Processing and properties of jute (Corchorus olitorius L.) fibres and their sustainable composite materials: a review

M. Ramesh *a and C. Deepa b
aDepartment of Mechanical Engineering, KIT-Kalaignarkarunanidhi Institute of Technology, Coimbatore-641402, Tamil Nadu, India. E-mail: mramesh97@gmail.com
bDepartment of Artificial Intelligence and Data Science, KIT-Kalaignarkarunanidhi Institute of Technology, Coimbatore-641402, Tamil Nadu, India

Received 12th September 2023 , Accepted 6th December 2023

First published on 14th December 2023


Abstract

Materials that are non-toxic and environmentally friendly are in demand due to the lack of resources and increasing contamination of the environment. In this case, fibre-reinforced composites (PFRCs) have become alternatives to synthetic fibre-reinforced composites (SFRCs) in many applications. Among the PFRCs, jute fibre-reinforced composites (JFRCs) have attracted attention from researchers and have been industrialized in many fields. This review focuses on the research prospects and challenges in jute plant cultivation, harvesting techniques, structure and properties of jute fibres, the effect of surface modification, fabrication methods, and properties of JFRCs. Furthermore, the structure–property relationships, finite element analysis (FEA), machining characteristics and life-cycle assessment (LCA) of JFRCs are discussed in detail. Also, a summary of the innovations in the field of JFRCs is provided. Finally, considering the promising future of this bio-material, several open questions and ideas for its transformation are discussed. According to the findings from the literature, it can be concluded that jute fibres are potential alternative reinforcing constituents to synthetic fibres.


1. Introduction

The use of composite materials has increased drastically in the last few decades.1,2 These materials are heterogeneous in nature and formed by the combination of reinforcing constituents such as fillers, particulates, and fibres with matrices.3,4 Furthermore, they have the ability to replace traditional materials such as metals and alloys by providing high strength and stiffness.5 Presently, the focus on protecting the environment has sparked interest in the research and fabrication of sustainable and bio-degradable materials.6–19 Accordingly, thin films, plastics and adhesives have been replaced by lightweight, convenient and attractive materials in diverse applications.20,21 Specifically, it is necessary to reduce the use of debris-creating pollution to maintain a safe and green environment.22–24 Extensive studies have proven the potential of natural fibres to replace synthetic fibres as reinforcement materials in composites,25–30 where several experimental trials have performed to replace synthetic fibres with natural fibres. Synthetic fibres, together with plant fibres such as flax, jute, sisal, hemp, ramie, and kenaf have been reported to show a good reinforcing effect.31–40

Natural fibre-reinforced composites (NFRCs) have become popular due to their degrading capability.41–43 Recent experimental studies demonstrated the fabrication of composite materials that are entirely biodegradable employing natural reinforcing fibres,44,45 where NFRCs have a positive effect on the environment and possess attractive properties.26,46–49 Natural fibres of different origins are readily available, allowing researchers to manufacture composites for housing and other applications.19,50,51 These fibres are light-weight and possess high stiffness and moderate mechanical and thermal properties, and thus can also be used as reinforcements in composites.52–65 Among the natural fibres, cellulose-based plant fibres have emerged as significant materials in recent years.66–70 The mechanical properties of plant fibres are dependent on the season of harvesting, area, local climate, rain, soil conditions, part of the plant harvested and plant maturity.22,23,71,72 Many research articles have explained the importance of several plant fibres and their composites.7,11,13,14,62,71,73–78 In this review, we focus on the processing, properties and applications of jute fibres and their composites. The structure of the review is presented in Scheme 1.


image file: d3ta05481k-s1.tif
Scheme 1 Structure of this review.

2. Background

2.1 Plant fibres

Plant fibres are beneficial raw materials for several industrial applications.78–80 They have numerous key advantages, including light-weight, high resistance, bio-degradability, CO2 neutralization, renewability, recyclability, moderate mechanical and physical strength, non-toxicity, strong degree of curvature without fracture, good insulation properties, durability, low wear during machining, good aspect ratio, acceptable modulus, increased vibration damping and easy surface modification.7,9,13,48,50,64,81–123 The fibres derived from the outer stem of plants possess very strong mechanical properties.124–127 However, the swelling of the fibres leads to micro-cracking, and ultimately a decrease in the mechanical strength of the composites.109,126,128 Many possible benefits of the use if fibres may include the global availability of diverse plant fibres, creation of employment opportunities in agriculture, improvement of the non-edible farm economy and low energy utilization.129–131 A description of plant fibres is presented in Table 1.
Table 1 Brief description of plant fibres
Plant Species Annual production (lakh ton) Major producer Advantages Disadvantages Ref.
Abaca Musa textilis 0.70 Philippines, Ecuador and Costa Rica Strong, hard and high tensile strength Low production 11, 132 and 133
Alfa Stipa tenacissima Southern Spain, North Africa Good modulus Low strength and elongation 132 and 134
Bamboo Bambuseae 300 Asia High tensile strength Low tensile modulus 11, 132, 133 and 135
Banana Musa acuminata India, Australia and South Africa Good tensile strength High moisture absorption 132
Coir Cocosnucifera 1.00 India and Sri Lanka Low density, high elongation and low thermal conductivity Low strength and modulus 11, 132, 133 and 136
Cotton Gossypium sp. 250.00 India, Pakistan, China and USA High production Low mechanical strength 11, 132, 133 and 137
Curaua Brazil Good elongation at break and tensile strength, and low density High moisture absorption 25 and132
Date palm Phoenix dactylifera India, Pakistan, United state, Middle East, Northern Africa and Canary Islands Good elongation Low modulus and strength 132
Flax Linum usitatissimum 8.30 France, China, Belgium, Ukraine, Canada, United state and India High specific mechanical properties Diverse properties due to environmental effects 11, 132, 133, 138 and 139
Hemp Cannabis sativa 2.14 China and Central Europe High growth rate, doesn't need of herbicides, pesticides and fertilizers low elongation at break 11, 132, 133, 139 and 140
Henequen Agave fourcroydes 0.30 Mexico, Asia and South America Good toughness and resiliency Low modulus and elongation art break 11, 132 and 133
Jute Corchorus capsularis 23.00 India, Bangladesh, Nepal, China, Indonesia, Thailand and Brazil Fine texture, heat and fire resistance properties Low elongation at break due to high lignin content, high moisture absorption and low tensile strength 11, 46, 132, 133, 141 and 142
Kenaf Hibiscus cannabinus 9.70 India, China and Thailand High growth rate (3 m in months) Brittle and difficult to process 11, 23, 132, 133 and 139
Nettle Urtica dioica Central Europe and UK Good tensile strength and modulus Slow growth rate 143
Oil palm Elaeis guineensis West Africa and Congo Basin High diameter Low tensile strength 11, 132 and 133
Pineapple fibre (PALF) Ananus comosus Brazil and Philippines Mechanical properties comparable to jute fibre Low microfibrillar angle 132
Ramie Boehmeria nivea 1.00 China and Eastern Asia Strong natural fibre and high specific tensile strength and modulus Stiff, brittle, and low elongation 11, 132, 133, 11, 132, 133 and 139
Roselle Hibiscus sabdariffa China, Sudan, Thailand, West Africa and India Good elongation Slow growth rate 132
Sisal Agave sisalana 3.78 China, Brazil, Kenya and Tanzania Good strength and durability Low tensile strength, modulus and toughness 11, 132, 136 and 144
Sugarcane bagasse 750 India Low growth time and high production Low strength and modulus 11, 136 and 145


Among the plant fibres, jute fibres appear to have great potential as a reinforcement material due to their high strength and rigidity.146–152 Furthermore, jute fibres can be used effectively in the production of composites given that they possess attractive physical and mechanical properties.153–161 Their characteristics significantly depend on the geographical region, surrounding climate and processing methods.156,162 However, they also undoubtedly have bottlenecks as a reinforcement in polymer matrices.25,52,56,163,164 In this case, jute fibres can be surface treated to enhance their properties, where chemical modification has been proven to be a favorable method for surface treatment.14,165–173 Several experiments have been conducted for the analysis of characteristics of jute produced in different ways.174–182

2.2 Plant fibre-reinforced composites

Recently, the use of PFRCs instead of SFRCs has been increasing due to their economic and environmental advantages in numerous practical applications.183–185 Recent studies have indicated the substantial progress in the application of renewable materials with enhanced support for global sustainability.11,25,82,90,186–188 Plant-based agro-waste and particulates originate from different parts of plants, which can be used as reinforcements in the development of bio-based materials.189 PFRCs have been applied in several fields and widely accepted by scientists because of their superior performances.190 Plant fibres have high/moderate thermal resistance and thermal stability when recycled, but they have issues when combined with pure bio-degradable polymers due to their low resistance and transition temperature.153,191–193 Fibres from bamboo, banana, abaca, roselle, sugar-cane bagasse, kusha grass, cantala, hemp, sisal, coconut, flax, kenaf, milkweed, curaua, coir, hay, cocos, Cordenka and jute can reinforce polymer matrices.52,64,110,112,115,194–216 Bird feathers, soybean seeds and oils, wheat straw, starch-based polymers and algae extracts are also used as composite reinforcements.26,217–229

Short-fibre composites have gained significant attention in the last two decades due to their processing advantages, anisotropic properties and rapid dispersion.25,58,156,230,231 Consequently, various PFRCs have been studies and their properties investigated.57,138,232–247 Many studies have extended their experimental findings related to different forms of plant fibres.100,248–251 A few research groups have devoted their efforts to determining how different factors influence the characteristics of PFRCs.252–256 The curiosity of materials investigators has led to growing interest in the review of PFRCs.13,26,71,74,75,77,78,90,140,186,257–259 Predominantly fibres from bast and leaf, and their reinforcements studied and compared to other fibres.23,54,100,138,260–277

3. Jute plants

3.1 History

Jute is among the most essential natural fibres, which is grown extensively in China, Thailand, UK, and Vietnam besides India and Bangladesh.278,279 Jute was discovered in Dunhuang, Gansu Province, China around BC-221AD. It is believed that jute was produced during the Western Han Dynasty.280–282 Historical records have shown the use of jute fibre products primarily in India during the Mughal Empire (1542–1605). Indian jute was merchandized by the British East India Company. During the 1800s, the jute fibre-processing industry was established in Dundee.283,284

3.2 Cultivation and harvesting

Jute, a crop cultivated in the rainy season, grows faster in humid and warm climates. This plant, which grows in alluvial, loam, sandy soils, and in places with a temperature above normal and more than 1000 mm of annual rainfall, can also survive in extreme water flooding. After planting, jute takes about 90–120 days to mature, and harvesting occurs amid blooms and blossoms, leading to the development of seeds. Once the plants are around 6 feet in height, they are harvested and graded according to colour, intensity and length.285Fig. 1(a and b) show jute plants. An ideal temperature range for good yields is 22 °C to 38 °C with 75% humidity. Throughout the season, jute plants are axed and preserved in water/dew retting for the separation process, where the fibres are separated from the outer stem of the plant. The stalks are tied into bundles and soaked in water for a few days. This process softens the outer cell wall and breaks the bonding between the outer and the inner fibre stick between hard pectin and it allows the separation of the fibres.285 Jute plant harvesting is described in Fig. 1(c–f).
image file: d3ta05481k-f1.tif
Fig. 1 (a and b) Dry and rainy season jute plants, respectively. (c–f) Harvesting of jute plants and (g–j) extraction of fibres from jute plants (https://www.jute.org).

3.3 Fibre extraction

Jute fibres are abundant, and hence considered to be a prospective source of reinforcements for composites.286 They are extracted from the stem of the jute plant through a combination of processes involving various steps such as cutting, retting, shredding, and drying.250 In most cases, jute fibres are extracted through the retting process. The stems are tied into small bundles and immersed in water for around 28 days, and during this time the outer bark is fully decomposed and fibres are exposed.197 The different types of retting include mechanical retting, water retting, chemical retting, microbial process and dew retting. The choice of retting process is based on the quality of water and cost. The water retting process is the traditional and common method for separating bast/stem fibres, where water is sprayed after the fibre is loosened, and then it is squeezed for dehydration. Subsequently, the separated fibres are washed with clean tap water and dried under sunlight. The obtained fibres are around 3–6 feet long and white or brown in colour, which are tied into bales for delivery to end-users. Upon drying, the bales are sent to jute mills to be processed into jute yarn and hessian.2 The extraction of jute fibres by the retting method is shown in Fig. 1(g–j).285Table 2 presents an overview of the fibre extraction methods, comparing various retting processes and mechanical extraction.
Table 2 Overview of fibre extraction process126,192,287–289
Extraction Process Retting Mechanical method
Dew retting process Water retting process
Procedure Stems are spread on grassy fields for reaction with bacteria, sunlight, and mist under normal air flow Plant stems are submerged in water (rivers, ponds, or tanks) and checked periodically (microbial retting) Fibres are separated using a hammer or/and decorticator
Days 14–21 days 7–14 days Depending on production of fibres
Benefits This process is applicable in heavy dew and water-scarce areas The fibres produced are uniform and of higher quality Produces large quantities of fibres in the minimum time
Disadvantages The separated fibres are very dark in colour with low strength Comparatively high processing cost, environmental concerns and low strength, but better than fibres produced through dew retting process High cost and acceptable quality fibres
For this process, the cultivating lands are occupied and the fibres are contaminated due to fungi


3.4 Jute fibres

Jute is a lignocellulosic fibre, having a low density, high strength and abundant availability. Consequently, researchers have shown interest in applying this fibre as a reinforcement in composites.290–292 Jute is the most flexible fibre with immense potential for the manufacture of composites due to its safety, strong, non-abrasive nature, viscoelasticity, biodegradability, combustibility, compostability, and good acoustic and thermal insulation.293–295

3.5 Fibre structure and composition

3.5.1 Structure. Plant fibres possess a complex micro-structure consisting of hundreds of micro-cells with a typical length of 2–5 mm and a nominal diameter of 0.2 mm, arranged with lignin and natural adhesive.296,297 Basically, plant fibres are multi-cellular in nature, possessing a variety of continuous cells of varying sizes, shapes, and fibre alignments. The central regions of many plant fibres may have a cavity, which is called a lacuna, surrounded by several lumens in their cellular structure. The orientation angle of jute fibre is lower than that of other plant fibres, endowing it with high strength but making it brittle in nature.298 The structural characteristics and high water absorption nature of plant fibres lead to inadequate bonding between fibres and matrices, resulting in fibre pullout during use.299–301 Microfibrils of jute fibre are bound with other plant fibres of identical structure.105 Jute is multi-celled in structure, as shown in Fig. 2(a),26 where its fibre cell wall consists of three layers with primary and secondary walls. Cellulose has the highest concentration, while lignin is the major portion of the cellulose-free middle lamella. Usually, the secondary layer is the thickest and governs the strength and characteristics of the fibres. The long cellulose chains are intertwined in bunches called microfibrils.26Fig. 2(b)61 shows the changes in the chemical structure of jute fibre upon alkaline treatment and the reaction occurring during its processing is presented in Fig. 2(c). Due to the presence of hemicelluloses, lignin, pectin and wax, the surface of the alkaline-treated fibres becomes more polar and rougher.
image file: d3ta05481k-f2.tif
Fig. 2 (a) Structure of a jute fibre. Reproduced from ref. 26 with permission from Elsevier, 2006. (b) Change in its structure by alkaline treatment. Reproduced from ref. 61 with permission from Elsevier, 2008. (c) Reaction of jute fibre with propionic anhydride.
3.5.2 Composition. Jute fibres contain large sub-categories of substances including cellulose, hemicelluloses, lignin, pectin, polyuronide, acetyl matter, fat, ash, nitrogen (N2) and some other components such as aqueous extracts and mineral substances.302–304 Jute fibres composed of lignocelluloses have been observed to possess strongly polarized hydroxyl groups.58,305,306 Details of the composition of the jute fibres are presented in Table 3. The bonding between the cellulose chains ensures their stability and solubility in water and chemical solutions.307 They exist as micro-fibrils in the plant cell-wall and their diameter varies in the range of 3–35 mm.308 The hydroxyl cellulose groups create fibrils by forming hydrogen bonds adjacent to and partially crystallizing other molecular cellulose chains.304 Hemicellulose is the second most important constituent, accounting for 15–30% of the cell-wall.309 Hemicelluloses are bound polysaccharides and closely connected with the cellulose microfibrils. The four major hemicellulose forms based on their structure are xylans, mannans, xyloglucans and glucomannan.291,310 The bonding of hemicellulose is very strong with cellulose fibrils presumably by hydrogen bonds, which is semi-soluble in water. Lignin is a highly complex amorphous, aromatic bio-polymer and has the least water absorption of the components of plant fibres. It is a naturally developed heterogeneous and irregular cross-linked polyphenol consisting of units of phenylpropane. These three elements are responsible for physical properties of the fibres.310–312 Another significant characteristic is that lignin is thermoplastic (i.e., it begins to soften at 90 °C and begins to melt at 170 °C).291
Table 3 Strength and composition of jute fibres
Density (g cm−3) Tensile strength (MPa) Tensile modulus (MPa) Elongation (%) Cellulose (%) Hemicellulose (%) Lignin (%) Ref.
1.4 400–800 10–30 1.8 50–57 8–10 153
256 ± 1.02 31 ± 1.34 0.78 ± 0.05 348
58–63 20–24 12–15 302
1.48 410–780 18–50 1.9 15–20 11–15 314 and 315
1.3 393–773 26.5 1.5–1.8 1
1.4 420–820 10–30 1.8 40–60 316
1.3–1.5 610–780 12–60 1.0–1.9 59–70 15–20 11–15 317
1.3–1.4 393–773 13–26.5 1.2–1.5 61–71.5 13.6–20.4 12–13 318
74.4 15 8.4 319
1.3–1.5 400–800 70 59–61 22.1 15.9 320
1.3 450–550 26–32 1.5–1.8 58–63 12 12–14 321
61–71.5 17.9–22.4 11.8–13 322
73.2 13.6 13.4 310


3.6 Properties of jute fibres

Among the plant fibres, jute stands out due its unique properties,323 which are strongly influenced by factors such as chemical composition, fibre structure, microfibrillar angle and cell-wall thickness.78,246 Variations are seen in these characteristics among plants, leading to significant variations in the properties of the fibres.26,324 The main factor determining the final properties of the fibres is their anatomical origin, given that fibres of different origins typically have different properties. However, factors such as age, growing trends, environmental conditions, and extraction techniques should be considered to affect the properties of the fibres.325 Jute is predominantly polar, owing to the availability of various polar groups on its backbone.326 Also, its tensile strength is in the range of 250–300 MPa, which is around seven-times lower than that of steel, but it has relatively high resistance.327

Research on the measurement of the tensile strength of jute fibres in dry and alkaline environments was carried out and it was found to be very high and increased when submerged in an alkali-solution ranging from 5–32%.328 Many sources have reported that the Young's modulus of jute fibre varies from 20–40 GPa, which is substantially lower than that of glass fibre of about 73 GPa. Discrepancies may occur due to circular cross-section assumptions, in addition to the inherent mechanical variability of natural resins.329–333 The decomposition of the constituents of jute begins with hemicellulose at 290 °C, then cellulose at 332 °C, and finally the lignin at 447 °C. Therefore, the processing of jute fibres was carried out in most experimental studies by compression or injection moulding process below 200 °C.334,335 The physical and chemical characteristics of jute fibres are presented in Tables 4 and 5 lists the main suppliers of jute fibres.

Table 4 Properties of jute fibres
Physical properties Chemical properties
Parameter Value/result Parameter Result
Abrasion resistance Relatively good Dye ability Good, easily dyed
Colour Yellow, brown, golden Effect of acid Destroyed by hot concentrated acids. Dilute acids cannot harm the fibre
Cross section Uneven, thick cell wall with lumen Effect of alkali Strong alkali destroys the fibre and reduces its strength
Diameter 18 μm Effect of bleaching Not affected by oxidizing and reducing bleaching agents
Dimensional stability Good Effect of insects Resistance is good
Heat resistance Good Effect of mildew Better than cotton and linen
Length (inch) 0.2–30 Electrical conductivity Moderate
Moisture regain 13.75% Organic solvent Resistant to organic solvents
Resiliency Not very good Thermal conductivity Poor
Specific gravity 1.48–1.5
Specific heat 0.324
Stretch and elasticity Not good and 2% elongation at break
Tenacity (g per den) 3.5–4.5


Table 5 Major suppliers/vendors of jute fibres
Name of the firm Country Ref.
Alam Fibre Impex Ltd Bangladesh 336–338
Amazon Region 296
Anil Ltd Turkey 339
Babu Bazaar Dhaka, Bangladesh 340
Bangladesh Jute Mills Corporation Dhaka, Bangladesh 341
Bangladesh Jute Research Institute (BJRI) Bangladesh 313 and 342–347
Carol Leigh's Hillcreek Fibre Studio USA 58 and 61
Chandra Prakash & Co. Pvt. Ltd Jaipur, India 317, 348 and 349
Composites Evolution UK 350
Cia Textile Castanhal Brazil 230
Dharmapuri District Tamil Nadu, India 153, 314 and 315
Extra Weave Private Ltd Kerala, India 351 and 352
Fatima-Alyaf Tala-e Jute Industries Ltd Gazipur, Bangladesh 41
Gloster Jute Mill Kolkata, India 353–358
HP Johannesson Trading Sweden 297
Hunan Shangke Co. Ltd China 359
Indarsen Shamlal Pvt. Ltd India 83
Indian Jute Industries Research Association (IJIRA) Kolkata, India 360–362
Janata Jute Mills Ltd Bangladesh 363 and 364
Kamarhatty Company Limited India 365
JB Plant Fibres Ltd UK 366
Kancheepuram District Tamil Nadu, India 367
Konark Jute Mills Orissa, India 56
Krishna Jute Industries India 368
Longtai Weaving Co., Ltd Changshu, China 369 and 370
Local Market Vijayawada, India 302
Local Market Bangladesh 371–374
Local Market India 375 and 376
Maranda Halli Market Tamil Nadu, India 316
Meena Fibres Puduchery, India 377 and 378
Materials Lab of UFSCar Brazil 379
Muszaki Konfekcio Kft Szeged, Hungary 182
Open Market Rajapalayam, Tamil Nadu, India 322
Redbud Textile Tech. Inc. China 380
Sakthi Fibres Pvt. Ltd Chennai, India 381
Schilgen GmbH & Co. Germany 292, 382 and 383
Sisalsul Brazil 384
Sonajute Casablanca, Morocco 385
Spinnerij Blancquaert NV Lokeren, Belgium 59 and 105
Stuart C. Hurlbert & Co. Framingham, MA 386 and 387
Textiles Divers Algerie Algeria 388
Tesac Co. Japan 81
Women Development Organization Uttarakhand, India 389–392
Yano Seikei 393
ZKK, Sdn Bhd Malaysia 321


4. Fibre surface modification

Perfect reinforcement and good interfacial bonding must be ensured to obtain durable composites for industrial applications, enabling the maximum use of the fibre reinforcement capacity.230,394,395 The strength of plant-based fibres can be enhanced through surface modification by the appropriate methods.14,165,168,169,352,396–398 However, plant fibres have some disadvantages, including hydrophilic behavior, poor thermal stability, and in particular poor adhesion with a hydrophobic resin, which results in poor interfacial bonding and low mechanical strengths.235,399 In this case, surface modification173,278,303,400–406 has the ability to solve most of these drawbacks. The limitations of PFRCs can be strengthened through the use of chemical or physical methods, treatment of fibres, or use of coupling agents.37,38,407–409 Slight leaching of the fibres leads to the degradation of hemicellulose, lignin and pectin, weakening their internal structure and causing the fibres to separate. A cleaner and rougher surface is beneficial for fibre/matrix bonding, which promotes mechanical inter-locking and bonding due to the exposed hydroxyl groups on the resin.371 Also physical and chemical treatments38,56,154,163,164,305,375,400,410 have attracted attention from researchers for further investigation due to their many benefits. Consequently, numerous surface treatments, for example, alkaline, acetylene, silane, peroxide and permanganate, benzoylation, electrical discharge, and cyano-ethylation treatments, have been developed by various researchers and are discussed in the literature.

4.1 Physical modification

Several physical modification methods can enhance the cleaning of the fibre surface, consequently improving the solid bonding at the interface.411 Many physical treatments such as plasma treatment,341,412–415 photo-oxidation by UV irradiation,35 ionizing radiation,416 gamma radiation,417 corona treatment,418–422 cold plasma treatment,423 ozone treatment,412 laser treatment,424 boiling under or without pressure,411 and high pressure plasma treatment425 have been used to enhance the compatibility of fibres. Table 6 presents the purpose and effect of various processes of physical treatment.
Table 6 Purpose and effect of physical treatment processes
Treatment type Purpose Effect Ref.
Plasma treatment To achieve an ablation or etching effect, resulting in a better interface with the matrices through mechanical interlocking A semi-industrial atmospheric pressure glow discharge plasma process in order to treat the fibre surface 341, 412–415 and 426
Ozone treatment To maintain mechanical strength Surface modified by placing the fibres in an ozone gas-filled closed container 412
Laser treatment To increase the structural stability of fibres Improving the surface roughness by evaporating the materials from fibre surface using a laser beam 424
Corona treatment To improve the acidity of the fibre surface Activating the cellulose in the fibre to improve the efficiency of the grafting process 419–422
Hornification treatment To increase the dimensional stability, which is known as hornification Surface modified by passing heat waves on the surface 319
UV radiation/ionising radiation To improve the adhesion, wettability, biodegradability and tribological properties 35 and 416


4.1.1 Plasma treatment. Gases such as oxygen (O2),413 helium (He),414 helium and nitrogen (N2),415 and air426 are employed for the modification of the fibre surface. The impact of plasma on the surface morphology of jute fibres and the structure–property relationships in JFRCs were studied. Jute fibres have been tested at atmospheric pressure for 60 s with dielectric plasma treatment.412 Other gases as well create an undesirable localized corona discharge rather than standard treatment. Three gas mixtures were employed in the treatment of jute fibres, as follows: (i) He at the flow rate of 14 L min−1; (ii) He and acetylene at 14 L min−1 and 0.7 L min−1; and (iii) He and N2 at the same flow rate. The fibres were tied to a roller, which moved the plasma source. A modification was performed for the roller of 5, 25, 50 and 100 revolutions. The fibres were treated at a speed of 0.424 s per rev using the plasma source. The equipment was operated at 970 W and the power supply frequency was 90 kHz.341
4.1.2 Ozone treatment. Ozone or oxy-fluorine gas was used in this process for the modification of the jute fibre surface. This treatment was done by placing the fibre in an air-tight jar completely filled with ozone for 1 h. An inter-connected jar with an ozone generator produced ozone gas continuously. Subsequently, the fibres were exposed to the gas at 20 °C in an open atmosphere by maintaining the flow rate of 50 L h−1 for 9 h.412
4.1.3 Laser treatment. Laser irradiation with a commercial CO2 pulse infra-red laser was performed on the surface of the fibre, producing laser beams. The sample was held in place of a concave–convex lens and a joule meter with the help of a sample holder. The parameters used in the determination of the laser marking intensity included speed, duty cycle, and frequency markings. The speed of the labeling was fixed at 200 bits per ms, service cycle at 50% and 5 kHz frequency. A convex–concave lens was employed to get a clear focus on the laser beam, which allowed the light to be altered in the irradiation area with a size in the range of 0.24–1.20 cm2. The beams were connected with the fibres through material evaporation, thermal decomposition and surface roughness modification.424
4.1.4 Corona treatment. This treatment has been reported to increase the bonding resistance of composites. Bataille et al.422 performed corona treatment as a type of pretreatment process to create active sites on the cellulose surface in the fibre and increase the effectiveness of its bonding. Research has demonstrated that the contact angle of a urea-formaldehyde droplet on a surface decrease by increasing the concentration of corona treatment.419 In this case, the improved wettability contributed to enhancing the bond strength. Corona treatment increases surface energy, which is mainly due to the increase in the content of carboxy and hydroxy groups. Belgacem et al.420 demonstrated the use of high-alpha cellulose, hard-wood fibre powder as a reinforcement material in polypropylene (PP), substantially improving the yield stress and elastic modulus of the composite. The extent of the transition depends on the treatment conditions, duration and intensity of corona. Dong et al.421 observed a decline in the property of polyethylene composites loaded with cellulose whether one or both compounds were corona treated.
4.1.5 Hornification treatment. Hornification is the process of stiffening the polymeric structure in the cells of fibres to improve their dimensional stability. It was reported that the fibres were kept in water at 22 °C for 3 h. Drying was performed at 80 °C in a furnace with a built-in electronic temperature monitor to measure the mass loss during the drying cycle. The furnace was programmed to heat to 80 °C at a rate of 1 °C min−1 and hold this temperature for 16 h. Then, it was cooled to 22 °C to prevent possible thermal shock to the fibres.319

4.2 Chemical modification

Plant fibres have hydrophobic surfaces and anisotropic internal structures, and therefore it is difficult for them to act as reinforcing materials in polymer matrices. Accordingly, chemical treatments have been developed for the modification of the surface of fibres.63,368,427–433 Several researchers reported the chemical treatment of jute fibres to enhance their hydrophilicity and properties of JFRCs.154,155,202,361,434–441 A few authors extensively investigated the effects of different forms of chemical treatment on plant fibres, in particular jute fibres.263,279,442 To date, various chemical treatments have been reported, using chemicals such as NaOH,38,58,61,63,230,302,361,371,382,386,387,441,443–451 silane,231,386,452–457 alkali and silane,458,459 potassium permanganate (KMnO4),356–358,460 maleic anhydride polypropylene (MAPP),58,305,356–358,375,461–466 stearic acid,356–358,467 bleaching,468 toluene diisocyanate (TDI),355–358 sodium silicate,417 UV radiation monomer grafting,383,469,470 maleic anhydride-grafted polypropylene (MAgPP),305,462,471,472 detergent washing, dewaxing, acetic acid treatment,443 graft co-polymerization,291,473–477 deposition of nano-SiO2 (ref. 359 and 478) and other chemical treatments.102,342,365,378,409,412,434,479–485Table 7 shows the different chemicals used and the effect of various chemical surface treatment methods.
Table 7 Chemicals used and effect of chemical surface treatment processes
Treatment type Chemical agents Effects
Alkaline treatment Sodium hydroxide Reduces fibre diameter
Mercerization Sodium hydroxide Mercerization of fibre-length with pre-evaluated tension or free shrinkage mode
Silane treatment Silane solution Strengthening of the interfacial relationship between fibres and matrix
Benzoylation treatment Benzoyl chloride Makes fibres hydrophobic
Bleaching/peroxide treatment Hydrogen peroxide Improves the adhesion of fibres with the matrix
Graft co-polymerization Methyl methacrylate, acrylonitrile, acrylamide, ethyl acrylate, and styrene Reduces absorption of moisture by fibres
Sodium chlorite treatment Sodium chlorite Removes moisture from fibres
Acrylation and acrylonitrile grafting Acrylic acid Bonding capacity and stress transfer of the interface increase
Esterification Acetic anhydride, dodecanoic acid, hexanoic acid, oleoyl chloride, octadecanoic acid, and docosanoic acid Improves fibre/matrix bonding
Permanganate treatment Potassium permanganate Improves wettability and thermal stability of fibres
Enzymatic treatment Pectinase, laccase, cellulase and xylanase Enhances the linking/meshing of fibers with the matrix
Stearic acid treatment Stearic acid Eliminates the extra-particles from the fibre surface


4.2.1 Alkaline treatment. This is a basic chemical surface modification process for plant fibres when reinforced with polymer matrix.63,157 The addition of NaOH to plant fibres facilitates the ionization of the hydroxyl group.157,400 In one study, NaOH treatment reduced the fibre diameter by 13.7% and weight by 21%.302 According to Arao et al.,81 alkali-treated JFRCs showed improved properties. NaOH treatment resulted in the removal of excess non-cellulosic materials, and thus weakened the fibre, reducing the strength of the composite.486 The influence of NaOH treatment on the properties of laminated JFRCs was investigated by Karabulut et al.122 The fibres were submerged in NaOH solution for 15 days at 25 °C and the results were analyzed using the Weibull distribution. It was observed that this treatment process had a significant influence on the mechanical properties of the fibres.

