composite under amorphous and crystallized states [212]. When a fiber is embedded into a semicrystalline polymer, the fiber may act as a nucleating site for the growth of spherulites [213]. If many nucleation sites exist along the fiber surface, the resulting spherulite growth will be restricted in the lateral direction and a columnar layer, known as trans-crystallinity, develops and encloses the fiber [213]. Although several J Mater Sci (2020) 55:829–892 studies have investigated the trans-crystallinity of conventional composites, only a few studies investigated the influence of the trans-crystallinity on the mechanical properties of natural fiber composites. Zafeiropoulos et al. [214, 215] studied the effect of trans-crystallinity on the interface of flax fiber/ polypropylene composite and found that the induced fiber trans-crystallinity from stearation surface modification resulted in 81% increase in the interfacial shear strength of the composite, owing to the improved interfacial adhesion between fiber and polymer by the presence of a trans-crystalline layer. Sawpan et al. [216] assessed the influence of the fiber trans-crystallinity from alkaline treatment on the interfacial shear strength of random-oriented hemp fiber/PLA composite manufactured by extrusion and injection molding techniques and reported that alkaline-treated fiber composite had 18% higher interfacial shear strength than that with no fiber treatment. They attributed their observation to the positive influence of the fiber crystallinity at the fiber/polymer interface, which is shown in Fig. 24 [216]. Simulation and modeling Analytical models Six analytical models have been used for prediction of the mechanical behavior of natural fiber composites with discontinuous fibers, including rules of mixture, series and parallel, Hirsch model, Halpin– Tsai equation, Bowyer–Bader model, and shear lag model [217]. The fundamental properties used in these models are elastic modulus of the fiber (Ef), elastic modulus of matrix (Em), tensile strength of fiber (rf), tensile strength of matrix (rm), fiber volume fraction (Vf), and matrix volume fraction (Vm). In these models, the micro-defects in the polymer and fiber/polymer interface are not considered [218]. Virk et al. [219] used rules-of-mixture model to predict the elastic modulus and tensile strength of unidirectional jute fiber/epoxy composite manufactured by resin infusion molding technique and reported that using the common method to measure the cross-sectional area of the fibers by linear measurements of fiber diameter and an assumption of circular cross section results in an overestimation of the cross-sectional area and hence leads to low values of key mechanical properties of natural fibers. They 857 J Mater Sci (2020) 55:829–892 Table 16 Effect of fiber/matrix treatment on mechanical properties of natural fiber composites Reference Treatment Composite Apparent fiber diameter (lm) Manufacturing method Observation Gassan and Gutowski [64] Fiber corona treatment Tossa jute fiber/epoxy composite – Compression molding Ragoubi et al. [167] Fiber ultraviolet (UV) treatment Fiber corona treatment 30% increase in the flexural strength under optimum corona treatment condition 30% increase in the flexural strength Random-oriented hemp fiber/polypropylene composite Flax and hemp fiber/tannin and hexamine composite Random-oriented miscanthus fiber/polylactic acid and polypropylene composite Random-oriented flax fiber/ polyester composite Random-oriented sisal fiber/ high-density polyethylene composite Woven jute fiber/highdensity polyethylene composite Random-oriented jute fiber/ polyester composite – Extrusion molding Improved Young’s modulus, stiffness, and tensile strength – Compression molding – Extrusion– compression molding Improved tensile strength and modulus by optimum duration of corona treatment of 10 min Improved mechanical properties 20 Compression molding Compression molding Improved permeation and mechanical properties 25% and 18% increases in the tensile strength and modulus, respectively 30a Compression molding 30a Hand lay-up 35% and 30% increases in the interlaminar shear strength and flexural strength, respectively 14% increase in the flexural strength 100a Hand lay-up Pizzi et al. [168] Ragoubi et al. [169] Marais et al. [67] Fiber plasma treatment Martin et al. [68] Seki et al. [170] Sinha and Panigrahi [171] Seki et al. [172] Woven jute fiber/polyester composite Jang et al. [173] – Random-oriented coir fiber/ – polylactic acid composite Random-oriented kraft fiber/ 25 polypropylene composites Jute fiber/vinylester – composite Compression molding Extrusion molding Compression molding Aziz et al. [78] Random-oriented hemp and kenaf fiber/polyester composite Compression molding Cao et al. [155] Random-oriented bagasse fiber/polyester composite Hemp: 180a Kenaf: 110a 390a Prasad et al. [174] Random-oriented coir fiber/ polyester composite 150–300a Hand lay-up Beg and Pickering [70] Ray et al. [74] Fiber-beating treatment Fiber alkaline treatment Compression molding 72% and 129% increases in the interlaminar shear strength through the low- and radio-frequency oxygen plasma treatment, respectively Improved mechanical as well as the thermal behavior 10% enhancement in the tensile strength 35%, 23%, and 19% increases in the flexural strength, modulus, and laminar shear strength, respectively 25% and 150% increases in the flexural strength of hemp and kenaf fiber composite, respectively. 13% and 65% increases in tensile and impact strength, respectively, with a 1% alkaline fiber treatment 40% increase in the impact and flexural strength by 5% alkaline treatment for 72 h 858 J Mater Sci (2020) 55:829–892 Table 16 continued Composite Apparent fiber diameter (lm) Manufacturing method Observation Ray et al. [175] Random-oriented jute fiber/ vinylester composite – Hand lay-up Onal et al. [176] Random-oriented carpet waste jute yarn composite Woven pina fiber/polylactic acid composite 10 Compression molding Compression molding 20% enhancement in the flexural strength by 5% alkaline treatment for 4 h Improved impact strength Reference Treatment Huda et al. [177] 50a Qin et al. [178] Unidirectional ramie fiber/cellulose composite – Compression molding Bledzki et al. [179] Unidirectional flax and hemp fiber/epoxy and polypropylene composite Random-oriented banana fiber/epoxy composite – Filament winding 150a Hand lay-up Random-oriented hemp fiber/polypropylene composite Banana fiber/polylactic acid composite – Compression molding – Compression molding Xu et al. [181] Random-oriented kenaf fiber/polystyrene composite 50–120a Compression molding Pothan and Thomas [182] Cantero et al. [183] Abaca fiber/polyester composite Random-oriented flax fiber/ polypropylene composite 50–250a Compression molding Extrusion molding Abdul Khalil and Ismail [184] Ismail and Abdul Khalil [185] Sgriccia et al. [186] Oil palm empty-fruit-bunch – and coir fiber/polyester composite 180a Oil palm empty-fruitbunch/natural rubber composite Hemp and kenaf fiber/epoxy 160a composite Eng et al. [187] Random-oriented oil palm fiber/polylactic acid composite Venkateshwaran et al. [180] Suardana et al. [79] Jandas et al. [157] Fiber silane treatment – – Compression molding Compression molding Compression molding Melt blending 26% improvement in the flexural strength through 40% alkaline treatment 23% improvement in the tensile strength through 9% alkaline fiber treatment for 1 h Improved flexural strength, tensile strength, tensile modulus, and elongation 20, 12, 132, and 131% increases in the flexural strength, flexural modulus, tensile strength, and tensile modulus through 1% alkaline treatment, respectively 2% and 4% increases in the flexural and tensile strength through 6% silane treatment, respectively 136% and 49% enhancement in the tensile and impact strength, respectively 45% and 25% increases in the storage modulus and tand (the ratio of viscous to elastic properties) through 5% silane fiber treatment, respectively Improved storage modulus 6% and 3% increases in the tensile and flexural strength through 2.5% silane treatment, respectively Improved tensile strength Improved tensile strength, modulus, tear strength, fatigue life, and hardness Silane-treated fiber composite had a higher flexural modulus than alkaline-treated composite and similar to glass fiber composite 18% improvement in the tensile and flexural strength 859 J Mater Sci (2020) 55:829–892 Table 16 continued Reference Treatment Composite Apparent fiber diameter (lm) Manufacturing method Observation Bledzki et al. [80] Joseph et al. [188] Fiber acetylation treatment Random-oriented flax fiber/ polypropylene composite Random-oriented abaca fiber/phenolic composite 100a Injection molding Compression molding Up to 35% increases in the tensile and flexural strength 81, 700, and 8% enhancement in the tensile strength, Young’s modulus, and impact strength through 50% acetylation treatment, respectively. Higher improvement in the tensile strength, Young’s modulus, flexural strength, and impact strength of the composite by acetylation treatment than those by silane treatment Improved stress transfer efficiency at the interface of fiber and resin by determining the optimum treatment time Decreases in the flexural properties 240a Hill and Abdul Khalil [189] Random-oriented coir and oil palm fiber/polyester composite Coir: 336a Oil Palm: 408a Compression molding Zafeiropoulos et al. [98, 190] Flax fiber/polypropylene composite 20 Compression molding Luz et al. [191] Random-oriented bagasse fiber/polypropylene composite Sisal fiber/polystyrene composite 10 Injection molding – Compression molding 91% improvement in the tensile strength – Manikandan et al. [192] Flax fiber/low-density polyethylene composite Random-oriented sisal fiber/ polystyrene composite 6% enhancement in the tensile strength Improved storage modulus and damping properties Joseph et al. [86] Fiber peroxide treatment Sreekala et al. [95] Sapieha et al. [193] Sisal fiber/polyethylene composite Random-oriented oil palm fiber/phenolic composite Cellulose fiber/polyethylene composite – Compression molding Combined compression/ extrusion molding Compression molding Compression molding Compression molding Abdul Razak et al. [194] Hong et al. [91] Kenaf fiber/polylactic acid composite Random-oriented jute fiber/ polypropylene composite – Manikandan et al. [84] Fiber benzoylation treatment Wang et al. [85] Yang et al. [195] Fiber treated by a maleated coupling agent Random-oriented rice husk fiber/polypropylene composite – 150–500a – 45a 209a Compression molding Compression molding Injection molding Improved tensile strength Improved impact resistance and flexural and tensile properties Dicumyl peroxide-treated composites exhibited superior mechanical properties than treatment benzoyl peroxide-treated composites Improved tensile and flexural strength Improvements in the Young’s modulus and dynamic storage modulus through maleic anhydride treatment Improved tensile strength through the positive influence of maleated polypropylenes on the interfacial bonding 860 J Mater Sci (2020) 55:829–892 Table 16 continued Reference Treatment Li et al. [196] Keener et al. [197] Liu et al. [198] Gassan and Bledzki [199] Wielage et al. [200] Li et al. [93] Fiber bleaching Misra et al. [201] Vilay et al. [94] Sreekala et al. [202] Joseph et al. [203] Paul et al. [204] Kalaprasad et al. [205] Torres et al. [206] Li et al. [100] Pickering et al. [102] Fiber acrylation and acrylonitrile grafting Composite Apparent fiber diameter (lm) Manufacturing method Observation Random-oriented hildegardia fiber/ polypropylene composite Random-oriented jute and flax fiber/polypropylene composite Random-oriented abaca fiber/high-density polyethylene composite Woven tossa jute fiber/ polypropylene composite – Injection molding Improved tensile strength through maleated polypropylenes 10–20 Injection molding 60% increase in the tensile and flexural strength – Injection molding – Compression molding Random-oriented flax and hemp fiber/polypropylene composite Random-oriented flax fiber/ polyethylene composite Random-oriented sisal fiber/ polyester composite – Injection molding – Injection molding Hand lay-up Improvement in the moduli and strength through maleic anhydridegrafted styrene treatment Improved dynamic modulus and specific damping capacity through maleic polypropylene anhydride treatment Improved dynamic mechanical properties through maleic acid anhydride-grafted treatment Improved tensile strength through sodium chlorite treatment Improved tensile, flexural, and impact properties through sodium chlorite treatment Superior tensile and flexural properties for acrylation-treated composite than those with alkaline treatment Improved tensile strength, Young’s modulus, and elongation at break of the acrylic acid-treated fibers Improved tensile properties 50–240 Random-oriented bagasse fiber/polyester composite 10 Vacuum bagging Random-oriented oil palm fiber/phenolic composite – Compression molding 100–300 Injection molding Compression molding Compression molding Improved tensile and flexural strength Improved tensile strength and Young’s modulus Compression molding Injection molding 23% increase in the shear strength through 3% stearic acid treatment Improved tensile strength through potassium permanganate treatment Injection molding 22% increase in the tensile strength through fungal treatment, 32% increase in the tensile strength through combined fungal/alkaline treatment Fiber isocyanate Random-oriented sisal fiber/ treatment polyethylene composite Fiber stearic acid Banana fiber/polypropylene treatment composite Random-oriented sisal fiber/ low-density polyethylene composite Sisal fiber/polyethylene composite Random-oriented flax fiber/ Fiber high-density polyethylene permanganate composite treatment Fiber fungal Random-oriented hemp treatment fiber/polypropylene composite 125–150a 100–300 223 – 20 861 J Mater Sci (2020) 55:829–892 Table 16 continued Reference Treatment Chabba and Resin treatment Netravali [103] by glutaraldehyde and vinyl alcohol a Composite Apparent fiber diameter (lm) Manufacturing method Observation Unidirectional flax fiber/soy protein composite – Compression molding Glutaraldehyde improved mechanical properties and vinyl alcohol improved the toughness Diameter of a fiber bundle proposed a fiber area correction factor to accurately predict the elastic modulus and tensile strength of the composite. Kalaprasad et al. [220] conducted a comparison between prediction of tensile properties of random-oriented sisal fiber/low-density polyethylene composite manufactured by injection 80 Axial Stress (MPa) Amorphous Polymer 60 40 0% Fiber 3% Fiber 5% Fiber 10% Fiber 20% Fiber 20 0 0 0.02 0.04 0.06 0.08 Axial Strain (a) 80 Axial Stress (MPa) Crystallized Polymer 60 40 0% Fiber 3% Fiber 5% Fiber 10% Fiber 20% Fiber 20 0 0 0.02 0.04 0.06 0.08 Axial Strain (b) Figure 23 Axial tensile stress–strain curves of pulp fiber/ polylactic acid composites under a amorphous and b crystallized states. molding technique obtained by series and parallel, Hirsch, shear lag, Halpin–Tsai, and Bowyer–Bader models and reported that all models except the series-and-parallel model showed good agreement with experimental tensile strength and modulus of longitudinally oriented composites. They also reported that Hirsch and Bowyer–Bader models had a good agreement with experimental results of randomly oriented composites. Facca et al. [221] predicted the Young’s modulus of random-oriented hemp fiber/ HDPE composite manufactured by compression molding technique using rules of mixture, Halpin– Tsai equation, and shear lag model. Their results showed that Young’ modulus predicted by Halpin– Tsai equation had a higher accuracy than that by other models because the efficiency factor in the Halpin–Tsai equation is normally derived from experimental measurements of the material to be modeled. Beckermann and Pickering [222] predicted the tensile strength of hemp fiber/polypropylene composite manufactured by extrusion and injection molding techniques by Bowyer–Bader model and reported that fiber aspect ratio and orientation are two main parameters that should be considered in the prediction of the mechanical behavior of natural fiber composites. Migneault et al. [223] predicted tensile elastic modulus of random-oriented pulp fiber/HDPE composite manufactured by injection molding technique using shear lag model and Halpin–Tsai equation and reported that the predictions considering an orientation factor in the modeling had a good agreement with the experimental results. Munde and Ingle [224] compared the modeling results of Hirsch model, Halpin–Tsai equation, and Bowyer–Bader model with the experimental results of the tensile strength and elastic modulus of 862 J Mater Sci (2020) 55:829–892 Figure 24 Optical light micrographs of hemp fiber/ polylactic acid composite with a untreated fiber and b alkaline-treated fiber [216]. 50 μm 50 μm (a) random-oriented coir fiber/polypropylene composite manufactured by compression molding technique. They found that Hirsch model developed a higher accuracy in prediction of the tensile strength and modulus of the composite than Halpin–Tsai and Bowyer–Bader models. Finite element simulation Finite element (FE) method has been used to model the properties of natural fiber composites, because it is possible to efficiently study the effect of different parameters, such as fiber reinforcement type, fiber volume fraction, fiber aspect ratio, and fiber orientation, on the properties of natural fiber composites [225]. Among the FE methods, representative volume element (RVE) FE, as the most popular homogenization-based multi-scale method and most effective and accurate method for mechanical properties of composite materials, has been considered for prediction of the mechanical properties of composites. Two RVE methods of direct RVE and orientation averaging methods have been used to simulate natural fiber composites with discontinuous fibers. In direct RVE method, matrix surrounds a number of natural fibers, whereas in the orientation averaging method, a single fiber is embedded in the matrix (as a block unit cell (UC)). A few studies predicted the mechanical properties of natural fiber composites with discontinuous fibers by FE method. Modniks and Andersons [226] used RVE orientation averaging method to predict the Young’s modulus of randomoriented flax fiber/polypropylene composite manufactured by extrusion molding technique (shown in Fig. 25). They assumed that the UC is transversely isotropic and determined the elastic constants of the composite from five loading nodes. They showed that (b) the modeling predictions were in a good agreement with the experimental results. Modniks and Andersons [227] also used RVE orientation averaging method for prediction of the nonlinear deformation of the random-oriented flax fiber/polypropylene composite (shown in Fig. 26). They first conducted the simulation under six load conditions of transverse tension, transverse compression, pure tension, axial tension, shear tension, and equi-biaxial tension to determine deformation parameters, and then they input the parameters into the RVE orientation averaging method for prediction of the nonlinear deformation of the composite. They reported a good agreement between the stress–strain curve predictions and experimental results. Kern et al. [225] developed a direct 3D RVE model for investigating the influence of fibers and micro-cellular voids on the tensile behavior of random-oriented wheat straw fiber/polypropylene composite (shown in Fig. 27). They found that the composite with voids had a shorter fracture path than that with no void. Their predictions were in a good agreement with the experimental results. Sliseris et al. [228] developed a direct 3D RVE model for simulation of the tensile behavior of random-oriented flax fiber/polypropylene composite considering the influence of the fiber aspect ratio, bundles, and defects. Flax fiber was modeled as a linear isotropic elastic material and polypropylene resin was modeled as a nonlinear plastic material. To consider the influence of the fiber defect in the modeling, they modeled the fibers and the interface between the fiber bundles as a brittle material with a continuum damage mechanic model (shown in Fig. 28). They found that plastic deformation of the composite started at the fiber endings and around fiber defects, which was in agreement with the experimental results. 863 J Mater Sci (2020) 55:829–892 Figure 25 Schematic of a the unit cell (comprising a fiber embedded in a block of matrix) and b loading nodes of the unit cell for prediction of the Young’s modulus by representative volume element (RVE) orientation averaging method [226]. Longitudinal modulus Transverse modulus Shear modulus Shear modulus Orthogonal cross-secons Poisson’s rao (a) Figure 26 Schematic of the unit cell for prediction of the nonlinear deformation of a random-oriented fiber composite [227]. Durability properties The low durability of natural fiber composites results in a decrease in the mechanical properties and the growth of fungus and bacteria in the composite [229]. Several studies have been conducted to investigate the key durability-related properties of water Figure 27 Meshing of arbitrarily distributed wheat straw fibers embedded in a solid polypropylene and b polypropylene with microcellular voids in a direct 3D representative volume element (RVE) model [225]. (b) absorption, wear resistance, and weathering in natural fiber composites. Table 17 presents a summary of the existing studies on the durability properties of natural fiber composites [230–240]. All natural fibers are hydrophilic in nature, and their saturation moisture content ranges from 3 to 13% [241]. The water absorption of the natural fiber in natural fiber composites is a serious concern for their potential outdoor application [241]. The absorption of moisture in natural fibers results in the swelling of fibers and a reduction of the adhesion between fibers and matrix, dimensional instability, matrix cracking, and weak mechanical properties of the composites due to debonding between the fibers and the matrix [139, 242]. The water absorption property of natural fiber composites depends on the fiber volume 864 J Mater Sci (2020) 55:829–892 Figure 28 Modeling of the failure mechanism considering the damage at fiber defects using 3D representative volume element (RVE) model for simulation of the tensile behavior of random-oriented flax fiber/polypropylene composite [228]. fraction, fiber orientation, temperature, exposed surface area, permeability of the fibers, void content, and hydrophilicity of the composite components. The water absorption increases with increasing natural fiber volume fraction of composite. All three dimensions of composites increase by the fiber water absorption. Pressurization at low fiber volume fractions can improve the water absorption of natural fiber composites. Chemical treatment of fiber surfaces can also remove the hydrophilic OH bonds [85, 243]. Symington et al. [244] found that the mechanical properties of flax and coir fiber composites decreased more than those of kenaf, jute, and abaca fiber composites under fully soaked conditions. They also showed that abaca and kenaf fibers had the highest capacity to absorb moisture (164% and 150% by weight, respectively), which resulted in severe damage to the structure of the fibers. Wear resistance refers to the ability of the composite materials to withstand scraping, skidding, rubbing, and sliding objects on its surface [233]. The extended exposure of natural fiber composites to sunlight and UV radiation results in physical and mechanical weakening of the composite [237]. Time-dependent properties Creep The extended exposure of natural fiber composites to stress results in a decrease in the strength and elastic modulus of composites [1]. The damage mechanisms in composites due to creep result in fiber fracture, fiber pull-out, matrix yielding, interfacial debonding, and crack coalescence. An imperfect interface leads to repeated frictional sliding in cyclic loading under fatigue conditions. Three important parameters that can affect the creep behavior of natural fiber composites are fiber loading, temperature, and styrene concentration [245]. A number of studies have assessed the effect of the creep on natural fiber composites. Alvarez et al. [246] investigated the influence of fiber content on the creep behavior of random-oriented sisal fiber/starch composites manufactured by injection molding technique through short-term flexural creep tests and found improved creep resistance by an increased fiber content in the matrix. Acha et al. [247] found the same result for a woven jute fiber/polypropylene composite. Xu et al. [248] showed that a random-oriented bagasse fiber/ PVC composite had better creep resistance than random-oriented bagasse fiber/HDPE composites manufactured by extrusion molding technique at low temperatures. Amiri et al. [249] presented a time– temperature superposition principal to generate a creep compliance master curve for a flax fiber/vinylester composite. Yang et al. [250] reported that the increase in banana fiber content up to 60% by weight in a PLA matrix in the random-oriented direction resulted in the highest creep resistance. The influence of fiber treatment on the creep behavior of woven jute fiber/epoxy composite manufactured by hand lay-up method was investigated by Militký and Jabbar [251] through a three-point bending test. They showed that a plasma-treated fiber composite exhibited a lower creep strain than an untreated fiber composite. 865 J Mater Sci (2020) 55:829–892 Table 17 Summary of existing studies on the durability properties of natural fiber composites Reference Composite Masoodi and Woven jute fiber/epoxy Pillai [27] composite Apolinario Unidirectional flax fiber/ et al. [230] polyester composite Assarar et al. Unidirectional flax fiber/ [231] epoxy composite Apparent fiber diameter (lm) Manufacturing method Observation 103a Compression molding Infusion molding Hand lay-up 26% increase in the water absorption of the composite by incorporating 40% by weight jute fiber 9% decrease in the water uptake of 1% silane-treated fiber composite under saturation conditions Water aging considerably decreased the elastic modulus (i.e., 40%) and increased the elongation (i.e., 61%) of the composite compared to the glass fiber composite Composites made with non-dry fibers (with moisture content of 10%) exhibited lower moisture absorption, degree of swelling, and shrinking than those made with dry fibers (dried in an oven at 80 °C) The use of fibers with length of 1–1.5 mm and diameter of 350 lm significantly reduced the specific wear rate (i.e., by about three to four times) of the neat polyester resin Longer fibers (* 5 mm) had higher wear resistance because less pulling-out of fibers occurs at that length. When the ends of the fibers in the matrix were perpendicular to the counterface and the anti-sliding direction (i.e., anti-parallel orientation), the composites exhibited higher wear resistance as compared to other directions Coir fiber composite had higher wear resistance than sugarcane and oil palm fiber composites Presence of kenaf fibers oriented in the normal direction enhanced the wear performance of the resin by * 85% 15 – Lu and Vuure [232] Unidirectional flax fiber/ polyester composite 10 Vacuumassisted resin infusion Yousif and El-Tayeb [233] El-Tayeb [234] Random-oriented oil palm fiber/polyester composite Random-oriented sugarcane fiber/ polyester composite 350 Hand lay-up 1200–2700a Hand lay-up Yousif [235] Coir fiber/polyester composite Unidirectional kenaf fiber/epoxy composite 215–304a Compression molding Vacuum bagging Random-oriented palm leaf fiber/polypropylene composite – Injection molding Singh and Gupta [238] Unidirectional jute fiber/ phenolic composite – Pultrusion technique Dash et al. [239] Joseph et al. [240] Unidirectional jute fiber/ polyester composite Random-oriented banana fiber/phenolic composite – Compression molding Compression molding Chin and Yousif [236] Abu-Sharkh and Hamid [237] a 250–400a – Reinforcement of polypropylene composite with palm leaf fiber increased the weathering stability of the composite as compared to composite without fiber, owing to the natural antioxidant properties of lignin, which increased the protection of the composite from UV radiation Exposing the composite without any treatment to weathering led to color fading, fibrillation, black spots, bulging, and resin cracking, as well as a 22% reduction in the flexural strength of the composite after 2 years Bleaching reduced the effects of weathering Alkalization and silane fiber modifications slowed the effects of weathering Diameter of a fiber bundle Shrinkage Rearranging and reorienting of resin molecules of composites in the liquid phase result in shrinkage [1]. The shrinkage of natural fibers generates internal stresses at the fiber/matrix interface and causes damage and degradation of the initial properties of the composite. The influential parameters on the 866 Other properties Thermal properties Fire resistance The flammability behavior of natural fiber composites has become one of the most important issues in recent years because of the use of natural fiber composites in different applications [1]. One of the weaknesses of natural fiber composites is the loss of strength and stiffness at high temperatures [257]. Thermal decomposition and combustion occur when natural fiber composites are exposed to fire or any other high-intensity heat sources [4]. The flammability of natural fiber composites is variable for different types of natural fibers because they have different chemical compositions and microstructures. By decreasing the cellulose content of natural fiber, increasing the crystallinity, and decreasing the polymerization of the composites, the fire resistance of the 3.5 Pure 5%, Untreated 5%, Treated 10%, Untreated 10%, Treated Shrinkage (%) 3 2.5 2 1.5 1 0.5 0 0 30 60 90 120 150 180 150 180 Time (min) (a) 6 Pure 5%, Untreated 5%, Treated 10%, Untreated 10%, Treated 5 Shrinkage (%) volumetric shrinkage of composites are the density, processing conditions, flow pattern, shape, residual internal stress, fiber content, crystallinity, and nonuniform cooling rates [4]. The use of fibers in the longitudinal direction, an increase in the fiber content, and a decrease in the residual internal stress, non-uniform cooling rates, and the crystallinity of the composites all decrease the shrinkage of natural fiber composites [252–254]. Figure 29 shows the shrinkage behavior of a random-oriented coir fiber/PLA composite manufactured by compression molding technique over time at 80 °C for the longitudinal and transverse directions of the composite, in which fibers were subjected to plasma treatment with 0%, 5%, and 10% fiber content by weight in the composite [173]. A limited number of studies investigated the shrinkage behavior of natural fiber composites. Santos et al. [255] reported that the shrinkage of a random-oriented bamboo fiber/HDPE composite manufactured by