In a study, the fibres were soaked in distilled water containing 0.5 wt% alkali solution and held for 2 h at 25 °C, then washed in water, allowed to dry at 27 °C, and then they were placed in a heater for 4 h at 80 °C.58 Also, fibres were immersed in 10% NaOH solution for 3 h at room temperature and then heated at 100 °C, where 1 mL of 10% alkali solution was added per 0.15 g of fibre. Then, they were rinsed with water after treatment and left to dry for 48 h.230 Subsequently, they were washed and dried at 60 °C for one day, pre-treated at 55 °C for 1.5 h with 0.1% H2SO4 solution, and then washed again. The washed fibres were subjected to the following processing: immersed in a solution of (i) 2% NaOH, 0.2% Na2SiO3 and Na2P3O10, as well as 0.15% Na2SO3 at 100 °C for one hour, (ii) 5% NaOH at 20 ± 1 °C for 4 h, and (iii) 20 ± 1 °C for 4 h in 1.5% catalyst and 98.5% ethanol solution. Finally, the fibres were rinsed and dried at 60 °C in an oven for one day.371,487

In one study, jute fibres were treated at 30 °C in 5% NaOH solution for 4 and 8 h and rinsed with pure water to eliminate the NaOH clinging to the fibre surface, and then neutralized with diluted hydrochloric/acetic acid. The fibres were dried for 48 h at 30 °C, followed by 6 h drying in an oven at 100 °C.441 Then, they were immersed for 2 h in 5% NaOH solution, cleaned with water and soaked for 1 h in a 2% acetic acid solution,386,387 dewaxed at 50 °C for 2 days, dried at 80 °C for 4 h, and then heated for 30 min at 100 °C. The dewaxed fibres were modified with 1% NaOH for 4 h, the residual alkali removed, rinsed with water again, and dried.382 Finally, the fibres were washed for 1 h with 4% NaOH solution, dried overnight and heated at 105 °C for 60 min after treatment.361

4.2.2 Mercerization. Mercerization is applied to strengthen the interfacial relationship between the fibres and the matrix.488,489 Comprehensive structural modifications may be caused by fibre mercerization, which is dependent on the treatment conditions such as solution concentration, treatment time, fibre tension, and temperature.490–492 The fibres were soaked at around 25 °C for 2 h in an NaOH solution.61 This mercerizing system enabled the systematic and continuous mercerization of the fibre length with pre-evaluated tension or mercerization in free shrinkage mode.493
4.2.3 Silane treatment. A silane agent was prepared and stirred for 5 min in methanol solution using a magnetic stirrer. Subsequently, the fibres were soaked in the prepared mixture for 60 min. The treated fibres were rinsed using water and dried for 12 h.386 In another study, the fibres were immersed in a 0.5 wt% silane solution for 1 min using an amino-silane binding solution and kept at 50 °C for 4 h.231,452 Also, the solvent was made by dissolving liquid oligomeric siloxane solution at 1 wt% and jute fabrics were alkalized, and later sonicated in an ultrasonic bath for 5 min. Then they were plunged in the oligomeric siloxane for 1 h and the treated jute fabrics were eventually dried for 24 h at 608 °C.339
4.2.4 MAPP/MAHgPP treatment. MAPP is the most effective coupling agent to enhance the interfacial adhesion between the fibres and matrix.99,105,494–508 Ultimately, the capability of MAPP to improve the strength of the composites relies on factors such as the fibre structure, matrix miscibility, and composite processing conditions.162,305,504–509 The mechanical properties of JFRCs increased through the treatment of the fibres with up to 6 wt% MAPP.463–466 The mechanical strength increased to 50%, but the energy at the JFRC interface decreased due to the impact loading when the fibres were treated with 0.1% MAPP solution at 100 °C for 5 min.305,462 The fibres were washed in toluene solvent at 100 °C for 5 min with 1% MAPP solution, and then rinsed in toluene and dried at 60 °C.375 Park et al.58 tested the interfacial properties of JFRCs and found that the modification of the fibres could enhance the performance of the composite. During the winding cycle, MAPP and acetone solutions were coated all the fibre layers, which were then left in an oven for 30 min at 60 °C to remove the acetone solution from the fibre surface.356–358
4.2.5 Graft co-polymerization. Some researchers extensively studied the development of vinyl-based polymers and the effects of methyl methacrylate, acrylonitrile, acrylamide, ethyl acrylate, and styrene treatments on jute fibres.473–477 Samal et al.473 observed an improvement in the absorption of moisture by the fibres on co-polymerization with styrene and methacrylate. Also, acrylonitrile, styrene or vinyl acetate graft co-polymerization on jute enhanced the condensation and thus improved the moisture sensitivity of the fibres.474 Mohanty et al.475 reported that the co-polymerization of acrylonitrile on jute fibres enhanced the thermal stability of the resulting composites. X-rays were used in the study of the co-polymerization of vinyl monomers on jute graft.476 Tripathy et al.477 reported the co-polymerization of methyl methacrylate on chemically modified fibres using the redox method of KMnO4 malonic acid. They successfully enhanced the inter-laminar strength by decreasing the amount of resin, filling higher fibre levels and decreasing the absorption of moisture in the composite with the use of these modifications. Then, the fibres were co-polymerized using different chemicals, resulting in a 10–30% increase in strength and modulus. Polymethylmethacrylate synthesis is also successful in this regard, albeit to a lesser degree.291
4.2.6 Esterification. Anhydride reaction of fibres is known as esterification. This modification involves alkali-treatment accompanied by esterification.510 Several chemicals, for example, acetic-anhydride, dodecanoic-acid, hexanoic-acid, oleoyl-chloride, octadecanoic-acid, and docosanoic-acid, have been employed for enhancing the fibre/matrix bonding during esterification.409,479,480 Jute fibres were soaked in 5% NaOH and mixed well at 27 °C for 1 h. The treated fibres were immersed in glacial acetic acid at 27 °C for 1 h. Then, the fibres were submerged again for 5 min in propionic anhydride containing concentrated H2SO4. Finally, the fibres were rinsed with water and dried for 12 h at 80 °C.345
4.2.7 Oxidation. Extracted jute fibres were dried in an air-circulation chamber for 5 h at 105 °C and soaked in a liquor-based sodium solution in the ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]16. Oxidation was performed at 70 °C for 5 h, and then the oxidized fibres were then cooled and isolated. Subsequently, they were washed thoroughly with running tap water. Lastly, the washed oxidized fibres were dried at atmospheric temperature, and again dried in an oven for 6 h at 105 °C.342
4.2.8 Enzymic treatment. Enzyme treatment can be used as an effective, low-cost and eco-friendly method for the processing of NFRCs.511 A solution was prepared by mixing 1% of Texazym DLG, 3% of Texazym BFE and 0.2 g Texawet DAF in distilled water and jute fibres were soaked in this enzymic mixture at 50 °C for 2 h. Then, the fibres were rinsed with clean water and dried for 48 h at 30 °C.412 In another experiment, jute fibres were treated through the use of laccase and multi-enzyme synergistic effect, resulting in an improvement in the mechanical properties of JFRCs due to improvement in the cross-coupling of lignin by the enzymes.370 Also, jute fibres were treated with enzyme solutions such as pectinase, laccase, cellulase and xylanase in different ratios. The enzymes were observed to destroy the structures of hemicelluloses, lignin and pectin, which resulted in a decrease in the diameter and increase in the aspect ratio of the fibres.511
4.2.9 Stearic acid treatment. During the winding cycle, a mixture of stearic acid in acetone solution was added to the layers of the matrix and jute fibres. Subsequently, the metal plate containing the treated fibres was held for 30 min at 60 °C to remove acetone from the surface. Then, it was kept in a mould for 9 min at 205 °C under the load of 0.2 MPa.356,357 Finally, acetone was added to eliminate the extra particles from the fibre surface during the winding of each layer.358
4.2.10 KMnO4 treatment. During the winding of fibres, a mixture of KMnO4 and acetone solution was prepared and added to the matrix and jute fibres. Then, the metal plate with the prepared matrix and fibres was held for 30 min in an air-circulating oven at 60 °C to remove acetone from its surface, and eventually moulded by compression moulding.356–358
4.2.11 Bleaching. Bleaching, which is an essential process for the creation of glossy fabric through use of hydrogen peroxide (H2O2), produces pure white colour fibres with reduced tensile strength.468 Bleaching employing peracetic acid produces less tenacity loss and whiteness without impairing the feel of the fabric. Therefore, bleaching in two steps is performed using both H2O2 and peracetic-acid, resulting in better properties. A change in the colour of jute fibres was seen from light golden brown to a dark colour due to the treatment and bleaching with H2O2 developed light yellow based brown fibres, showing the partial removal of lignin, wax and fatty portions.448 The effect of this treatment curing time was observed at different temperatures, and the fibres were observed to be well separated and isolated when treated for 6 h due to the formation of irregularities on their surface.449 An increase in the crystallinity of the fibres was seen only after 6 h. An increase in the modulus of jute fibre by 12% was seen within 4 h, after which it increased by 68% and 79% when treated for 8 h.
4.2.12 Toluene diisocyanate treatment. A solution of toluene diisocyanate in chloroform was prepared during winding and applied to the jute fibres and matrix. The prepared matrix and fibres were kept in a hot chamber at 100 °C for clearing the chloroform particles from the fibre surface and moulded by compression moulding process.356–358
4.2.13 Nano-SiO2 deposition. The deposition of nanoparticles on the surface of fibres is one of the most promising methods to enhance their interfacial bonding.512,513 Liu et al.478 used a sol–gel coating of nano-SiO2 on jute fibres and found that nano-SiO2 reacted with the jute fibres to create a stable interface with the hydroxyl groups. Then, the fibres were immersed in ethanol containing ammonia at a concentration of 0.15 to 0.75 mol L−1. This solution was sonicated for one hour and stirred mechanically at 40 °C for 5 h. Lastly, the nano-SiO2-coated jute fibres were washed in water to remove the excess material and then dried for 2 h at 70 °C.359

4.3 Advantages and disadvantages of chemical treatment process

4.3.1 Advantages. The objective of the surface treatment of plant fibres is to increase their bonding strength, and thus the stress transferability in the reinforced materials. The significance of surface-treated plant fibres was seen through the improvement of mechanical strength and dimensional stability of the resultant composites when compared with the untreated fibre-reinforced composites.77 An improvement in the degree of crosslinking in the interface region and strong bonding can be achieved with the help of chemical treatments. Surface modification results in the breaking of bundles of composite fibres into micro-fibrils. This results in a decrease in the fibre diameter and roughness on the fibre surface, further resulting in better fibre–matrix interfacial adhesion with improved mechanical properties.411 Given that chemical treatment results in the removal of lignin and hemicellulose from the plant fibers, it affects their chemical composition, molecular orientation of the cellulose crystallites and degree of polymerization. Chemical surface modification can activate hydroxyl groups, which can create a strong bond with the resin.297,433,434 Effective modification is done through debonding of the hydrogen structure in the fibre, thereby enhancing the surface roughness of the fibre.400 The chemical treatment process removes impurities and hemicelluloses on the fibre surface and enhances the reactivity of the hydroxyl groups with the coupling agent. This treatment results in a finer fibre surface, increased crystallinity, reduction in defects, excellent bonding and hydrophobic nature.154,514 Chemical treatments are beneficial for the removal of large concentrations of wax, lignin, and oil that cover the outer layer of the fibre cell-wall, which decomposes cellulose and reveals crystallites. Also, this treatment induces fibrillation, which increases the effective contact area for interaction with the resin. Thus, the creation of an uneven surface texture and the improvement of the aspect ratio provide good interfacial bonding, which lead to a significant improvement in the properties of the composite.154,515,516
4.3.2 Disadvantages. Chemical treatment decreases the spiral angle, and consequently the molecular orientation improves the elastic fibre package.435,446,517 The number of cellulose hydroxyl groups in the fibre–matrix interface is reduced by the chemical treatment process.518 Acetylation involves the reaction of the cell wall hydroxyl groups of lignocellulosic materials with acetic or propionic anhydride at elevated temperatures, which increases the hydrophobicity of the plant fibers. The hydroxyl groups of lignin and hemicelluloses react with the reagents, whereas the hydrogen bonding on the closely packed hydroxyl groups of crystalline cellulose prevent the diffusion of the reagent, and thus results in very low rates of reaction.518

5. JFRCs

JFRCs consist of jute fibres as a reinforcement with a matrix at different interfaces. The fibres and matrix maintain their physical structure and chemical compositions during this process.519–524 The reinforcement carries the load, while the matrix material holds the fibres in an exact position and direction, acts as a medium of load transfer to the fibres and protects them from environmental damage due to temperature and humidity.291 JFRCs have high impact resistance and medium tensile and flexural strengths compared to other PFRCs.525 Studies have been reported on JFRCs, especially those containing polymer matrices.331,384,518,526–528 Not only the polymer matrix but also the reinforcement should be reusable for entirely bio-based composites. Therefore, the use of jute fibres together with polymer matrices in the manufacturing of composites at low cost has attracted attention from researchers.464 Contemporary studies recommended that jute fibres are excellent alternative reinforcement materials in polymer composites to replace synthetic fibres, which are costly and non-renewable.183,450Table 8 presents the published works on JFRCs.
Table 8 Reported experimental works on JFRCs
Reinforcement Matrix Surface treatment Fabrication method Ref.
Tossa jute Polymer latex Styrene butadiene Mechanical mixing 353
Jute + glass Isophthalic polyester Hand lay-up 153 and 317
Jute PLLA Hot pressing 41
Jute Epoxy Stacking method 313
Jute Epoxy Vacuum-assisted resin infiltration 343
Jute Epoxy H2SO4, H2O2, Na2SiO3, Na3PO4 Hand lay-up 434
Jute Epoxy NaOH Vacuum infusion 451
Jute Epoxy NaOH + silane Vacuum-assisted resin infusion 432
Jute Epoxy (MGS RIM 135) Vacuum infusion 529
Jute Epoxy NaOH/3-aminopropyl-triethoxy-silane solution/3-phenyl-aminopropyl-trimethoxy-silane 382
Jute Epoxy Compression moulding 315 and 381
Jute Epoxy Hand layup 83, 311, 314, 391 and 530–532
Jute Epoxy, polyester NaOH 5
Jute + clay Epoxy NaOH Mould method 385
Jute PLA Injection moulding, compounding and extrusion 81
Jute PLA Compression moulding 363 and 533
Jute PLA NaOH Injection moulding 302
Jute PLA NaOH + silane Melt mixing 349
Jute Polyester Pultrusion 336 and 338
Jute PP Compression moulding 389, 392 and 534
Jute PP Melt mixing 390
Jute PP Injection moulding 535
Jute + banana Cashew nut shell liquid (CNSL) Hand layup 536
Jute Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) NaOH + acetic acid Compression moulding 443
Jute Portland pozzolana fly ash cement + carboxylated styrene butadiene NaOH Mixing 327
Jute Polyester (crystic P9901) Pultrusion 337
Jute PP Sodium periodate Compounding and extrusion 342
Jute Soluble starch NaOH + dimethyl sulphoxide (DMSO) Solution casting method 376
Jute Poly butylene succinate (PBS) NaOH + Na2SiO3 + Na2P3O10 + Na2SO3 Mixing method 371
Jute Vinylester (HPR 8711) NaOH Hand layup 441
Jute Elekeiroz S.A and Resana S.A. NaOH Mould method 230
Jute PP Hot pressing 537
Jute PP Screw mixing 538
Jute + hemp PP NaOH + MAPP 58
Jute PP MAPP Injection moulding 59 and 99
Jute PP MAPP Compounding and extrusion 348 and 539
Jute PP MAPP Pultrusion 292
Jute Epoxy Infrared laser, ozone, enzyme and plasma Hand layup and compression molding 412
Jute Glucofuranoside based trifunctional EP monomer Methyl hexahydrophthalic-acid (MHHPA) Compression moulding 182
Jute Epoxy NaOH Vacuum infusion 386 and 387
Jute Polyester NaOH Compression moulding 540
Bangla Tossa Jute PP NaOH Compression moulding 347 and 515
Jute PP Alkaline + silane 61
Jute Methacrylated soybean oil NaOH Compression moulding 297
Jute Polylactide (GCS PLA 4320) NaOH + benzoyl peroxide Solvent casting 368
Jute PP KMnO4, MAPP, TDI and ST Commingling 356–358
Jute PP KMnO4, MAPP, TDI and ST Compression moulding 355
Jute Polyester NaOH Hand layup 367 and 541
Jute + glass Polyester Resin transfer moulding 542
Jute Polyester Alkali, micro-emulsion silicon and fluorocarbon Compression moulding 543
Jute Polyester Film stacking 341
Jute Vinylester Compression moulding 361 and 544
Tossa jute Epoxy Corona + UV treated 383
Jute PP MAHgPP Hot pressing 545
Jute PP MAHgPP Compounding and extrusion 105
Jute HDPE MAPE Melt mixing 362
Jute Carboxylated styrene butadiene NaOH 353
Jute LDPE 2-Hydroxyethyl methacrylate Compression moulding 344
Jute Polyethylene NaOH, H2SO4 + CH3COOH (acetic acid) Hot pressing 345
Jute + glass Epoxy Hand layup 316
Jute + banana + glass Epoxy Hand layup 377
Jute Epoxy Succinic Hand layup 378
Anhydride + phthalic
Anhydride
Jute Epoxy (diglycidyl ether of the bisphenol A) Mould method 384
Jute Epoxy resin Kinetix R240 Hand layup 24
Jute Green epoxy resin CHS-Epoxy G520 Sulphuric acid (H2SO4) and NaOH Hand layup 546
Jute + glass + Kevlar Epoxy Hand layup 547
Jute + basalt Epoxy resin SX8 EVO NaCl Vacuum assisted resin infusion 350
Jute Epoxy (LY 556) Hand layup 548 and 549
Jute + Al2O3 Epoxy (LY 556) NaOH + ethanol Hand layup and compression moulding 550
Jute LPDE NaOH + oligomeric siloxane (Z-6173) Compression moulding 339
Jute Tetraethyl orthosilicate (TEOS) and ethanol NaOH + nano SiO2 359
Jute + carbon Polyvinyl butyral Hot press technique 321
Jute Recycled PP Hot press technique 388
Jute + copper slag Polyester NaOH + calcium hydroxide Compression moulding 322
Jute Polyester NaOH + PLA coating Hand layup and compression moulding 551
Jute Epoxy NaOH Compression moulding 487
Jute HDPE Hot pressing 364
Jute + Kenaf HDPE Microwave assisted compression moulding 552
Jute Epoxy DTE 1100 NaOH + SiC Hand layup and vacuum bagging 553
Jute PP MAPP Compression moulding 379
Jute Polyurethane NaOH + silane Extrusion 374
Jute PP Phosphoric acid (H3PO4) Extrusion 554
Jute LDPE NaOH Hot pressing 346
Jute Polyester Hand lay-up 555


5.1 Processing methods

5.1.1 Hand lay-up method. Hand lay-up is an ancient method, which is the easiest and most extensively used technology for the manufacture of composites.547 Employing this method, fibres were impregnated in a matrix and layered in a mould over each other and pressed for 1 h. The resin and the hardener were mixed in the ratio of 100[thin space (1/6-em)]:[thin space (1/6-em)]32, before lay-up. Then, the mixture was applied to the fibres and spread uniformly using a roller, and the extra resin was squeezed out.548,549 An insulating material was placed between the mould plates and the releasing agent was applied to the inner surface of the mould. The composite was taken from the mould, kept at 30 °C for 48 h and cured for 16 h at 60 °C in an air-circulating oven.34 The pressure was maintained, and the thickness of the material was dependent on the wt% of the fibres.556 A releasing gel was sprayed on the mould to enable the non-sticky removal of the materials.548 After 24 h of curing the resin at room temperature, it was heated for 2 h in an oven at 100 °C.393

JFRCs containing untreated and NaOH-modified jute fibres were produced in the shape of circular rods. The fibres were kept at 100 °C for 4 h, dipped in resin and pulled by hand through a cylindrical tube. The fibres inside the tube were cured at 30 °C for one day followed by 4 h at 80 °C.441 The composite together with releasing sheets was placed between mould plates at 120 °C for 1 h and a uniform load applied.412,546 The composites were produced by chopping the fibres into plies and applying resin to them. The fibre laminates inside the mould were allowed to cure for 24 h at 30 °C. Then, the laminates were left out for 15 days to obtain the samples for testing.293 The load was applied to the top of the mould and transferred to the bottom portion of the composite laminate, and then it was allowed to settle for 48 h at 30 °C. Subsequently, the composites were taken out and placed in an airtight container for 48 h for further testing.311

5.1.2 Film stacking method. The development of JFRCs using the film-stacking technique is explained in Fig. 3(a). The fibres are stacked on each side of multiple layers of resin along their length at 90°. The lay-ups are subjected to quick press consolidation involving the following steps: (i) pre-compression, (ii) heating under vacuum, (iii) quick transfer to a press for consolidation and cooling, and (iv) removal from the press. In a standard experiment, a lay-up of PLA films and jute fibres was placed in a metal frame, pre-pressed at 3.3 MPa for 15 s, and then transferred to the heating stage. The lay-up was moved immediately, the heating section was automatically sealed, and the pressure was reduced to 400 Pa within 1–2 min. The pre-compressed lay-ups were heated at 180–220 °C for 3–10 min. Following the completion of the heating, the lay-up was moved back to the press, where the assembly was stabilized by applying a load of 3.3 MPa.366
image file: d3ta05481k-f3.tif
Fig. 3 Process for the fabrication of JFRCs: (a) film stacking process and (b) pultrusion process. Reproduced from ref. 129 with permission from Elsevier, 2013. (c) Injection molding. Reproduced from ref. 587 with permission from Elsevier, 2020. (d) Mould method. Reproduced from ref. 385 with permission from Springer, 2017.
5.1.3 Compression moulding. The compression moulding technique is used for the fabrication of composites without any damage to the fibre during processing.53,366,410,557 In this process, the resin is dissolved in chloroform and stirred for 4 h for the preparation of a homogeneous mixture at room temperature. Then, the mixture is poured into a mould for the preparation of a film and dried at 60 °C to evaporate the excess chloroform. Finally, the dried films are kept in a press and a loading of 13.34 kN applied at 166 °C to obtain a uniform laminate.443 A hot compression moulding machine was used for the production of JFRCs from continuous fibres. The matrices were melted by heating and quickly wetted with jute fibres.363 The fibres were dried at 80 °C for 2 h under 50 mbar before composite preparation. Then, these laminates were pressed with a hydraulic pressure of 200 bar to get a uniform thickness.182

Jute fibres were cleaned, dried and immersed for 1 h in NaOH solution, washed, and then dried again at 50 °C. Subsequently, 3% nano-clay was mixed with resin and stirred for 2 h, and then put in an oven for the elimination of voids such as air bubbles. The jute fibres were cut into dimensions of 300 × 300 mm2, and the resin fibre ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]3 was considered. Two layers of jute fibres were employed for the preparation of the laminate together with 1.5 wt% resin, accelerator and catalyst.540 The mould was kept in a compression moulding machine at 200 °C and a load of 0.2 MPa was applied.355 The fibre was submerged in a benzene peroxide solution and allowed to dry for 2 h at 80 °C. Composites were prepared with the addition of three layers of low-density polyethylene sheets and two layers of jute. The sheets were processed by heating the polymeric granules at 115 °C and kept under pressure for 5 min. The laminates were placed at 135 °C and a hydraulic press was used for moulding them.344 The jute fibre was initially washed with water to clean the dust and other particles found in the fibres. The first layer of the specimen was placed in the die, then resin was spread over the fibre, and allowed to dry for 10–15 min. Prior to drying the resin, it was layered with a second layer, and then all the remaining layers were layered. Finally, the composites were kept in a moulding machine to achieve the perfect samples through application of a load in the range of 250–400 N.315

5.1.4 Pultrusion. Pultrusion is an automatic and consistent process for producing composite components of high volume with uniform cross-sections, in which the wetted fibres are drawn through a die at a constant velocity. This method offers good efficiency and ease of processing, but there is a limited number of reports on the pultrusion of PFRCs, and in particular their resistance to moisture.337,558–561 There are many parameters to consider during this process, among which the drawing speed and temperature are the most important, which were taken as 180 mm min−1 and 85 °C in a report.336 Previous studies337,560 demonstrated circular cross-section rods pultruded by reinforcing jute fibres. Fig. 3(b) shows a schematic of the pultrusion process for the fabrication of composites.129

Jute fibres were impregnated in a resin tank filled with polyester resin. To obtain the desired form, a pulling system was used to move the resin-mixed jute fibres through a die. The curing was performed in the die, which was accurately designed to give the shape of the end product.562,563 The pultrusion technique was applied using an extruder fitted with a die assembly to cover fibres with a resin and MAPP mixture. Specifically, 3 wt% MAPP was applied to the pure resin. The extruder and die were heated at a maximum temperature of 200 °C. The coated fibres were washed with water and cut into 4 mm long granules. The granular particles were dried overnight at 110 °C. Under the same conditions, they were once again extruded with the same extruder to homogenize the fibre/matrix mixture.292