injection molding technique was reduced up to 58% by an increased fiber content in the composite. Tan et al. [256] assessed the shrinkage behavior of random-oriented coir fiber/polypropylene composites manufactured by injection molding technique and found a 13% reduction in the volumetric shrinkage of the composite with a 40% increase in the fiber content by weight. J Mater Sci (2020) 55:829–892 4 3 2 1 0 0 30 60 90 120 Time (min) (b) Figure 29 Shrinkage behavior of random-oriented untreated and plasma-treated-coir fiber/polylactic acid composite manufactured by compression molding technique at 80 °C over time for a longitudinal and b transverse directions of the composite. natural fiber composites can be improved [257]. The use of coatings and additives, such as ceramic, intumescent, silicone, phenolic, ablative, and glass mats, also improves the fire resistance of natural fiber composites. Heating intumescent materials beyond a specific temperature results in the formation of the cellular and charred surface by initiation of foaming and expanding, which in turn can protect the underlying surface from heat and flame. By adding talc and nanoparticles as fillers to the natural fiber composites, the fire resistance of the composites increases by the formation of heat barriers. Several researchers have studied the behavior of natural fiber composites at high temperatures. Some studies reported that natural fibers generally start to degrade at approximately 240 °C [31, 258–260]. However, lignin starts to degrade at approximately 200 °C and hemicellulose and cellulose start to degrade at higher temperatures [261]. Table 18 presents the decomposition temperature of major natural fibers, and 867 Table 18 Decomposition temperature of major natural fibers [259, 262] Fiber Decomposition temperature (°C) Bagasse Bamboo Cotton Hemp Jute Kenaf Rice straw Rice husk Pina 232 224 232 215 215 229 238 234 240 Table 19 presents glass transition and melting point of major resins used in natural fiber composites. Thermal conductivity The use of heat-insulating materials is an appropriate approach for reducing energy costs and improving manufacturing efficiency in industries such as automotive, construction, and packaging [1]. Natural fibers contain lumens, hollow portions filled with air. As a result, the thermal conductivity of composites decreases with an increased fiber content. By an increased natural fiber volume fraction of the composites, the amount of air contained in the composite increases, which results in heat insulation. Several studies have been performed to investigate the thermal conductivity of natural fiber composites. Pujari et al. [264] studied the thermal conductivity of random-oriented jute and banana fibers/epoxy Table 19 Glass transition temperature and melting point of major resins [243, 263] Thermal conductivity (w/m°k) J Mater Sci (2020) 55:829–892 0.4 0.363 0.3 0.251 0.243 0.239 0.230 0.231 0.228 0.2 0.1 0 Epoxy resin Jute 30% Banana 30% Jute 40% Banana 40% Jute 60% Banana 60% Figure 30 Thermal conductivity of epoxy resin, and jute and banana fiber/epoxy composites with different fiber contents. composites manufactured by hand lay-up method and reported that the thermal conductivities of the jute and banana fiber composites were 0.231 and 0.228 W/m K in the - 20 °C to 300 °C range and at the maximum fiber volume fraction, respectively. Figure 30 shows the thermal conductivity of natural fiber composites with different fiber contents [264]. Paul et al. [204] reported that an increase in fiber loading decreased the thermal conductivity of a random-oriented banana fiber/polypropylene composite manufactured by compression molding technique. The thermal conductivity of a randomoriented flax fiber/HDPE composite manufactured by extrusion molding technique was assessed by Li et al. [265]. They found that the thermal conductivity of the composite reduced with increasing fiber content. Mangal et al. [266], Osugi et al. [267], Mounica et al. [268], and Ramanaiah et al. [269] found the same result for random-oriented pina fiber/phenolic manufactured by compression molding technique, unidirectional hemp fiber/epoxy manufactured by Resin Glass transition temperature (°C) Melting point (°C) Polypropylene Low-density Polyethylene High-density Polyethylene Polystyrene Nylon 6 Nylon 6,6 Polyester Vinylester Epoxy Phenolic Starch PLA PHA 0.9–1.55 - 120 - 80 100–135 40 50 60 104–121 70–167 170 60 58 2–15 160–176 105–116 120–140 – 222 250–269 250–300 – – – 110–115 150–162 160–175 868 J Mater Sci (2020) 55:829–892 Thermal analysis Polylactic acid Sisal fiber 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite (a) Polypropylene Sisal fiber 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite (b) Figure 31 TGA curves of sisal fiber composites with a polylactic acid and b polypropylene resin [272]. compression molding technique, unidirectional bamboo fiber/polyester manufactured by hand layup method, and unidirectional fish tail palm fiber/ polyester composites manufactured by hand lay-up method, respectively. Agarwal et al. [270] assessed the influence of fiber treatment on the thermal conductivity of an oil palm fiber/phenolic composite. They found that alkali and silane treatments increased the thermal conductivity of the composite more than acetylation treatment, meaning that acetylation treatment was more suitable than alkali and silane treatments for improving the heat-insulating characteristics of the composite. Thermogravimetry analysis (TGA), differential scanning calorimetry (DSC), and differential thermal analysis (DTA) were used to analyze the thermal behavior of natural fiber composites. Feng et al. [271] assessed the thermal behavior of random-oriented kenaf fiber/polypropylene composite manufactured by injection molding technique through DSC and reported that maleated coupling agent reduced the melting temperature of the composite. Joseph et al. [87] used TGA and DSC to study the thermal behavior of random-oriented sisal fiber/polypropylene composite. They found that incorporation of sisal fiber with fiber weight fraction of 20% treated by maleic anhydride-modified polypropylene, urethane derivative of polypropylene glycol, and KMnO4, respectively, caused 12%, 11%, and 9% increase in the crystallization temperature of the composite compared to that with no fiber, which was because the surface of the treated fibers acted as nucleating sites for the crystallization of the polymer. Mofokeng et al. [272] assessed the thermal stability of random-oriented sisal fiber/PLA and sisal fiber/polypropylene composites manufactured by injection molding technique through TGA and their crystallization through DSC. Figure 31 shows the TGA curves of composites with fiber weight content of 1, 2, and 3%. They found that the thermal stability of both polymers increased with an increased fiber content, with a more significant improvement in the case of polypropylene. Figure 32 shows the DSC curves of composites. The DSC results in the figure illustrated a significant influence of the fibers on the cold crystallization and melting behavior of PLA. They also reported that fiber content had a negligible influence on the melting characteristics of the polypropylene, but it had a significant influence on the non-isothermal crystallization temperature range. In 2013, Azwa and Yousif [273] studied the thermal stability of random-oriented untreated and 6% alkaline-treated kenaf fiber/ epoxy composite. Figure 33 shows TGA and DTA curves presented in that study [273]. They reported that the addition of the kenaf fiber into the epoxy slightly improved both the charring and the thermal stability of the composite. They also reported that untreated kenaf fiber/epoxy composite started to lose weight earlier compared to the other composites, which was because of the higher moisture content of the untreated fibers. 869 J Mater Sci (2020) 55:829–892 Polylactic acid 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite Viscoelastic properties Melting temperature Glass transition temperature (a) Polypropylene 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite Melting temperature (b) Crystallization temperature Polypropylene 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite (c) Figure 32 DSC curves of sisal fiber composites with a polylactic acid (heating curves), b polypropylene (heating curves), and c polypropylene (cooling curves) resin [272]. Polymers are normally viscoelastic fluids that behave as a viscous or elastic material depending on the rate of flow or deformation [274]. Dynamic mechanical analysis has been widely utilized to investigate the viscoelastic properties (storage modulus, loss modulus, and damping factor) of composites [11]. Dynamic mechanical analysis is a technique that a small deformation is cyclically applied to a sample under different temperatures, and the elastic and viscous response of the sample is then monitored. In this technique, a sinusoidal force is applied to the sample at a set frequency and the stiffness and damping of the sample are measured [275]. The in-phase component, out-of-phase component, and the ratio of the loss of the storage are defined as storage modulus, loss modulus, and damping factor, respectively. Pothan et al. [275] investigated the influence of the fiber content on the dynamic mechanical properties of random-oriented banana fiber/polyester composite manufactured by compression molding technique. They showed that an increase in the fiber volume fraction up to 40% led to a decrease in the storage modulus of the composite below the glass transition temperature, but an increase in the storage modulus above the glass transition temperature. They also reported that, at 40% fiber content, the peak loss modulus got broadened and peak damping factor lowered, indicating the improved fiber/matrix adhesion. Idicula et al. [276] assessed the viscoelastic properties of random-oriented banana/sisal hybrid fiber/polyester composite manufactured by compression molding technique and reported that the storage modulus of the composite enhanced above the glass transition temperature of the polymer with an increase in the fiber volume fraction up to 40% and then decreased. Pan et al. [277] studied the storage modulus of random-oriented kenaf fiber/PLA composite manufactured by melt mixing technique and reported that the storage modulus increased by 28% by reinforcement of the composite with 30% kenaf fiber by weight. Mofokeng et al. [272] investigated the viscoelastic properties of random-oriented sisal fiber/PLA and sisal fiber/polypropylene composites manufactured by injection molding technique through dynamic mechanical analysis. Figures 34 and 35 show the storage modulus (E0 ), loss modulus (E00 ), and damping factor (tand) curves of composites with PLA and polypropylene, respectively [272]. 870 J Mater Sci (2020) 55:829–892 Figure 33 a TGA and b DTA curves of random-oriented untreated and 6% alkalinetreated kenaf fiber/epoxy composites [273]. They reported that the storage and loss modulus of the PLA decreased with an increase in the fiber content (shown in Fig. 34a, b). They also reported that the cold crystallization of the PLA was around 110 °C and the existence of the fibers in the composite led to an increase in the modulus between the cold crystallization and melting of the polymer (shown in Fig. 34b). The larger area under the a-relaxation peak in Fig. 34c indicates that the molecular chains exhibited a higher degree of mobility, therefore better damping properties. The presence of fibers increased the modulus and glass transition temperature of polypropylene composite (shown in Fig. 35a, b), which was because of the restriction in the segmental motion. The existence of fibers had no impact on the a-relaxation of the composite (shown in Fig. 35c), which was related to the crystalline fraction of the polypropylene. Mylsamy and Rajendran [278] performed dynamic mechanical analysis on untreated and alkaline-treated unidirectional agave fiber/ epoxy composite manufactured by compression molding technique and reported that the composite with poor interfacial bonding tended to dissipate 871 J Mater Sci (2020) 55:829–892 more energy compared to those with good interfacial bonding. Glass transition temperature Morphology α-relaxation region Cold crystallization region Polylactic acid 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite (a) Glass transition temperature Polylactic acid 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite (b) Polylactic acid 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite α-relaxation peak Cold crystallization region (c) Figure 34 Dynamic mechanical analysis of random-oriented sisal fiber/polylactic acid composite manufactured by injection molding technique: a storage modulus, b loss modulus, c damping factor [272]. Optical microscopy, SEM, and atomic force microscopy (AFM) are three main methods for morphological studies of composites. Optical microscopy is ideal for general inspection purposes, and SEM provides detailed topographical and compositional information of the composite by high magnification view at micro-scale [277]. On the other hand, AFM is a useful method to determine the surface roughness of fibers at nanoscale by measuring the interatomic forces between the sample surface and the measurement tip [10]. A large number of studies investigated the morphology of natural fiber composites (e.g., [275, 277, 279]). Some of these studies are presented in this review paper. Pothan et al. [275] assessed the influence of the fiber content on the morphology of random-oriented banana fiber/polyester composite manufactured by compression molding technique through SEM. Figure 36a–c shows SEM micrographs of composites with fiber volume fractions of 10, 20, and 40%, respectively. They reported that there was a good fiber/matrix bonding in the composite with 40% fiber content, whereas fiber/matrix debonding was evident in composites with 10% and 20% fiber content. They also showed that there was no gap between the fibers and matrix in 40% fiber composite owing to the strong bonding. However, when the fiber concentration was lower, the packing of the fibers was not efficient, leading to an easier failure of the bonding at the interfacial region. Pan et al. [277] studied the morphology of randomoriented kenaf fiber/PLA composite manufactured by melt mixing technique using SEM. Figure 37 shows the SEM micrographs of kenaf fiber [277]. It can be seen in the figure that the diameter of the fibers ranged from 5 to 50 lm and their surface was rough with the adhesion of some smaller fibers. This was attributed to different sources and processing history of the kenaf fibers. Figure 38 shows the SEM micrographs of pure PLA and composite with 30% kenaf fiber by weight [277]. The fractured surface of the pure PLA was irregular and rough because of the brittle behavior of the resin. It is also shown in the figure that most fibers (indicated by ‘‘A’’) were tight connected to the matrix, and a lot of fibers were 872 J Mater Sci (2020) 55:829–892 Polypropylene 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite (a) Glass transition temperature Polypropylene 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite (b) Polypropylene 1% Fiber/Composite 2% Fiber/Composite 3% Fiber/Composite α-relaxation Glass transition temperature (c) Figure 35 Dynamic mechanical analysis of random-oriented sisal fiber/polypropylene composite manufactured by injection molding technique: a storage modulus, b loss modulus, c damping factor [272]. broken during the tensile testing. It can be seen in Fig. 38 that there are some fibers in the composite (indicated by ‘‘B’’) separated from the matrix during the tensile deformation, indicating that the interface interaction between the matrix and fibers further improved. Aggregation of kenaf fibers (indicated by ‘‘C’’) and the presence of a small amount of air bubbles (indicated by ‘‘D’’) are also visible in the figure, which was because of the rough surface of kenaf fiber and high viscosity of the PLA. Hebel et al. [279] investigated the morphology of unidirectional bamboo fiber/epoxy composite manufactured by compression molding technique produced by a combination of heat and pressure treatment through optical microscopy. Figure 39a–c shows the optical micrographs of the fractured surface of the autoclave/compression molded composites produced at 100 °C with pressures of 15, 20, and 25 MPa [279]. It should be noted that the composite produced by pressure of 20 MPa exhibited a higher tensile strength than that by 15 and 25 MPa. They reported that the surface of the composite at 20 MPa pressure was rather smooth and homogenously covered by the resin. At this pressure, the polymer network formed a very thin layer that efficiently interacted with the surface of the fiber and penetrated into the fibers due to its viscosity. They also reported that reducing the pressure to 15 MPa resulted in the formation of large resin beads on the fiber surface, which led to the reduction in the wetting and homogenous coverage of the fibers. At high pressures of 25 MPa, the epoxy formed larger crystal-like beads than those at 15 MPa and infiltration inhibited by a propagated carbonization of the fibers (as shown by the darker color strip). Lee et al. [280] used AFM to characterize the kraft fiber treated by Trichoderma reesei enzyme. Figure 40 shows tapping mode AFM images of treated kraft fiber and its cellulose aggregate fibrils. The cellulose microfibril bundles are evident in Fig. 40a, which can be a potential source for cellulose aggregate fibrils. Several structural imperfections such as kinks and twists are evident in Fig. 40b that were because of the conformability of fibers in the treatment process by the removal of the constraining primary layer of the kraft fiber. George et al. [281] used AFM to investigate the influence of the fiber treatment on the nanostructure of the hemp fiber. They reported that fiber treatment rendered the surface topography of the hemp fiber clean and exposed 873 J Mater Sci (2020) 55:829–892 Figure 36 SEM micrographs of random-oriented hemp fiber/polyester composite with fiber content of a 10%, b 20%, and c 40% manufactured by compression molding technique [275]. Fiber/matrix debonding Fiber pull-out Fiber/matrix debonding Matrix cracking (a) (b) (c) Figure 37 SEM micrographs of kenaf fiber with magnification of a 9 100 and b 9 500 in random-oriented kenaf fiber/polylactic acid composite manufactured by melt mixing technique [277]. 50μm 100μm (a) the individual fiber bundles. They also showed that hemp fibers treated by 10% NaOH, xylanase enzyme, and laccase enzyme had 250%, 218%, and 16% higher surface adhesion forces than those with no treatment. Spectroscopic characterization Fourier transform infrared spectroscopy (FTIR) is an effective technique to determine the chemical composition of the composite and functional groups (b) interacting within natural fibers, characterizing their covalent bonding information [26]. Table 20 presents the FTIR analysis peaks for different untreated natural fibers. Lu et al. [293] investigated the spectroscopic characteristics of the benzylated-treated sisal fiber by FTIR and reported that peaks at 1250 cm-1 and 1363 cm-1 related to lignin and hemicellulose partly removed by treatment of the sisal fiber. Mofokeng et al. [272] used FTIR analysis to determine the existence and type of interfacial interaction in 874 J Mater Sci (2020) 55:829–892 Figure 38 SEM micrographs of fractured surface of a pure polylactic acid, and randomoriented kenaf fiber/polylactic acid composite with magnification of b 9 200, c 9 500, and d 9 500 manufactured by melt mixing technique [277]. D C 50μm (a) 100μm (b) A B 50μm (c) random-oriented sisal fiber/PLA composite manufactured by injection molding technique (shown in Fig. 41). Peaks at 1780–1680 cm-1, 3600–3000 cm-1, and 1180 cm-1 corresponded to C=O, O–H, and C– O–C stretches, respectively. They reported that the O–H band of the composite became more pronounced and broader with an increase in the fiber content, which was because of the existence of free OH groups in the fiber. They also reported that a new peak at 1650 cm-1 was created corresponded to OH, which was oriented from bending of the unresolved OH group of the absorbed water by cellulose. Alavudeen et al. [294] conducted FTIR analysis on the alkaline- and sodium lauryl sulfate (SLS)-treated random-oriented banana fiber/epoxy composite manufactured by hand lay-up method. They reported that fiber treatment removed most of the lignin and hemicellulose components on the fibers, which helped to improve the mechanical properties of the composite. They also found that an increase in the concentration of alkaline and SLS beyond 10% 50μm (d) damaged the fiber surface, leading to the poor fiber adhesion and worse mechanical properties of the composite. Faruk Hossen et al. [295] by FTIR analysis on the propionic anhydride-treated random-oriented jute fiber/polyethylene composite manufactured by compression molding technique reported that the band at 1641 cm-1 indicated absorbed water in crystalline cellulose and the band disappeared upon the fiber treatment. They also reported that the band at 1512 cm-1 was because of the existence of the aromatic rings in lignin. Incorporation of nanoparticles in natural fiber composites Nanoparticles have been utilized as nanofiller materials to improve the properties of natural fiber composites [296]. Owing to the high surface area (higher than 500 m2/g) and pore filling properties of nanoparticles, the properties of natural fiber composites were enhanced by addition of a small amount J Mater Sci (2020) 55:829–892 875 Figure 39 Optical micrographs of the fracture surface of autoclave/compression molded unidirectional bamboo fiber/ epoxy composite produced at 100 °C with pressure of a 15 MPa, b 20 MPa, and c 25 MPa [279]. (up to 5% by weight) of nanofillers in natural fiber composites [297]. A few studies investigated the properties of natural fiber composites containing nanofillers. Table 21 presents a summary of the existing studies of the influence of nanofillers on the properties of natural fiber composites [298–302]. Application of natural fiber composites Natural fiber composites have been utilized in the automotive, construction, aircraft components, packaging, sporting equipment, electrical parts, and biomedical industries owing to the eco-friendliness, lightweight, good mechanical behavior, sound attenuation, and vibration damping of them [15, 17, 243, 303]. However, their low durability may exclude them from being an alternative to glass fiberreinforced composites in wet areas such as in piping, boats, and kayaks [243]. Henry Ford made the first attempt to use natural fibers in the automotive industry using flax and hemp fibers in 1941 [43]. Many automotive companies are now using natural fiber composites made with polyester, polypropylene, flax, hemp, and sisal [304]. In structural applications, natural fiber composites have been utilized to 876 J Mater Sci (2020) 55:829–892 Figure 40 Tapping mode AFM images of a kraft fiber treated by Trichoderma reesei enzyme and b its cellulose aggregate fibrils [280]. develop load-bearing elements, such as beam, roof, wall, and pedestrian bridge [305, 306]. Jute fiber/ polypropylene composite has been used in primary structural applications, such as indoor elements in housing [1, 28], temporary low-cost outdoor housing for defense and rehabilitation [29, 30], and transportation [307]. Yousif and Ku [308] investigated the behavior of a liquid storage tank made with coir fiber/polyester composite and reported that the interfacial adhesion strength between coir fiber and polyester resin soaked in water and diesel oil was 107 MPa. They suggested the potential application of the coir fiber/polyester composite in design and manufacture of the body of liquid storage tanks. Natural fiber composites have also been used in functional applications of protective medical apparels, such as baby diapers, feminine hygiene products, and adult incontinence items, plasters, bike frames, golf stick, tennis rackets, and helmets [309]. According to the reports by MarketsandMarkets Inc. [310] and Grand View Research Inc. [311], the natural fiber composite market was valued at $16.5 billion in 2016 and is projected to reach $46.3 billion by 2025, because of the governmental regulations for the use of environmentally friendly products in industry. The global natural fiber composite market share by the end-use industries in 2016 was 35%, 27%, 9%, 13%, and 16% for transportation, building and construction, consumer goods, electrical and electronics, and other applications, respectively [311]. Based on this report and recent reports by Dicker et al. [243] and Gurunathan et al. [3], the automotive and construction industries were the biggest market for natural fiber composites in 2016–2017 to manufacture seat backs, headliners, truck liners, door panels, decking, railing, window frames, dash boards, and frames. Because of the rapid growth of the use of natural fiber composites in the automotive market, high-density glass fibers are being replaced with low-density natural fiber composites; lightweight vehicles reduce the amount of the fuel required, which in turn leads to a reduction of CO2 emission [3]. The study by Yang et al. [312] revealed that a 25% reduction in the weight of an automobile would result in a 220 billion pound reduction in CO2 emission per annum. Germany is a leader in the use of natural fiber composites in automotive industry. The automotive companies in Germany, such as 877 J Mater Sci (2020) 55:829–892 Table 20 FTIR peaks of different natural fibers (in cm-1) Reference Fiber OH stretching C–H stretching C=O CH2 stretching symmetric bending C–O stretching C-vibration of acetyl group in lignin and hemicellulose bBending vibration of C– glycosidic linkages H and C–O groups of aromatic ring in polysaccharides C–OH bending Mwaikambo and Ansell [207] Keshk et al. [282] Bilba et al. [283] Spinace et al. [284] Saha et al. [285] Sawpan et al. [286] De Rosa et al. [287] Fiore et al. [288] Arrakhiz et al. [289] Reddy et al. [290] Neto et al. [291] Al–Maadeed et al. [292] Sisal 3447 – 1737 1384 – – – – Kenaf 3397 2902–2918 – 1416 – 1036 – – Banana 3600–3100 2910 1416 – 1036 – – Curaua 3200–3500 2750–2900 1740 – 1240 – – – Jute 3200–3600 2910 1739 1430 1294 – – – Hemp 3410 1732 1425 1247 – 892 – Okra 3100–3600 2854–2925 1734 1430 1243–1384 1320–1370 894 598 2916 – Artichoke 3342 2854–2923 1737 1422 – 1318–1372 892 589 Coir 3330 2910 1420 1235 – – – Borassus 3435 2851–2920 1746 – 1254 1058 893 – Pina 3450 2900 1730 – 1260 1050 – – Palm leaf 3300 2946 1735 – – – – – 1725 Mercedes, BMW, Audi, and Volkswagen, have taken the initiative to use natural fiber composite for interior and exterior applications because natural fiber composite components are lightweight, impact resistant, nontoxic, and no-sharp edged fracture in the case of a crash. Asia Pacific (especially China, India, South Korea, Thailand, and Indonesia) is expected to have the fastest-growing market for natural fiber composites during the next 5 years owing to the increasing demand from the transportation, building, and construction [310]. The European Union’s ban on nonrecyclable plastic is expected to drive the market in the Europe region. Based on the report by Global Biocomposite Market in 2018 [313], some of the current leading companies in the global natural fiber composite market are FlexForm Technologies (USA), Tecnaro GmbH (Germany), Trex Company Inc. (USA), Fiberon LLC (USA), Meshlin Composites ZRT (Hungary), UPM (Finland), Jelu-Werk J. Ehrler GmbH & Co. KG (Germany), Green Bay Decking (USA), Universal Forest Products Inc. (USA), Toray Industries Inc. (Japan), and Owens Corning (USA). Table 22 summarizes the applications of natural fiber composites based on different properties. 878 J Mater Sci (2020) 55:829–892 Life-cycle assessment of natural fiber composites Figure 41 FTIR spectra of random-oriented sisal fiber/polylactic acid composite manufactured by injection molding technique [272]. Figure 42 shows the life cycle of a natural fiber composite [19]. The fully environmental superiority of natural fiber composites than that of synthetic composites is still questionable due to the relative excessive processing requirements of natural fiber composites. Therefore, careful life-cycle assessment (LCA) is necessary before natural fiber composite is used commercially [4]. LCA is part of ISO14000 series of Environmental Management System (EMS) and determines the overall potential impact of natural fiber composites on the environment throughout the whole life cycle from extraction of raw materials to end-of-life stage. A number of LCA studies have been conducted to date on natural fiber composites. The first LCA study on natural fiber composites was performed by Wotzel et al. [314] on hemp fiber/ epoxy composite and acrylonitrile butadiene styrene (ABS) for production of automotive side-panel component of passenger cars through cradle-to-grave lifecycle approach. They reported that the cumulative Table 21 Summary of existing studies on the effect of nanofiller on the properties of natural fiber composites Reference Composite Apparent fiber diameter (lm) Manufacturing method Observation Huang and Netravali [298] Unidirectional flax fiber/soy protein composite 237a Resin transfer molding Sen and Kumar [299] Random-oriented coir fiber/epoxy composite – Compression molding 13%, 4%, 8%, 24%, and 17% increase in tensile strength, tensile strain, tensile modulus, flexural strength, and flexural modulus of composite with fiber volume fraction of 48% incorporating 5% nanoclay, respectively The smoke density and limiting oxygen index (as fire retardancy properties), respectively, 28% decreased and 26% increased when the composite contained 1.6% coir nanofiller 19%, 27%, and 34% decrease in water absorption incorporating 1%, 3%, and 5% nanoclay (by weight), respectively. 7%, 3%, and 23% enhancement in the flexural strength, modulus, and toughness incorporating 5% nanoclay, respectively 163% increase in flexural strength incorporating 3% nanoclay. 