5.1.5 Hot pressing. The fabrication of composite by the hot pressing process is done in two stages. In a reported study, in the first stage, resin films with a thickness of 0.2–0.3 mm were fabricated by applying a load of 250 kN in a hot pressing machine at 250 °C. Then, the sheets were kept at 180 °C and 50 kN pressure for 10 min and cooled by running water. In the next stage, the jute fibres were dried for 24 h at 60 °C. A mould with a size of 15 × 15 cm2 was used, filled by woven jute mat between the resin films and kept in a hot press.41 In another study, oven-dried jute fibres, polyethylene pellets, and nanoparticles were mixed, the mixture was moulded and kept in a press at 180 °C and a load of 7 MPa applied for 1 h. Then, the mould was cooled at 30 °C and the test specimens were prepared.345
5.1.6 Melt mixing method. JFRCs were processed using the melt mixing method in a setup with roller blades and a volumetric chamber. The fibres and resin were premixed at different wt% of fibre and kept in a pre-heated chamber at 150 °C. Mixing was achieved at a speed of 25 rpm for 10 min. Then, the chemically modified fibres were combined with resin at 160 °C and maintaining the same rotor speed. Subsequently, these pre-mixes were moulded using equipment at 150 °C and a uniform load applied.362 Jute fibres were laid in a mould and Teflon coatings were applied on the top and bottom of the mould for the easy removal of laminates. Then, the assembly was hot-pressed by applying a load of 4 MPa at 165 °C. The polymer melted and impregnated the jute fibres at this temperature and compaction occurred. Subsequently, the pressure was increased to 6 MPa, and then the composite was left for 10 min. Then the composite was taken out from the mould and samples were prepared for testing.
5.1.7 Compounding and extrusion. Gunning et al.564 examined the processing conditions for the fabrication of composites through extrusion and found that the composite produced using lower temperature profiles exhibited good properties at higher temperatures due to the reduction in the thermal degradation of the fibres during the process. They manufactured composites with different fibre loadings using jute, hemp and lyocell fibres. While the average residual fibre length was short, the composites showed enhanced mechanical properties. Also, the extrusion temperature, screw speed, and configuration were found to have a dominant impact on the consistency of the composite regarding the operation employing a twin-screw extruder. Degradation occurred at 180 °C, particularly for the natural fibre, which was higher than the melting temperature of the resin. Thus, the temperature was set to 180 °C to minimize the degradation of the jute fibre during fabrication. The compound was actually extruded a temperature 10–20 °C higher than the set temperature. Furthermore, the screw velocity was operated at 150 rpm to reduce the thermal degradation by shear heating.81

Initially, jute fibres were mixed with PP granules and the mixtures were passed through an extruder unit at 135 °C. Using a grinding machine, all the bits were broken into smaller granules with a size of 3 mm to 0.5 mm.342 A twin-screw extruder was employed for the direct thermoplastic processing of long fibres.328 The fibres were washed and dried at 60 °C for 48 h before compounding, until a uniform weight was reached. Then, the fibres were integrated in the side feeder of the extruder, which fed them directly into the melting of polymers. Also, the content of the fibres in wt% was regulated by the speed of the screw, which was determined using the targeted material, fibre weight, length, and roving time.348 In another study, composite samples were compounded on a twin-screw extruder. Dog-bone shaped specimens were made from compound granules using an injection moulding machine.105 Reinforcing jute fibres were dehydrated for one day at 50 °C and 4 h at 80 °C before extrusion to remove moisture. The constituent materials were fed into the extruder. The screw speed of the extruder was fixed at 145 rpm and the temperature range was 165–195 °C. The composites were pelletized to obtain granules, which were dried at 85 °C for 12 h and used for the fabrication of the samples.539

5.1.8 Injection moulding. The mass production of composites is usually performed by the injection moulding process,302 which has the benefits of high processability and ability of the fibre to degrade during the process.81 In this process, the fibres are washed and chopped to 3 mm.565 Yang et al.566 manufactured JFRCs using an injection moulding process. They observed an increase in the modulus after the addition of the jute fibre; however, there was a decrease in the tensile strength. An extruder compounded the jute fibre, resin, and alkaline solution at a temperature in the range of 165–180 °C. Following immersion in a water bath, the fibres were cooled and cut into granules. For processing, the granules were kept at 100 °C for 4 h and injected into composites. The injection moulding process had three-zone temperature profiles ranging from 160–195 °C.59Fig. 3(c) shows the standard manufacturing cycle for JFRCs by injection moulding.565
5.1.9 Vacuum infusion. The method of vacuum infusion was designed for the manufacture of NFRCs.567 The fibres were arranged in layers and pre-compacted by applying a pressure of 1200 kPa in a hydraulic press. The fibres were layered on a mould and coated with silicone gel after compaction. Two vacuum tubes were connected, one with a pump, while the other was inserted in a resin bottle. The tubes were mounted on the laminate layers, and a sealant tape was used for sealing the entire system. Using a vacuum pump, 3 kPa vacuum was drawn. The laminate was kept under vacuum for 1 h to ensure the complete evaporation of moisture.386,387 Resin and hardener were mixed at a ratio of 24[thin space (1/6-em)]:[thin space (1/6-em)]100 and degassed in a desiccator. The product was drawn by the differential pressure produced by the vacuum through the laminate via the resin inlet tube, and the stacked fabric plies were completely wetted. After the laminates were fully wetted, the vacuum was reduced to 35 kPa by placing a dead weight on top of the laminate and it was left overnight for setting. Then, the panel was removed from the vacuum and heated at 50 °C for 3 h and 80 °C for another 3 h.
5.1.10 Solvent casting method. The solvent casting process is desirable for the manufacture of composites with improved strength due to the improvement in wettability and better bonding of the fibre with the matrix. The resin is dissolved in chloroform and impregnated with jute fibres and the solvent is gradually evaporated. Subsequently, the pre-pregs are kept at 60 °C for one day to remove the chloroform completely. Finally, the pre-pregs are layered in the mould and pressed with a load of 5 MPa at 170 °C for 5 min.368
5.1.11 Mould method. Samples were prepared via the impregnation of jute fibres with resin and different clay materials in a mould. The mould was kept in a hydraulic press until it achieved the desired thickness. Then, the compressed system was maintained for 24 h and demoulding was done. The preparation of JFRCs by the mould method is shown in Fig. 3(d).385
5.1.12 JFRCs with adhesives. Composite laminates with jute fibres and Betamate 2096 adhesive were fabricated using the hand lay-up and hot press compression moulding technique using a steel mould and a heated plate hydraulic press. The composites were fabricated in a way to have a stacking sequence consisting of 5 layers of jute. The fibre loading was kept at around 30 wt% and the adhesive at 70 wt% of the final fabricated composite. The curing time was of 8 h at 70 °C and the adherends were cut from the composite plates.568
5.1.13 Characteristics of fabrication methods. The methods for the fabrication of PFRCs depend mainly on factors such as the characteristics of the matrices and reinforcements, the shape and size of the part produced, and the users. The characteristics of the fabrication technique varies from product to product. To obtain good adhesion, it is necessary to address issues such as material preforms, processing parameters, and post-processing techniques.569

5.2 Properties of JFRCs

PFRCs are highly dependent on major factors such as properties of the reinforcing fibre, matrix, and interfacial bonding between them.246,569,570 Composites are exposed to the environment during processing, which may affect their properties. The properties of the composites have been studied by several researchers, resulting in a substantial improvement in their properties.62,74–77,257,258,345 Heat treatment of fibres can reduce the water content in the fibres and improve the properties of the resulting composites by enhancing the compatibility between the fibres and polymer.571 The mechanical properties and hardness of jute fibre-reinforced composites were evaluated and an enhancement in properties was observed with the addition of fibres. Furthermore, the composites exhibited good thermal stability and decomposition.572 In another study, an enhancement in the properties of JFRCs was observed through the incorporation of graphene powder during the processing of the composites.451

5.3 Physical properties

5.3.1 Hardness. The hardness is measured using a Shore-D type hardness tester357 and a Rockwell hardness tester5,342 according to the ASTM D785 standard.5,342,357 In one report, this test was conducted using a 1/16′′ ball indenter and applying a load of 1000 N. The tester was placed vertically and pushed onto the specimen for alignment of the presser foot to the specimen.375 With oxidation of jute fibre and fibre packing, there was an increase in average hardness due to the flexibility and stiffness of the composites.573 Compared to the untreated and oxidized fibres, the treated JFRCs yielded much greater hardness (Fig. 4(a)). This was due to the uniform fibre dispersion in the resin with reduced voids and good bonding between them.342 Also, the micro-hardness of the composites was measured using a Leitz micro-hardness tester. In this test, a diamond indenter was pushed into the specimen at 136° between opposite faces and subjected to load. The indentation mark left on the specimen was determined after the load was removed, and the arithmetic mean was calculated.574 The durability of the Vickers number was determined using the equation:
 
image file: d3ta05481k-t1.tif(1)
where P is the applied load, L is the diagonal of the indenter and X and Y are horizontal and vertical lengths, respectively.

image file: d3ta05481k-f4.tif
Fig. 4 Influence of the surface modifications of fibres on the (a) hardness and (b and c) dielectric constant values of JFRCs.
5.3.2 Chemical resistance test. The chemical absorption characteristics of composites have been evaluated to determine their resistance to chemical substances according to ASTM D543-87 requirements, where an increase in weight suggests that the material has low chemical resistance.575 The influence of chemicals such as benzene, toluene, CCl4 and purified water; acids such as HCl, H2SO4 and acetic-acid and alkali solutions such as NaOH, Na2CO3 and NH4OH has been tested. For example, specimens were soaked in chemical solutions for a day, taken out and rinsed with water, and cured at ambient temperature. Then, the weight of the specimen was measured and the weight gain/loss percentage (WCR) was calculated using the following equation:345
 
WCR = {(WfWi)/Wi} × 100(2)
where Wi and Wf are the weight of the samples before and after immersion, respectively.
5.3.3 Dielectric properties. Dielectric experiments are conducted to measure the dielectric constant, loss factor, dissipation factor, resistivity and conductivity of a material. In composites, their dielectric properties depend on the polarization of their constituents and mainly influenced by inter-laminar, dipolar, and atomic polarization.303,576–580 The polarization of the interfaces at very low frequencies has a greater effect on the dielectric properties. Interfacial polarization occurs when charge carriers are trapped at the intersection of heterogeneous systems, primarily by dipoles in the case of NFRCs. In an external field, these dipoles may get turned to some degree, and thus contribute to the polarization of the material. However, there is a decrease in the magnitude of the effect of interfacial polarization with an increase in frequency. In plant fibres, the availability of –OH groups contributes to their polarization.303 The dielectric constant is calculated using eqn (3), as follows:
 
image file: d3ta05481k-t2.tif(3)
where C is the parallel capacitance, t is the width, A is the cross-sectional area, and ε0 is the free space permittivity (8.85 × 10−12 F m−1). The resistivity and conductivity of the sample can be calculated using the following equations:
 
image file: d3ta05481k-t3.tif(4)
 
Conductivity = 1/ρ(5)
where R is the parallel resistance and ρ is the resistivity.

In a study, the test samples were mounted in parallel plates between electrodes and the measurements were performed in the frequency range of 100 and 2 × 106 Hz. Higher temperatures were not investigated, given that resin softening occurred at 160 °C.355 The dielectric constants increased at all frequencies with an increase in the fibre wt%, as shown in Fig. 4(b). Both types of polarization contribute to the polarizability of non-polar molecules, hence displaying lower dielectric constants at all frequencies. Jute fibres are polar in nature, and therefore were incorporated in a non-polar resin, which resulted in an increase in the amount of polar groups in the material, resulting in dipole and orientational polarization. JFRCs are heterogeneous materials, and thus their interfacial polarization often effectively leads to a dielectric constant. Hence, the dielectric constant increased at all frequencies for the fibre materials. A full dipole orientation was possible at low frequency, leading to an increase in the dielectric constant. However, at high frequency, the vibrations of the particles were strong, and thus there was no full orientation of the dipole. There was a decrease in the dielectric constant with an increase in frequency.355 The dielectric constant of the composites was determined using the following equation:

 
ε = C/Co(6)
where C is the capacitance and Co is the capacitance under vacuum. Co is determined from the sample dimensional information using the following equation:
 
Co = 8.8510 − 12A/t(7)
where t is the thickness.

The influence of chemical modifications on the dielectric constant of JFRCs is presented in Fig. 4(c), which clearly shows that the raw fibre-reinforced sample has the maximum dielectric constant compared to the chemically treated fibre samples. This proved the reduction in the dielectric constant due to the chemical treatment of the fibres. The treatments minimized the hydrophilicity of the fibres, thereby reducing the absorption of moisture and number of polar groups in the composite. This resulted in a decrease in the polarization orientation, resulting in low dielectric constant values. The surface treatments improved the adhesion, resulting in a reduction in the amount of voids and defects and a reduction in the absorption of moisture, which caused a reduction in the dielectric constant. Treatment with stearic acid resulted in only slight topographical changes on the surface of the fibre, with a difference between the dielectric constants of the chemically treated and untreated fibre composites.355

5.3.4 Dimensional stability. The dimensional stability of composites undergoes changes in dimensional factors, namely, length, thickness, and width, when they are exposed to different environmental conditions such as humidity and dry and wet conditions. In this case, the stability of composites is important for determining their shelf life and end-user applications. To determine the dimensional stability, characterization tests such as density, thickness swelling, water absorption, weathering and cyclic tests are conducted.581 Dimensional stability can also be analysed via FTIR, XRD, TGA, and DSC analysis.

5.4 Mechanical properties

Impregnation and good bonding must be ensured for the production of composites for industrial applications and the exploitation of fibres as reinforcement materials, given that the final mechanical properties depend on the bonding between the fibre and resin.156,582,583 Natural fillers have become significant over the past decade, given that they improve the strength of composites.525,581,584–587 The factors influencing the strengths of NFRCs are as follows: (i) type and strength of reinforcements, (ii) type of resin and mixing ratio, and (iii) method by which composites are blended, casted and cured.588 Among them, the fibre–matrix compatibility, which leads to a uniform distribution of fibres, is one of the prime factors affecting the strength of composites.589 Many researchers have studied the mechanical properties of NFRCs98,100,331,435,517,590,591 and the results showed that the tensile and flexural strength of root portion fibre-based composites is higher than that of tip-portion fibre-based composites. Alternatively, the modulus of tip-portion fibre composites was observed to be higher than that of root-portion fibre composites.592 Various researchers have conducted studies on the mechanical properties of jute fibres and JFRCs.103,156,202,230,278,331,383,593–606
5.4.1 Tensile strength. Tensile testing involves the application of load along the axis of a specimen until it fractures, and the elongation of the specimen against the load is noted based on the increase in gauge length.260 The test specimens are prepared according to the ASTM D3039,5,83,315,329,336,343,379,387,393,529,540,574 ASTM D638,99,230,302,313,314,317,339,340,342,345,346,362,363,366,368,375,377,378,381,390,392,539,544,546,550,607–609 ASTM D882,41,376 ASTM D2256-02,443 ASTM D3822,327 ASTM C1557,319 ISO 527,311,351,607 GB1447-83,371 DIN EN ISO527-4 (ref. 365) and DIN 53455 (ref. 56) standards. Experiments were conducted using a universal testing machine (UTM),5,41,81,83,311,314,315,317,327,340,342,345,346,351,362,377,379,540 including Instron models 4206,357 1026,56 3385,368 3382,539 1122,607 1342,608 1195,550 8821S,385 3369,313 6025,366 and 4400R,387 Zwick Z020,182 Krystal-Elmec,378 Mecmesin,609 MTS 370,546 Shimadzu AG-X,319 Shimadzu AUTOGRAPH AG339,529 and Zwick Roell.443 Samples were prepared with the dimensions of 300 × 25 × 3 mm,540 140 × 10 × 2 mm,182 255 × 25 × 2 mm,387 110 × 50 × 10 mm,41 165 × 13 × 3 mm,362 180 × 25 × 15 mm,363,366 148 × 10 × 4 mm,342 115 × 6 × 3.1 mm,345,346 200 × 20 × 4 mm,546 150 × 15 × 3 mm,550 165 × 25 × 3 mm,339 250 × 25 × 3 mm,379 165 × 19 × 3.2 mm (ref. 392) and 120 × 84 × 20 mm.83 The test was performed with a cross-head speed in the range of 1–10 mm min−1, gauge length of 5–50 mm and load-cell capacity of 5–100 kN.34,41,81,83,182,313,327,336,339,340,342,345,346,357,366,368,371,375,378,379,387,393,443,544,546,607–609 Many researchers have performed experiments under atmospheric conditions of 20–35 °C and 65% relative humidity. The test specimen geometry is indicated in Fig. 5(a).315
image file: d3ta05481k-f5.tif
Fig. 5 Tensile testing of JFRCs: (a) scheme of the tension test specimen according to the ASTM standards. Reproduced from ref. 315 with permission from Springer, 2016. (b) Typical tensile stress vs. strain curves.

The tensile strength and modulus can be evaluated using the following equations:

 
Tensile strength σt = P/bh(8)
where P is load, b is the specimen width, and h is the specimen thickness.
 
Young's modulus (E) = stress/strain(9)

Fig. 5(b) displays the typical uniaxial tensile test results of jute fibres. The strength and modulus of NaOH-treated fibres increased by 13% and 17% compared to the untreated fibres. The NaOH treatment induced the removal of non-cellulosic products, such as hemicellulose, lignin, and pectin from the inter-fibrillary regions610 and resulted a higher tensile strength and modulus.611 An exposure period of 2 h with 5% NaOH showed acceptable results.443

5.4.2 Flexural/bending strength. Flexural strength is a combination of both compressive and tensile strength and directly varies with the shear strength. Specific processes, such as stress, stretching, and shearing, occur simultaneously during flexural testing.612 The flexural test of JFRCs has been carried out using UTM,5,362,377,379,381 Hounsfield UTM,41,354,361,613 Zwick/Roell Z005,350 Zwick/Roell Z010,336,443,542 Zwick Z020,182 Instron 4303,361,441 Instron 3382,539 Instron 1195,550 Mecmesin,609 Shimadzu AGS-J,546 Shimadzu AUTOGRAPH AG-G Series,339 and LLOYD LR30K341 machines. The sample dimensions for the flexural test were obtained according to the ASTM D790,5,41,99,153,313–315,317,336,339,341–343,350,357,361–363,368,377–379,381,390,392,441,443,539,542,546,608 ASTM 348–80,614 ASTM D5943-96,550 ASTM CISO 14125:1998,311,351 GB 1449–1983,371 DIN 53452,56 DIN EN ISO14125 (ref. 365) and DIN EN63 (ref. 383) standards. The test was carried out using samples with dimensions of 79 × 10 × 1 mm,41 56 × 20 × 3.6 mm,363 100 × 10 × 2 mm,182 79 × 10 × 4.1 mm,357 80 × 15 × 1.5 mm,362,383 79 × 10 × 4.1 mm,342 110 × 20 × 20 mm,354,361,613 160 × 12.7 × 4 mm,546 100 × 15 × 3 mm,550 13 × 12.7 × 5.5 mm,379 116 × 24 × 13 mm,530 76 × 25 × 3.2 mm (ref. 392) and 110 × 20 × 20 mm.613 A span width in the range of 30–140 mm and cross-head speed in the range of 1–5 mm min−1 have been considered by various researchers.41,182,329,341,350,357,361,362,368,371,441,443,542,608,609,613 The samples were screened at room temperature fitted with a 2 kN load-cell.379 The flexural strength can be calculated using the following formula:
 
Flexural strength σf = 3PL/2bh2(10)
 
Flexural strain = 6sh/L2(11)
where P is load, L is the length, b is the width, h is the thickness, and s is the deflection.

The details of the flexural strength of JFRCs are presented in Fig. 6. The results showed that the treated composites possessed improved strength and modulus. However, the brittleness in the composites was due to the addition of nano-clay, which was evident from the decrease in the strain value at high stress. No fibre fracture in both nano-phased and traditional composites was found. It was also noted that JFRCs exhibited high flexural strain, which suggests the presence of ductility in these bio-composites.443 In the case of the treated composites, high flexural strength and modulus were obtained compared with the untreated composites. The nano-clay-infused samples also performed better in terms of flexural characteristics compared to traditional composites. Nevertheless, following the absorption of moisture, a reduction in the flexural intensity and modulus was observed. The highest decrease in flexural strength was observed in the untreated composites and the lowest found in the samples containing 4 wt% nano-clay. The flexural modulus values were 21% and 6–7%, respectively. This is an example of the moisture-arresting behaviour of nano-clay in the composites.443 The typical flexural stress vs. strain curves are shown in Fig. 6(a–c). With a shorter immersion duration, a dramatic increase in the average flexural load was observed. It was assumed that the cellulose content in the jute fibre was reduced and that the plasticization effect caused the jute fibres to become more versatile. The flexible nature of the material obtained was due to the absorbed water molecules, which were stored in the cavities and cracks, acting as a plasticizer.337 The flexural modulus can be calculated using the following equation:34,361

 
Flexural modulus = L3m/4bt3(12)
where L is the support length, m is the slope of the load vs. deflection curve, b is the width and t is the thickness.


image file: d3ta05481k-f6.tif
Fig. 6 Flexural testing of JFRCs: (a–c) typical flexural stress vs. strain curves and (d–f) compressive stress vs. strain curves after exposure to distilled water; sea water; and acidic solution, respectively.
5.4.3 Compressive strength. Compression test specimens were prepared according to the ASTM D695 (ref. 337 and 540) standard. The size of the rectangular specimen was 13 × 13 × 3 mm,540 while that of the circular specimen was 12.7 mm and 25.4 mm.337 Tests were conducted on a 400 kN capacity UTM,540 Instron 3367 (ref. 337) and hydraulic UTM.354,613 The tests were carried out at a cross-head speed of 5 mm min−1 and loading rate of 13 kN min−1. The cold crushing strength (CCS) was determined using the following relation:354,613
 
CCS = F/A(13)
where F is the fracture load and A is the face area of the cube.

Fig. 6(d–f) show the associated compressive stress vs. strain curves for immersion in fresh water, sea water, and acidic solution, respectively. It was observed that the maximum pressure increased with an increase in the soaking time after exposure to aqueous solution.337

5.4.4 Interlaminar/interfacial shear strength. Although several researchers have manufactured and studied different types of NFRCs due to their different properties, to date few experiments have been reported on the interfacial strength of JFRCs.7,155,158,437,446,590,614,615 Based on the ASTM D2344 (ref. 329, 339 and 550) and ASTM D3846 (ref. 387) standards, the ILSS test was conducted. The test was carried out using Instron 5569,387 Instron 3385,369 Instron 1195 (ref. 550) and Shimadzu AUTOGRAPH AG-G339 tensile testing machines with an attached load-cell. A narrow square cross-section beam of 45 mm was supplied at a rate of 1.3 mm min−1 under 3-point bending, as shown in Fig. 7(a). When the loading cylinder produced a downward force, normal and transverse shear stress was exerted on the specimen. The use of a short rod is believed to be sufficient for a reduction in stress due to bending, which results in an ILS fracture due to the cracking between the laminate on a horizontal plane.329
image file: d3ta05481k-f7.tif
Fig. 7 (a) ILSS test configuration, (b) ILSS of JFRCs and (c and d) multi-axial fatigue test curves: normalized stiffness degradation and total stiffness degradation. Reproduced from ref. 293 with permission from Elsevier, 2016.

The shear properties of JFRCs were determined using the 3-point bending mode on a UTM at a velocity of 20 mm min−1. The measurements were performed on composite samples with a size of 30 × 20 mm. The composite stress vs. strain curves were plotted and the breakage power, modulus, and elongation details were obtained.369 In another experiment, a UTM AG-G series was used with a cross-head speed of 1.3 mm min−1. The test specimens had a size of 26.3 × 6.4 mm.339 Using a micrometer,315,317 the specimen was set in the UTM. The experiment was conducted at a cross-head speed of 1.4 mm min−1. The ILSS was estimated as the ratio of maximum load to the shear area.387 The IFSS of the JFRCs was tested using a micro-droplet testing process. The fibres were kept in a steel frame and the resin micro-droplets were shaped on each fibre using a tip pin. The resin micro-droplets were placed in a the hot chamber at 220 °C and shaped in a circular cross-section due to the decrease surface strength. A specially built micrometer was used for testing the micro-droplet specimen. The laminates used for the various stacking sequences are shown in Fig. 7(b). ILS failure did not occur in the middle of the laminate, whereas shear mode failure occurred along the interface. The IFSS was determined using the pull-out force F, which was estimated using the following equation:

 
image file: d3ta05481k-t4.tif(14)
where Df is the fibre diameter and L is the length of the embedded fibre in the matrix.58

5.4.5 Fatigue strength. Despite the fatigue behavior of SFRCs, NFRCs517,616–619 has received little attention from researchers. Identification of the deterioration of the stiffness of a material was used to determine its damage from fatigue, which was defined by measuring the shift in load encountered under constant pressure, showing the stiffness deterioration value of the material during each loading period.620 Quaresimin et al.621 analyzed the fatigue strength of composite materials and highlighted the major impact of shear stress on the fatigue life of their components. They also noted that the load on a sample under multi-axial charge arose from the rebuilding of the non-uniform stress distribution on the sample. Amijima et al.622 interpreted data from multi-directional tension–tension tests as the ratio of pure application results for each loading form and combined the ratios to represent the multi-axial loading results.

A uniaxial tension–tension-type fatigue test was conducted in compliance with ASTM D3479M standards using a servo-hydraulic system powered by a 25 kN unit. The experiments were performed in the frequency range of 0.5 to 2 Hz with a constant load. The frequency was set below 2 Hz to avoid heat, which leads to a shortened fatigue life, and the constant stress ratio was kept at 0.1. Cycles for failure were developed using software, which predicted the fatigue strength.348 Katogi et al.198 tested the fatigue of unidirectional JFRCs under tensile loading. They recorded a tensile strength of around 61 MPa and 55% endurance limit. Katogi et al.596 conducted experiments on the influence of fibre type, shape, fibre wt%, surface modification, and matrix type on the damping capacity of a material under tension–tension fatigue tests. In the case of JFRCs, treatment with silane on the jute surface resulted in the need for higher loads given that progressive damage to was observed. Shah et al.597 examined the impact of fibre size, shape and strain ratio on the fatigue life of JFRCs and compared the results with that of other NFRCs. The values showed that the JFRCs displayed the highest static tensile and fatigue strength. JFRCs also survived millions of cycles at a stress ratio of 45% under fatigue charge. The fatigue properties of JFRCs were evaluated by Gassan and Bledzki154 by measuring the strength and modulus of the composites as well as their damping factor. They found that the reinforcement with chemically modified fibres resulted in a lower fatigue value and damage to the composite propagation. Fatigue studies on JFRCs were performed by reinforcing untreated and chemically modified jute fibres.231 The defects in the JFRCs were identified to play a major role. Higher concentrations of defects were used to absorb high impact energy and enhanced the fatigue strength of the composite.