180% increase in flexural modulus incorporating 5% nanoclay 60%, 17%, and 12% increase in tensile strength, flexural strength, and damage index of the composite incorporating 5% nanosilica, respectively Alamri and Random-oriented recycled 200a Low [300] cellulose fiber/epoxy composite Lim et al. [301] Ashik et al. [302] a Random-oriented Napier grass fiber/epoxy composite Cross-plied jute fiber/ epoxy composite Diameter of a fiber bundle Compression molding – Resin infusion – Compression molding 879 J Mater Sci (2020) 55:829–892 Table 22 Application areas of natural fiber composites based on their properties [13, 243] Properties Suitable application High strength and low weight Renewability Transportation (such as aircraft, automobile), mobile electronics, sport equipment, construction and building, storage devices (such as grain storage silos, bio-gas containers) Short-life-span products such as plastic cutlery, packaging, toys, and electronic products of computers, phones, faxes, radios, stereos, CD players Toys, hobbyist-built items, consumer-handled items Medical devices and implants Competitive consumer products such as window and door frames, furniture, railroad sleepers, automotive panels, gardening items, shelves, noise insulting panels, and packaging Dry-use products Short-life-span products, limited exposure to harsh environments Non-toxicity Biocompatible Low cost High water absorption Weak durability energy demand of hemp fiber/epoxy composite was 45% lower than that of ABS. Table 23 presents the summary of the LCA studies on natural fiber composites. The existing LCA studies revealed that natural fiber composites have superior environmental performance than composites with synthetic fiber or resin. However, cradle-to-gate LCA studies of flax, as an agrochemical-based material, and glass by Dissanayake et al. [321, 322] reported that flax sliver is comparable in energy terms to glass fiber mat, and continuous glass fiber reinforcement appears to be Figure 42 Life cycle of a typical natural fiber-reinforced composite [19]. superior from an environmental energy point of view to spun flax yarn. Conclusions The properties of easy accessibility, low density, low price, good thermal and acoustic insulation, ecofriendliness, recyclability, renewability, and satisfactory mechanical properties make natural fiber composites an attractive alternative to carbon, glass, and other synthetic fiber composites. Natural fiber composites are currently being used in the construction 880 J Mater Sci (2020) 55:829–892 Table 23 Summary of some of the studies on life-cycle assessment (LCA) of natural fiber composites References Composite Application Life-cycle approach Results Side-panel component for passenger car (Audi A3) Bus body component Cradle-to-grave Cumulative energy demand of hemp fiber/ epoxy composite was 45% lower than that of ABS Cradle-to-grave Composite with PTP had 65% and 39% lower energy demand and global warming potential than those with polyester, respectively Curaua fiber/polypropylene composite had 75% and 66% lower climate change and ozone depletion potential than those of glass fiber/polypropylene, respectively Hemp fiber/PTP composite had 54% and 44% lower global warming and climate change potential than those of glass fiber/polyester, respectively Sugarcane bagasse fiber/polypropylene had 20% and 25% lower energy demand and global warming impact than those of talc/ polypropylene, respectively Kenaf fiber/soy-based resin composite had 58%, 13%, 97%, and 22% lower global warming, ozone depletion, hedgehog cancer, and ecotoxicity potential than those of glass fiber/polyester composite, respectively Hybrid glass-hemp fiber/epoxy composite had 18%, 19%, 20%, and 20% lower global energy requirement, global warming potential, ecotoxicity, and life cycle cost than those of glass fiber/epoxy composite, respectively Wotzel et al. [314] Hemp fiber/epoxy versus ABS Mussig et al. [315] Hemp fiber/Triglycerides and polycarbon acid anhydrides (PTP) versus hemp fiber/ polyester Curaua fiber/polypropylene versus glass fiber/ polypropylene Automotive internal component Cradle-to-grave Schmehl et al. [317] Hemp fiber/PTP versus glass fiber/polyester Bus body component Cradle-to-grave Luz et al. [318] Sugarcane bagasse fiber/ polypropylene versus talc/ polypropylene Cradle-to-grave Wang et al. [319] Kenaf fiber/soy-based resin versus glass fiber/polyester Vehicle interior aesthetic covering component Sheet molding compound material LaRosa et al. [320] Hybrid glass-hemp fiber/epoxy versus glass fiber/epoxy Zah et al. [316] Cradle-to-gate Elbow fittings for Cradle-to-grave the seawater cooling pipeline and building, packaging, transportation, military, sporting equipment, and medical industries in an effort to reduce greenhouse gas emissions, and it is anticipated that this type of composite will be extensively used in other applications in the near future. However, their drawbacks, such as poor interfacial adhesion between the natural fibers and the matrix, moisture absorption, poor fire resistance, low impact strength, and low durability, have at least partly prevented them from being considered a reliable alternative for conventional composites. This paper has presented an up-to-date and all-encompassing review on the properties of natural fiber composites. It has been shown that the use of modification techniques can lead to improved physical, mechanical, durability-related, and time-dependent properties of natural fiber composites, which points to the feasibility of using these composites as an alternative material in a wide range of applications. The review presented here will help researchers to better understand various properties of natural fiber composites to facilitate the development of new green materials with improved performance. Based on this review, further investigations are recommended to investigate the influence of matrix modification and fiber length on the properties of natural fiber composites. Compliance with ethical standards Conflict of interest Both authors declare that they have no conflict of interest. 881 J Mater Sci (2020) 55:829–892 References [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] Campilho RDSG (2015) Natural fiber composites. CRC Press, Boca Raton Alix S, Colasse L, Morvan C, Lebrun L, Marais S (2014) Pressure impact of autoclave treatment on water sorption and pectin composition of flax cellulosic-fibres. Carbohydr Polym 102:21–29 Gurunathan T, Mohanty S, Nayak SK (2015) A review of the recent developments in biocomposites based on natural fibres and their application perspectives. Compos Part A Appl Sci Manuf 77:1–25 Faruk O, Bledzki AK, Fink H-P, Sain M (2012) Biocomposites reinforced with natural fibers: 2000–2010. Prog Poly Sci 37(11):1552–1596 Madsen B, Gamstedt EK (2013) Wood versus plant fibers: similarities and differences in composite applications. Adv Mater Sci Eng 2013:1–14 Rana AK, Mandal A, Bandyopadhyay S (2003) Short jute fiber reinforced polypropylene composites: effect of compatibiliser, impact modifier and fiber loading. Compos Sci Technol 63(6):801–806 Jawaid M, Khalil HPSA (2011) Cellulosic/synthetic fibre reinforced polymer hybrid composites: a review. Carbohydr Polym 86(1):1–18 Stamboulis A, Baillie CA, Peijs T (2001) Effects of environmental conditions on mechanical and physical properties of flax fibers. Compos Part A Appl Sci Manuf 32(8):1105–1115 Summerscales J (2018) Bast fibres and their composites. University of Plymouth, UK, https://www.fose1.plymou th.ac.uk/sme/mats347/bast_book.htm, Accessed 10 Aug 2019 George J, Sreekala MS, Thomas S (2001) A review on interface modification and characterization of natural fiber reinforced plastic composites. Polym Eng Sci 41(9):1471–1485 John MJ, Thomas S (2008) Biofibres and biocomposites. Carbohydr Polym 71(3):343–364 Zini E, Scandola M (2011) Green composites: an overview. Polym Compos 32(12):1905–1915 La Mantia FP, Morreale M (2011) Green composites: a brief review. Compos Part A Appl Sci Manuf 42(6):579–588 Ku H, Wang H, Pattarachaiyakoop N, Trada M (2011) A review on the tensile properties of natural fiber reinforced polymer composites. Compos Part B Eng 42(4):856–873 Dittenber DB, GangaRao HVS (2012) Critical review of recent publications on use of natural composites in infrastructure. Compos Part A Appl Sci Manuf 43(8):1419–1429 [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] Khalil HPSA, Bhat AH, Yusra AFI (2012) Green composites from sustainable cellulose nanofibrils: a review. Carbohydr Polym 87(2):963–979 Koronis G, Silva A, Fontul M (2013) Green composites: a review of adequate materials for automotive applications. Compos Part B Eng 44(1):120–127 Nguong CW, Lee SNB, Debnath S (2013) A review on natural fibre reinforced polymer composites. World Acad Sci Eng Technol 7:1123–1130 Faruk O, Bledzki AK, Fink HP, Sain M (2014) Progress report on natural fiber reinforced composites. Macromol Mater Eng 299(1):9–26 Satyanarayana KG (2015) Recent developments in ‘green’ composites based on plant fibers-preparation, structure property studies. J Bioprocess Biotech 5(206):1–12 Mohammed L, Ansari MN, Pua G, Jawaid M, Islam MS (2015) A review on natural fiber reinforced polymer composite and its applications. Int J Polym Sci 2015:1–15 Pickering KL, Efendy MGA, Le TM (2016) A review of recent developments in natural fibre composites and their mechanical performance. Compos Part A Appl Sci Manuf 83:98–112 Sanjay MR, Madhu P, Jawaid M, Senthamaraikannan P, Senthil S, Pradeep S (2018) Characterization and properties of natural fiber polymer composites: a comprehensive review. J Clean Prod 172:566–581 Kalia S, Dufresne A, Cherian BM, Kaith BS, Avérous L, Njuguna J, Nassiopoulos E (2011) Cellulose-based bio-and nanocomposites: a review. Int J Polym Sci 2011:1–35 Mokhtar M (2007) Characterization and treatments of pineapple leaf fibre thermoplastic composite for construction application. Universiti Teknologi Malaysia, Johor Bahru Alamri H, Low IM (2012) Mechanical properties and water absorption behaviour of recycled cellulose fibre reinforced epoxy composites. Polym Test 31(5):620–628 Masoodi R, Pillai KM (2012) A study on moisture absorption and swelling in bio-based jute-epoxy composites. J Reinf Plast Compos 31(5):285–294 Khalil HPSA, Bhat IUH, Jawaid M, Zaidon A, Hermawan D, Hadi YS (2012) Bamboo fibre reinforced biocomposites: a review. Mater Des 42:353–368 Chandramohan D, Marimuthu K (2011) A review on natural fibers. Int J Res Rev Appl Sci 8(2):194–206 Ramamoorthy SK, Skrifvars M, Persson A (2015) A review of natural fibers used in biocomposites: plant, animal and regenerated cellulose fibers. Polym Rev 55(1):107–162 Mohanty AK, Misra M, Drzal LT (2001) Surface modifications of natural fibers and performance of the resulting 882 [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] J Mater Sci (2020) 55:829–892 biocomposites: an overview. Compos Interfaces 8(5):313–343 Foulk JA, McAlister Iii DD (2002) Single cotton fiber properties of low, ideal, and high micronaire values. Text Res J 72(10):885–891 Mohanty AK, Misra M (1995) Studies on jute composites—a literature review. Polym Plast Technol Eng 34(5):729–792 Biswas S, Srikanth G, Nangia S (2001) Development of natural fibre composites in India. In: Proceedings of annual convention & trade show, Composites, Composites Fabricator’s Association at Tampa, Florida, USA, October 03–06 Yan L, Chouw N, Jayaraman K (2014) Flax fibre and its composites–a review. Compos Part B Eng 56:296–317 Shahzad A (2012) Hemp fiber and its composites—a review. J Compos Mater 46(8):973–986 Li Y, Mai Y-W, Ye L (2000) Sisal fibre and its composites: a review of recent developments. Compos Sci Technol 60(11):2037–2055 Akil H, Omar MF, Mazuki AAM, Safiee S, Ishak ZAM, Bakar AA (2011) Kenaf fiber reinforced composites: a review. Mater Des 32(8–9):4107–4121 Jayavani S, Deka H, Varghese TO, Nayak SK (2016) Recent development and future trends in coir fiber-reinforced green polymer composites: review and evaluation. Polym Compos 11(37):3296–3309 Pandey SN (2007) Ramie fibre: part I. Chemical composition and chemical properties. A critical review of recent developments. Text Prog 39(1):1–66 Balaji A, Karthikeyan B, Raj CS (2014) Bagasse fiber–the future biocomposite material: a review. Int J Cemtech Res 7(1):223–233 Sanadi AR, Caulfield DF, Jacobson RE, Rowell RM (1995) Renewable agricultural fibers as reinforcing fillers in plastics: mechanical properties of kenaf fiber-polypropylene composites. Ind Eng Chem Res 34(5):1889–1896 Mohanty AK, Misra M, Drzal LT (2005) Natural fibers, biopolymers, and biocomposites. CRC Press, Boca Raton Hattalli S, Benaboura A, Ham-Pichavant F, Nourmamode A, Castellan A (2002) Adding value to Alfa grass (Stipa tenacissima L.) soda lignin as phenolic resins 1. Lignin characterization. Polym Degrad Stab 76(2):259–264 Hoareau W, Trindade WG, Siegmund B, Castellan A, Frollini E (2004) Sugar cane bagasse and curaua lignins oxidized by chlorine dioxide and reacted with furfuryl alcohol: characterization and stability. Polym Degrad Stab 86(3):567–576 Malkapuram R, Kumar V, Negi YS (2009) Recent development in natural fiber reinforced polypropylene composites. J Reinf Plast Compos 28(10):1169–1189 [47] [48] [49] [50] [51] [52] [53] [54] [55] [56] [57] [58] [59] [60] [61] [62] [63] Martı́-Ferrer F, Vilaplana F, Ribes-Greus A, BeneditoBorrás A, Sanz-Box C (2006) Flour rice husk as filler in block copolymer polypropylene: effect of different coupling agents. J Appl Polym Sci 99(4):1823–1831 Rowell RM (2008) Natural fibres: types and properties. In: Pickering K (ed) Properties and performance of naturalfibre composites. Elsevier, Amsterdam, pp 3–66 Baeurle SA, Hotta A, Gusev AA (2006) On the glassy state of multiphase and pure polymer materials. Polymer 47(17):6243–6253 Lokensgard E (2016) Industrial plastics: theory and applications, 6th edn. Cengage Learning, p 544 Sinha R (2004) Outlines of polymer technology: manufacture of polymers. Prentice Hall of India Private Limited, India, New Delhi Mohammad NAB (2007) Synthesis, characterization, and properties of the new unsaturated polyester resins for composite applications. MARA University of Technology, pp 45–56 Iijima T, Tochimoto T, Tomoi M (1991) Modification of epoxy resins with poly (aryl ether ketone) s. J Appl Polym Sci 43(9):1685–1692 Mukherjee RN, Pal SK, Sanyal SK, Phani KK (1984) Role of interface in fibre reinforced polymer composites with special reference to natural fibres. J Polym Mater 1:69–81 Pritchard G (2012) Developments in reinforced plastics—4. Elsevier Applied Science Publishers, New York Sarkar BK, Roy R, Ray D (1997) Emerging dominance of composites as a structural material. J Inst Eng India Part MM Metall Mater Sci Div 78:31–37 Bastioli C (2001) Global status of the production of biobased packaging materials. Starch-Stärke 53(8):351–355 Edgar KJ, Buchanan CM, Debenham JS, Rundquist PA, Seiler BD, Shelton MC, Tindall D (2001) Advances in cellulose ester performance and application. Prog Polym Sci 26(9):1605–1688 TiaGe Volova (2004) Polyhydroxyalkanoates—plastic materials of the 21st century: production, properties, applications. Nova Publishers, New York Sharma SK, Mudhoo A (2011) A handbook of applied biopolymer technology: synthesis, degradation and applications. Royal Society of Chemistry, London PolymerOhio (2015) Bio-based resins: a growing opportunity for bioplastics packaging. Polymers PolymerOhio Kuruppalil Z (2011) Green plastics: an emerging alternative for petroleum-based plastics. Int J Eng Res Innov 3(1):59–64 Lilholt H, Lawther JM (2000) Natural organic fibers. In: Kelly A, Zweben C (eds) Comprehensive composite materials, vol 1. Elsevier Science, Amsterdam 883 J Mater Sci (2020) 55:829–892 [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75] [76] Gassan J, Gutowski VS (2000) Effects of corona discharge and UV treatment on the properties of jute-fibre epoxy composites. Compos Sci Technol 60(15):2857–2863 Bataille P, Dufourd M, Sapieha S (1994) Copolymerization of styrene on to cellulose activated by corona. Polym Int 34(4):387–391 Mukhopadhyay S, Fangueiro R (2009) Physical modification of natural fibers and thermoplastic films for composites—a review. J Thermoplast Compos Mater 22(2):135–162 Marais S, Gouanvé F, Bonnesoeur A, Grenet J, PoncinEpaillard F, Morvan C, Métayer M (2005) Unsaturated polyester composites reinforced with flax fibers: effect of cold plasma and autoclave treatments on mechanical and permeation properties. Compos Part A Appl Sci Manuf 36(7):975–986 Martin AR, Manolache S, Mattoso LHC, Rowell RM, Denes F Plasma