Fig. 7(c) demonstrates the typical JFRC load degradation during tensile and multi-axial fatigue test cycles. The curve of the degradation of tensile fatigue stiffness value was normalized through application of load during the cycle. The curve of the deterioration of multi-axial fatigue stiffness was built by integrating the normalized deterioration of stiffness resulting from tensile and torsional loads.293Fig. 7(d) shows the weakening of the rigidity arising from multi-axial fatigue testing. The same findings were also observed for these curves in the stiffness deterioration of the fatigue tests, where the curves correspond to 75% and 65% strain rates. The results indicated an incremental failure compared to the tensile loading. Failure of the JFRCs under axial loading occurred at 75% after 90[thin space (1/6-em)]000 cycles and at 70% straining point after 260[thin space (1/6-em)]000 cycles. The resistance limit for the composites under these loads was found at a stress level of 65%, where the JFRC lost 50% of its resistance to rigidity.293

5.5 Impact strength

The impact strength of JFRCs can be assessed by a range of testing procedures, for example, weight loss single or repeat test, weighted pendulum test, projectile test, explosion test, constant strain test, Split-Hopkinson bar test, and instrumented pendulum test.623–625 The specimen placed between two conical tapered cups was repeatedly hit by a hammer-type pendulum attached to a revolving circular disc. When the release knob set at a fixed angle was touched, it released the pendulum, allowing it to fall. The hanging pendulum was taken up again after each impact and falling was continued until the specimen was broken.361
5.5.1 Charpy impact test. The Charpy impact test is carried out to assess the energy storing ability of materials against sudden loads. This test is performed by taking a material off a spinning pendulum in one blow.5 Notched specimens are prepared according to the ASTM D6110,41,315–317,342 ASTM D256 (ref. 384) and ASTM A370 (ref. 5) standards. The absorbed energy indicates the impact load-carrying capacity of the material. A hammer shatters the test component and the energy consumed indicates the resistance of the material to shock loads.5 The specimens are prepared with different sizes including 55 × 10 × 10 mm,5 79 × 10 × 4.1 mm (ref. 342) and 120 × 12 × 10 mm.384 A schematic illustration of the Charpy impact test scheme is illustrated in Fig. 8(a).315 The energy absorption with respect to time is shown in Fig. 8(b).556 Charpy measurements of composite materials consisting of jute fibres are shown in Fig. 8(c). The low strength was recorded because of the fibre content in the composite structure, which could initiate crack propagation, and eventually possible fracture. The high fibre content also increased the chances of fibre agglomeration, which could create a stress concentration region that needs minimal energy to cause a crack.292
image file: d3ta05481k-f8.tif
Fig. 8 Impact testing process of JFRCs: (a) schematic illustration of the Charpy impact test specimen according to ASTM standards. Reproduced from ref. 315 with permission from Springer, 2016. (b) Energy versus time during impact tests. Reproduced from ref. 556 with permission from Scielo, 2016. (c) Charpy impact strength. Reproduced from ref. 292 with permission from Elsevier, 2009.
5.5.2 Izod impact test. The Izod impact test is conducted according to the ASTM D256,99,357,362,366,368,381,550 ASTM D4812,540 JIS K7110 (ref. 81) and DIN 53433 (ref. 56) standards using an Izod impact tester, Impactometer 6545 and Zwick test machine. The specimens have a cross-section of 65 × 12 × 3 mm,540 60 × 10 × 2 mm (ref. 357) and 63.5 × 12.7 × 3 mm.362 Test samples having a notch of 45° and depth of 2.5 mm were prepared for experiments.362 Also, a specimen was tested using a 15 J hammer pendulum at an impact speed of 3.8 m s−1.357
5.5.3 Low-velocity impact test. The low-velocity impact test is carried out using a weight-balancing machine, CEAST 9350,556 Instron/Ceast 9340 (ref. 350) and Zwick/Roell HIT230F model machines with a constant impactor mass of 23 kg at a height of 110 mm and diameter of 19.8 mm.297 The experiments are carried out according to the ASTM D3763 and ISO 6603-2:2000 standards and the impactor stroke occurs in the middle of the samples through the use of a support with a diameter of 40–75 mm. The energy stored is chosen to allow the calculation of the damaged area but without encouraging perforation for most specimens.350,556 The height of the drop is taken to produce 25 J of impact energy to fracture specimens with a thickness of 1–3 mm.

5.6 Fracture behaviour

5.6.1 Torsion testing. The torsion test is conducted on a rheometer using a torsion specimen with a rectangular shape, and the thickness is defined by the number of layers in the composite laminate. The torsion modulus was obtained at a strain rate of 0.002 through small amplitude torsion tests performed in the frequency range of 0.1–40 Hz.385 The complex shear modulus (G) was obtained from the dynamic frequency analysis using the following equation:626
 
G = E2 + E′′2(15)
where E′ and E′′ are the storage and loss modulus, respectively.
5.6.2 Creep behavior. Creep is an interesting end-use factor for any composite material, especially in the case of JFRCs given that their properties are dependent on time and temperature. As a result of the viscoelasticity of the polymer matrix,627 temperature- and time-dependent degradation in strength and modulus can occur over time. Different methods of mathematical modeling were applied for the examination of the creep characteristics of composites.628–630 However, the details of the creep behaviors were barely considered due to the adhesive nature of the fibre/matrix.628,631,632 The experiments were carried out in the temperature range of 40–100 °C using a TA instruments Q800 DMA412 and PerkinElmer dynamic mechanical analyzer.633 The temperature range of the composites was chosen to be below their glass transition. The stress was applied for 30 min at the center of the sample after balancing at the optimal condition and the creep strain was measured. Then, the stress was measured, where the composites were identified as a linear viscoelastic region, confirming that the performance of the tests was linear.412 The experiments were performed using a 15 mm long specimen. The applied static stress for the creep was 4000 kPa, and then the material was allowed to recover below 4 kPa. Relatively low stress was applied during testing to confirm the linearity of the viscoelastic behavior.633 The experimental temperature was set at 50 °C. Details of the creep activity of JFRCs at different temperatures with and without fibre treatment are presented in Fig. 9. The curves in the figure indicate good agreement with the experimental results.412
image file: d3ta05481k-f9.tif
Fig. 9 Creep curves of (a) untreated, (b) enzyme-treated, (c) laser-treated, (d) ozone-treated, and (e) plasma-treated JFRCs. Reproduced from ref. 412 with permission from Elsevier, 2015.
5.6.3 Fracture toughness. The measure of resistance against crack growth is called fracture toughness of a material.634 The resistance of NFRCs to crack propagation or fracture toughness is based on various parameters including matrix toughness, additives, fibre characteristics, resin dispersion, fibre distribution, fibre orientation and duration after processing. Significant resilience of the fracture was observed to be influenced by the applied stress, fracture mode, matrix fractures, fibre debonding, fibre breakage, fibre pullout and bridging.635,636 Crack propagation depends on fibre debonding at the intersection of the fibre and matrix failure.94,637,638 In NFRCs, the stress applied is transmitted through their interface from the matrix to the fibre; therefore, the adhesion between the fibre/matrix plays a vital role in crack propagation.639–641

5.7 Thermal properties

The thermal properties of NFRCs significantly limit their industrial applications due to the degradation of their fibre component at higher temperatures.45 In NFRCs, thermal stability is their main property, which can be calculated using different thermal analyses. It is possible to apply thermal analysis to assess the moisture content and other volatile matter available in fibres that can induce the degradation of the fabricated composite materials. Most plant fibres appear to lose mass at about 160 °C.494 Among the important fibre constituents, lignin first starts to degrade at about 210 °C, whereas other constituents, mostly cellulose, are oxidized and deteriorated at elevated temperatures. Previous studies have demonstrated that the crystallinity index (CI) and cross-linking of cellulose have an influence on the thermal behavior of composites.642–644 Sinha and Rout645 reported that chemical surface modification increases the thermal stability of PFRCs. It was reported that the thermal resistance of JFRCs shows better stability than pure polymers and other fibres. A decrease in resistance was observed with an increase in the content of fibres (wt%), which is consistent with the higher amount of polymer resin than jute fibres.59 Also, the thermal properties of composites were investigated such as thermo-gravimetric analysis (TGA),56,59,61,353,354,356,368,376,560 differential scanning calorimetry (DSC),353,356,363,376 dynamic mechanical analysis (DMA), heat deflection temperature (HDT),292,381 dynamic mechanical thermal analysis (DMTA),182,383,646–665 thermal aging,311 fire flow test311 by various researchers.
5.7.1 Thermal conductivity. Thermal conductivity, diffusivity, and specific heat capacity are the three major thermal-based properties of composites, which are required for effective heat transfer. Thermal conductivity is assessed according to the ASTM-E1530 standards using a Guarded Heat Flow Meter. The test sample with dimensions of 50 × 4 mm is mounted and a load of 10 psi applied. The heating element provided at the top and bottom maintains uniform heat transfer. A calorimeter is positioned to act as a transducer of the heat flow and the samples are tested in the temperature range of 30–130 °C. The difference in temperature is measured from the output of the heat flow transducer548 after thermal equilibrium is reached. The formula to determine one-dimensional heat conduction is as follows:
 
image file: d3ta05481k-t5.tif(16)
where k is the thermal conductivity, T1T2 is the temperature difference, and D is the sample thickness. The thermal resistance and conductivity of composites can calculated using the following equations:
 
image file: d3ta05481k-t6.tif(17)
 
image file: d3ta05481k-t7.tif(18)

The specific heat capacity of the composites is determined by differential scanning calorimetry at 10 °C min−1 according to the ASTM E1269-11 standards.548 Thermal diffusivity is determined using the following equation:

 
image file: d3ta05481k-t8.tif(19)
where k, ρ and Cp are the conductivity, density and specific heat capacity, respectively.

5.7.2 TGA. TGA is conducted to assess the thermal stability and degradation profiles of composites. The thermogravimetric behavior of plant fibres directly depends on their chemical constituents.45 This method can be employed to characterize composites that display weight loss or gain due to dehydration or decomposition.381 The TGA of JFRCs was conducted using TG209F1 NETZSCH,353,354 Q50 TGA,368 Thermal Analyst 2100,61 TA Q500,59,356 SDT Q600,319 Mettler Toledo,345 Shimadzu DSC 50,379 Universal V1.12E56 and PerkinElmer346,362,376 instruments according to the ASTM E1131 standards345 in a complex O2/N2 atmosphere in the temperature range of 30–1000 °C at a heating rate of 10 °C min−1.319,346,353,354,376,379 TGA was conducted under an N2 atmosphere with a sample of weight 10–65 mg. The specimens were kept in a furnace and processing was done at the rate of 5–20 °C min−1 in the temperature range of 0–900 °C (ref. 56, 59, 356 and 362) and the corresponding weight loss was recorded. The degradation nature of the JFRCs is presented in Fig. 10(a and b). The presence of jute fibres had a significant effect on the thermal degradation behavior, which shifted from 315 °C to 322 °C. In another sample, the degradation process was delayed, and the residue left at 600 °C was also higher than that of the other bio-composites, indicating the higher stability of jute fibre films than other combinations.376
image file: d3ta05481k-f10.tif
Fig. 10 Characterization of JFRCs: (a) TGA, (b) DTGA thermograms, (c) effect of fibre content on DSC, (d) effect of chemical treatment on DSC, (e–g) DMA curves of E′, E′′, and tan[thin space (1/6-em)]δ, (h–j) dependence of E′, tan[thin space (1/6-em)]δ, and adhesion factor on temperature, and (k) HDT curves, respectively.
5.7.3 DSC analysis. DSC is conducted to analyze the melting properties and crystallization nature of composites.356 The instruments employed include a DSC200PC NETZSCH,353 PerkinElmer Diamond DSC,376 DSC7 PerkinElmer,363 DSC8000 PerkinElmer,346 TGA/DSC1 STAR, Mettler Toledo,345 Shimadzu DSC50 (ref. 379) and TA Q1000 (ref. 356) for the DSC analysis of JFRC samples. DSC is performed according to the ASTM D3418 standards345 under an N2 atmosphere in the temperature range of −50 °C and 600 °C at the heating rate of 5–10 °C min−1. Also, 5–10 mg sample in aluminum discs is covered and fed in the calorimeter heating equipment.353,376,379 The heat flow is measured using indium before testing. DSC curves are obtained from the thermal cycle, which offer reliable results in terms of the melting (Tm) and crystallization (Tc) temperature.356,363

Fig. 10(c and d) show the impact of fibre quantity and surface modifications on the DSC curves for the crystallization of a jute/PP composite. In the case of PP, crystallization occurred at 107 °C, while the temperature of the onset of crystallization was 118 °C. When jute fibres were incorporated in the resin, the crystallization temperature and onset crystallization of the composite increased to 111 °C and 119 °C, respectively. In the case of the untreated jute fibre composite, the crystallization temperature increased to 112.5 °C. This increased impact was due to the addition of the jute fibres, which acted as reinforcing agents and induced or favored the process of crystallization, leading to an increase in the crystallization temperature. The jute fibre reinforcement also favoured the formation of a transcrystalline layer on the surface, which increased the crystallization temperature, thereby supporting the crystallization. This was preceded by a decrease in the crystallization (DHc) value.356

5.7.4 DMA. DMA is performed over a broad temperature range for the determination of the viscoelastic properties, and particularly the glass transition temperature (Tg).362 DMA is also employed for the evaluation of the relative stiffness and the damping properties of composites. DMA tests are conducted to track the storage and loss moduli and loss factor as a function of temperature. The test is conducted using a Rheometrics RDA-III,362 Perkin DMA8000,515,556 DMA2980,358 DMA983,544 Q800 DMA,341,368,369,546 Mettler Toledo83 and Eplexor® 150N628 according to the DIN 53457,556 ASTM D4065 (ref. 515) and ASTM D4065-01 (ref. 443) standards by various researchers. In the literature, DMA was performed with specimens possessing dimensions of 27.4 × 3.1 × 3 mm,362 50 × 10 × 3 mm,628 60 × 6 × 3 mm (ref. 544) and 28 × 7 × 3 mm.515 The temperature was set in the range of −120 °C and 200 °C at the heating rate of 2–3 °C min−1 under an N2 flow.546,556 Static and dynamic strain rates in the range of 0.05% to 0.2% (ref. 83, 358, 362, 412, 515, 544 and 628) were used at 3 °C min−1 and 5 Hz.341,368

The storage modulus (E′), loss modulus (E′′) and tan[thin space (1/6-em)]δ of JFRC specimens were evaluated by DMA at the heating rate of 5 °C min−1 in the temperature range of 30 °C and 120 °C at a frequency of 1 Hz and 15 μm.443E′′ is the viscous component contribution of the material, representing the energy dissipated by the device as heat during the process due to viscous movements in the material. E′ was found to be high at room temperature, decreasing with an increase in temperature (Fig. 10(e)). In the case of the treated composites, E′ was observed to be higher than the untreated composite. This was attributed to the stronger inter-laminar fibre/matrix bonding and chemical treatment. The higher E′ with an increase in the content of nano-clay was observed to be an indicator of the higher crystallinity and brittle nature of the composite. A slight bump was observed at around 50 °C for the samples without nano-clay. The specimen entered the cold crystallization temperature range during the cycle, where there was an increase in the CI and a decrease in the E′ value.610

The shift in E′ of different JFRCs is depicted in Fig. 10(e), showing that the E′ values of all the specimens are almost the same, which decreased with an increase in temperature due to the change from the glassy to rough state. However, there was an increase in E′ with an increase in the nano-cellulose content in both regions, which showed the superior strengthening effects of jute fibres compared against the above-mentioned temperature. In the glassy area, the constituents were highly immovable, compact, packed tightly and had strong intermolecular forces, which resulted in a high E'; however, the intermolecular forces decreased with an increase in temperature and the constituents became mobile and lost their compact structure, resulting in a loss of stiffness, and thus E′. The uncoated JFRCs had the lowest E′ value in the given temperature range, as shown in Fig. 10(e). The JFRC had an E′ value of 4.05 GPa at 30 °C, whereas the composites with other combinations showed an average value of 5.5 GPa, resulting in a 16% to 56% increase in E′. In addition, E′ increased from 0.3 GPa to 0.94 GPa at 150 °C, representing a 71% to 235% increase respectively. This shows the increase in the stiffness of the reinforcement in the rough surface region with an increase in the concentration of nano-cellulose above the content of jute fibres.546

Fig. 10(f) presents the E′′ value as a function of temperature, which increased, and then decreased. The highest heat dissipation occurred at the highest E′′ value is higher, reflecting the Tg. The sudden increase in E′′ suggests the rapid increase in the free movement of the polymeric chains at higher temperatures, which is not possible at lower temperatures. This figure also shows an increase in E′′ with an increase in the concentration of the nano-cellulose coating, reflecting the maximum energy dissipation with an increase in wear rate.546 The tan[thin space (1/6-em)]δ of the coated JFRC is shown in Fig. 10(g), where peak tan[thin space (1/6-em)]δ value was observed for high energy dissipation, while a decrease in the tan[thin space (1/6-em)]δ peak was seen with an increase in the nano-cellulose concentration. The maximum peak of 0.59 for the composite was reduced to 0.42, reflecting a decrease of 40%, which suggests strong inter-molecular and physical interactions. This leads to excellent interfacial restrictions and low energy dissipation, given that E′ was affected more by the increase in nano-cellulose concentration than E′′. Also, the width of the tan[thin space (1/6-em)]δ peak was wider for the nano-cellulose-coated JFRCs. The increase in the width indicated an increase in the interface volume, reducing the movability of polymer particles in the device and maximum segments following an increase in the content of nano-cellulose in the JFRCs.546

The E′ of the JFRCs as a function of temperature is presented in Fig. 10(h), which shows a steady decrease in the E′ of all the JFRCs. In the glassy region, a small variation in E′ was seen, with the treated composites having higher values in the rub environment, which could be attributed to improvement in the fibre/matrix bonding at the interface, reduced molecular weight of the polymer particles, and excellent reinforcement of the surface-modified fibres. This resulted in an improvement in the thermal stability and strength of the material at higher temperatures.412 The tan[thin space (1/6-em)]δ of JFRCs as a function of temperature is presented in Fig. 10(i). The unmodified JFRC was observed to have a higher peak than the surface-modified JFRCs. The tendency of the composites having weak fibre/matrix connectivity to dissipate high energy was seen compared to the composites with strong inter-laminar connectivity, i.e., weak interface bonding. This resulted in higher damping,666,667 indicating the phenomenon of molecular relaxation in the interfacial region of the composites.412Fig. 10(j) shows the low adhesion of the surface-treated JFRCs compared to the untreated JFRC, which shows an enhancement in the fibre–matrix bonding.412

5.7.5 DMTA. DMTA was performed on SFRCs to investigate the effect of incorporating fillers, surface modifiers, chemical agents, compatibilizers, etc. on the fibre/matrix interface.646–657 Similar experiments were performed for natural fibres, given that there were no comprehensive studies on reinforcing filler.658–661 DMTA experiments were performed in the temperature range of 0 °C to 200 °C at a heating rate of 3 °C min−1.383 The size of the specimen was 55 × 10 × 2 mm (ref. 182) and 53 × 13 × 1.5 mm (ref. 383) at a frequency of 1 Hz, and a relative strain of 0.1%. Ray et al.441 examined the DMT characteristics of vinylester-based JFRCs and observed a change in Tg. Some researchers662,663 reported a similar phenomenon for compatible jute fibres and epoxy resin. Some other authors also reported the presence of a secondary transition at a temperature higher than Tg.664,665 This increase was thought to be correlated with the micro-Brownian movement of the immobilized polymeric molecules in close proximity to the outer surface. According to the test, the E′ and tan[thin space (1/6-em)]δ values are calculated at different temperatures and Tg values of the DMTA system.182
5.7.6 HDT. The HDT test, which is also known as the heat distortion temperature test, is carried out according to the ASTM D648 (ref. 381) and DIN EN ISO75 (ref. 292) standards. Fig. 10(k) shows the HDT of the composites in relation of the fractions of jute fibre. The HDT value for 25% jute fibre increased to above 100 °C, and simultaneously, the impact energy loss was around 6%, retaining a value of 9 kJ m−2. The enhancement in HDT was small for a higher concentration of jute, while the force of impact was appreciable. Jute fibres decreased the HDT to 22 kJ m−2 due to the beneficial impact of their internal structure and mercerization on the HDT due to the enhanced properties.292
5.7.7 Thermal ageing. The thermal ageing of composites is due to its increasing use for structural applications and changes in environmental conditions. Micro-cracking is observed, which is often a damaging characteristic that can be induced in polymers exposed to cryogenic conditions. Tensile strength tests were performed on specimens kept at 75 °C for 10 h and subjected to deep freezing at −75 °C for 6. Delamination growth, matrix cracking, and fibre failure are the most common modes of damage in thermal ageing. The cryogenic condition introduced matrix cracking and/or debonding at the interfaces. The fibre/matrix bonding was weak during cryogenic conditioning. At low temperatures, the first type of damage was usually matrix micro- and inter-laminar cracks. This is one of the reasons for the decrease in the tensile strength of composites when exposed to low heat. It is well known that heat conditioning at higher temperatures provides better bonding, and therefore enhanced mechanical properties are achieved, given that the fibre cross-linking is very high during heating at higher temperatures, increasing the mechanical properties of the composites. Both NFRCs and SFRCs exhibit similar behavior in thermal ageing processes.311
5.7.8 Fire retardant test. The fire-retardant characteristics of composites would easily diminish due to their ignitability. This is important for a reduction in the combustible properties of composites.554 In this case, a flow test is conducted according to the ASTM D635 standard and the rate of burning is evaluated from the test. Fire retardants such as Al2O3, MgO, and H3PO4 were used for enhancing the fire-retardant characteristics of JFRCs.554 NFRCs showed low burning rates compared to SFRCs; hence, the fire-retardant properties of NFRCs were observed to be much better than SFRCs. The diameter of natural fibres is higher than that of any synthetic fibre; therefore, the composite layer acts as an insulating material, thus slowing down the rate of burning of the materials.311

5.8 Acoustical properties

An acoustic emission (AE) test was conducted on JFRCs by positioning sensors at both ends of the specimen to allow linear localization of the acoustic signals. The experiments were conducted based on the ASTM E976 standard.336 The detection point was localized between the sensors to avoid disturbances from the detection of the acoustic signals. The sensors used for the AE resonated at a frequency of 150 kHz. The sensors were mounted on the specimen surface to allow linear localization, while the two other sensors acted as guard sensors to distinguish AE and noise signals.542 Implementation of the fragmentation approach with non-destructive AE helped to determine the micro-failure mechanism of the specimen. Fig. 11 displays the AE signal strength and energy of untreated, alkaline, and silane-treated JFRCs. The AE signals indicated that the energy was very low before the boiling water test due to the occurrence of a micro-failure cycle at the swelled fibrils after the boiling water test. With an increase in the AE energy, its amplitude and signal strength were seen as a result of the chemical surface modification of the jute fibres. The results obtained for silane-modified jute fibre after the boiling water test were compared to that of other combinations. The AE signal strength and its amplitude were high compared to the other combinations after the boiling water test.61
image file: d3ta05481k-f11.tif
Fig. 11 AE amplitude of untreated and chemically treated JFRCs. Reproduced from ref. 61 with permission from Elsevier, 2008.

5.9 Wear properties

The objective of the Green Tribology Congress held in Japan in 2009 urged materials science researchers to focus on reducing the dependency on synthetic materials and minimizing the wastage of materials due to wear.16,668 In this case, PFRCs can be used as components, which exhibit good tribological properties in many cases due to their effect on wear and tear, depending on the material selection and processing conditions.669–674 An abrasive test was conducted to study of the strength of the bonding at the interface of composites on their performance through use of a pin-on-disc apparatus according to the ASTM G65 standard.675 Wear test specimens with dimensions of 3 × 3 × 25 mm were abraded against a spinning disc, on which an abrasive paper was pasted. The experiments were carried out in various cycles with the sliding distance corresponding to 6–200 m by various researchers. A sliding speed of 0.418 m s−1 was used, and the weight loss was calculated at the end of each series of 100 cycles using a physical balance.368

Abrasive powder was fed into the contact surface, and then examined. The sample was loaded against the wheel using a lever and fine sand particles were added.549 The disc rotated with the aid of a DC motor at a speed of 600 rpm. The disc surface acted as a counterpart and was finished with a roughness of 0.6–0.7 μm by abrasion from a 1200 SiC layer. Measurement of the abrasion loss was performed using an electronic weighing machine with a precision of ±0.001 mg.676 In another experimental study, an abrasion tester was used for two-body abrasive wear tests.677 A 45 × 35 × 5 mm specimen was made to slide against a spinning drum, on which an abrasive sheet was placed, and readings were taken for four period sets. A sliding velocity of 2.56 m min−1 was applied, and the abrasion loss was estimated when 100 cycles per set were completed.375 The pre- and post-experiment weight of the specimen was reported via an electronic mini-balance that had a minimum count of 0.01 mg.389 The wear rate of the specimen was calculated using the following equation:

 
W = V/ρD (mm3 mm−1)(20)
where V is the weight loss, ρ is the density of the material and D is the sliding distance. The specific wear rate was as follows:
 
image file: d3ta05481k-t9.tif(21)
where Δm is the mass loss, ρ is the density, L is the sliding distance, and Fn is the applied load.

The wear rate vs. sliding velocity of the JFRCs is presented in Fig. 12(a), showing an increase in the common wear rate of the JFRCs with an increase in the sliding velocity. A decrease in the wear loss was observed with an increase in the sliding speed due to the surface contact between the specimen and wheel.327 The JFRC with 36 wt% fibre loading showed lower wear loss than other combinations. The effect of load vs. specific wear rate is presented in Fig. 12(b), which shows that under all operating conditions of the plain epoxy specimen, it displayed a higher specific wear than the other specimens, increasing with an increase in the content of fibres (wt%). However, the impact of load on the wear rate of the JFRCs showed an irregular trend. The same 36 wt% composite exhibited better wear characteristics than the other samples under the standard load conditions, where the minimum specific wear rate was obtained. The wear rate of the JFRCs was observed to be high at a specific load due to the fact that the resin was supported by a low volume of fibres, and the depth penetration of the abrasive particles could also be a reason for this. Ploughing or wedge forming could be the potential wear mechanism for reduced wear loss and addition of the fibre to the resin allowed a decrease in losses due to wear.313 The 36 wt% fibre-loaded composite showed the minimum wear rate at various sliding speeds and regular load under stable conditions due to the strong fibre–matrix interfacial bonding. The average wear rate of the 48 wt% composite was higher than that of the 36 wt% sample due to the inadequate fibre impregnation in the former, resulting in poor interfacial bonding.


image file: d3ta05481k-f12.tif
Fig. 12 Wear properties of JFRCs (a and b) effect of sliding velocity and normal load on the specific wear rate.

5.10 Water/moisture absorption properties

The major drawback of NFRCs is their poor resistance to moisture absorption.678 It is known that these composites are sensitive to moisture, which results in the swelling of their surface.679 This is even more challenging in plant fibres, which have low water-absorption resistance456,680–683 due to their hydrophilic characteristics and chemical composition, causing them to retain hydrogen water molecules.132,160 The intensity of moisture absorption depends on the fibre surface, substrate, fibre/matrix interface, relative humidity and surrounding temperature. The hybridization of plant fibres with less water-intake conventional fibres has an effective impact on the resulting composite materials,84,684–686 thereby allowing a reasonable performance. Factors such as fibre volume, reinforcing type, fibre composition, processing temperature, exposed surface area, polarity, degree of cross-linking and crystallinity affect their water-absorption nature.9,687–690

Moisture absorption tests on JFRCs were conducted according to the ASTM D570-98,381,548,550,556,574,691 ASTM D570-99,340,342,346 ASTM D570,374 ASTM D2495-07,443 and IS:2380-77 (ref. 608) standards. The samples were prepared with dimensions of 100 × 13 × 3 mm,443 40 × 10 × 0.5 mm,340 39 × 10 × 4.1 mm (ref. 342) and 80 × 80 × 2 mm (ref. 59) for water absorption studies. The samples were dried for 5 h at 105 °C and their weight recorded. Distilled water was poured into a desiccator for the creation a humid atmosphere and the humidity was found to be 98% after 12 h. Then, the samples were placed in the desiccator and their weight recorded at regular intervals.443 Subsequently, the specimens were dried at 105 °C, cooled with silica gel in a desiccator, and their weight recorded. Next, they were soaked in distilled water for 2 h. Afterwards, the excess water on their surface was removed using a soft cotton cloth. Finally, the weight the specimens was recorded.342

In another study, the specimens were first dried at 70 °C and their weight recorded using a weighing balance at a precision of ±0.01 g. Then, the specimens were immersed in water and left for a day. They were taken out at frequent intervals and their wet weight recorded.340 In another study, jute fibre was cured for 4 h at 99 °C, and then its moisture absorption was measured at 23 °C.105 The weight gain of the composite specimens submerged in water at 23 °C was taken for water absorption.381 The JFRCs were subjected to water absorption experiments through immersion in sea water, clean water, and acidic water. Different parameters were investigated including the diffusion coefficient and overall moisture quantity using the absorption curves.337 The strength in each case was found to be inversely proportional to the percentage water absorption.560 A water absorption analysis was performed for jute fibres and for the JFRCs. The samples were kept in water for 5 h at 23 °C, taken out and their weights recorded.317 The difference in weight was taken for the absorption of moisture.574 The moisture content or percentage increase in weight was determined as follows:315,336

 
Water absorption (%) = (W1W2/W2) × 100(22)
where W1 is the weight after immersion and W2 is initial weight of the specimen. The Fickian diffusion coefficient was determined using the following equation:
 
image file: d3ta05481k-t10.tif(23)
where D is the diffusion coefficient, Mm is the moisture content, h is the thickness of the laminate and k is the slope of the plot of M(t) versus t1/2, as determined using the following equation:
 
image file: d3ta05481k-t11.tif(24)

The corrected diffusion coefficient (Dc) is expressed as:

 
image file: d3ta05481k-t12.tif(25)
where h is the thickness; l is the length and n is the width of the specimen.

The curves for the absorption of water by the JFRC are shown in Fig. 13(a), displaying the proportion of water consumed versus immersion time. It can be observed that the water-absorbing nature was very quick until reaching a saturation point, after which the absorption rate showed a decreasing trend.336 The water absorption curves for the JFRCs reported by other researchers are shown in Fig. 13(b) and (c). The samples were submerged in normal water, water with pH 8, and acidic water with pH 3. The curves showed an increase in absorbed water content following an increase in immersion time. The JFRCs followed pseudo-Fickian behaviour, where the gain in weight after the initial increase never reached equilibrium.692 These composites are known to have a higher coefficient of diffusion and overall moisture content if submerged in purified water, followed by pH 3 acidic water. In contrast with sterile water, a decrease in the value of D and a decrease in moisture content were observed when the specimen was kept in water with pH 8 and pH 3 acidic solution. This is due to the presence of large molecules in pH 8 water, which delay the diffusion time, resulting in the minimum kinetic absorption parameter.693 Acidic water at pH 3 did not result in greater corrosion than purified water.694


image file: d3ta05481k-f13.tif
Fig. 13 (a) Water absorption curves of JFRCs and (b and c) weight gain of JFRCs after exposure to distilled water, sea water and acidic solution after flexural and compression tests, respectively.

5.11 FEA

Modeling and simulation are important processes used for the enhancement of design and analysis of the performance of NFRCs.19 A significant numerical and confirmatory approach is numerical analysis or FEA for the optimal solution to a variety of design problems.338 This was performed to determine the natural frequency of JFRCs553,695 under the pure opening mode, mixed mode, and pure shearing mode fracture power. This analysis was carried out using the ANSYS Work-bench commercial FEA programme. Rizal et al.695 performed FEA on JFRC plates and found that the frequency of their vibrations is critical for the identification of their Young's modulus, rigidity modulus and Poisson's ratio. The vibration modes and the natural frequency of the JFRCs are presented in Fig. 14,695 showing the natural frequency of 297.77 Hz and 222.25 Hz in the flexural and torsional mode, respectively.
image file: d3ta05481k-f14.tif
Fig. 14 Simulation results of vibration characteristics from FEA: (a) flexural mode and (b) torsional mode. Reproduced from ref. 695 with permission from AIP Publishing, 2019.

ANSYS was used to the test the thermal conductivity of the composites. Coding was done in APDL (ANSYS Programming Design Language), which delivered the output for the conductive heat transfer of the composites. 3D cylinders-in-cube physical models were used for the simulation of the unit cells of the composites at various contents of the fibres (wt%). A one-axis aligned orthogonal coordinate scheme with the orientation of the fibre was used in the study of the micro-mechanics of fibre-reinforced materials. The diameter of the fibre was measured, which matched the fibre load from 0–40 wt%. For the discretization of the constituents, a 3D quadratic brick element, SOLID90, was used in the evaluation of the thermal properties.548 FE models for jute/glass fibre hybrid composites were created using the ANSYS19 software for the determination of the J integral and stress intensity factor. The FEA results were observed to be closely correlated with the experimental results.530 The thermal expansion coefficient of JFRCs was evaluated by FEA using the representative unit cell method. Also, a hexagonal array unit cell was used for the study of the contiguity of JFRCs.338 The results showed that the coefficient of thermal expansion increases with decrease in the fibre content.

A continuum damage mechanics-based mesoscale model was developed to describe the in-plane tensile response in NFRCs. These composites demonstrate a nonlinear response, with an initial rapid stiffness degradation rate, which eventually decreases, unlike SFRCs, where their stiffness degradation tends to be constant and linear. The elementary fibres were treated as composites, idealised as cylindrical tubes of varying diameters, and reinforced by microfibrils at variable orientations, as governed by the classical laminate theory.696 Andersons et al.697 proposed a semi-empirical model, whereby tensor-based orthotropic stress–strain relationships were made to fit experimental observations of NFRCs, and the classical laminate theory was employed to simulate the laminate tensile response. The authors showed that a purely macroscale analytical approach can reasonably reproduce the nonlinear behaviour in NFRC unidirectional laminates. Panamoottil et al.698 demonstrated a ‘hierarchical’ approach to simulate the tensile response of a single resin-impregnated natural fibre. Analytical microscale models were developed separately for elementary fibres, resin, and the fibre–matrix inter-phase layer, which were then combined in a finite element based ‘unit model’ of a fibre yarn impregnated and surrounded by matrix.

Poilane et al.699 developed a viscoelastoplastic model for fibre-direction-only tensile response in a single-ply NFRC. The strain response was considered split into pure elastic, viscoelastic, and viscoplastic components. Based on observations from creep and repeated load-unload tests, the authors proposed free energy and dissipation potentials that capture plastic yielding, temperature dependence and strain rate effects in the fibre direction. Sliseris et al.700 proposed micromechanical models within a thermodynamic framework for NFRCs under tensile loading. In this model, the fibre length and diameter were randomly distributed, and separate constitutive laws were defined such as elementary fibre, defected regions of fibre, and regions between elementary fibres and resin. The fibres were modelled simply as linear elastic, but the defected fibre and intra-bundle regions were modelled as brittle materials with linearly degrading stiffness after a specified threshold. The matrix was modelled with constant stiffness and von Mises plasticity with isotropic hardening. In the case of the woven fabric model, both the fibre and matrix were governed by nonlinear, isotropic hardening plasticity laws, but with no state variable laws that would permit any degradation of the material properties. Both models were executed via a finite element-based representative volume element loaded in tension. The models captured the initiation and progressive evolution of damaged zones, and the representative volume element response closely reproduced the experimentally observed nonlinear stress–strain response.

5.11.1 Dynamic relaxation method. In the dynamic relaxation method (DRM), the equilibrium equation is converted into the dynamic equation. The damping term is expressed in the finite difference, and the solution is obtained through the iteration process. The time increment and the optimum damping coefficient are used to stabilize the solution. The stiffness matrix, applied load, boundary conditions, and mesh size of the specimen are the main variables that affect the solution.

5.12 LCA

The aim of LCA is to determine the environmental impact of using plant fibres on the surroundings.701 The consumption of energy during the fabrication of bio-composites increases their environmental impact. This can be rectified through assessment of the life-cycle and service life simultaneously during the fabrication of the composite. It is also useful in minimizing the environmental effects during the processing of materials.120 LCA was developed for the analysis of the processing of PFRCs and improvements in the environmental impact of the reaction were found.702 Detailed analysis outside the system boundaries is required for the consequences of LCA if government policies permit the implementation of bio-based materials. A comparative LCA for a bio-based versus oil-based polymer composites was conducted by Rosa et al.703 This study was useful for the analysis of composite processing, ecological effects induced by fibre extraction and the consideration of bio-based and polymer-based resins. LCA on processing of poly(3-hydroxybutyrate) was conducted for the evaluation of its green-house gas emission, acidification and eutrophication potential. The results showed that parameters such as raw material selection and choice of methodology have an effect on the variability of the process.704

Alves et al.705 explained various ecological, social and economic merits of replacing synthetic fibres with jute fibres in an automotive structural front bonnet through LCA. The feasibility study of this application for environmental improvements was made as a case study designed by them. Environmental characterization was carried out with the use of LCA and a significant improvement in the environmental performances was seen. They also performed a quantitative analysis on the social and economic impacts and found that JFRCs performed better in both categories compared to SFRCs. Pandita et al.706 conducted LCA experiments on the processing of jute/glass fibre composites using the GaBi 4.3 software. The energy consumed during the processing of the jute/glass fibres was determined to be around 33 MJ kg−1 of fibres. With the positive energy adjustment as a result of the use of plant fibres, the energy consumed to process 1 kg of jute was subtracted with the lowest energy balance of 8.7 MJ kg−1.707

5.13 Spectroscopy studies

5.13.1 FTIR spectroscopy. FTIR spectroscopy is a useful tool for the identification of the functional groups present in composites based on the wavenumber of the peak of the surface molecular layer with respect to its transmittance.708 FTIR spectrometers including Tensor 27 Bruker,327,353 Shimadzu FTIR81001,342,345,356 PerkinElmer,346,539 PerkinElmer 1720X,362,378 Nexus 870,354,371,613 ABB Bomen FTLA2000-102,385 Nicolet 6700,368 Nicolet iN10 (ref. 319) and Nicolet 10DX56 instruments were used in the analysis of composites. The spectrum of the sample was recorded with scans in the range of 32–64.319,327,346,353,354,362,378,539,613 The FTIR spectra of KBr-coated JFRC samples were examined in the wavenumber range of 4000–400 cm−1 at a resolution of 2 cm−1. The fibres were cured at 85–105 °C for one hour, and then the samples were coated with KBr for testing.354,613 FTIR measurements on oven-dried JFRC pellets were carried out at 105 °C for 1 h. Also, the FTIR spectra of untreated and chemically treated JFRCs were recorded.342 Analyses of silane-treated fibres showed a peak value of 1100 cm−1. The peaks near 1100 cm−1 indicated the formation of Si–O–C bonds, showing the reaction of the silane coupling agent with the hydroxyl groups in the fibres.386

The FTIR spectrum of jute fibres presented in Fig. 15(a) shows the –OH stretching bond, indicating the presence of hydroxyl elements in the fibres.373 The peak at around 2900 cm−1 is attributed to the C–H bonds, while the peaks at 1600, 1400, and 1060 cm−1 represent –OH bonding, –CH2 bonding, –C–O pyranose, and glycosidic C–H bonding, respectively. A broad absorption band was typical in all the spectra in the range of 3200–3600 cm−1, corresponding to the O–H stretching (Fig. 15(b)).709 Alkali-treatment caused a reduction in the O–H stretching, resulting in more O–H groups accessible for the acetic-acid reaction. At around 2950 cm−1,710 the C–H stretching frequency of the methyl groups and methylene was observed. However, the intensity of this peak decreased upon surface treatment. Hemicelluloses are the key elements of fibre alteration, showing a peak at around 1730 cm−1 for the C[double bond, length as m-dash]O bonding of carboxylic acid and hemicellulose constituents. In the case of the alkali-treated fibres, they were dissolved in the chemicals, and thus no peaks were observed for the C[double bond, length as m-dash]O stretching. Lignin611 is responsible for the absorbance band at around 1464 cm−1. This peak was reduced by chemical treatment, indicating the extraction of lignin from the surface-modified fibres.


image file: d3ta05481k-f15.tif
Fig. 15 FTIR spectrum of (a) raw and (b) chemically modified jute fibres. Reproduced from ref. 709 with permission from Elsevier, 2008. (c) ATR-FTIR spectra of raw and chemically modified jute fibres and (d) XRD patterns of jute fibers and cellulose nano-fibrils.

An absorption band at around 3430 cm−1, which is characteristic of the stretching vibration of –OH bonded with hydrogen, was common in all the spectra. In the spectrum of the detergent washed and dewaxed jute, the peak at 1740 cm−1 can be attributed to the C[double bond, length as m-dash]O stretching vibration of hemicellulose. Alternatively, the sharp intensification the ester group appeared in the case of the MMA-grafted jute at 1743 cm−1. The grafted jute fibre showed an intense peak at 2244 cm−1, corresponding to the nitrile group, whereas the cyano-ethylated fibres show a low-intensity peak. This can be explained by the migration of many acrylonitrile monomers in the chain in the AN-grafted jute with the addition of only one AN molecule in the cyano-ethylated samples per group of cellulose macro-molecules in –OH. The positive Lassaigne Test confirmed the presence of N2 in the cyano-ethylated specimens.711 A similar observation was recorded for cyano-ethylated jute.436

5.13.2 Attenuated total reflectance (ATR) FTIR spectroscopy. ATR-FTIR was applied for the identification of the functional groups in jute fibres. The ATR technique was used to test samples in different states without any preparation.356 This testing practice in combination with infra-red light permits the direct examination of composites without prior preparation.358,712,713 A variety of materials is employed in ATR testing, including the widely used germanium, diamond, platinum and ZnSe.357 ATR-mode spectrometers including Bruker Vertex70,341,374 Shimadzu FTIR81001,357,358 and Nicolet 6700DX443 instruments have been used by different researchers. A total of 64 scans was performed in the wavenumber range of 400–4000 cm−1 at a resolution of 4 cm−1.341,356–358,374,443 The spectra were standardized at 895 cm−1 from the vibration provided there was no subsequent shift in the peaks.341 At intervals of 60 min each spectrum was captured using 32 scans.443 The ATR-FTIR spectra of jute fibres are presented in Fig. 15(c), where the increments in the absorption peaks for NaOH-treated jute fibres can be clearly observed. The silane-treated jute fibres were produced a lower transmittance due to the changes in the hydroxyl groups of cellulose and lignin. These results proved the occurrence of changes on the surface of the fibres due to the chemical modifications.374
5.13.3 XRD studies. The structural features of jute fibres and JFRC samples were investigated via XRD analysis. The dried samples were filled in the rectangular hollow region of the sample holder.353 The XRD patterns were recorded using an Ultima,327,353,613 Bruker D8 Focus,319 PANalytical X'Pert PRO MPD546 and Bruker AXS390 X-ray diffractometer in different studies. These instruments produce X-rays using an Ni filter from CuKα radiation. The CuKα radiation was used to record the XRD patterns at a scan rate of 1° min−1 (ref. 327 and 613) for 2 h in the 2θ range of 10 °C to 60 °C. A TG analyzer was used to investigate the thermal decomposition in the temperature range of 30–500 °C at a heating rate of 10 °C min−1. The prepared specimens were measured at 200 mL min−1 in fresh ambient air. A sample mass of 10 mg was taken for every experiment.390 The size of the crystallite was calculated using the following formula:
 
Crystallite size = 0.9λ/(β[thin space (1/6-em)]cos[thin space (1/6-em)]θ)(26)

The percentage CI was estimated using the following relation:

 
% of CI = (I002Iam)/I002 × 100(27)
where I002 represents the intensity of the (002) reflection from the crystalline cellulose lattice and Iam represents the intensity of the peak from the amorphous portion of the fibre.714 The XRD patterns of jute fibres presented in Fig. 15(d) display diffraction characteristics of semicrystalline materials with small amorphous peaks. Typical well-defined peaks were observed at 2θ values of 16° and 23°.546

5.13.4 X-ray photoelectron spectroscopy. XPS assessments are performed using a spectrophotometer fitted with a monochromatic X-ray source. Several spectrophotometers operate at 1486.6 eV using AlKα radiation. The energy scale is set to a value of 83.98 eV with the help of Au photoelectrons. The spectra are captured at 160 eV in the binding energy range of 0–1400 eV. This determination is done using the ratio of the supplied oxygen to carbon. The chemical states on the surface of the sample are determined by fitting the curve of carbon and oxygen. The model is equipped with a function of the Lorentzian–Gaussian mix Voigt profile, and a program is used for fitting the non-linear least squares curve.341

5.14 Morphological property analysis

5.14.1 Scanning/field emission scanning electron microscopy analysis. Scanning electron microscopy (SEM) is a method for the examination of the fibre/matrix adhesion based on surface morphology at the microscopic level. Specifically, the surface morphology of broken composites shows the degree of fibre/matrix adhesion.715 The SEM specimens are sputtered with gold/palladium/platinum before their evaluation to reduce/avoid electro-static charging during examination.41,56,81,296,302,315,319,327,336,339,344,345,348,353–355,362,368,381,387,389,390,412,515,539,544,546,556,574,607,608,613,628 In the literature, images were captured using Philips FEG XL30,41,389,390 XL20,56 Philips 501,608 Hitachi S-4500,99 TM3000,296,319,379 TM3030,346 S-3400N,378 SU1510,369 Jeol JSM7600F,340 JSM-5600LV,356–358,371 JSM-6480LV,315,549,550,574 JSM-5610,387 JSM-6701F,342 35CF,375,628 JSM5800,362 JSM6060,339 JSM840,385 JSM6400,344,374 JSM7001FD,81 TM3030,345 LEO 440,515 Leo Supra 50VP,83 Quanta 200F,539 Vega LSV,353,354,613 Vega-Tescan TS5130,546 Zeiss EVO MA,302,348,368 FESEM Zeiss Auriga,336 Zeiss DMS960,607 Siemens XL30,556 DXS-10ACKT,381 TS5130-Tescan,412 and LEICA S440 (ref. 544) microscopes by various researchers. The morphologies of the composites were observed at a 5 kV,339,379,556 10 kV,81,344,368,381,515 15 kV,296,348,385,387,539 20 kV,56,412,544,574 25 kV (ref. 302, 336 and 607) and 30 kV (ref. 41, 378 and 546) acceleration voltage.

Fig. 16(a–f) show the SEM images of jute fibre morphology in the longitudinal direction and its failure modes. Generally, the longitudinal shapes of the jute fibres are not uniform along their length.61Fig. 16(c–e) demonstrate the micro-failure mode under axial load for raw, chemically treated, and silane-treated jute fibres, respectively. This can be affected by tensile stress prior to the failure in each case before the boiling water test, while the micro-failure mode after the test can be affected by the low tensile strength due to the fibrils swelling in the fibres. In the silane-treated sample, the fibres did not show any swelling following the water absorption due to the strong bonding between their adjacent layers.61 The surface defect due to the concentration of chemicals is shown in Fig. 16(f).343Fig. 16(g) shows bundles of jute fibres combined by their constituents. Following surface modification, the fibres were observed to split due to the degradation of the surface mesh structure, and whole portions of the cellulose and non-cellulose materials were removed, as observed in Fig. 16(h). The fibres were further isolated into individual fibres due to the chemical treatment and surface fibrillation of the fibres was possible, as shown in Fig. 16(i), which may be due to their further realignment following bleaching treatment. The morphological changes occurring after the nanocellulose coating on the jute surface are depicted in Fig. 16(j–l). The nanocellulose precipitated on the fibre, forming a coating. The thickness of the coating slowly increased with the an increase in the concentration of nanocellulose, as evident from the images.


image file: d3ta05481k-f16.tif
Fig. 16 Morphology of jute fibre: (a) axial direction. Reproduced from ref. 61 with permission from Elsevier, 2008. (b) Longitudinal direction. Reproduced from ref. 348 with permission from Elsevier, 2016. (c) Failure of fibre in axial direction. Reproduced from ref. 61 with permission from Elsevier, 2008. (d) Failure of alkali-treated fibre in axial direction. Reproduced from ref. 61 with permission from Elsevier, 2008. (e) Failure of silane-treated fibre in axial direction. Reproduced from ref. 61 with permission from Elsevier, 2008. (f) Surface defects. Reproduced from ref. 343 with permission from Elsevier, 2013. (g) Raw fibre, (h) alkali-treated, (i) bleached, and surface coated jute fibres at: (j) 0 wt%, (k) 2 wt%, and (l) 5 wt% of nano-cellulose.

The changes in the fibre surface following various surface treatments are described in Fig. 17(a–f), demonstrating major texture changes in surface after treatments. Fig. 17(a) shows the raw and fragmented surface morphology of the enzyme-treated fibres, arising as a result of the removal of the binding constituents from the fibres. Fig. 17(b) shows the thermal degradation following the laser-treatment of the fibres, resulting in a rough and porous surface. The crack growth and increase in surface roughness of the jute fibres treated with ozone are visible in Fig. 17(c). Plasma treatment induced a small increase in the surface roughness of the fibre (Fig. 17(d)).412Fig. 17(e) displays the surface of the KMnO4-treated jute fibre. The images clearly show that the fibre surface become rough after KMnO4 treatment compared to the raw fibre surface. This rough surface initiated the formation of fibre–matrix interlocking, which yielded better properties.357Fig. 17(f) displays the micrograph of the chemically treated jute fibres soaked in cement paste. As the surface of the jute fibres became degraded, the key explanation for the decomposition of the fibres in chemical medium is Ca2+ fastening on the fibre surface, whereas when the surface for the fibre is coated with polymer, the Ca2+ fastening tends to be reduced.327 The surfaces of the chemically modified fibres analyzed through SEM images showed comparatively rough surfaces due to the treatment (Fig. 17(g–i)). The chemically treated fibres provided cleaner and rougher surfaces due to the removal of impurities, non-cellulosic constituents, inorganic substances, waxes, etc. The rough surface texture and defibrillation were also a result of the stronger fibre-to-matrix interaction with a wider surface area.443 The removal of waxes and other impurities made the fibre cleaner and increases its bonding strength, as confirmed by the images presented in Fig. 17(j–l).371Fig. 18 displays the results of the cross-sectional micro structural analysis carried out on jute fibre. It highlights an irregular geometric cross-section characterizing jute fibre, which is usually not circular, and the diameter is not uniform throughout the length of the fibre. Moreover, along its longitudinal axis, the fibre cross-section is not constant.348


image file: d3ta05481k-f17.tif
Fig. 17 SEM images of modified jute fibres: (a) enzyme, (b) laser, (c) ozone, (d) plasma, (e) KMnO4, (f) immersed in cement paste for 90 days, (g) detergent washed, (h) ethanol washed, (i) detergent + ethanol washed jute fibres, (j) 3% NaOH, (k) 5% NaOH, and (l) 7% NaOH treated fibres.

image file: d3ta05481k-f18.tif
Fig. 18 SEM images of the cross-section jute fibres at their failure modes. Compiled from ref. 61 with permission from Elsevier, 2008,296 with permission from Elsevier, 2016,343 with permission from Elsevier, 2008, and348 with permission from Elsevier, 2016.

The fractured surface of the unaltered JFRCs is presented in Fig. 19(a). The fibre that emerged from the matrix confirmed the weak fibre–polymer bonding, resulting in the formation of voids in the matrix. The better adhesion prevented fibre pull-out and debonding. The SEM micrograph in Fig. 19(b) demonstrates the increased adherence of the fibres to the resin as a result of their successful bonding at the interface, resulting in reduced fibre pull-out by KMnO4 treatment.357 The fibre breakage and pull-out from the matrix are shown in Fig. 19(c), with no resin binding to the fibres (Fig. 19(d)). Clear adhesion on the surface of the jute fibre between the latex and the resin is evident in Fig. 19(e), where the latex-coated jute fibres are strongly adherent to the resin.544 The SEM images of the surface-modified JFRCs with different fibre contents are shown in Fig. 19(f–h).371 The SEM findings, as observed from the fractured specimens after the tensile examination in Fig. 19(i), support the enhanced interfacial adhesion.607


image file: d3ta05481k-f19.tif
Fig. 19 SEM micrographs of fractured surfaces of JFRCs: (a) raw fibre, (b) KMnO4 treated, (c) fibre breakage from the matrix, (d) resin layer on the fibres, (e) bonding between the fibre and the matrix, (f–h) effects of surface modification with different fibre contents on tensile properties and (i) fractured surfaces of JFRCs.

The water absorbed in fibres establishes hydrogen bonds, which minimize the fibre–matrix interactions. The interfaces of the virgin composites tended to be intact, showing pull-out fibres (Fig. 20(a)). Fibre/matrix failure was evident in the images after ageing (Fig. 20(b–e)), where the fibres removed during tensile testing are visible. Fibre degradation as the result of moisture content was confirmed by the existence of fibrillation and micro-cracks (Fig. 20(f–g)). It was also observed that the lumen size of the fibres in the JFRCs appeared to be fully occupied with the matrix (Fig. 20(d and e)).336Fig. 20(h–l) display the SEM images of the fractured JFRC surface. The SEM image of the KMnO4-treated surfaces of the broken composites is depicted in the Fig. 20(i), showing a decrease in fibre pull-out after chemical surface modification.358 A small gap between the fibre and resin and slight adherence of the resin on fibre were also seen. This is an indication of the saturation of the fibre surface –OH groups, requiring a higher concentration of compatibilizer, and accordingly the interface strength can be improved further.628 This is due to the changes in the moisture content and the stress produced at the joint due to the thermal expansion coefficient of the resin and failure at the interface of the fibre/resin bonding.716


image file: d3ta05481k-f20.tif
Fig. 20 SEM micrographs of fractured surfaces of (a–g) JFRCs, (h) KMnO4-treated and (i–l) Epolene G3003-treated JFRCs.
5.14.2 Scanning probe microscopy analysis. Topographic images of jute fibres were obtained through use of silicon nitride pyramid tip probes with a spring constant of 0.12 N m−1 and a contact mode digital instrument. Single fibres were taken and fixed on both ends on a glass slide. The test was conducted at five different spots on the fibre surface to capture the fibre surface topography.341 The typical photographs of the position of the jute fibres are shown in Fig. 21. The untreated fibres displayed some degree of ruggedness, as seen in Fig. 21(a). Alternatively, the jute fibres treated with He and acetylene plasma showed with a smoother surface. However, an increase in roughness was observed when He/N2 plasma treatment was performed. The roughness was found to have a very broad statistical variability due to the inhomogeneous existence of the jute fibres.341
image file: d3ta05481k-f21.tif
Fig. 21 Typical SEM contact mode images of jute fibres: (a) raw, (b) He treated, (c) He/acetylene treated and (d) He/N2 treated. Reproduced from ref. 341 with permission from Elsevier, 2011.
5.14.3 Optical microscopy analysis. The micrographs captured through use of optical microscopy are shown in Fig. 22. The images for the samples of coated jute fibres showed better matrix blends and enhanced impregnation within and outside the jute fibres compared to the non-coated specimens.363
image file: d3ta05481k-f22.tif
Fig. 22 Optical images of jute micro-braid JFRC specimens: (a, c and e) uncoated and (b, d and f) PP/PVA coated. Reproduced from ref. 363 with permission from Elsevier, 2006.
5.14.4 Polarized optical microscopy (POM) analysis. The reinforcement type in a composite has a major effect on its surface morphology. Jute fibres were reinforced with PP films and processed using a precisely controlled iso-thermal process, and then cooled.494Fig. 23 shows the POM photographs of iso-thermally crystallized samples,356 indicating an increase in the thickness of the transcrystalline layer with time. The growth of the transcrystalline layer eventually comes into contact with the spherulite boundary in the bulk. Another explanation for the formation of these layers is based on the thermal stress produced at the interface. The load-induced polymer chains serve as nucleation points.717
image file: d3ta05481k-f23.tif
Fig. 23 POM images of: (a) neat PP, (b) jute fibre in resin melt, (c) nuclei-growth initiating and (d) formation of transcrystalline layer. Reproduced from ref. 356 with permission from Elsevier, 2013.
5.14.5 Atomic force microscopy (AFM) analysis. The AFM technique is very sensitive to the tip–sample interaction due to the bonding difference, and also the contact area of samples, and thus can map topographical features. The 3D surface topography and root-mean squares of jute fibres were examined using AFM analysis. The fibres were immobilized, and scanning was performed in tapping mode with a silicon cantilever. The scanning rate was set at 5000 × 5000 nm2 and at a frequency of 2.0 Hz.369 Using an atomic force microscope at a resonance frequency of 300 kHz with a voltage of 220 mV, fracture surfaces were examined after the tests.105 A microscope with a Nanoscope controller was used to observe the dispersion of jute nano-fibrils in the matrix. The images were captured using a phosphorous-doped silicone tip in tapping mode with a nominal frequency of 312 kHz.376 Investigation of the jute nano-fibril dispersion into composites was done by AFM analysis which was observed to be uniform, as shown in Fig. 24.
image file: d3ta05481k-f24.tif
Fig. 24 AFM images of (a) 2D, (b) 3D, (c) untreated, and (d) laccase-treated jute fibres.

The jute fibre surface was rough, irregular and covered with impurities, as visible from the magnified images (Fig. 24(c)). The AFM image presented in Fig. 24(d) shows the natural surface of the jute fibres, with the roughness minimized possibly due to the enzymatic dislocation and redistribution of lignin. Several lignins in the jute fibres were reduced to low-molecular-weight fragments at the start of laccase treatment and dissolved in the aqueous media. Laccase-mediated lignin polymerization was predominant at the later stage of the reaction, and the dissolved lignins were bound on the fibre surface.369 Similar results were also seen in the AFM topography of the fibres on a smaller scale (Fig. 25). The phase picture also displays some 200 nm structures. The images show the details of the topography due to the surface nature of the fibre or larger topographical structures.382 An AFM study was carried out on the broken surface of the fibres. The picture shows a waxy surface with compact structures. The occurrence of failure in the inner layer of the composites was likely due to the covering constituents such as fat, lignin, pectin and hemicellulose.382


image file: d3ta05481k-f25.tif
Fig. 25 AFM images of (a) untreated, (b) NaOH-treated and (c) NaOH/(APS + ED)-treated jute fibres. AFM topography and phase images of (d) untreated, (e) NaOH-treated and (f) NaOH/(APS + ED)-treated jute fibre surfaces.

5.15 Structure-process–property relationships

The structure of plant fibres is based on the hierarchical arrangement of fibrils and the composition of the constituents present in the fibre with different proportions.70 Structure–property relationships were developed for a description of the characteristics of jute fibres.155 A relationship between structure and property was defined by the thermal stability of the material.45 The relationship between the mechanical property and structure of plant-based fibres can be described using the following relation:155
 
σ α Cfri(28)
where C is the crystallinity ratio, fr is Herman's factor and i is an exponent.

An increase in the tenacity of the chemically modified jute fibres with an increase in physical parameters was found according to the above-mentioned equation. Correlation with a factor of 87% was achieved. The degree of polymerization of the matrix has little influence on the structural parameters. The structure–property relationship of polymer-based JFRCs was analyzed by Sengupta.555 The analysis showed the random orientation of non-woven JFRCs, indicating better properties than woven JFRCs. The impact strength and moisture-absorbing nature of multi-layered JFRCs were also observed to be higher compared to single-layered JFRCs.