modification of sisal and high-density polyethylene composites: effect on mechanical properties. In, 2000 2000 di Benedetto RM, Gelfuso MV, Thomazini D (2015) Influence of UV radiation on the physical-chemical and mechanical properties of banana fiber. Mater Res 18:265–272 Beg MDH, Pickering KL (2008) Mechanical performance of Kraft fibre reinforced polypropylene composites: influence of fibre length, fibre beating and hygrothermal ageing. Compos Part A Appl Sci Manuf 39(11):1748–1755 Cao Y, Sakamoto S, Goda K (2007) Effects of heat and alkali treatments on mechanical properties of kenaf fibers. In: 16th international conference on composite materials, Kyoto, Japan, vol 1, pp 8–13 Rong MZ, Zhang MQ, Liu Y, Yang GC, Zeng HM (2001) The effect of fiber treatment on the mechanical properties of unidirectional sisal-reinforced epoxy composites. Compos Sci Technol 61(10):1437–1447 Singh B, Gupta M, Verma A (1996) Influence of fiber surface treatment on the properties of sisal-polyester composites. Polym Compos 17(6):910–918 Ray D, Sarkar BK, Rana AK, Bose NR (2001) The mechanical properties of vinylester resin matrix composites reinforced with alkali-treated jute fibres. Compos Part A Appl Sci Manuf 32(1):119–127 Iannace S, Ali R, Nicolais L (2001) Effect of processing conditions on dimensions of sisal fibers in thermoplastic biodegradable composites. J Appl Polym Sci 79(6):1084–1091 Walker JCF (2006) Primary wood processing: principles and practice. Springer, Berlin [77] [78] [79] [80] [81] [82] [83] [84] [85] [86] [87] [88] [89] [90] Brı́gida AIS, Calado VMA, Gonçalves LRB, Coelho MAZ (2010) Effect of chemical treatments on properties of green coconut fiber. Carbohydr Polym 79(4):832–838 Aziz SH, Ansell MP (2004) The effect of alkalization and fibre alignment on the mechanical and thermal properties of kenaf and hemp bast fibre composites: part 1–polyester resin matrix. Compos Sci Technol 64(9):1219–1230 Suardana NPG, Piao Y, Lim JK (2011) Mechanical properties of hemp fibers and hemp/pp composites: effects of chemical surface treatment. Mater Phys Mech 11(1):1–8 Bledzki AK, Mamun AA, Lucka-Gabor M, Gutowski VS (2008) The effects of acetylation on properties of flax fibre and its polypropylene composites. Express Polym Lett 2(6):413–422 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 Tserki V, Zafeiropoulos NE, Simon F, Panayiotou C (2005) A study of the effect of acetylation and propionylation surface treatments on natural fibres. Compos Part A Appl Sci Manuf 36(8):1110–1118 Zafeiropoulos NE, Baillie CA (2007) A study of the effect of surface treatments on the tensile strength of flax fibres: part II. Application of Weibull statistics. Compos Part A Appl Sci Manuf 38(2):629–638 Manikandan Nair KC, Diwan SM, Thomas S (1996) Tensile properties of short sisal fiber reinforced polystyrene composites. J Appl Polym Sci 60(9):1483–1497 Wang B, Panigrahi S, Tabil L, Crerar W (2007) Pre-treatment of flax fibers for use in rotationally molded biocomposites. J Reinf Plast Compos 26(5):447–463 Joseph K, Mattoso LHC, Toledo RD, Thomas S, De Carvalho LH, Pothen L, Kala S, James B (2000) Natural fiber reinforced thermoplastic composites. Nat Polym Agrofibers Compos 159:159–201 Joseph PV, Joseph K, Thomas S, Pillai CKS, Prasad VS, Groeninckx G, Sarkissova M (2003) The thermal and crystallisation studies of short sisal fibre reinforced polypropylene composites. Compos Part A Appl Sci Manuf 34(3):253–266 Aranberri-Askargorta I, Lampke T, Bismarck A (2003) Wetting behavior of flax fibers as reinforcement for polypropylene. J Colloid Interface Sci 263(2):580–589 Bismarck A, Aranberri-Askargorta I, Springer J, Lampke T, Wielage B, Stamboulis A, Shenderovich I, Limbach HH (2002) Surface characterization of flax, hemp and cellulose fibers; surface properties and the water uptake behavior. Polym Compos 23(5):872–894 Fuqua MA, Ulven CA (2008) Characterization of polypropylene/corn fiber composites with maleic anhydride 884 grafted polypropylene. J Biobased Mater Bioenergy 2(3):258–263 [91] Hong CK, Kim N, Kang SL, Nah C, Lee YS, Cho BH, Ahn JH (2008) Mechanical properties of maleic anhydride treated jute fibre/polypropylene composites. Plast, Rubber Compos 37(7):325–330 [92] Sun JX, Sun XF, Zhao H, Sun RC (2004) Isolation and characterization of cellulose from sugarcane bagasse. Polym Degrad Stab 84(2):331–339 [93] Li X, Panigrahi S, Tabil LG (2009) A study on flax fiberreinforced polyethylene biocomposites. Appl Eng Agric 25(4):525–531 [94] Vilay V, Mariatti M, Taib RM, Todo M (2008) Effect of fiber surface treatment and fiber loading on the properties of bagasse fiber—reinforced unsaturated polyester composites. Compos Sci Technol 68(3–4):631–638 [95] Sreekala MS, Kumaran MG, Joseph S, Jacob M, Thomas S (2000) Oil palm fibre reinforced phenol formaldehyde composites: influence of fibre surface modifications on the mechanical performance. Appl Compos Mater 7(5–6):295–329 [96] George J, Janardhan R, Anand JS, Bhagawan SS, Thomas S (1996) Melt rheological behaviour of short pineapple fibre reinforced low density polyethylene composites. Polymer 37(24):5421–5431 [97] Paul SA, Oommen C, Joseph K, Mathew G, Thomas S (2010) The role of interface modification on thermal degradation and crystallization behavior of composites from commingled polypropylene fiber and banana fiber. Polym Compos 31(6):1113–1123 [98] Zafeiropoulos NE, Baillie CA, Hodgkinson JM (2002) Engineering and characterisation of the interface in flax fibre/polypropylene composite materials. Part II. The effect of surface treatments on the interface. Compos Part A Appl Sci Manuf 33(9):1185–1190 [99] Corrales F, Vilaseca F, Llop M, Girones J, Mendez JA, Mutje P (2007) Chemical modification of jute fibers for the production of green-composites. J Hazard Mater 144(3):730–735 [100] Li X, Panigrahi SA, Tabil LG, Crerar WJ (2004) Flax fiberreinforced composites and the effect of chemical treatments on their properties. In: ASAE/CSAE North Central regional annual meeting, Manitoba, Canada, September 24–25 2004. ASAE, pp 1–10 [101] Zadorecki P, Flodin P (1985) Surface modification of cellulose fibers. II. The effect of cellulose fiber treatment on the performance of cellulose–polyester composites. J Appl Polym Sci 30(10):3971–3983 [102] Pickering KL, Li Y, Farrell RL, Lay M (2007) Interfacial modification of hemp fiber reinforced composites using J Mater Sci (2020) 55:829–892 [103] [104] [105] [106] [107] [108] [109] [110] [111] [112] [113] [114] [115] [116] fungal and alkali treatment. J Biobased Mater Bioenergy 1(1):109–117 Chabba S, Netravali AN (2004) ‘Green’composites using modified soy protein concentrate resin and flax fabrics and yarns. JSME Int J Ser A Solid Mech Mater Eng 47(4):556–560 Kumar R, Zhang L (2008) Water-induced hydrophobicity of soy protein materials containing 2, 2-diphenyl-2-hydroxyethanoic acid. Biomacromol 9(9):2430–2437 Kumar R, Zhang L (2009) Soy protein films with the hydrophobic surface created through non-covalent interactions. Ind Crops Prod 29(2–3):485–494 Newill JR, McKnight SH, Hoppel CP, Cooper GR (1999) Effects of coatings on moisture absorption in composite materials. Army Research Lab Aberdeen Proving Ground MD, US Doherty W, Halley P, Edye L, Rogers D, Cardona F, Park Y, Woo T (2007) Studies on polymers and composites from lignin and fiber derived from sugar cane. Polym Adv Technol 18(8):673–678 Stark NM (1999) Wood fiber derived from scrap pallets used in polypropylene composites. For Prod J 49:39–46 Araujo JR, Waldman WR, De Paoli MA (2008) Thermal properties of high density polyethylene composites with natural fibres: coupling agent effect. Polym Degrad Stab 93(10):1770–1775 Yildiz S, Gezer ED, Yildiz UC (2006) Mechanical and chemical behavior of spruce wood modified by heat. Build Environ 41(12):1762–1766 Chand N, Dwivedi UK (2006) Effect of coupling agent on abrasive wear behaviour of chopped jute fibre-reinforced polypropylene composites. Wear 261(10):1057–1063 Corradini E, Ito EN, Marconcini JM, Rios CT, Agnelli JAM, Mattoso LHC (2009) Interfacial behavior of composites of recycled poly (ethyelene terephthalate) and sugarcane bagasse fiber. Polym Test 28(2):183–187 Elkington M, Bloom D, Ward C, Chatzimichali A, Potter K (2015) Hand layup: understanding the manual process. Adv Manuf Polym Compos Sci 1(3):138–151 Kumar AP, Singh RP, Sarwade BD (2005) Degradability of composites, prepared from ethylene–propylene copolymer and jute fiber under accelerated aging and biotic environments. Mater Chem Phys 92(2–3):458–469 Mulinari DR, Voorwald HJC, Cioffi MOH, Da Silva MLCP, da Cruz TG, Saron C (2009) Sugarcane bagasse cellulose/HDPE composites obtained by extrusion. Compos Sci Technol 69(2):214–219 Richardson MOW, Zhang ZY (2000) Experimental investigation and flow visualisation of the resin transfer mould filling process for non-woven hemp reinforced phenolic 885 J Mater Sci (2020) 55:829–892 [117] [118] [119] [120] [121] [122] [123] [124] [125] [126] [127] [128] [129] [130] composites. Compos Part A Appl Sci Manuf 31(12):1303–1310 Rouison D, Sain M, Couturier M (2004) Resin transfer molding of natural fiber reinforced composites: cure simulation. Compos Sci Techno 64(5):629–644 Rouison D, Sain M, Couturier M (2006) Resin transfer molding of hemp fiber composites: optimization of the process and mechanical properties of the materials. Compos Sci Technol 66(7–8):895–906 Santos PA, Spinacé MAS, Fermoselli KKG, De Paoli M-A (2007) Polyamide-6/vegetal fiber composite prepared by extrusion and injection molding. Compos Part A Appl Sci Manuf 38(12):2404–2411 Scherübl B, Hintermann M (2005) Application of natural fibre reinforced plastics for automotive exterior parts, with a focus on underfloor systems. In: Proceedings of the 8th international AVK-TV conference, Baden, Germany, 2005. p D5 Van Voorn B, Smit HHG, Sinke RJ, De Klerk B (2001) Natural fibre reinforced sheet moulding compound. Compos Part A Appl Sci Manuf 32(9):1271–1279 Zou Y, Huda S, Yang Y (2010) Lightweight composites from long wheat straw and polypropylene web. Bioresour Technol 101(6):2026–2033 Dai D, Fan M (2014) Wood fibres as reinforcements in natural fibre composites: structure, properties, processing and applications. In: Hodzic A, Shanks R (Eds) Natural fibre composites. Elsevier, Amsterdam, pp 3–65 Sebe G, Cetin NS, Hill CAS, Hughes M (2000) RTM hemp fibre-reinforced polyester composites. Appl Compos Mater 7(5–6):341–349 Summerscales J, Searle TJ (2005) Low-pressure (vacuum infusion) techniques for moulding large composite structures. Proc Inst Mech Eng Part L J Mater Des Appl 219(1):45–58 Francucci G, Rodrı́guez ES, Vázquez A (2010) Study of saturated and unsaturated permeability in natural fiber fabrics. Compos Part A Appl Sci Manuf 41(1):16–21 Masoodi R, Pillai KM, Grahl N, Tan H (2012) Numerical simulation of LCM mold-filling during the manufacture of natural fiber composites. J Reinf Plast Compos 31(6):363–378 Goren A, Atas C (2008) Manufacturing of polymer matrix composites using vacuum assisted resin infusion molding. Arch Mater Sci Eng 34(2):117–120 Ashby MF (2012) Materials and the environment: eco-informed material choice. Elsevier, Amsterdam Lystrup A, Løgostrup T, Knudsen H, Vestergaard T, Lilleheden L, Vestergaard J (1998) Hybrid yarn for [131] [132] [133] [134] [135] [136] [137] [138] [139] [140] [141] [142] [143] [144] thermoplastic fiber composites. Summary of technical results. Final report for MUP2 Framework Program, Denmark Baley C, Lan M, Davies P, Cartié D (2015) Porosity in ocean racing yacht composites: a review. Appl Compos Mater 22(1):13–28 Madsen B, Lilholt H (2003) Physical and mechanical properties of unidirectional plant fibre composites—an evaluation of the influence of porosity. Compos Sci Technol 63(9):1265–1272 Vaxman A, Narkis M, Siegmann A, Kenig S (1989) Void formation in short-fiber thermoplastic composites. Polym Compos 10(6):449–453 Bowles KJ, Frimpong S (1992) Void effects on the interlaminar shear strength of unidirectional graphite-fiber-reinforced composites. J Compos Mater 26(10):1487–1509 Dong C, Takagi H (2014) Flexural properties of cellulose nanofibre reinforced green composites. Compos Part B Eng 58:418–421 Judd NCW, Wright WW (1978) Voids and their effects on the mechanical properties of composites—an appraisal. SAMPE J 14:10–14 Ghiorse SR (1993) Effect of void content on the mechanical properties of carbon/epoxy laminates. SAMPE Q 24:54–59 Zampaloni M, Pourboghrat F, Yankovich SA, Rodgers BN, Moore J, Drzal LT, Mohanty AK, Misra M (2007) Kenaf natural fiber reinforced polypropylene composites: a discussion on manufacturing problems and solutions. Compos Part A Appl Sci Manuf 38(6):1569–1580 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. Compos Sci Technol 62(14):1857–1868 Biswas S, Kindo S, Patnaik A (2011) Effect of fiber length on mechanical behavior of coir fiber reinforced epoxy composites. Fibers Polym 12(1):73–78 Gao S-L, Mäder E (2006) Jute/polypropylene composites I. Effect of matrix modification. Compos Sci Technol 66(7–8):952–963 Karmaker AC, Shneider JP (1996) Mechanical performance of short jute fibre reinforced polypropylene. J Mater Sci Lett 15(3):201–202 Shibata S, Cao Y, Fukumoto I (2005) Press forming of short natural fiber-reinforced biodegradable resin: effects of fiber volume and length on flexural properties. Polym Test 24(8):1005–1011 Sharma NK, Kumar V (2013) Studies on properties of banana fiber reinforced green composite. J Reinf Plast Compos 32(8):525–532 886 [145] Liu Q, Hughes M (2008) The fracture behaviour and toughness of woven flax fibre reinforced epoxy composites. Compos Part A Appl Sci Manuf 39(10):1644–1652 [146] Zhang Y, Li Y, Ma H, Yu T (2013) Tensile and interfacial properties of unidirectional flax/glass fiber reinforced hybrid composites. Compos Sci Technol 88:172–177 [147] Li Y, Chen C, Xu J, Zhang Z, Yuan B, Huang X (2015) Improved mechanical properties of carbon nanotubescoated flax fiber reinforced composites. J Mater Sci 50(3):1117–1128. https://doi.org/10.1007/s10853-014-866 8-3 [148] Kafi AA, Magniez K, Fox BL (2011) Effect of manufacturing process on the flexural, fracture toughness, and thermo-mechanical properties of bio-composites. Compos Part A Appl Sci Manuf 42(8):993–999 [149] Kinloch AJ, Taylor AC, Techapaitoon M, Teo WS, Sprenger S (2015) Tough, natural-fibre composites based upon epoxy matrices. J Mater Sci 50(21):6947–6960. http s://doi.org/10.1007/s10853-015-9246-z [150] Hsieh TH, Kinloch AJ, Masania K, Lee JS, Taylor AC, Sprenger S (2010) The toughness of epoxy polymers and fibre composites modified with rubber microparticles and silica nanoparticles. J Mater Sci 45(5):1193–1210. https://d oi.org/10.1007/s10853-009-4064-9 [151] Agunsoye JO, Aigbodion VS (2013) Bagasse filled recycled polyethylene bio-composites: morphological and mechanical properties study. Results Phys 3:187–194 [152] Wong KJ, Zahi S, Low KO, Lim CC (2010) Fracture characterisation of short bamboo fibre reinforced polyester composites. Mater Des 31(9):4147–4154 [153] Muralidhar BA (2013) Tensile and compressive behaviour of multilayer flax-rib knitted preform reinforced epoxy composites. Mater Des 49:400–405 [154] Bledzki AK, Gassan J, Zhang W (1999) Impact properties of natural fiber-reinforced epoxy foams. J Cell Plast 35(6):550–562 [155] Cao Y, Shibata S, Fukumoto I (2006) Mechanical properties of biodegradable composites reinforced with bagasse fibre before and after alkali treatments. Compos Part A Appl Sci Manuf 37(3):423–429 [156] Dhakal HN, Zhang ZY, Richardson MOW, Errajhi OAZ (2007) The low velocity impact response of non-woven hemp fibre reinforced unsaturated polyester composites. Compos