6. Machining characteristics of JFRCs

The need for machining operation of a certain degree is important in the production cycle of composites. Removal of the excess materials for the establishment of the exact shape and size of the finished goods is a subtractive process. One of the main machining operations performed in composite laminates718 is drilling, while milling is a significant machining operation that can produce the required shape, size and tolerances.529 Machining of NFRCs is a complex process due to their high abrasiveness, anisotropic nature and heterogeneous structure.719 During machining,720 issues such as fibre peel-up, pull-out, debonding, spalling, and fuzzing could be identified. An analysis conducted by Nassar et al.721 on the machinability of NFRCs helped in the identification of the key factors affecting the quality of machined parts. The correlation among cutting velocity, thrust force and feed rate for the drilling of PFRCs was reported by Chandramohan and Marimuthu.392 They found no remarkable effect of velocity on thrust force, and with an increase in cutting speed at lower feed rates, but a decrease at a higher feed rate. The following equation can help measure the surface damage of a drilled hole (delamination).
 
Delamination factor F = Dmax/D(29)
where Dmax is the maximum hole diameter and D is the diameter of the drill.

Yallew et al.718 conducted experiments on the impact of machining parameters on the output responses during drilling of JFRCs. They observed a close relation between the geometries of the tool and the delamination of the hole, and parabolic drills produced holes of better quality than solid drills with lower thrust forces. Celik et al.529 examined the impact of cutting parameters on the thrust force, delamination and surface roughness during milling of JFRCs and found that the cutting speeds and feed rates were highly affected by the thrust force, delamination and surface roughness. The effect of cutting parameters on the thrust force and torque in JFRC milling was investigated and it was found that the cutting speed and feed rate were highly influencing parameters, while the depth and feed rate were the dominant variables for torque.722,723

Using the grey relational analysis (GRA) technique in combination with the principle of component analysis, Sankar et al.724 optimized parameters such as speed, feed, cut depth and fibre angle in the machining of JFRCs. They also found that the optimal solution was obtained by the use of combined techniques. The surface roughness and material removal rate were optimized in the machining of jute/flax composites by using the Taguchi technique and ANOVA. The results showed that the spindle velocity and drill diameter are the most influential parameters for surface roughness and material removal rate.725 GRA and ANOVA were used in the test of the influence of treatment on the delamination factor of JFRCs.367 It was observed that alkali-treatment decreased the delamination at the exit and the feed rate was the most influential factor. Celik and Alp726 investigated the importance of milling parameters on the cutting force, delamination, vibration intensity and surface roughness. They found that an increased spindle speed produced an increase in the vibration and delamination responses and reduction in surface roughness. They also found an improvement in all responses with an increase in the feed rate.

7. Applications of jute fibres and JFRCs

7.1 Applications of jute fibres

➢Bulk packaging for agricultural commodities.

➢Food, garment, and non-textile industries as raw materials.

➢Making raw cotton bales, wrapping cloth and coarse cloth.

➢Making window curtains, chair covers, floor carpets, floor mats, hessian cloth, handbags, ornaments, fancy items, technical clothing, etc.

➢Use of jute plant butts for weaving low-priced cloth.

➢Store seeds, prevention of weeds and application of slow fertilizer to the seedlings.291

7.2 Applications of JFRCs

Bio-based composite materials are emerging as viable alternative reinforcements to SFRCs in several engineering applications.297,727 NFRCs are cost-effective products; hence, numerous applications in various fields in the form of furniture, roof panels, ceilings, partition walls, packaging materials, automotive, and railway compartment interiors and storage silos have been reported. These composites are also used in non-structural auto body parts such as door panels, package trays, hat racks, instrument panels, internal engine covers, sun visors, boot liners, oil air filters, seat backs and exterior underfloor panelling.568 PFRCs are potential materials for biomedical and packaging applications.728,729 A report on Audi car door panels replacing conventional breaking panels with NFRCs showed a decrease in the energy consumption by 45% and CO2 and methane emissions.82 Plywood, fibre boards of medium density, artificial boards, building components and plush doors were made from JFRCs.291 These composites have been applied in the biomedical field such as tissue engineering, diagnostics, drug delivery, and wound healing.730 Other applications of JFRCs include lampshades, shopping bags, paper-weights, helmets, toilet closets, bath devices, electrical fittings, covers, hoses, household post-boxes, milk-boxes, fancy tiles, food containers, temporary ceiling panels, bio-gas storage devices, and movable and pre-made buildings for use in times of natural disasters.5,731–733

One of the promising applications of JFRCs is based on their ability to be shaped into a variety of complex parts by harnessing their potential as reinforcement, i.e., making various products such as building components,520,537 food grain storage units,520 models of low-cost housing units,520,521 roofing,519 wood replacements,521,522 and pipes.734 In ancient times, JFRCs were used for making cords, belts, hessians, sack cloths, mats and tapes.291 They were also used in the manufacture of building materials, thin sheets, window frames, vehicle shutters, goods packaging, geo-textiles, printed circuit boards, absorbent material, household accessories, bio-degradable shopping bags, etc. Some of the applications of JFRCs are presented in Fig. 26.


image file: d3ta05481k-f26.tif
Fig. 26 Applications of JFRCs: (a) doors, (b) furniture, (c) partition boards, (d) fishing boat, (e) rugs, (f) boxes, (g) curtains, (h) mats, (i) tray, (j) stationery, (k) footwear and (l) bag.

7.3 Comparison of properties and applications

Plant-based fibres are environmentally sustainable substitutes to synthetic fibres, and thus used as reinforcements in polymer matrix composites. However, the mechanical properties of these fibres and their composites in load-bearing components and structural applications are limited. A comparison of the properties and applications of different plant-based natural fibres is presented in Table 9.
Table 9 Comparison of the properties and applications of jute fibres with other plant-based natural fibres
Fibre Density (g cm−3) Tensile strength (MPa) Tensile modulus (GPa) Specific modulus (approx) Elongation (%) Applications Ref.
Abaca 1.5 400–980 6.2–20 9 1.0–10 Twines, ropes, etc. 11, 132 and 735
Alaska 271–733 3.2–49.5 0.6–1.7 736
Alfa 0.89 35 22 25 5.8 Paper, construction, etc. 737
Areca 0.7–0.8 147–322 1.12–3.15 10.2–13.15 78, 132 and 134
Bagasse 1–1.5 200–300 16–28 19 1 Bio-fuel, pulp, paper, construction, fencing, panels, railing systems, decking and window frame 11, 136, 145 and 738
Bamboo 0.5–1 200–750 10–30 24 2.2–3.5 Construction, furniture, textile and paper 132, 134 and 739
Banana 1.4 450–550 10 10 2–8 Food container, textile, etc. 132 and 140
B. mori silk 1.5 200–210 6 18–19 78
Coir 1–1.5 100–220 3–7 4.5 10–50 Floor mats, door mats, storage tanks, packaging, helmets, mirror casing, roofing sheets, brushes and mattresses 11, 132, 133, 136 and 140
Cotton 1.2–1.8 300–790 5–13 8 4–12 Textile, goods and cordage and furniture 11, 132, 133, and 140
Curaua 1.5 100–1000 12–100 40 1–5 Car interior, furniture and acoustic insulation components 25, 132 and 740
Elephant grass 817 185 7.4 2.5 741
Flax 1–1.4 350–1850 28–100 50 1–3.5 Composite, panels, fencing, fork, laptop cases, window frame, tennis racket, bicycle frame and textile 11, 132, 133 and 138–140
Hemp 1.2–1.8 300–870 25–85 48 1.1–4 Construction, food, papers, medicine, fuel and textiles, cordage, furniture, packaging, geo-textiles and electronics 11,132,133,139 and140
Henequen 1–1.3 450–600 10–17 10.5 3.5–6 Textile industries 11,132,133 and140
Helicteres isora 544–1111 23.3–71.7 0.6–1.7 736
1–1.5 480–620 4–6.5 736
Jute 1.2–1.5 300–920 10–80 32 1.2–2 Construction, automotive, textile, building panels, packaging, door frames, geo-textiles and door shutters 11, 46, 132, 133 and 140–142
Kenaf 1.55 200–900 15–50 25 1.4–3 Construction, paper, mobile cases, bags, insulations, packaging, and textiles 11, 23, 132, 133, 139 and 140
Nettle 680 40 2 Food, medicals, cosmetics, and drug industries 143 and742
Oil palm 1–1.6 80–248 0.5–3.2 2 17–25 Palm wine, food, medicine, boiler fuel, roofing, building materials, door frame and windows 11, 132, 133 and 743
Niagara 400–741 16.7–45.6 0.6–1.7 736
Oliver 461–899 20.9–55.5 0.8–1.7 736
PALF 0.8–1.6 180–1627 1.44–82.5 35 1.6–14.5 Textiles, carpets, mops and curtains 132, 140 and 743
Petiole 690 185 15 2.1 741
Piassava 1.4 134–143 1.07–4.59 2 7.8–21.9 744
Rachilla 650 61 2.8 8.1 741
Rachis 610 73 2.5 13.5 741
Ramie 1.0–1.55 180–1627 1.44–82.5 35 1.6–14.5 Textiles, packaging, fishing nets, household furnishings and paper manufactures 11, 132, 133, 139 and 140
Root 1150 157 6.2 3 745
Sea grass 1.5 453–692 3.1–3.7 13–26.6 741
Sansevieria ehrenbergii 0.88 50–585 1.5–7.67 2.8–21.7 741
Sansevieria trifasciata 0.89 526–598 13.5–15.3 741
Sansevieria cylindrica 0.915 585–676 0.2–11.2 11–14 741
Sisal 1.33–1.5 363–700 9.0–38 17 2.0–7.0 Construction, automobile, doors, roofing sheets, paper and pulp, and shutting plates 11, 132, 136, 140, 144 and 743
Snake grass 0.887 278.82 9.71 2.87 741
Spata 0.69 75.6 3.1 6 741
Spider silk 870–970 12 17.5 78
Talipot 890 143–294 9.3–13 2.7–5 745
Tussah silk 1.3 250 6 34 78


8. Opportunities and challenges

In recent times, several challenges have emerged with the exploitation of petroleum products such as high cost and degradation of the environment. Consequently, researchers have focused on developing bio-composites as a replacement for petroleum-based products. The extensive overview of jute fibres and their composite processing, properties and characterization described in this review offer insight into opportunities for the researchers in this field. Future research on jute fibres and their biocomposites should focus on the following aspects:

➢Significant research is needed for the elimination of deficiencies seen in JFRCs including fibre uniformity; mechanical strength; interfacial bonding and reduced susceptibility to moisture; bio-degradability characteristics; manufacturing methods to increase fibre content and fibre orientation; and delamination behavior to optimize efficiency.

➢Jute fibres have a hexagonal cross-section, varying chemical composition and fibre structure, resulting in poor bonding to typical resins. Swelling and microbial attack make jute fibres vulnerable to dimensional instability. Exploration of the possibilities of using surface treatments to reduce these limitations of jute fibres is necessary. To make cost-effective and eco-friendly JFRCs, surface modification techniques must be improved. Also, research must be carried out to predict the appropriate modification techniques for suitable composite application.

➢The anisotropic nature of jute fibres and their composites may cause dimensional changes during machining. Care is required in the selection of the cutting conditions, machining parameters, tool dimensions, tool geometry, etc.

➢Modern methods for providing inter-ply fibre reinforcement have been developed using techniques such as knitting, 3D weaving or fabric-wide weaving and pinning. However, these methods are not economical, inefficient and have design-specific limitations, which can be solved by creating new material processing techniques.

➢The durability of JFRCs is a major unpredictable factor. Thus, new techniques or methods should be developed to investigate and evaluate their durability and biodegradability properties.

➢At the laboratory level, existing pre- and post-treatment JFRCs are appropriate, but they are not very suitable for commercial products. The length of these procedures needs to be reduced for use in large-scale manufacture.

➢Extensive research is needed to overcome impediments such as water intake and moisture absorption for long-term stability in outdoor applications, specifically in extreme weather conditions.

➢A composite comprised of three entities is vulnerable to reinforcement, matrix and interface failure. The failure of one may initiate failure in the other and the process occurring can be identified by the load needed to activate individual fibres. Composite failure should usually be controlled by the mechanism triggered by stress. Focus on important major aspects such as fibre treatment and resin/matrix bonding is imperative for location of possible fields for the application of JFRCs.

➢JFRCs are used in various sectors such as automobiles, aerospace, and electronics. However, more research has to be carried out for their application in the biomedical field as scaffolds. The further exploration of this topic of research and the industrial use of JFRCs will open a new pathway.

9. Summary and concluding remarks

The desire for more environmentally friendly materials in the modern world has caused researchers to focus on natural cellulosic fibers, which have successfully replaced synthetic fibers in a variety of diverse uses. The main aim of this review was to get the attention of materials researchers to explore the potentiality of jute fibres as an alternative to replace synthetic reinforcements in composites. Jute fibres show exceptional properties and their ecological considerations are attractive features for stimulation of the urge to use jute as a reinforcement. This review is the collective source of preceding experimental effort and meticulous view of jute fibres from their processing and properties to their product level for various applications. Their processing methods have been well recognized with adequate analysis of several key properties. According to this review, they can be summarized as follows:

➢Jute fibres obtained by the retting process have higher strength compared to other processes. Hence, the retting process is recommended for extracting the fibres from plants.

➢Jute fibre has been chemically treated with various reagents for a reduction in cellulose hydrophilicity. These treatments have been found to strengthen the adhesion between the fibres themselves, as well as the fibres and matrix.

➢Surface treatment can help in the removal of the hemicellulose, lignin, and pectin that cover the outer surface of fibre cell walls, leading to the breakdown of the cell structure and the separation of the fibers into smaller filaments.

➢Composites with jute fibre reinforcement can be produced using several methods. Among them, many researchers have used hand lay-up, compression moulding, vacuum bag moulding, resin transfer infusion and pultrusion processes.

➢Correct selection and monitoring of process parameters, for example, working temperature, load, time, relative and humidity, are necessary pre-requisites for the successful manufacture of JFRCs.

➢The fibre surface decomposition, resin flow pattern, void formation, cross-linking of fibres, and wetting conditions are the major parameters that have an effect on the properties of the prepared composites.

➢Substantial improvements in the properties of surface-treated JFRCs relative to untreated JFRCs are viable.

➢The mechanical strengths of JFRCs increase initially with the addition of the fibres, and then decrease when the fibre content exceeds 50%.

➢Tensile and flexural strength and their moduli impact the energy absorbed and the hardness of JFRCs can be improved with an increase in fibre content.

➢The addition of jute fibres to polymer matrices increases the wear characteristics of the composites by a reduction of 3.5–45% in the friction coefficient.

➢DMA tests confirm the modification in the fibre/matrix interphase by the addition of compatibilizers and chemical treatment of the fibres.

➢Breakdown of the jute fibres under stress and voids between the fibre and matrix was shown by electron microscopy observation, suggesting the probability of enhancement of the reliability fibre/matrix interface.

➢LCA results show the use of resin as the most influential factor on environmental impact. The comparison of the environmental cautiousness in JFRCs showed that they are better than SFRCs.

➢FEA results show the occurrence of high correlation between experimental and theoretical results.

➢The conclusion from the machining of JFRCs is that responses such as thrust force, torque, and surface roughness are highly affected by the speed of cutting, depth of cut, feed rate, geometry of the tool and material.

Abbreviations

AEAcoustic emission
AFMAtomic force microscopy
ATRAttenuated total reflectance
DSCDifferential scanning calorimetry
DMADynamic mechanical analysis
DMTADynamic mechanical thermal analysis
FEAFinite element analysis
FTIRFourier transform infrared spectroscopy
HDTHeat deflection temperature
HeHelium
IFSSInterfacial shear strength
ILSSInterlaminar shear strength
JFRCsJute fibre-reinforced composites
LCALife cycle assessment
MAgPPMaleic anhydride grafted polypropylene
NFRCsNatural fibre-reinforced composites
N2Nitrogen
O2Oxygen
PFRCsPlant fibre-reinforced composites
POMPolarized optical microscopy
PLAPolylactic acid
PPPolypropylene
KMnO4Potassium permanganate
SEMScanning electron microscopy
NaOHSodium hydroxide
TGAThermogravimetric analysis
TEMTransmission electron microscopy
UTMUniversal testing machine
XPSX-ray photoelectron spectroscopy

Conflicts of interest

There are no conflicts to declare.

References

  1. M. S. Charan, A. K. Naik and N. Kota, et al., Review on developments of bulk functionally graded composite materials, Int. Mater. Rev., 2022, 67, 797–863 CrossRef.
  2. V. Lotfi, H. Li and D. V. Dao, et al., Natural fiber-reinforced composites: A review on material, manufacturing and machinability, J. Thermoplast. Compos. Mater., 2021, 34, 238–284 CrossRef.
  3. M. Ramesh, L. Rajeshkumar and N. Srinivasan, et al., Influence of filler material on properties of fiber-reinforced polymer composites: A review, e-Polym., 2022, 22, 898–916 CrossRef CAS.
  4. A. Gallos, G. Paes and V. Allais, et al., Lignocellulosic fibers: a critical review of the extrusion process for enhancement of the properties of natural fiber composites, RSC Adv., 2017, 7, 34638–34654 RSC.
  5. A. Gopinath, M. Senthilkumar and A. Elayaperumal, Experimental investigations on mechanical properties of jute fibre reinforced composites with polyester and epoxy resin matrices, Procedia Eng., 2014, 97, 2052–2063 CrossRef CAS.
  6. L. Zamora-Mendoza, E. Guamba, K. Miño, M. P. Romero, A. Levoyer, J. F. Alvarez-Barreto, A. Machado and F. Alexis, Antimicrobial Properties of Plant Fibers, Molecules, 2022, 27(22), 7999 CrossRef CAS.
  7. T. C. Mokhena and M. J. John, Cellulose nano-materials: new generation materials for solving global issues, Cellulose, 2020, 27, 1149–1194 CrossRef CAS.
  8. K. Adekunle, D. Akesson and M. Skrifvars, Mechanical and viscoelastic properties of soybean oil thermoset reinforced with jute fabrics and carded lyocell fibre, J. Appl. Polym. Sci., 2011, 122(5), 2855–2863 CrossRef CAS.
  9. H. N. Dhakal, Z. Y. Zhang and M. W. Richardson, Effect of water absorption on the mechanical properties of hemp fibre reinforced unsaturated polyester composites, Compos. Sci. Technol., 2007, 67(7–8), 1674–1683 CrossRef CAS.
  10. H. N. Dhakal, Z. Y. Zhang and M. W. Richardson, Creep behaviour of hemp fibre reinforced unsaturated polyester composites, J. Biobased Mater. Bioenergy, 2009, 3, 232–237 CrossRef CAS.
  11. O. Faruk, A. K. Bledzki and H. P. Fink, et al., Biocomposites reinforced with natural fibres: 2000-2010, Prog. Polym. Sci., 2012, 37, 1552–1596 CrossRef CAS.
  12. J. P. Correa, J. M. M. Navarrete and M. A. H. Salazar, Carbon footprint considerations for biocomposite materials for sustainable products: A review, J. Cleaner Prod., 2019, 208, 785–794 CrossRef CAS.
  13. F. P. La Mantia and M. Morreale, Green composites: a brief review, Composites, Part A, 2011, 42, 579–588 CrossRef.
  14. M. J. John and S. Thomas, Review bio-fibres and biocomposites, Carbohydr. Polym., 2008, 71, 343–364 CrossRef CAS.
  15. K. R. Sumesh, K. Kanthavel and V. Kavimani, Peanut oil cake-derived cellulose fiber: Extraction, application of mechanical and thermal properties in pineapple/flax natural fiber composites, Int. J. Biol. Macromol., 2020, 150, 775–785 CrossRef CAS.
  16. V. Chaudhary and F. Ahmad, A review on plant fiber reinforced thermoset polymers for structural and frictional composites, Polym. Test., 2020, 91, 106792 CrossRef CAS.
  17. B. P. Chang, A. K. Mohanty and M. Mishra, Studies on durability of biobased composites: a review, RSC Adv., 2020, 10, 17955–17999 RSC.
  18. A. Sailesh, R. A. Kumar and S. Saravanan, Mechanical properties and wear properties of kenaf–aloe vera–jute fiber reinforced natural fiber composites, Mater. Today: Proc., 2018, 5, 7184–7190 CAS.
  19. V. Alhijazi, Q. Zeeshan and Z. Qin, et al., Finite element analysis of natural fibers composites: A review, Nanotechnol. Rev., 2020, 9, 853–875 CrossRef.
  20. F. Catania, et al., Thin-film electronics on active substrates: review of materials, technologies and applications, J. Phys. D: Appl. Phys., 2022, 55, 323002 CrossRef.
  21. D. Chandramohan, B. Murali and P. Vasantha-Srinivasan, et al., Mechanical, moisture absorption and abrasion resistance properties of bamboo-jute-glass fiber composites, J. Bio- Tribo-Corros., 2019, 5, 66 CrossRef.
  22. R. Velmurugan and V. Manikandan, Mechanical properties of palmyra/glass fiber hybrid composites, Composites, Part A, 2007, 38(10), 2216–2226 CrossRef.
  23. H. M. Akil, M. F. Omar and A. A. M. Mazuki, et al., Kenaf fiber reinforced composites: A review, Mater. Des., 2011, 32(8–9), 4107–4121 CrossRef CAS.
  24. M. R. Bambach, Compression strength of natural fibre composite plates and sections of flax, jute and hemp, Thin-Walled Struct., 2017, 119, 103–113 CrossRef.
  25. P. Wambua, J. Ivens and I. Verpoest, Natural fibres: can they replace glass in fibre reinforced plastics?, Compos. Sci. Technol., 2003, 63, 1259–1264 CrossRef CAS.
  26. M. Y. Khalid, et al., Natural fiber reinforced composites: Sustainable materials for emerging applications, Results Eng., 2021, 11, 100263 CrossRef CAS.
  27. S. H. Kamarudin, M. S. Mohd Basri, M. Rayung, F. Abu, S. Ahmad, M. N. Norizan, S. Osman, N. Sarifuddin, M. S. Z. M. Desa and U. H. Abdullah, et al., A review on natural fiber reinforced polymer composites (NFRPC) for sustainable industrial applications, Polymers, 2022, 14(17), 3698 CrossRef CAS PubMed.
  28. J. P. Shebaz Ahmed, et al., A comprehensive review on recent developments of natural fiber composites synthesis, processing, properties, and characterization, Eng. Res. Express, 2023, 5, 032001 CrossRef.
  29. S. B. R. Devireddy and S. Biswas, Physical and mechanical behavior of unidirectional banana/jute fiber reinforced epoxy based hybrid composites, Polym. Compos., 2017, 38, 1396–1403 CrossRef CAS.
  30. F. I. Mahir, K. N. Keya and B. Sarker, et al., A brief review on natural fiber used as a replacement of synthetic fiber in polymer composites, Mater. Eng. Res., 2019, 1(2), 88–99 CrossRef.
  31. S. Panthapulakkal and M. Sain, Studies on the water absorption properties of short hemp-glass fibre hybrid polypropylene composites, J. Compos. Mater., 2007, 41(15), 1871–1883 CrossRef CAS.
  32. K. John and S. V. Naidu, Chemical resistance of sisal/glass reinforced unsaturated polyester hybrid composites, J. Reinf. Plast. Compos., 2007, 26(4), 373–376 CrossRef CAS.
  33. E. M. F. Aquino, L. P. S. Sarmento and W. Oliveira, et al., Moisture effect on degradation of jute/glass hybrid composites, J. Reinf. Plast. Compos., 2007, 26(2), 219–233 CrossRef CAS.
  34. K. S. Ahmed and S. Vijayarangan, Elastic property evaluation of jute–glass fibre hybrid composite using experimental and CLT approach, Indian J. Eng. Mater. Sci., 2006, 13(5), 435–443 CAS.
  35. A. A. Kafi, M. Z. Abedin and M. D. H. Beg, et al., Study on the mechanical properties of jute/glass fibre-reinforced unsaturated polyester hybrid composites: effect of surface modification by ultraviolet radiation, J. Reinf. Plast. Compos., 2006, 25(6), 575–588 CrossRef.
  36. R. S. Rana, A. kumre, S. Rana and R. Purohit, Characterization of properties of epoxy sisal/glass fiber reinforced hybrid composite, Mater. Today: Proc., 2017, 4, 5445–5451 Search PubMed.
  37. P. A. Sreekumar, R. Saiah, J. M. Saiter, N. Leblanc, K. Joseph, G. Unnikrishnan and S. Thomas, Effect of chemical treatment on dynamic mechanical properties of sisal fiber-reinforced polyester composites fabricated by resin transfer molding, Compos. Interfaces, 2008, 15, 263–279 CrossRef CAS.
  38. L. Y. Mwaikambo and M. P. Ansell, Chemical modification of hemp, sisal, jute, and kapok fibres by alkalization, J. Appl. Polym. Sci., 2002, 84, 2222–2234 CrossRef CAS.
  39. S. D. Salman, Partial replacement of synthetic fibers by natural fibers in hybrid composites and its effect on monotonic properties, J. Ind. Text., 2021, 51, 258–276 CrossRef CAS.
  40. Y. S. D. Girijappa, S. M. Rangappa and J. Parameswaranpillai, et al., Natural fibers as sustainable and renewable resource for development of eco-friendly composites: A comprehensive review, Front. Mater., 2019, 6, 226 CrossRef.
  41. G. M. A. Khan, M. Terano and M. A. Gafur, et al., Studies on the mechanical properties of woven jute fabric reinforced poly(L-lactic acid) composites, J. King Saud Univ., Eng. Sci., 2016, 28, 69–74 CrossRef.
  42. O. P. Balogun, K. K. Alaneme, A. A. Adediran, I. O. Oladele, J. A. Omotoyinbo and K. F. Tee, Evaluation of the physical and mechanical properties of Short Entada mannii-glass fiber hybrid composites, Fibers, 2022, 10(3), 30 CrossRef CAS.
  43. A. A. A. Rashdi, S. M. Sapuan and M. M. H. M. Ahmad, et al., Combined effects of water absorbtion due to water immersion, soil buried and natural weather on mechanical properties of kenaf fibre unsaturated polyester composites (KFUPC), Int. J. Mech. Manuf. Eng., 2010, 5(1), 11–17 Search PubMed.
  44. B. Baghaei, M. Skrifvars and L. Berglin, Manufacture and characterization of thermoplastics composites made from PLA/hemp co-wrapped hybride yarn prepregs, Composites, Part A, 2013, 50(1), 93–101 CrossRef CAS.
  45. M. Asim, M. T. Paridah and M. Chandrasekar, et al., Thermal stability of natural fibers and their polymer composites, Iran. Polym. J., 2020, 29, 625–648 CrossRef CAS.
  46. C. Alves, P. M. C. Ferrao and A. J. Silva, et al., Ecodesign of automotive components making use of natural jute fibre composites, J. Cleaner Prod., 2010, 18, 313–327 CrossRef CAS.
  47. X. U. Yuqiong, Y. O. U. Min and Q. U. Jinping, Melt rhelogy of poly (lactic acid) plasticized by epoxidixed soybean oil, Wuhan University, J. Nat. Sci., 2009, 12, 349–354 Search PubMed.
  48. K. G. Satyanarayana, G. C. A. Gregorio and F. Wypych, Biodegradable composites based on lignocellulosic fibres–an overview, Prog. Polym. Sci., 2009, 34, 982–1021 CrossRef CAS.
  49. S. Johnson, L. Kang and H. M. Akil, Mechanical behavior of jute hybrid bio-composites, Composites, Part B, 2016, 91, 83–93 CrossRef CAS.
  50. X. Zhou, S. H. Ghaffar and W. Dong, Fracture and impact properties of short discrete jute fibre reinforced cementitious, Mater. Des., 2013, 49, 35–47 CrossRef CAS.
  51. A. Gholampour and T. Ozbakkaloglu, A review of natural fiber composites: properties, modification and processing techniques, characterization, applications, J. Mater. Sci., 2020, 55, 829–892 CrossRef CAS.
  52. S. H. Lee and S. Wang, Biodegradable polymers–bamboo fibre biocomposite with bio-based coupling agent, Composites, Part A, 2006, 37, 80–91 CrossRef CAS.
  53. R. V. Patel, A. Yadav and J. Winczek, Physical, mechanical, and thermal properties of natural fiber-reinforced epoxy composites for construction and automotive applications, Appl. Sci., 2023, 13(8), 5126 CrossRef CAS.
  54. Z. Sun, Y. Duan, H. An, X. Wang, S. Liang and N. Li, Research progress and application of natural fiber composites, J. Nat. Fibers, 2023, 20, 2206591 CrossRef.
  55. C. F. Kuan, C. C. M. Ma and H. C. Kuan, et al., Preparation and characterization of the novel water crosslinked cellulose reinforced poly(butylene succinate) composites, Compos. Sci. Technol., 2006, 66, 2231–2241 CrossRef CAS.
  56. S. B. Hosseini, G. Milan, L. Haitao and H. David, Effect of fiber treatment on physical and mechanical properties of natural fiber-reinforced composites: A review, Rev. Adv. Mater. Sci., 2023, 62, 20230131 Search PubMed.
  57. A. Arbelaiz, B. Fernández and J. A. Ramos, et al., Mechanical properties of short flax fibre bundle/polypropylene composites: Influence of matrix/fibre modification, fibre content, water uptake and recycling, Compos. Sci. Technol., 2005, 65(10), 1582–1592 CrossRef CAS.
  58. J. M. Park, S. T. Quang and B. S. Hwang, et al., Interfacial evaluation of modified jute and hemp fibres/polypropylene (PP)–maleic anhydride polypropylene copolymers (PP-MAPP) composites using micromechanical technique and nondestructive acoustic emission, Compos. Sci. Technol., 2006, 66, 2686–2699 CrossRef CAS.
  59. T. T. L. Doan, H. Brodowsky and E. Mader, Jute fibre/polypropylene composites II. Thermal, hydrothermal and dynamic mechanical behaviour, Compos. Sci. Technol., 2007, 67(13), 2707–2714 CrossRef CAS.
  60. B. Madsen, P. Hoffmeyer and A. B. Thomsen, et al., Hemp yarn reinforced composites-I. Yarn characteristics, Composites, Part A, 2007, 38(10), 2194–2203 CrossRef.
  61. J. M. Park, P. G. Kim and J. H. Jang, et al., Interfacial evaluation and durability of modified jute fibers/polypropylene (PP) composites using micromechanical test and acoustic emission, Composites, Part B, 2008, 39(6), 1042–1061 CrossRef.
  62. N. Sgriccia, M. C. Hawley and M. Misra, Characterization of natural fiber surfaces and natural fiber composites, Composites, Part A, 2008, 39(10), 1632–1637 CrossRef.
  63. P. Sahu and M. K. Gupta, A review on the properties of natural fibres and its bio-composites: Effect of alkali treatment, Proc. Inst. Mech. Eng., Part L, 2019, 234(1), 198–217 Search PubMed.
  64. K. Hariprasad, K. Ravichandran and V. Jayaseelan, et al., Acoustic and mechanical characterization of polypropylene composites reinforced by natural fibres for automotive applications, J. Mater. Res. Technol., 2020, 9, 14029–14035 CrossRef CAS.
  65. G. K. Sathishkumar, M. Ibrahim and M. M. Akheel, et al., Synthesis and mechanical properties of natural fiber reinforced epoxy/polyester/polypropylene composites: a review, J. Nat. Fibers, 2022, 19, 3718–3741 CrossRef CAS.
  66. H. N. Dhakal, Z. Y. Zhang and M. O. W. Richardson, et al., The low velocity impact response of non-woven hemp fibre reinforced unsaturated polyester composites, Compos. Struct., 2007, 81, 559–567 CrossRef.
  67. D. Ray, S. Sengupta and S. P. Sengupta, et al., A study of the mechanical and fracture behaviour of jute–fabric–reinforced clay-modified thermoplastic starch-matrix composites, Macromol. Mater. Eng., 2007, 292(10–11), 1075–1084 CrossRef CAS.
  68. A. Pietak, S. Korte and E. Tan, et al., A novel technique for characterizing the surface of natural fibres, Appl. Surf. Sci., 2007, 253, 3627–3635 CrossRef CAS.
  69. S. Leduc, J. R. G. Urena and R. Gonzalez-Nunez, et al., LDPE/Agave fibre composites: effect of coupling agent and weld line on mechanical and morphological properties, Polym. Polym. Compos., 2008, 16(2), 115–124 CrossRef CAS.
  70. M. Ramesh, Flax (Linum usitatissimum L.) fibre reinforced polymer composite materials: A review on preparation, properties and prospects, Prog. Mater. Sci., 2018, 102, 109–166 CrossRef.
  71. G. Koronis, A. Silva and M. Fontul, Green composites: a review of adequate materials for automotive applications, Composites, Part B, 2013, 44, 120–127 CrossRef CAS.
  72. A. O'Donnell, M. A. Dweib and R. P. Wool, Natural fibre composites with plant oilbased resin, Compos. Sci. Technol., 2004, 64, 1135–1145 CrossRef.
  73. G. Bogoeva-Gaceva, M. Avella and M. Malinconico, et al., Natural fiber eco-composites, Polym. Compos., 2007, 28(1), 98–107 CrossRef CAS.
  74. H. Ku, H. Wang and N. Pattarachaiyakoop, et al., A review on tensile properties of natural fiber reinforced polymer composites, Composites, Part B, 2011, 42(4), 856–873 CrossRef.
  75. J. George, M. S. Sreekala and S. Thomas, A review on interface modification and characterization of natural fiber reinforced plastic composites, Polym. Eng. Sci., 2001, 41(9), 1471–1485 CrossRef CAS.
  76. H. Y. Heung, M. P. H. Ho and K. T. Lau, et al., Natural fibre-reinforced composites for bioengineering and environmental engineering applications, Composites, Part B, 2009, 40(7), 655–663 CrossRef.
  77. M. M. Kabir, H. Wang and K. Lau, et al., Chemical treatments on plant-based natural fibre reinforced polymer composites: An overview, Composites, Part B, 2012, 43(7), 2883–2892 CrossRef CAS.
  78. M. Ramesh, K. Palanikumar and K. H. Reddy, Plant fibre based bio-composites: Sustainable and renewable green materials, Renewable Sustainable Energy Rev., 2017, 79, 558–584 CrossRef.
  79. A. K. Jakubowska, E. Bogacz and M. Zimniewska, Review of natural fibers. Part I-vegetable fibers, J. Nat. Fibers, 2012, 9, 150–167 CrossRef.
  80. K. Palanikumar, M. Ramesh and K. H. Reddy, Experimental investigation on the mechanical properties of green hybrid sisal and glass fiber reinforced polymer composites, J. Nat. Fibers, 2016, 13(3), 321–331 CrossRef CAS.
  81. Y. Arao, T. Fujiura and S. Itani, et al., Strength improvement in injection-molded jute-fibre-reinforced polylactide green-composites, Composites, Part B, 2015, 68, 200–206 CrossRef CAS.
  82. S. V. Joshi, L. T. Drzal and A. K. Mohanty, et al., Are natural fibre composites environmentally superior to glass fibre reinforced composites?, Composites, Part A, 2004, 35, 371–376 CrossRef.
  83. M. Jawaid, H. P. S. A. Khalil and A. Hassan, et al., Effect of jute fibre loading on tensile and dynamic mechanical properties of oil palm epoxy composites, Composites, Part B, 2013, 45, 619–624 CrossRef CAS.
  84. S. Mishra, A. Mohanty and L. Drzal, et al., Studies on mechanical performance of biofibre/glass reinforced polyester hybrid composites, Compos. Sci. Technol., 2003, 63(10), 1377–1385 CrossRef CAS.
  85. S. N. Monteiro, F. P. D. Lopes and D. C. O. Nascimento, et al., Processing and properties of continuous and aligned curaua fibres incorporated polyester composites, J. Mater. Res., 2013, 2(1), 2–9 CAS.
  86. M. Q. Zhang, M. Z. Rong and X. Lu, Fully biodegradable natural fibre composites from renewable resources: all-plant fibre composites, Compos. Sci. Technol., 2005, 65, 2514–2525 CrossRef CAS.
  87. V. K. Mathur, Composite materials from local resources, Constr. Build. Mater., 2006, 20, 470–477 CrossRef.
  88. X. Xu, K. Jayaraman and C. Morin, et al., Life cycle assessment of wood-fibre-reinforced polypropylene composites, J. Mater. Process. Technol., 2008, 198, 168–177 CrossRef CAS.
  89. K. Singh Rajesh, H. R. Murty and S. K. Gupta, et al., An overview of sustainability assessment methodologies, Ecol. Indic., 2009, 9, 189–212 CrossRef.
  90. J. Summerscales, N. Dissanayake and A. Virk, et al., A review of bast fibres and their composites. Part 2 – composites, Composites, Part A, 2010, 41(10), 1336–1344 CrossRef.
  91. H. Gangarao and R. Liang, Advanced fibre reinforced polymer composites for sustainable civil infrastructures, International Symposium on Innovation & Sustainability of Structures in Civil Engineering, Xiamen University, China, 2011 Search PubMed.
  92. R. A. Kurien, M. M. Anil, S. L. S. Mohan and J. A. Thomas, Natural fiber composites as sustainable resources for emerging applications- a review, Mater. Today: Proc., 2023 DOI:10.1016/j.matpr.2023.04.363.
  93. M. Bedrich, S. Janouskova and H. Tomas, How to understand and measure environmental sustainability: indicators and targets, Ecol. Indic., 2012, 17, 4–13 CrossRef.
  94. M. James, E. Richard and R. C. Stuart, et al., Natural fibre composite energy absorption structures, Compos. Sci. Technol., 2012, 72, 211–217 CrossRef.
  95. K. Begum and M. A. Islam, Natural fibre as a substitute to synthetic fibre in polymer composites: a review, Res. J. Eng. Sci., 2013, 2(3), 46–53 Search PubMed.
  96. T. Lopez-Lara, J. B. Hernandez-Zaragoza and J. Horta, et al., Sustainable use of tepetate composite in earthen structure, Adv. Mater. Sci. Eng., 2013, 6, 806387 CAS.
  97. G. Cicala, G. Cristaldi and G. Recca, et al., Composites based on natural fibre fabrics, Woven Fabr. Eng., 2015, 317–342 Search PubMed.
  98. J. Rout, M. Misra and S. S. Tripathy, et al., The influence of fibre treatment of the performance of coir-polyester composites, Compos. Sci. Technol., 2001, 61(9), 1303–1310 CrossRef CAS.
  99. A. K. Rana, A. Mandal and S. Bandyopadhyay, Short jute fibre reinforced polypropylene composites: effect of compatibiliser, impact modifier and fibre loading, Compos. Sci. Technol., 2003, 63, 801–806 CrossRef CAS.
  100. Y. Li, Y. W. Mai and L. Ye, Sisal fibre and its composites: a review of recent developments, Compos. Sci. Technol., 2000, 60, 2037–2055 CrossRef CAS.
  101. S. Wong, R. Shanks and A. Hodzic, Interfacial improvements in poly(3-hydroxybutyrate)-flax fibre composites with hydrogen bonding additives, Compos. Sci. Technol., 2004, 64, 1321–1330 CrossRef CAS.
  102. I. V. Weyenberg, J. Ivens and A. D. Coster, et al., Influence of processing and chemical treatment of flax fibres on their composites, Compos. Sci. Technol., 2003, 63(9), 1241–1246 CrossRef.
  103. J. Gassan, A study of fibre and interface parameters affecting the fatigue behaviour of natural fibre composites, Composites, Part A, 2002, 33, 369–374 CrossRef.
  104. M. Baiardo, E. Zini and M. Scandola, Flax fibre–polyester composites, Composites, Part A, 2004, 35, 703–710 CrossRef.
  105. T. T. L. Doan, S. L. Gao and E. Mader, Jute/polypropylene composites I. Effect of matrix modification, Compos. Sci. Technol., 2006, 66, 952–963 CrossRef CAS.
  106. A. C. H. Barreto, M. A. Esmeraldo and D. S. Rosa, et al., Cardanol biocomposites reinforced with jute fibre: microstructure, biodegradability, and mechanical properties, Polym. Compos., 2010, 31(11), 1928–1937 CrossRef CAS.
  107. I. M. Low, M. McGrath and D. Lawrence, et al., Mechanical and fracture properties of cellulose-fibre-reinforced epoxy laminates, Composites, Part A, 2007, 38(3), 963–974 CrossRef.
  108. B. N. Dash, A. K. Rana and H. K. Mishra, et al., Novel low-cost jute-polyester composites. Part 1: Processing, mechanical properties and SEM analysis, Polym. Compos., 1999, 20(1), 62–71 CrossRef CAS.
  109. M. Jawaid, H. P. S. A. Khalil and A. A. Bakar, Woven hybrid composites: Tensile and flexural properties of oil palm-woven jute fibres based epoxy composites, Mater. Sci. Eng., A, 2011, 528(15), 5190–5195 CrossRef CAS.
  110. N. Abilash and M. Sivapragash, Environmental benefits of eco-friendly natural fiber reinforced polymeric composite materials, Int. J. Appl. Innov. Eng. Manag., 2013, 2, 53–59 Search PubMed.
  111. D. Cho, H. S. Lee and S. O. Han, Effect of fiber surface modification on the interfacial and mechanical properties of kenaf fiber-reinforced thermoplastic and thermosetting polymer composites, Compos. Interfaces, 2009, 16, 711–729 CrossRef CAS.
  112. F. A. Silva, R. D. T. Filho and J. A. M. Filho, et al., Physical and mechanical properties of durable sisal fibre–cement composites, Constr. Build. Mater., 2010, 24, 777–785 CrossRef.
  113. R. Gujjala, S. Ojha and S. K. Acharya, et al., Mechanical properties of woven jute–glass hybrid-reinforced epoxy composite, J. Compos. Mater., 2013, 48, 3445–3455 CrossRef.
  114. P. S. Latha, M. V. Rao and V. V. K. Kumar, et al., Evaluation of mechanical and tribological properties of bamboo–glass hybrid fiber reinforced polymer composite, J. Ind. Text., 2015, 46, 3–18 CrossRef.
  115. Z. Sun, Progress in the research and applications of and applications of natural fiber-reinforced polymer matrix composites, Sci. Eng. Compos. Mater., 2018, 25(5), 835–846 CrossRef CAS.
  116. S. Nunna, P. R. Chandra and S. Shrivastava, et al., A review on mechanical behaviour of natural fiber based hybrid composites, J. Reinf. Plast. Compos., 2012, 29(11), 759–769 CrossRef.
  117. H. N. Dhakal, Z. Y. Zhang and R. Guthrie, et al., Development of flax/carbon fibre hybrid composites for enhanced properties, Carbohydr. Polym., 2013, 96(1), 1–8 CrossRef CAS.
  118. Y. Zhang, Y. Li and H. Ma, et al., Tensile and interfacial properties of unidirectional flax/glass fiber reinforced hybrid composites, Compos. Sci. Technol., 2013, 88, 172–177 CrossRef CAS.
  119. M. Le Guen, R. H. Newman and A. Fernyhough, et al., Tailoring the vibration damping behaviour of flax fibre-reinforced epoxy composite laminates via polyol additions, Composites, Part A, 2014, 67, 37–43 CrossRef CAS.
  120. M. Ramesh, C. Deepa and L. R. Kumar, et al., Life-cycle and environmental impact assessments on processing of plant fibres and its bio-composites: A critical review, J. Ind. Text., 2022, 51(4), 5518S–5542S CrossRef.
  121. J. Mussig and N. Graupner, Test methods for fibre/matrix adhesion in cellulose fibre reinforced thermoplastic composite materials: A critical review, Rev. Adhes. Adhes., 2020, 8(2), 68–129 CrossRef.
  122. N. Karabulut, M. Aktas and H. E. Balcioglu, Surface modification effects on the mechanical properties of woven jute fabric reinforced laminated composites, J. Nat. Fibers, 2019, 16, 629–643 CrossRef CAS.
  123. G. Ekundayao and S. Adejuyigbe, Reviewing the development of natural fiber polymer composite: A case study of sisal and jute, Am. J. Mech. Mater. Eng., 2019, 3(1), 1–10,  DOI:10.11648/j.ajmme.20190301.11.
  124. W. Liu, L. T. Drzal and A. K. Mohanty, et al., Influence of processing methods and fiber length on physical properties of kenaf fiber reinforced soy based biocomposites, Composites, Part B, 2007, 38, 352–359 CrossRef.
  125. M. Zampaloni, F. Pourboghrat and S. A. Yankovich, et al., Kenaf natural fiber reinforced polypropylene composites: A discussion on manufacturing problems and solutions, Composites, Part A, 2007, 38, 1569–1580 CrossRef.
  126. M. T. Paridah, A. B. Basher and S. Saiful Azry, et al., Retting process of some bast plant fibres and its effect on fibre quality: A review, BioResources, 2011, 6(4), 5260–5281 Search PubMed.
  127. M. Ramesh and S. Nijanthan, Mechanical property analysis of kenaf–glass fibre reinforced polymer composites using finite element analysis, Bull. Mater. Sci., 2016, 39(1), 147–157 CrossRef CAS.
  128. R. Liu, Y. Peng and J. Cao, et al., Water absorption, dimensional stability, and mold susceptibility of organically-modified-montmorillonite modified wood flour/polypropylene composites, BioResources, 2014, 9(1), 54–65 CrossRef CAS PubMed.
  129. A. Memona and A. Nakai, Mechanical properties of jute spun yarn/PLA tubular braided composite by pultrusion molding, Energy Procedia, 2013, 34, 818–829 CrossRef.
  130. N. M. Salehudiin, N. Salim and R. Roslan, et al., Improving the properties of kenaf reinforced polypropylene composite by alkaline treatment, Mater. Today: Proc., 2023, 75, 156–162 Search PubMed.
  131. L. Zamora-Mendoza, F. Gushque, S. Yanez, N. Jara, J. F. Álvarez-Barreto, C. Zamora-Ledezma, S. A. Dahoumane and F. Alexis, Plant fibers as composite reinforcements for biomedical applications, Bioengineering, 2023, 10(7), 804 CrossRef CAS.
  132. M. A. Fuqua, S. Huo and C. A. Ulven, Natural fibre reinforced composites, Polym. Rev., 2012, 52, 259–320 CrossRef CAS.
  133. L. Yan, B. Kasal and L. Huang, A review of recent research on the use of cellulosic fibres, their fibre fabric reinforced cementitious, geo-polymer and polymer composites in civil engineering, Composites, Part B, 2016, 92, 94–132 CrossRef CAS.
  134. M. C. Paiva, I. Ammar and A. R. Campos, et al., Alfa fibres: Mechanical, morphological and interfacial characterization, Compos. Sci. Technol., 2007, 67, 1132–1138 CrossRef CAS.
  135. S. M. Hejazi, M. Sheikhzadeh and S. M. Abtahi, et al., A simple review of soil reinforcement by using natural and synthetic fibres, Constr. Build. Mater., 2012, 30, 100–116 CrossRef.
  136. L. Mwaikambo, Review of the history, properties and application of plant fibres, Afr. J. Sci. Technol., 2006, 7(2), 120–133 Search PubMed.
  137. U. S. Bongarde and V. D. Shinde, Review on natural fibre reinforcement polymer composite, Int. J. Eng. Sci. Innov. Technol., 2014, 3(2), 431–436 Search PubMed.
  138. L. Yan, N. Chouw and K. Jayaraman, Flax fibre and its composites – a review, Composites, Part B, 2014, 56, 296–317 CrossRef CAS.
  139. T. Sen and H. N. Reddy, Various industrial applications of hemp, kenaf, flax and ramie natural fibres, Int. J. Innov. Manag. Technol., 2011, 2, 192–198 Search PubMed.
  140. R. Malkapuram, V. Kumar and Y. S. Negi, Recent developments in natural fiber reinforced polypropylene composites, J. Reinf. Plast. Compos., 2009, 28(10), 1169–1189 CrossRef CAS.
  141. P. K. Sahoo, G. C. Sahu, P. K. Rana and A. K. Das, Preparation, characterization, and biodegradability of jute-based natural fiber composite superabsorbents, Adv. Polym. Technol., 2005, 24, 208–214 CrossRef CAS.
  142. N. Karthi, K. Kumaresan, G. Rajeshkumar, S. Gokulkumar and S. Sathish, Tribological and thermo-mechanical performance of chemically modified musa acuminata/corchorus capsularis reinforced hybrid composites, J. Nat. Fibers, 2022, 19, 4640–4653 CrossRef CAS.
  143. C. R. Vogl and A. Hartl, Production and processing of organically grown fibre nettle (Urtica dioica L.) and its potential use in the natural textile industry: A review, Am. J. Alternative Agric., 2003, 18, 119–128 CrossRef.
  144. Z. Samouh, O. Cherkaoui, D. Soulat, A. R. Labanieh, F. Boussu and R. E. moznine, Identification of the physical and mechanical properties of moroccan sisal yarns used as reinforcements for composite materials, Fibers, 2021, 9(2), 13 CrossRef CAS.
  145. S. D. Asagekar and V. K. Joshi, Characteristics of sugarcane fibres, Indian J. Fibre Text. Res., 2014, 39, 180–184 CAS.
  146. M. L. Sanyang, S. M. Sapuan and M. Jawaid, et al., Recent developments in sugar palm (Arenga Pinnata) based biocomposites and their potential industrial applications: A review, Renewable Sustainable Energy Rev., 2016, 54, 533–549 CrossRef CAS.
  147. N. Sultana, et al., Short jute fiber preform reinforced polypropylene thermoplastic composite: Experimental investigation and its theoretical stiffness prediction, ACS Omega, 2023, 8(27), 24311–24322 CrossRef CAS.
  148. A. C. Karmaker and G. Hinrichsen, Processing and characterization of jute fibre reinforced thermoplastic polymers, Polym.-Plast. Technol. Eng., 1991, 30(5–6), 609–629 CrossRef CAS.
  149. A. Belaadi, H. Laouici and M. Bourchak, Mechanical and drilling performance of short jute fibre-reinforced polymer biocomposites: statistical approach, Int. J. Adv. Manuf. Technol., 2020, 106, 1989–2006 CrossRef.
  150. B. Pradeepa, M. Malathi and A. V. Kiruthika, Properties of jute fibre reinforced polymer composites – a review, Eur. J., Eng. Sci., Tech., 2023, 6(1), 10–29 CrossRef.
  151. Z. Zhang, S. Cai and Y. Li, et al., High performances of plant fiber reinforced composites-A new insight from hierarchical microstructures, Compos. Sci. Technol., 2020, 194, 108151 CrossRef CAS.
  152. D. Gupta, P. K. Sahu and R. Banerjee, Forecasting jute production in major contributing countries in the world, J. Nat. Fibers, 2009, 6, 127–137 CrossRef.
  153. M. Ramesh, K. Palanikumar and K. H. Reddy, Mechanical property evaluation of sisal–jute–glass fibre reinforced polyester composites, Composites, Part B, 2013, 48, 1–9 CrossRef CAS.