Struct 81(4):559–567 [157] Jandas PJ, Mohanty S, Nayak SK (2013) Mechanical properties of surface-treated banana fiber/polylactic acid biocomposites: a comparative study of theoretical and experimental values. J Appl Polym Sci 127(5):4027–4038 [158] Liang S, Guillaumat L, Gning P-B (2015) Impact behaviour of flax/epoxy composite plates. Int J Impact Eng 80:56–64 J Mater Sci (2020) 55:829–892 [159] Clarinval AM, Halleux J (2005) Classification of biodegradable polymers. In: Smith R (Ed) Biodegradable polymers for industrial applications. Elsevier, Amsterdam, pp 3–31 [160] Bowen CR, Dent AC, Stevens R, Cain M, Stewart M (2005) Determination of critical and minimum volume fraction for composite sensors and actuators. In: First international conference on multi-material micro manufacture, Cardiff, UK, 2005. Elsevier, pp 1–4 [161] Facca AG, Kortschot MT, Yan N (2007) Predicting the tensile strength of natural fibre reinforced thermoplastics. Compos Sci Technol 67(11–12):2454–2466 [162] Ma X, Yu J, Kennedy JF (2005) Studies on the properties of natural fibers-reinforced thermoplastic starch composites. Carbohydr Polym 62(1):19–24 [163] We˛cławski BT, Fan M, Hui D (2014) Compressive behaviour of natural fibre composite. Compos Part B Eng 67:183–191 [164] Afroughsabet V, Ozbakkaloglu T (2015) Mechanical and durability properties of high-strength concrete containing steel and polypropylene fibers. Constr Build Mater 94:73–82 [165] Ismail MA (2007) Compressive and tensile strength of natural fibre-reinforced cement base composites. AL Rafdain Eng J 15(2):42–51 [166] Gozde Ozerkan N, Ahsan B, Mansour S, Iyengar SR (2013) Mechanical performance and durability of treated palm fiber reinforced mortars. Int J Sustain Built Environ 2(2):131–142 [167] Ragoubi M, Bienaimé D, Molina S, George B, Merlin A (2010) Impact of corona treated hemp fibres onto mechanical properties of polypropylene composites made thereof. Ind Crops Prod 31(2):344–349 [168] Pizzi A, Kueny R, Lecoanet F, Massetau B, Carpentier D, Krebs A, Loiseau F, Molina S, Ragoubi M (2009) High resin content natural matrix–natural fibre biocomposites. Ind Crops Prod 30(2):235–240 [169] Ragoubi M, George B, Molina S, Bienaimé D, Merlin A, Hiver JM, Dahoun A (2012) Effect of corona discharge treatment on mechanical and thermal properties of composites based on miscanthus fibres and polylactic acid or polypropylene matrix. Compos Part A Appl Sci Manuf 43(4):675–685 [170] Seki Y, Sever K, Sarikanat M, Güleç HA, Tavman IH (2009) The influence of oxygen plasma treatment of jute fibers on mechanical properties of jute fiber reinforced thermoplastic composites. In: 5th international advanced technologies symposium (IATS’09), Karabuk, Turkey 2009. pp 1–4 J Mater Sci (2020) 55:829–892 [171] Sinha E, Panigrahi S (2009) Effect of plasma treatment on structure, wettability of jute fiber and flexural strength of its composite. J Compos Mater 43(17):1791–1802 [172] Seki Y, Sarikanat M, Sever K, Erden S, Gulec HA (2010) Effect of the low and radio frequency oxygen plasma treatment of jute fiber on mechanical properties of jute fiber/polyester composite. Fibers Polym 11(8):1159–1164 [173] Jang JY, Jeong TK, Oh HJ, Youn JR, Song YS (2012) Thermal stability and flammability of coconut fiber reinforced poly (lactic acid) composites. Compos Part B Eng 43(5):2434–2438 [174] Prasad SV, Pavithran C, Rohatgi PK (1983) Alkali treatment of coir fibres for coir-polyester composites. J Mater Sci 18(5):1443–1454. https://doi.org/10.1007/BF01111964 [175] Ray D, Sarkar BK, Rana AK, Bose NR (2001) Effect of alkali treated jute fibres on composite properties. Bull Mater Sci 24(2):129–135 [176] Onal L, Karaduman Y (2009) Mechanical characterization of carpet waste natural fiber-reinforced polymer composites. J Compos Mater 43(16):1751–1768 [177] Huda MS, Drzal LT, Mohanty AK, Misra M (2008) Effect of chemical modifications of the pineapple leaf fiber surfaces on the interfacial and mechanical properties of laminated biocomposites. Compos Interfaces 15(2–3):169–191 [178] Qin C, Soykeabkaew N, Xiuyuan N, Peijs T (2008) The effect of fibre volume fraction and mercerization on the properties of all-cellulose composites. Carbohydr Polym 71(3):458–467 [179] Bledzki AK, Fink HP, Specht K (2004) Unidirectional hemp and flax EP-and PP-composites: influence of defined fiber treatments. J Appl Polym Sci 93(5):2150–2156 [180] Venkateshwaran N, Perumal AE, Arunsundaranayagam D (2013) Fiber surface treatment and its effect on mechanical and visco-elastic behaviour of banana/epoxy composite. Mater Des 47:151–159 [181] Xu Y, Kawata S, Hosoi K, Kawai T, Kuroda S (2009) Thermomechanical properties of the silanized-kenaf/polystyrene composites. Express Polym Lett 3(10):657–664 [182] Pothan LA, Thomas S (2003) Polarity parameters and dynamic mechanical behaviour of chemically modified banana fiber reinforced polyester composites. Compos Sci Technol 63(9):1231–1240 [183] Cantero G, Arbelaiz A, Llano-Ponte R, Mondragon I (2003) Effects of fibre treatment on wettability and mechanical behaviour of flax/polypropylene composites. Compos Sci Technol 63(9):1247–1254 [184] Abdul Khalil HPS, Ismail H (2000) Effect of acetylation and coupling agent treatments upon biological degradation of plant fibre reinforced polyester composites. Polym Test 20(1):65–75 887 [185] Ismail H, Khalil HPSA (2000) The effects of partial replacement of oil palm wood flour by silica and silane coupling agent on properties of natural rubber compounds. Polym Test 20(1):33–41 [186] Sgriccia N, Hawley MC, Misra M (2008) Characterization of natural fiber surfaces and natural fiber composites. Compos Part A Appl Sci Manuf 39(10):1632–1637 [187] Eng CC, Ibrahim NA, Zainuddin N, Ariffin H, Yunus WM, Wan Z (2014) Impact strength and flexural properties enhancement of methacrylate silane treated oil palm mesocarp fiber reinforced biodegradable hybrid composites. Sci World J 2014:1–8 [188] Joseph S, Koshy P, Thomas S (2005) The role of interfacial interactions on the mechanical properties of banana fibre reinforced phenol formaldehyde composites. Compos Interfaces 12(6):581–600 [189] Hill CAS, Abdul Khalil HPS (2000) Effect of fiber treatments on mechanical properties of coir or oil palm fiber reinforced polyester composites. J Appl Polym Sci 78(9):1685–1697 [190] Zafeiropoulos NE, Williams DR, Baillie CA, Matthews FL (2002) Engineering and characterisation of the interface in flax fibre/polypropylene composite materials. Part I. Development and investigation of surface treatments. Compos Part A Appl Sci Manuf 33(8):1083–1093 [191] Luz SM, Del Tio J, Rocha GJM, Gonçalves AR, Del’Arco AP Jr (2008) Cellulose and cellulignin from sugarcane bagasse reinforced polypropylene composites: effect of acetylation on mechanical and thermal properties. Compos Part A Appl Sci Manuf 39(9):1362–1369 [192] Manikandan Nair KC, Thomas S, Groeninckx G (2001) Thermal and dynamic mechanical analysis of polystyrene composites reinforced with short sisal fibres. Compos Sci Technol 61(16):2519–2529 [193] Sapieha S, Allard P, Zang YH (1990) Dicumyl peroxidemodified cellulose/LLDPE composites. J Appl Polym Sci 41(9–10):2039–2048 [194] Abdul Razak N, Ibrahim N, Zainuddin N, Rayung M, Saad W (2014) The influence of chemical surface modification of kenaf fiber using hydrogen peroxide on the mechanical properties of biodegradable kenaf fiber/poly (lactic acid) composites. Molecules 19(3):2957–2968 [195] Yang H-S, Kim H-J, Park H-J, Lee B-J, Hwang T-S (2007) Effect of compatibilizing agents on rice-husk flour reinforced polypropylene composites. Compos Struct 77(1):45–55 [196] Li X, Zhang J, He J, Jeevan Prasad Reddy D, Varada Rajulu A (2010) Tensile properties of hildegardia fibers reinforced polypropylene biocomposites. J Compos Mater 44(14):1681–1687 888 [197] Keener TJ, Stuart RK, Brown TK (2004) Maleated coupling agents for natural fibre composites. Compos Part A Appl Sci Manuf 35(3):357–362 [198] Liu H, Wu Q, Zhang Q (2009) Preparation and properties of banana fiber-reinforced composites based on high density polyethylene (HDPE)/Nylon-6 blends. Bioresour Technol 100(23):6088–6097 [199] Gassan J, Bledzki AK (2000) Possibilities to improve the properties of natural fiber reinforced plastics by fiber modification–Jute polypropylene composites–. Appl Compos Mater 7(5–6):373–385 [200] Wielage B, Lampke T, Utschick H, Soergel F (2003) Processing of natural-fibre reinforced polymers and the resulting dynamic—mechanical properties. J Mater Process Technol 139(1–3):140–146 [201] Misra S, Misra M, Tripathy SS, Nayak SK, Mohanty AK (2002) The influence of chemical surface modification on the performance of sisal-polyester biocomposites. Polym Compos 23(2):164–170 [202] Sreekala MS, Kumaran MG, Thomas S (2002) Water sorption in oil palm fiber reinforced phenol formaldehyde composites. Compos Part A Appl Sci Manuf 33(6):763–777 [203] Joseph K, Thomas S, Pavithran C (1996) Effect of chemical treatment on the tensile properties of short sisal fibre-reinforced polyethylene composites. Polymer 37(23):5139–5149 [204] Paul SA, Boudenne A, Ibos L, Candau Y, Joseph K, Thomas S (2008) Effect of fiber loading and chemical treatments on thermophysical properties of banana fiber/ polypropylene commingled composite materials. Compos Part A Appl Sci Manuf 39(9):1582–1588 [205] Kalaprasad G, Francis B, Thomas S, Kumar CR, Pavithran C, Groeninckx G, Thomas S (2004) Effect of fibre length and chemical modifications on the tensile properties of intimately mixed short sisal/glass hybrid fibre reinforced low density polyethylene composites. Polym Int 53(11):1624–1638 [206] Torres FG, Cubillas ML (2005) Study of the interfacial properties of natural fibre reinforced polyethylene. Polym Test 24(6):694–698 [207] Mwaikambo LY, Ansell MP (2002) Chemical modification of hemp, sisal, jute, and kapok fibers by alkalization. J Appl Polym Sci 84(12):2222–2234 [208] Guan J, Hanna MA (2006) Selected morphological and functional properties of extruded acetylated starch–cellulose foams. Bioresour Technol 97(14):1716–1726 [209] Joonobi M, Harun J, Tahir PM, Zaini LH, SaifulAzry S, Makinejad MD (2010) Characteristic of nanofibers extracted from kenaf core. BioResources 5(4):2556–2566 J Mater Sci (2020) 55:829–892 [210] Barone JR (2005) Polyethylene/keratin fiber composites with varying polyethylene crystallinity. Compos Part A Appl Sci Manuf 36(11):1518–1524 [211] Zafeiropoulos NE, Baillie CA, Matthews FL (2001) An investigation of the effect of processing conditions on the interface of flax/polypropylene composites. Adv Compos Lett 10(6):293–297 [212] Suryanegara L, Nakagaito AN, Yano H (2009) The effect of crystallization of PLA on the thermal and mechanical properties of microfibrillated cellulose-reinforced PLA composites. Compos Sci Technol 69(7–8):1187–1192 [213] Quan H, Li Z-M, Yang M-B, Huang R (2005) On transcrystallinity in semi-crystalline polymer composites. Compos Sci Technol 65(7–8):999–1021 [214] Zafeiropoulos NE, Baillie CA, Matthews FL (2001) A study of transcrystallinity and its effect on the interface in flax fibre reinforced composite materials. Compos Part A Appl Sci Manuf 32(3–4):525–543 [215] Zafeiropoulos NE, Baillie CA, Matthews FL (2001) The effect of transcrystallinity on the interface of green flax/ polypropylene composite materials. Adv Compos Lett 10(5):229–236 [216] Sawpan MA, Pickering KL, Fernyhough A (2011) Effect of fibre treatments on interfacial shear strength of hemp fibre reinforced polylactide and unsaturated polyester composites. Compos Part A Appl Sci Manuf 42(9):1189–1196 [217] Xiong X, Shen SZ, Hua L, Liu JZ, Li X, Wan X, Miao M (2018) Finite element models of natural fibers and their composites: a review. J Reinf Plast Compos 37(9):617–635 [218] Ilczyszyn F, Cherouat A, Montay G Effect of hemp fibre morphology on the mechanical properties of vegetal fibre composite material. In, 2014 2014. Trans Tech Publ, pp 485-489 [219] Virk AS, Hall W, Summerscales J (2012) Modulus and strength prediction for natural fibre composites. Mater Sci Technol 28(7):864–871 [220] Kalaprasad G, Joseph K, Thomas S, Pavithran C (1997) Theoretical modelling of tensile properties of short sisal fibre-reinforced low-density polyethylene composites. J Mater Sci 32(16):4261–4267. https://doi.org/10.1023/A: 1018651218515 [221] Facca AG, Kortschot MT, Yan N (2006) Predicting the elastic modulus of natural fibre reinforced thermoplastics. Compos Part A Appl Sci Manuf 37(10):1660–1671 [222] Beckermann GW, Pickering KL (2009) Engineering and evaluation of hemp fibre reinforced polypropylene composites: micro-mechanics and strength prediction modelling. Compos Part A Appl Sci Manuf 40(2):210–217 [223] Migneault S, Koubaa A, Erchiqui F, Chaala A, Englund K, Wolcott MP (2011) Application of micromechanical J Mater Sci (2020) 55:829–892 [224] [225] [226] [227] [228] [229] [230] [231] [232] [233] [234] [235] [236] models to tensile properties of wood–plastic composites. Wood Sci Technol 45(3):521–532 Munde YS, Ingle RB (2015) Theoretical modeling and experimental verification of mechanical properties of natural fiber reinforced thermoplastics. Proc Technol 19:320–326 Kern WT, Kim W, Argento A, Lee EC, Mielewski DF (2016) Finite element analysis and microscopy of natural fiber composites containing microcellular voids. Mater Des 106:285–294 Modniks J, Andersons J (2010) Modeling elastic properties of short flax fiber-reinforced composites by orientation averaging. Comput Mater Sci 50(2):595–599 Modniks J, Andersons J (2013) Modeling the non-linear deformation of a short-flax-fiber-reinforced polymer composite by orientation averaging. Compos Part B Eng 54:188–193 Sliseris J, Yan L, Kasal B (2016) Numerical modelling of flax short fibre reinforced and flax fibre fabric reinforced polymer composites. Compos Part B Eng 89:143–154 Stamboulis A, Baillie CA, Garkhail SK, Van Melick HGH, Peijs T (2000) Environmental durability of flax fibres and their composites based on polypropylene matrix. Appl Compos Mater 7(5–6):273–294 Apolinario G, Ienny P, Corn S, Léger R, Bergeret A, Haudin J-M (2016) Effects of water ageing on the mechanical properties of flax and glass fibre composites: degradation and reversibility. In: Natural fibres: advances in science and technology towards industrial applications. Springer, Berlin, pp 183–196 Assarar M, Scida D, El Mahi A, Poilâne C, Ayad R (2011) Influence of water ageing on mechanical properties and damage events of two reinforced composite materials: flax– fibres and glass–fibres. Mater Des 32(2):788–795 Lu MM, Van Vuure AW (2019) Improving moisture durability of flax fibre composites by using non-dry fibres. Compos Part A Appl Sci Manuf 123:301–309 Yousif BF, El-Tayeb NSM (2007) The effect of oil palm fibers as reinforcement on tribological performance of polyester composite. Surf Rev Lett 14(06):1095–1102 El-Tayeb NSM (2008) A study on the potential of sugarcane fibers/polyester composite for tribological applications. Wear 265(1–2):223–235 Yousif BF (2009) Frictional and wear performance of polyester composites based on coir fibres. Proc Inst Mech Eng Part J J Eng Tribol 223(1):51–59 Chin CW, Yousif BF (2009) Potential of kenaf fibres as reinforcement for tribological applications. Wear 267(9–10):1550–1557 889 [237] Abu-Sharkh BF, Hamid H (2004) Degradation study of date palm fibre/polypropylene composites in natural and artificial weathering: mechanical and thermal analysis. Polym Degrad Stab 85(3):967–973 [238] Singh B, Gupta M (2005) Performance of pultruded jute fibre reinforced phenolic composites as building materials for door frame. J Polym Environ 13(2):127–137 [239] Dash BN, Rana AK, Mishra HK, Nayak SK, Tripathy SS (2000) Novel low-cost jute–polyester composites. III. Weathering and thermal behavior. J Appl Polym Sci 78(9):1671–1679 [240] Joseph S, Oommen Z, Thomas S (2006) Environmental durability of banana-fiber-reinforced phenol formaldehyde composites. J Appl Polym Sci 100(3):2521–2531 [241] Bledzki AK, Gassan J (1999) Composites reinforced with cellulose based fibres. Prog Polym Sci 24(2):221–274 [242] Sarani Z, Lee KP (2002) Polystyrene-benzoylated EFB reinforced composites. Polym Plast Technol Eng 41(5):951–962 [243] Dicker MPM, Duckworth PF, Baker AB, Francois G, Hazzard MK, Weaver PM (2014) Green composites: a review of material attributes and complementary applications. Compos Part A Appl Sci Manuf 56:280–289 [244] Symington MC, Banks WM, West OD, Pethrick RA (2009) Tensile testing of cellulose based natural fibers for structural composite applications. J Compos Mater 43(9):1083–1108 [245] Raghavan J, Meshii M (1997) Creep rupture of polymer composites. Compos Sci Technol 57(4):375–388 [246] Alvarez VA, Kenny JM, Vázquez A (2004) Creep behavior of biocomposites based on sisal fiber reinforced cellulose derivatives/starch blends. Polym Compos 25(3):280–288 [247] Acha BA, Reboredo MM, Marcovich NE (2007) Creep and dynamic mechanical behavior of PP–jute composites: effect of the interfacial adhesion. Compos Part A Appl Sci Manuf 38(6):1507–1516 [248] Xu Y, Wu Q, Lei Y, Yao F (2010) Creep behavior of bagasse fiber reinforced polymer composites. Bioresour Technol 101(9):3280–3286 [249] Amiri A, Hosseini N, Ulven CA (2015) Long-term creep behavior of flax/vinyl ester composites using time-temperature superposition principle. J Renew Mater 3(3):224–233 [250] Yang T-C, Wu T-L, Hung K-C, Chen Y-L, Wu J-H (2015) Mechanical properties and extended creep behavior of bamboo fiber reinforced recycled poly (lactic acid) composites using the time–temperature superposition principle. Constr Build Mater 93:558–563 [251] Militký J, Jabbar A (2015) Comparative evaluation of fiber treatments on the creep behavior of jute/green epoxy composites. Compos Part B Eng 80:361–368 890 [252] Annicchiarico D, Alcock JR (2014) Review of factors that affect shrinkage of molded part in injection molding. Mater Manuf Process 29(6):662–682 [253] Hakimian E, Sulong AB (2012) Analysis of warpage and shrinkage properties of injection-molded micro gears polymer composites using numerical simulations assisted by the Taguchi method. Mater Des 42:62–71 [254] Nian S-C, Wu C-Y, Huang M-S (2015) Warpage control of thin-walled injection molding using local mold temperatures. Int Commun Heat Mass Transf 61:102–110 [255] Santos JD, Fajardo JI, Cuji AR, Garcı́a JA, Garzón LE, López LM (2015) Experimental evaluation and simulation of volumetric shrinkage and warpage on polymeric composite reinforced with short natural fibers. Front Mech Eng 10(3):287–293 [256] Tan H-S, Yu Y-Z, Xing L-X, Zhao L-Y, Sun H-q (2013) Density and shrinkage of injection molded impact polypropylene copolymer/coir fiber composites. Polym Plast Technol Eng 52(3):257–260 [257] Kozłowski R, Władyka-Przybylak M (2008) Flammability and fire resistance of composites reinforced by natural fibers. Polym Adv Technol 19(6):446–453 [258] John MJ, Anandjiwala RD (2008) Recent developments in chemical modification and characterization of natural fiberreinforced composites. Polym Compos 29(2):187–207 [259] Joseph PV (2001) Studies on short sisal fibre reinforced isotactic polypropylene composites. Ph.D. Thesis, Mahatma Gandhi University, India [260] Satyanarayana KG, Arizaga GGC, Wypych F (2009) Biodegradable composites based on lignocellulosic fibers— an overview. Prog Polym Sci 34(9):982–1021 [261] Wielage B, Lampke T, Marx G, Nestler K, Starke D (1999) Thermogravimetric and differential scanning calorimetric analysis of natural fibres and polypropylene. Thermochim Acta 337(1–2):169–177 [262] Yao F, Wu Q, Lei Y, Guo W, Xu Y (2008) Thermal decomposition kinetics of natural fibers: activation energy with dynamic thermogravimetric analysis. Polym Degrad Stab 93(1):90–98 [263] Grand AF, Wilkie CA (2000) Fire retardancy of polymeric materials. CRC Press, New York [264] Pujari S, Ramakrishna A, Padal KTB (2017) Investigations on thermal conductivities of jute and banana fiber reinforced epoxy composites. J Inst Eng (India) Ser D 98(1):79–83 [265] Li X, Tabil LG, Oguocha IN, Panigrahi S (2008) Thermal diffusivity, thermal conductivity, and specific heat of flax fiber—HDPE biocomposites at processing temperatures. Compos Sci Technol 68(7–8):1753–1758 J Mater Sci (2020) 55:829–892 [266] Mangal R, Saxena NS, Sreekala MS, Thomas S, Singh K (2003) Thermal properties of pineapple leaf fiber reinforced composites. Mater Sci Eng, A 339(1–2):281–285 [267] Osugi R, Takagi H, Liu K, Gennai Y (2009) Thermal conductivity behavior of natural fiber-reinforced composites. In: Proceedings of the Asian pacific conference for materials and mechanics, Yokohama, Japan, November 13–16 2009, pp 1–3 [268] Mounika M, Ramaniah K, Prasad AVR, Rao KM, Reddy KHC (2012) Thermal conductivity characterization of bamboo fiber reinforced polyester composite. J Mater Environ Sci 3(6):1109–1116 [269] Ramanaiah K, Prasad AVR, Chandra Reddy KH (2013) Mechanical and thermo-physical properties of fish tail palm tree natural fiber—reinforced polyester composites. Int J Polym Anal Charact 18(2):126–136 [270] Agrawal R, Saxena NS, Sreekala MS, Thomas S (2000) Effect of treatment on the thermal conductivity and thermal diffusivity of oil-palm-fiber-reinforced phenolformaldehyde composites. J Polym Sci, Part B: Polym Phys 38(7):916–921 [271] Feng D, Caulfield DF, Sanadi AR (2001) Effect of compatibilizer on the structure-property relationships of kenaffiber/polypropylene composites. Polym Compos 22(4):506–517 [272] Mofokeng JP, Luyt AS, Tábi T, Kovács J (2012) Comparison of injection moulded, natural fibre-reinforced composites with PP and PLA as matrices. J Thermoplast Compos Mater 25(8):927–948 [273] Azwa ZN, Yousif BF (2013) Characteristics of kenaf fibre/ epoxy composites subjected to thermal degradation. Polym Degrad Stab 98(12):2752–2759 [274] Costa C, Fonseca AC, Serra AC, Coelho JFJ (2016) Dynamic mechanical thermal analysis of polymer composites reinforced with natural fibers. Polym Rev 56(2):362–383 [275] Pothan LA, Oommen Z, Thomas S (2003) Dynamic mechanical analysis of banana fiber reinforced polyester composites. Compos Sci Technol 63(2):283–293 [276] 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. Compos Sci Technol 65(7–8):1077–1087 [277] Pan P, Zhu B, Kai W, Serizawa S, Iji M, Inoue Y (2007) Crystallization behavior and mechanical properties of biobased green composites based on poly (l-lactide) and kenaf fiber. J Appl Polym Sci 105(3):1511–1520 [278] Mylsamy K, Rajendran I (2011) The mechanical properties, deformation and thermomechanical properties of alkali J Mater Sci (2020) 55:829–892 [279] [280] [281] [282] [283] [284] [285] [286] [287] [288] [289] [290] treated and untreated Agave continuous fibre reinforced epoxy composites. Mater Des 32(5):3076–3084 Hebel DE, Javadian A, Heisel F, Schlesier K, Griebel D, Wielopolski M (2014) Process-controlled optimization of the tensile strength of bamboo fiber composites for structural applications. Compos Part B Eng 67:125–131 Lee JM, Heitmann JA, Pawlak JJ (2007) Local morphological and dimensional changes of enzyme-degraded cellulose materials measured by atomic force microscopy. Cellulose 14(6):643–653 George M, Mussone PG, Abboud Z, Bressler DC (2014) Characterization of chemically and enzymatically treated hemp fibres using atomic force microscopy and spectroscopy. Appl Surf Sci 314:1019–1025 Keshk S, Suwinarti W, Sameshima K (2006) Physicochemical characterization of different treatment sequences on kenaf bast fiber. Carbohydr Polym 65(2):202–206 Bilba K, Arsene M-A, Ouensanga A (2007) Study of banana and coconut fibers: botanical composition, thermal degradation and textural observations. Bioresour Technol 98(1):58–68 Spinacé MAS, Lambert CS, Fermoselli KKG, De Paoli M-A (2009) Characterization of lignocellulosic curaua fibres. Carbohydr Polym 77(1):47–53 Saha P, Manna S, Chowdhury SR, Sen R, Roy D, Adhikari B (2010) Enhancement of tensile strength of lignocellulosic jute fibers by alkali-steam treatment. Bioresour Technol 101(9):3182–3187 Sawpan MA, Pickering KL, Fernyhough A (2011) Effect of various chemical treatments on the fibre structure and tensile properties of industrial hemp fibres. Compos Part A Appl Sci Manuf 42(8):888–895 De Rosa IM, Kenny JM, Maniruzzaman M, Moniruzzaman M, Monti M, Puglia D, Santulli C, Sarasini F (2011) Effect of chemical treatments on the mechanical and thermal behaviour of okra (Abelmoschus esculentus) fibres. Compos Sci Technol 71(2):246–254 Fiore V, Valenza A, Di Bella G (2011) Artichoke (Cynaracardunculus L.) fibres as potential reinforcement of composite structures. Compos Sci Technol 71(8):1138–1144 Arrakhiz FZ, El Achaby M, Kakou AC, Vaudreuil S, Benmoussa K, Bouhfid R, Fassi-Fehri O, Qaiss A (2012) Mechanical properties of high density polyethylene reinforced with chemically modified coir fibers: impact of chemical treatments. Mater Des 37:379–383 Reddy KO, Maheswari CU, Shukla M, Song JI, Rajulu AV (2013) Tensile and structural characterization of alkali treated Borassus fruit fine fibers. Compos Part B Eng 44(1):433–438 891 [291] Neto ARS, Araujo MAM, Souza FVD, Mattoso LHC, Marconcini JM (2013) Characterization and comparative evaluation of thermal, structural, chemical, mechanical and morphological properties of six pineapple leaf fiber varieties for use in composites. Ind Crops Prod 43:529–537 [292] AlMaadeed MA, Kahraman R, Khanam PN, Al-Maadeed S (2013) Characterization of untreated and treated male and female date palm leaves. Mater Des 43:526–531 [293] Lu X, Zhang MQ, Rong MZ, Shi G, Yang GC (2003) Selfreinforced melt processable composites of sisal. Compos Sci Technol 63(2):177–186 [294] Alavudeen A, Rajini N, Karthikeyan S, Thiruchitrambalam M, Venkateshwaren N (2015) Mechanical properties of banana/kenaf fiber-reinforced hybrid polyester composites: effect of woven fabric and random orientation. Mater Des 1980–2015(66):246–257 [295] Faruk Hossen M, Hamdan S, Rahman MR, Islam MS, Liew FK, hui Lai JC, Rahman MM (2016) Effect of clay content on the morphological, thermo-mechanical and chemical resistance properties of propionic anhydride treated jute fiber/polyethylene/nanoclay nanocomposites. Measurement 90:404–411 [296] Saba N, Tahir P, Jawaid M (2014) A review on potentiality of nano filler/natural fiber filled polymer hybrid composites. Polymers 6(8):2247–2273 [297] Denault J, Labrecque B (2004) Technology group on polymer nanocomposites–PNC-Tech. Industrial Materials Institute. National Research Council Canada, vol 75. Québec, Canada [298] Huang X, Netravali A (2007) Characterization of flax fiber reinforced soy protein resin based green composites modified with nano-clay particles. Compos Sci Technol 67(10):2005–2014 [299] Sen A, Kumar S (2010) Coir-fiber-based fire retardant nano filler for epoxy composites. J Therm Anal Calorim 101(1):265–271 [300] Alamri H, Low IM (2013) Effect of water absorption on the mechanical properties of nanoclay filled recycled cellulose fibre reinforced epoxy hybrid nanocomposites. Compos Part A Appl Sci Manuf 44:23–31 [301] Lim KH, Majid MSA, Ridzuan MJM, Basaruddin KS, Afendi M (2017) Effect of nano-clay fillers on mechanical and morphological properties of Napier/epoxy Composites. In: International conference on applications and design in mechanical engineering (ICADME 2017), Shenzhen, China, 2017. IOP Publishing, pp 1–9 [302] Ashik KP, Sharma RS, Raghavendra N (2017) Effect of filler on mechanical properties of natural fiber reinforced composites. Asian J Chem 29(8):1697–1701 892 [303] Afroughsabet V, Biolzi L, Ozbakkaloglu T (2016) Highperformance fiber-reinforced concrete: a review. J Mater Sci 51(14):6517–6551. https://doi.org/10.1023/A: 1018651218515 [304] Zimniewska M, Wladyka-Przybylak M (2016) Natural fibers for composite applications. In: Fibrous and textile materials for composite applications. Springer, Berlin, pp 171–204 [305] Dweib MA, Hu B, Shenton Iii HW, Wool RP (2006) Biobased composite roof structure: manufacturing and processing issues. Compos Struct 74(4):379–388 [306] Ticoalu A, Aravinthan T, Cardona F (2010) A review of current development in natural fiber composites for structural and infrastructure applications. In: 2010. Engineers Australia, pp 113–117 [307] Khondker OA, Ishiaku US, Nakai A, Hamada H (2005) Fabrication mechanical properties of unidirectional jute/PP composites using jute yarns by film stacking method. J Polym Environ 13(2):115–126 [308] Yousif BF, Ku H (2012) Suitability of using coir fiber/ polymeric composite for the design of liquid storage tanks. Mater Des 1980–2015(36):847–853 [309] Bilisik K, Karaduman NS, Bilisik NE (2016) Fiber architectures for composite applications. In: Fibrous and textile materials for composite applications. Springer, Berlin, pp 75–134 [310] MarketsandMarkets (2017) Biocomposites market by fiber (wood fiber and non-wood fiber), polymer (synthetic and natural), product (hybrid and green), end-use industry (transportation, building & construction and consumer goods), and region—global forecast to 2022, US [311] GrandViewResearch (2018) Biocomposites market size, share & trends analysis report by fiber type (wood, nonwood), by polymer type (natural, synthetic), by product type (green, hybrid), by end use, and segment forecasts, 2018–2025, US [312] Yang Y, Boom R, Irion B, van Heerden D-J, Kuiper P, de Wit H (2012) Recycling of composite materials. Chem Eng Process Process Intensif 51:53–68 [313] TransparencyMarketResearch (2018) Global biocomposites market to benefit from increased demand in construction J Mater Sci (2020) 55:829–892 [314] [315] [316] [317] [318] [319] [320] [321] [322] and automotive industry, to exhibit 9.46% CAGR by 2025 New York, US Wötzel K, Wirth R, Flake M (1999) Life cycle studies on hemp fibre reinforced components and ABS for automotive parts. Die Angewandte Makromolekulare Chemie 272(1):121–127 Müssig J, Schmehl M, Von Buttlar HB, Schönfeld U, Arndt K (2006) Exterior components based on renewable resources produced with SMC technology—considering a bus component as example. Ind Crops Prod 24(2):132–145 Zah R, Hischier R, Leão AL, Braun I (2007) Curauá fibers in the automobile industry–a sustainability assessment. J Clean Prod 15(11–12):1032–1040 Schmehl M, Müssig J, Schönfeld U, Von Buttlar HB (2008) Life cycle assessment on a bus body component based on hemp fiber and PTPÒ. J Polym Environ 16(1):51–60 Luz SM, Caldeira-Pires A, Ferrão PMC (2010) Environmental benefits of substituting talc by sugarcane bagasse fibers as reinforcement in polypropylene composites: ecodesign and LCA as strategy for automotive components. Resour Conserv Recycl 54(12):1135–1144 Wang J, Shi SQ, Liang K (2013) Comparative life-cycle assessment of sheet molding compound reinforced by natural fiber vs. glass fiber. J Agric Sci Technol B 3:493–502 La Rosa AD, Cozzo G, Latteri A, Recca A, Björklund A, Parrinello E, Cicala G (2013) Life cycle assessment of a novel hybrid glass-hemp/thermoset composite. J Clean Prod 44:69–76 Dissanayake NPJ, Summerscales J, Grove SM, Singh MM (2009) Life cycle impact assessment of flax fibre for the reinforcement of composites. J Biobased Mater Bioenergy 3(3):245–248 Dissanayake NPJ, Summerscales J, Grove SM, Singh MM (2009) Energy use in the production of flax fiber for the reinforcement of composites. J Nat Fibers 6(4):331–346 Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.