  154. J. Gassan and A. K. Bledzki, Possibilities for improving the mechanical properties of jute/epoxy composites by alkali treatment of fibres, Compos. Sci. Technol., 1999, 59, 1303–1309 CrossRef CAS.
  155. J. Gassan and A. K. Bledzki, Alkali treatment of jute fibres: Relationship between structure and mechanical properties, J. Appl. Polym. Sci., 1999, 71, 623–629 CrossRef CAS.
  156. T. M. Gowda, A. C. B. Naidu and R. Chhaya, Some mechanical properties of untreated jute fabric-reinforced polyester composites, Composites, Part A, 1999, 30, 277–284 CrossRef.
  157. X. Y. Liu and G. C. Dai, Surface modification and micromechanical properties of jute fibre mat reinforced polypropylene composites, eXPRESS Polym. Lett., 2007, 1(5), 299–307 CrossRef CAS.
  158. D. Ray and B. K. Sarkar, Characterization of alkali-treated jute fibres forphysical and mechanical properties, J. Appl. Polym. Sci., 2001, 80, 1013–1020 CrossRef CAS.
  159. A. Stocchi, B. Lauke and A. Vazquez, et al., A novel fibre treatment applied to woven jute fabric/vinyl ester laminates, Composites, Part A, 2007, 38, 1337–1343 CrossRef.
  160. W.-m Wang, Z.-s Cai and J.-y Yu, Study on the chemical modification process of jute fibre, J. Eng. Fibers Fabr., 2008, 3(2), 1–11 CAS.
  161. P. Wongsriraksa, K. Togashi and A. Nakai, et al., Continuous natural fibre reinforced thermoplastic composites by fibre surface modification, Adv. Mech. Eng., 2013, 685104 CrossRef.
  162. X. Chen, Q. Gu and Y. Mi, Bamboo fiber-reinforced polypropylene composites: a study of the mechanical properties, J. Appl. Polym. Sci., 1998, 69, 1891–1899 CrossRef CAS.
  163. A. P. Kumar, R. P. Singh and B. D. Sarwade, Degradability of composites, prepared from ethylene–propylene copolymer and jute fibre under accelerated aging and biotic environments, Mater. Chem. Phys., 2005, 92, 458–469 CrossRef CAS.
  164. M. A. Khan, M. M. Hassan and L. T. Drzal, Effect of 2-hydroxyethyl methacrylate (HEMA) on the mechanical and thermal properties of jute–polycarbonate composite, Composites, Part A, 2005, 36, 71–81 CrossRef.
  165. A. Campos, J. M. Marconcini and S. M. Martins-Franchetti, et al., The influence of UV-C irradiation on the properties of thermoplastic starch and polycaprolactone biocomposite with sisal bleached fibres, Polym. Degrad. Stab., 2012, 97(10), 1948–1955 CrossRef CAS.
  166. G. Cicala, G. Cristaldi, G. Recca, et al., Composites based on natural fibre fabrics, in Woven Fabric Engineering, ed. P. Dobnik Dubrovski, InTech, 2010,  DOI:10.5772/10465, ISBN978-953-307-194-7.
  167. M. T. Zafar, S. N. Maiti and A. K. Ghosh, Effect of surface treatment of jute fibers on the interfacial adhesion in poly(lactic acid)/jute fiber biocomposites, Fibers Polym., 2016, 17, 266–274 CrossRef CAS.
  168. S. Kaewkuk, W. Sutapun and K. Jarukumjorn, Effect of heat treated sisalfibre on physical properties of polypropylene composites, Adv. Mater. Res., 2010, 123–125, 1123–1126 CAS.
  169. A. C. Milanese, M. O. H. Cioffi and H. J. C. Voorwald, Mechanical behavior of natural fibre composites, Procedia Eng., 2011, 10, 2022–2027 CrossRef CAS.
  170. R. Samson, S. Mani and R. Boddey, et al., The potential of C4 perennial grasses for developing a global BIO-HEAT industry, Crit. Rev. Plant Sci., 2005, 24, 461–495 CrossRef.
  171. M. Jacob and S. Thomas, Studies on Lignocellulosic Fibre Reinforced Natural Rubber Composites, 2010, http://shodhganga.inflibnet.ac.in/bitstream/10603/509/12/12part3.pdf Search PubMed.
  172. P. A. Sreekumar, S. P. Thomas and J. M. Saiter, et al., Effect of fibre surface modification on the mechanical and water absorption characteristics of sisal/polyester composites fabricated by resin transfer moulding, Composites, Part A, 2009, 40, 1777–1784 CrossRef.
  173. C. Hong, I. Hwang and N. Kim, et al., Mechanical properties of silanized jute polypropylene composites, J. Ind. Eng. Chem., 2008, 14(1), 71–76 CrossRef CAS.
  174. S. Pujari, A. Ramakrishna and K. T. B. Padal, Comparison of ANN and regression analysis for predicting the water absorption behaviour of jute and banana fiber reinforced epoxy composites, Mater. Today: Proc., 2017, 4, 1626–1633 Search PubMed.
  175. B. Ozdemir, A. Yardimeden, E. Bahce, E. Kilickap and E. Emir, Analysis of drilling behaviour in jute fibres reinforced natural composites, J. Nat. Fibers, 2023, 20, 2159608 CrossRef.
  176. H. Song, J. Liu, K. He and W. Ahmad, A comprehensive overview of jute fiber reinforced cementitious composites, Case Stud. Constr. Mater., 2021, 15, e00724 Search PubMed.
  177. V. K. Mahakur, R. Paul, S. Kumar, P. K. Patowari, and S. Bhowmik, A study on mechanical properties and tribological behaviour of jute filler composites, in Techno-Societal, ed. P. M. Pawar, et al., Springer, Cham, 2022,  DOI:10.1007/978-3-031-34644-6_75, ICATSA 2022. 2024.
  178. P. Ravikumar, G. Rajeshkumar, P. Manimegalai, K. R. Sumesh, M. R. Sanjay and S. Siengchin, Delamination and surface roughness analysis of jute/polyester composites using response surface methodology: Consequence of sodium bicarbonate treatment, J. Ind. Text., 2022, 51, 360S–377S CrossRef CAS.
  179. A. K. Chattapadhyaya and P. S. Das, Estimation of growth rate: A critical analysis with reference to West Bengal agriculture, Indian J. Agric. Econ., 2000, 55, 116–135 Search PubMed.
  180. U. K. De, Nature and causes of inter district variations in yield of rice in West Bengal, 1970–71 to 1994–95, Indian J. Agric. Econ., 1999, 59(4), 554–565 Search PubMed.
  181. C. Sarkar, and P. K. Sahu, Statistical account of the growth in jute and aus paddy-Two competing crops in West Bengal agriculture, Proceedings of the 90th Indian Science Congress, Lucknow, India, January 3–7, 2001 Search PubMed.
  182. P. Niedermann, G. Szebenyi and A. Toldy, Characterization of high glass transition temperature sugar-based epoxy resin composites with jute and carbon fibre reinforce, Compos. Sci. Technol., 2015, 117, 62–68 CrossRef CAS.
  183. H. Singh, J. P. Singh and S. Singh, et al., A brief review of jute fibre and its composites, Mater. Today: Proc., 2018, 5, 28427–28437 CAS.
  184. A. Benkhelladi, H. Laouici and A. Bouchoucha, Tensile and flexural properties of polymer composites reinforced by flax, jute and sisal fibres, Int. J. Adv. Manuf. Technol., 2020, 108, 895–916 CrossRef.
  185. A. Bourmaud, J. Beaugrand and D. U. Shah, et al., Towards the design of high-performance plant fibre composites, Prog. Mater. Sci., 2018, 97, 347–408 CrossRef.
  186. J. Summerscales, N. P. J. Dissanayake and A. S. Virk, et al., A review of bast fibres and their composites. Part 1 – fibres as reinforcements, Composites, Part A, 2010, 41(10), 1329–1335 CrossRef.
  187. D. Shah, Developing plant fibre composites for structural applications by optimising composite parameters: a critical review, J. Mater. Sci., 2013, 48(18), 6083–6107 CrossRef CAS.
  188. M. Misnon, M. M. Islam and J. A. Epaarachchi, et al., Potentiality of utilising natural textile materials for engineering composites applications, Mater. Des., 2014, 59, 359–368 CrossRef CAS.
  189. D. Verma, K. L. Goh and V. Vimal, Interfacial studies of natural fiber-reinforced particulate thermoplastic composites and their mechanical properties, J. Nat. Fibers, 2022, 19(6), 2299–2326 CrossRef CAS.
  190. B. Madsen, A. Thygesan and H. Lilholt, Plant fibre composites–porosity and stiffness, Compos. Sci. Technol., 2009, 69, 1057–1069 CrossRef CAS.
  191. M. Syduzzaman, M. A. A. Faruque and K. Bilisik, et al., Plant-based natural fibre reinforced composites: A review on fabrication, properties and applications, Coatings, 2020, 10(10), 973 CrossRef CAS.
  192. M. R. Sanjay, S. Siengchin and J. Parameswaranpillai, et al., A comprehensive review of techniques for natural fibers as reinforcement in composites: Preparation, processing and characterization, Carbohydr. Polym., 2019, 207, 108–121 CrossRef.
  193. W. Woigk, C. A. Fuentes and J. Rion, et al., Interface properties and their effect on the mechanical performance of flax fibre thermoplastic composites, Composites, Part A, 2019, 122, 8–17 CrossRef CAS.
  194. C. Asasutjarit, J. Hirunlabh and J. Khedari, et al., Development of coconut coir-based lightweight cement board, Constr. Build. Mater., 2007, 21, 277–288 CrossRef.
  195. C. Onesippe, N. Passe-Coutrin and F. Toro, et al., Sugar cane bagasse fibres reinforced cement composites: thermal considerations, Composites, Part A, 2010, 41, 549–556 CrossRef.
  196. Z. Li, X. Wang and L. Wang, Properties of hemp fibre reinforced concrete composites, Composites, Part A, 2006, 37, 497–505 CrossRef.
  197. R. L. B. Cardoso, J. da Silva Rodrigues, R. P. B. Ramos, A. de Castro Correa, E. M. Leão Filha, S. N. Monteiro, A. C. R. da Silva, R. T. Fujiyama and V. S. Candido, Use of yarn and carded jute as epoxy matrix reinforcement for the production of composite materials for application in the wind sector: a preliminary analysis for the manufacture of blades for low-intensity winds, Polymers, 2023, 15(18), 3682 CrossRef CAS.
  198. M. A. Ismail, Compressive and tensile strength of natural fibre-reinforced cement base composites, Al-Rafidain Eng. J., 2007, 15, 42–51 Search PubMed.
  199. B.-H. Lee, H.-J. Kim and W.-R. Yu, Fabrication of long and discontinuous natural fibre reinforced polypropylene biocomposites and their mechanical properties, Fibres Polym., 2009, 10(1), 83–90 CrossRef CAS.
  200. N. Lu and S. Oza, Thermal stability and thermo-mechanical properties of hemp high density polyethylene composites: effect of two different chemical modifications, Composites, Part B, 2013, 44(1), 484–490 CrossRef CAS.
  201. H. Alamri, I. M. Low and Z. Alothman, Mechanical, thermal and microstructural characteristics of cellulose fibre reinforced epoxy/organoclay nanocomposites, Composites, Part B, 2012, 43(7), 2762–2771 CrossRef CAS.
  202. F. Vilaseca, J. A. Mendez and A. Pelach, et al., Composite materials derived from biodegradable starch polymer and jute strands, Process Biochem., 2007, 42, 329–334 CrossRef CAS.
  203. S. Ochi, Development of high strength biodegradable composites using manila hemp fibre and starch based biodegradable resin, Composites, Part A, 2006, 37, 1879–1883 CrossRef.
  204. E. Bodros, I. Pillin and N. Montrelay, et al., Could biopolymers reinforced by randomly scattered flax fibre be used in structural applications, Compos. Sci. Technol., 2007, 67, 462–470 CrossRef CAS.
  205. T. Nishino, K. Hirao and M. Kotera, et al., Kenaf reinforced biodegradable composite, Compos. Sci. Technol., 2003, 63, 1281–1286 CrossRef CAS.
  206. A. K. Bledzki, A. Jaszkiewicz and D. Scherzer, Mechanical properties of PLA composites with man-made cellulose and abaca fibres, Composites, Part A, 2009, 40, 404–412 CrossRef.
  207. A. Gomes, T. Matsuo and K. Goda, et al., Development and effect of alkali treatment on tensile properties of caraua fibre green composites, Composites, Part A, 2007, 38, 1811–1820 CrossRef.
  208. R. Iovino, R. Zullo and M. A. Rao, et al., Biodegradation of polylactic acid/starch coir bio composites under controlled composting conditions, Polym. Degrad. Stab., 2008, 93, 147–157 CrossRef CAS.
  209. B. Bax and J. Mussig, Impact and tensile properties of PLA/Cordenka and PLA/Flax composites, Compos. Sci. Technol., 2008, 68, 1601–1607 CrossRef CAS.
  210. A. Ramzy, D. Beermann and L. Steuernagel, et al., Developing a new generation of sisal composite fibres for use in industrial applications, Composites, Part B, 2014, 66, 287–298 CrossRef CAS.
  211. B. T. Weclawski, M. Fan and D. Hui, Compressive behaviour of natural fibre composite, Composites, Part B, 2014, 67, 183–191 CrossRef CAS.
  212. A. Abdal-hay, N. P. G. Suardana and D. Y. Jung, et al., Effect of diameters and alkali treatment on the tensile properties of date palm fiber reinforced epoxy composites, Int. J. Precis. Eng. Manuf., 2012, 13(7), 1199–1206 CrossRef.
  213. V. Manikandan, P. Authakkannan and V. A. Prabu, Review on natural fiber composites: Banana/hemp fiber and its hybrid composites, Int. J. Compos. Mater., 2015, 1, 1–14 Search PubMed.
  214. P. V. Senthil and A. Sirsshti, Studies on material and mechanical properties of natural fiber reinforced composites, Int. J. Eng. Sci., 2014, 3(11), 18–27 Search PubMed.
  215. C. Ververis, K. Georghiou and D. Danielidis, et al., Cellulose, hemicelluloses, lignin and ash content of some organic materials and their suitability for use as paper pulp supplements, Bioresour. Technol., 2007, 98, 296–301 CrossRef CAS PubMed.
  216. P. K. Bellairu, S. Bhat and E. V. Gijo, Modelling and optimization of natural fibre reinforced polymer nanocomposite: application of mixture design technique, Multidiscip. Model. Mater. Struct., 2021, 17, 507–521 CrossRef CAS.
  217. X. Huang and A. Netravali, Biodegradable green composites made using bamboo micro/nano-fibrils and chemically modified soy protein resin, Compos. Sci. Technol., 2009, 69, 1009–1015 CrossRef CAS.
  218. N. A. El-Wakil, R. E. Abou-Zeid and Y. Fahmy, et al., Modified wheat gluten as a binder in particle board made from reed, J. Appl. Polym. Sci., 2007, 106, 3592–3599 CrossRef CAS.
  219. S. Chabba and A. N. Netravali, Green composites Part 1: characterization of flax fabric and glutaraldehyde modified soy protein concentrate composites, J. Mater. Sci., 2005, 40, 6263–6273 CrossRef CAS.
  220. Z. Liu, S. Z. Erhan and D. E. Akin, et al., Green composites from renewable sources: preparation of epoxidized soyabean oil and flax fibre composites, J. Agric. Food Chem., 2006, 54, 2134–2137 CrossRef CAS PubMed.
  221. Y. Zou, S. Huda and Y. Yang, Lightweight composites from long wheat straw and polypropylene web, Bioresour. Technol., 2010, 101, 2026–2033 CrossRef CAS PubMed.
  222. S. Huda and Y. Yang, Composites from ground chicken quill and polypropylene, Compos. Sci. Technol., 2008, 68, 790–798 CrossRef CAS.
  223. S. Huda and Y. Yang, Chemically extracted fibres as reinforcement in lightweight poly(propylene) composites, Macromol. Mater. Eng., 2008, 293, 235–243 CrossRef CAS.
  224. S. Huda and Y. Yang, Feather fibre reinforced light-weight composites with good acoustic properties, J. Polym. Environ., 2009, 17, 131–142 CrossRef CAS.
  225. S. Huda and Y. Yang, A novel approach of manufacturing light-weight composites with polypropylene web and mechanically split cornhusk, Ind. Crops Prod., 2009, 30(1), 17–23 CrossRef CAS.
  226. A. L. Duigou, P. Davies and C. Baley, Interfacial bonding of flax fibre/poly(L-lactide) bio-composites, Compos. Sci. Technol., 2010, 70(2), 231–239 CrossRef.
  227. N. Reddy and Y. Yang, Biocomposites developed using water-plasticized wheat gluten as matrix and jute fibres as reinforcement, Polym. Int., 2011, 60, 711–716 CrossRef CAS.
  228. M. W. Lee, S. O. Han and Y. B. Seo, Red algae fibre/poly (butylene succinate) biocomposites: the effect of fibre content on their mechanical and thermal properties, Compos. Sci. Technol., 2008, 68(6), 1266–1272 CrossRef CAS.
  229. Y. Z. Wan, H. Luo and F. He, et al., Mechanical, moisture absorption, and biodegradation behaviours of bacterial cellulose fibre-reinforced starch biocomposites, Compos. Sci. Technol., 2009, 69(7–8), 1212–1217 CrossRef CAS.
  230. A. C. Albuquerque, K. Joseph and L. H. Carvalho, et al., Effect of wettability and ageing conditions on the physical and mechanical properties of uniaxially oriented jute-roving-reinforced polyester composites, Compos. Sci. Technol., 2000, 60, 833–844 CrossRef.
  231. D. Ray, B. K. Sarkar and N. R. Bose, Impact fatigue behaviour of vinylester resin matrix composites reinforced with alkali treated jute fibres, Composites, Part A, 2002, 33, 233–241 CrossRef.
  232. A. Bourmaud and C. Baley, Rigidity analysis of polypropylene/natural fibre composites after recycling, Polym. Degrad. Stab., 2009, 94, 297–305 CrossRef CAS.
  233. A. Le Duigou, P. Davies and C. Baley, Seawater ageing of flax/poly (lactic acid) biocomposites, Polym. Degrad. Stab., 2009, 94, 1151–1162 CrossRef CAS.
  234. M. S. Islam, K. L. Pickering and N. J. Foreman, Influence of hygrothermal ageing on the physico-mechanical properties of alkali treated industrial hemp fibre reinforced polylactic acid composites, J. Polym. Environ., 2010, 18, 696–704 CrossRef CAS.
  235. M. Assarar, D. Scida and A. El Mahi, et al., Influence of water ageing on mechanical properties and damage events of two reinforced composite materials: flax-fibres and glass-fibres, Mater. Des., 2011, 32, 788–795 CrossRef CAS.
  236. D. Scida, M. Assarar and C. Poilâne, et al., Influence of hygrothermal ageing on the damage mechanisms of flax-fibre reinforced epoxy composite, Composites, Part B, 2013, 48, 51–58 CrossRef CAS.
  237. I. M. D. Rosa, C. Santulli and F. Sarasini, Acoustic emission for monitoring the mechanical behavior of natural fibre composites: a literature review, Composites, Part A, 2009, 40, 1456–1469 CrossRef.
  238. M. Assarar, D. Scida and W. Zouari, et al., Acoustic emission characterization of damage in short hemp-fiber reinforced polypropylene composites, Polym. Compos., 2016, 37, 1101–1112 CrossRef CAS.
  239. S. Phillips, J. Baets and L. Lessard, et al., Characterization of flax/epoxy prepregs before and after cure, J. Reinf. Plast. Compos., 2013, 32(11), 777–785 CrossRef.
  240. M. Kersani, S. V. Lomov and A. W. V. Vuure, et al., Damage in flax/epoxy quasi-unidirectional woven laminates under quasi-static tension, J. Compos. Mater., 2015, 49(4), 403–413 CrossRef CAS.
  241. A. J. P. Silva, F. A. R. Lahr and A. L. Christoforo, et al., Properties of sugar cane bagasse to use in OSB, Int. J. Mater. Eng., 2012, 2, 50–56 CrossRef.
  242. L. J. D. Silva, T. H. Panzera and V. R. Velloso, et al., Hybrid polymeric composites reinforced with sisal fibres and silica microparticles, Composites, Part B, 2012, 43, 3436–3444 CrossRef.
  243. C. Santulli and A. P. Caruso, Effect of fibre architecture on the falling weight impact properties of hemp/epoxy composites, J. Biobased Mater. Bioenergy, 2009, 3, 291–297 CrossRef CAS.
  244. D. S. D. Vasconcellos, F. Sarasini and F. Touchard, et al., Influence of low velocity impact on fatigue behaviour of woven hemp fibre reinforced epoxy composites, Composites, Part B, 2014, 66, 46–57 CrossRef.
  245. A. V. R. Prasad and K. M. Rao, Mechanical properties of natural fibre reinforced polyester composites: Jowar, sisal and bamboo, Mater. Des., 2011, 32, 4658–4663 CrossRef CAS.
  246. M. Li, Y. Pu and V. M. Thomas, et al., Recent advancements of plant-based natural fiber–reinforced composites and their applications, Composites, Part B, 2020, 200, 108254 CrossRef CAS.
  247. T. Gurunathan, S. Mohanty and S. K. Nayak, A review of the recent developments in biocomposites based on natural fibres and their application perspectives, Composites, Part A, 2015, 77, 1–25 CrossRef CAS.
  248. M. Bouakba, A. Bezazi and K. Boba, et al., Cactus fibre/polyester biocomposites: manufacturing, quasi-static mechanical and fatigue characterisation, Compos. Sci. Technol., 2013, 74, 150–159 CrossRef CAS.
  249. A. Belaadi, A. Bezazi and M. Bourchak, et al., Tensile static and fatigue behaviour of sisal fibres, Mater. Des., 2013, 46, 76–83 CrossRef.
  250. F. D. A. Silva, N. Chawla and R. D. T. Filho, An experimental investigation of the fatigue behavior of sisal fibres, Mater. Sci. Eng., A, 2009, 516, 90–95 CrossRef.
  251. Y. Dobah, M. Bourchak, A. Bezazi, et al., Static and fatigue strength characterization of sisal fibre reinforced polyester composite material, 9th Int Conf on Compos Sci Technol: 2020 Scientific and Industrial Challenges, 2013 Search PubMed.
  252. A. N. Towo and M. P. Ansell, Fatigue of sisal fibre reinforced composites: constant-life diagrams and hysteresis loop capture, Compos. Sci. Technol., 2008, 68, 915–924 CrossRef CAS.
  253. S. Sangthong, T. Pongprayoon and N. Yanumet, Mechanical property improvement of unsaturated polyester composite reinforced with admicellar-treated sisal fibres, Composites, Part A, 2009, 40, 687–694 CrossRef.
  254. J. K. D. Santos, R. A. D. Cunha and R. C. T. S. Felipe, et al., Composite friend sisal/polyester treated in surface, HOLOS, 2011, 3, 102–111 CrossRef.
  255. H. J. Kim and D. W. Seo, Effect of water absorption fatigue on mechanical properties of sisal textile-reinforced composites, Int. J. Fatigue, 2006, 28, 1307–1314 CrossRef CAS.
  256. G. Basu, A. N. Roy and K. K. Satapathy, et al., Potentiality for value-added technical use of Indian sisal, Ind. Crops Prod., 2012, 36, 33–40 CrossRef.
  257. Y. J. Xie, C. A. S. Hill and Z. F. Xiao, et al., Silane coupling agents used for natural fibre/polymer composites: a review, Composites, Part A, 2010, 41, 806–819 CrossRef.
  258. D. B. Dittenber and H. V. S. Gangarao, Critical review of recent publications on use of natural composites in infrastructure, Composites, Part A, 2012, 43(8), 1419–1429 CrossRef.
  259. Q. T. Shubhra, A. K. M. M. Alam and M. A. Quaiyyum, Mechanical properties of polypropylene composites: A review, J. Thermoplast. Compos. Mater., 2013, 26(3), 362–391 CrossRef.
  260. M. Ramesh, Kenaf (Hibiscus cannabinus L.) fiber based bio-materials: A review on processing and properties, Prog. Mater. Sci., 2016, 78–79, 1–92 CrossRef CAS.
  261. A. Shahzad, Hemp fiber and its composites: A review, J. Compos. Mater., 2012, 46(8), 973–986 CrossRef CAS.
  262. A. K. Mohanty and M. Misra, Studies on jute composites: A literature review, Polym.-Plast. Technol. Eng., 1995, 34(5), 729–792 CrossRef CAS.
  263. J. Zamboni Schiavon and J. José de Oliveira Andrade, Comparison between alternative chemical treatments on coir fibers for application in cementitious materials, J. Mater. Res. Technol., 2013, 25, 4634–4649 CrossRef.
  264. K. Z. M. A. Motaleb, A. Ahad, G. Laureckiene and R. Milasius, Innovative Banana fiber nonwoven reinforced polymer composites: pre- and post-treatment effects on physical and mechanical properties, Polymers, 2021, 13(21), 3744 CrossRef CAS PubMed.
  265. M. A. S. Spinace, C. S. Lambert and K. K. G. Fermoselli, et al., Characterization of lignocellulosic curaua fibres, Carbohydr. Polym., 2009, 77(1), 47–53 CrossRef CAS.
  266. R. V. Silva and E. M. F. Aquino, Curaua fiber: A new alternative to polymeric composites, J. Reinf. Plast. Compos., 2008, 27(1), 103–112 CrossRef CAS.
  267. J. S. Caraschi and A. L. Leato, Characterization of curaua fiber, Mol. Cryst. Liq. Cryst., 2000, 353(1), 149–152 CrossRef CAS.
  268. K. Joseph, R. D. T. Filho and B. James, et al., A review on sisal fiber reinforced polymer composites, Rev. Bras. Eng. Agric. Ambient., 1999, 3(3), 367–379 CrossRef.
  269. R. M. S. James, J. Gisip and N. M. Yusof, Effect of chemical treatment on physical and mechanical properties of coir fibre-polypropylene composites, Sci. Res. J., 2023, 20, 145–157 CrossRef.
  270. D. H. Mueller and A. Krobjilowski, New discovery in the properties of composites reinforced with natural fibers, J. Ind. Text., 2003, 33(2), 111–130 CrossRef CAS.
  271. S. Shinoj, R. Visvanathan and S. Panigrahi, et al., Oil palm fiber (OPF) and its composites: A review, Ind. Crops Prod., 2011, 33(1), 7–22 CrossRef CAS.
  272. D. Verma, P. C. Gope and M. K. Maheswari, et al., Bagasse fiber composites: A review, J. Mater. Environ. Sci., 2012, 3(6), 1079–1092 Search PubMed.
  273. Y. R. Loh, D. Sujan and M. E. Rahman, et al., Sugarcane bagasse–The future composite material: A literature review, Resour., Conserv. Recycl., 2013, 75, 14–22 CrossRef.
  274. N. Justiz-Smith, G. Junior Virgo and V. E. Buchanan, Potential of Jamaican banana, coconut coir and bagasse fibres as composite materials, Mater. Charact., 2008, 59(9), 1273–1278 CrossRef CAS.
  275. S. N. Monteiro, R. J. S. Rodriquez and M. V. De Souza, et al., Sugar cane bagasse waste as reinforcement in low cost composites, Adv. Perform. Mater., 1998, 5(3), 183–191 CrossRef CAS.
  276. D. Liu, J. Song and D. P. Andersen, et al., Bamboo fiber and its reinforced composites: structure and properties, Cellulose, 2012, 19(5), 1449–1480 CrossRef CAS.
  277. A. K. Bledzki, S. Reihmane and J. Gassan, Thermoplastics reinforced with wood fillers: A literature review, Polym.-Plast. Technol. Eng., 1998, 37(4), 451–468 CrossRef CAS.
  278. S. H. Mahmud, S. C. Das and M. Z. I. Mollah, et al., Thermoset-polymer matrix composite materials of jute and glass fibre reinforcements: Radiation effects determination, J. Mater. Res. Technol., 2023, 26, 6623–6635 CrossRef CAS.
  279. A. B. M. Abid Hossen Bhuiyan, M. F. Hossain, M. S. Rana and M. S. Ferdous, Impact of fiber orientations, stacking sequences and ageing on mechanical properties of woven jute-kevlar hybrid composites, Results Mater., 2023, 20, 100477 CrossRef.
  280. A. K. Rana and K. Jayachandran, Jute fiber for reinforced composites and its prospects. Molecular Crystals and Liquid Crystals Science and Technology Section A, Mol. Cryst. Liq. Cryst., 2000, 353(1), 35–45 CrossRef CAS.
  281. L. Guo-Qing, Archeological evidence for the use of chu-nam on the 13th century Quanzhou ship, Fujian Province, China, Int. J. Naut. Archaeol., 1989, 18(4), 277–283 CrossRef.
  282. T. H. Tsien, Raw materials for old papermaking in China, J. Am. Orient. Soc., 1973, 93(4), 510–519 CrossRef.
  283. S. Shahinur, M. M. A. Sayeed, M. Hasan, A. S. M. Sayem, J. Haider and S. Ura, Current development and future perspective on natural jute fibers and their biocomposites, Polymers, 2022, 14(7), 1445 CrossRef CAS PubMed.
  284. D. Chakrabarty, Rethinking Working-Class History: Bengal, 1890–1940, Princeton University Press, Princeton, NJ, 2000 Search PubMed.
  285. P. B. Shelar and U. N. Kumar, A short review on jute fibre reinforced composites, Mater. Sci. Forum, 2021, 1019, 32–43 Search PubMed.
  286. A. Razmi and M. M. Mirsayar, On the mixed mode I/II fracture properties of jute fiber reinforced concrete, Constr. Build. Mater., 2017, 148, 512–520 CrossRef.
  287. P. Manimaran, P. Senthamaraikannan and K. Murugananthan, et al., Physicochemical properties of new cellulosic fibers from Azadirachta indica plant, J. Nat. Fibers, 2018, 15(1), 29–38 CrossRef CAS.
  288. P. Manimaran, P. Senthamaraikannan and M. R. Sanjay, et al., Study on characterization of Furcraea foetida new natural fiber as composite reinforcement for lightweight applications, Carbohydr. Polym., 2018, 181, 650–658 CrossRef CAS PubMed.
  289. T. P. Sathishkumar, P. Navaneethakrishnan and S. Shankar, et al., Characterization of new cellulose sansevieria ehrenbergii fibers for polymer composites, Compos. Interfaces, 2013, 20(8), 575–593 CrossRef CAS.
  290. National Jute Board, Development of jute fibre reinforced cement concrete composites, Final Project Report No. MDC/JTM/MM-IV/7.1/2008, Dated: 31.3.2008, Material Science Center, Indian Institute of Technology, Kharagpur, June, 2011 Search PubMed.
  291. D. Gon, K. Das and P. Paul, et al., Jute composites as wood substitute, Int. J. Textil. Sci., 2012, 1(6), 84–93 CrossRef.
  292. M. A. Khan, J. Ganster and H. P. Fink, Hybrid composites of jute and man-made cellulose fibres with polypropylene by injection moulding, Composites, Part A, 2009, 40, 846–851 CrossRef.
  293. Y. Dobah, M. Bourchak and A. Bezazi, et al., Multi-axial mechanical characterization of jute fibre/polyester composite materials, Composites, Part B, 2016, 90, 450–456 CrossRef CAS.
  294. A. N. Netravali and S. Chabba, Composites get greener, Mater. Today, 2003, 6(4), 22–29 CrossRef.
  295. M. A. Ashraf, M. Zwawi and M. T. Mehran, et al., Jute based bio and hybrid composites and their applications, Fibers, 2019, 7, 77 CrossRef CAS.
  296. S. R. Ferreira, E. Martinelli and M. Pepe, et al., Inverse identification of the bond behavior for jute fibers in cementitious matrix, Composites, Part B, 2016, 95, 440–452 CrossRef CAS.
  297. H. N. Dhakal, M. Skrifvars and K. Adekunle, et al., Falling weight impact response of jute/methacrylated soybean oil bio-composites under low velocity impact loading, Compos. Sci. Technol., 2014, 92, 134–141 CrossRef CAS.
  298. J. L. Thomason, Dependence of interfacial strength on the anisotropic fibre properties of jute reinforced composites, Polym. Compos., 2010, 31(9), 1525–1534 CrossRef CAS.
  299. T. T. L. Doan and E. Mader, Performance of jute fibre reinforced polypropylene, in 7th International AVK-TV Conference, Baden-Baden, September 28–29, 2004 Search PubMed.
  300. M. Lucka, A. K. Bledzki, and J. Michalski, Influence of the hydrophobisation of flax fibres on the water sensitivity, biological resistance and electrical properties of flax– polypropylene composites, in 5th Global Wood and Natural Fibre Composites Symposium, Kassel, Germany, 27–28 April, 2004, pp. 1–7 Search PubMed.
  301. A. Filho, S. Parveen and V. Rana, et al., Mechanical and micro-structural investigation of multi-scale cementitious composites developed using sisal fibres and microcrystalline cellulose, Ind. Crops Prod., 2020, 158, 112912 CrossRef CAS.
  302. G. Rajesh and A. V. R. Prasad, Tensile properties of successive alkali treated short jute fibre reinforced PLA composites, Prog. Mater. Sci., 2014, 5, 2188–2196 CAS.
  303. M. Raimo, Structure and morphology of cellulose fibers in garlic skin, Sci. Rep., 2020, 10, 2635 CrossRef CAS.
  304. A. K. Mohanty, M. Misra and L. T. Drzal, Surface modifications of natural fibres and performance of the resulting biocomposites: an overview, Compos. Interfaces, 2001, 8(5), 313–343 CrossRef CAS.
  305. J. Gassan and A. K. Bledzki, The influence of fibre surface treatment on the mechanical properties of jute–polypropylene composites, Composites, Part A, 1997, 28, 1001–1005 CrossRef.
  306. C. Baley, Analysis of the flax fibres tensile behaviour and analysis of the tensile stiffness increase, Composites, Part A, 2002, 33, 939–948 CrossRef.
  307. A. Etale, A. J. Onyianta, S. R. Turner and S. J. Eichhorn, Cellulose: A review of water interactions, applications in composites, and water treatment, Chem. Rev., 2023, 123(5), 2016–2048 CrossRef CAS.
  308. D. Klemm, B. heublein and H. P. Fink, et al., Cellulose: Fascinating biopolymer and sustainable raw material, Angew. Chem., Int. Ed., 2005, 44(22), 3358–3393 CrossRef CAS PubMed.
  309. H. V. Scheller and P. Ulvskov, Int. J. Sustainable Built Environ., 2010, 61(1), 263–289 CAS.
  310. M. Rajinipriya, M. Nagalakshmaiah and M. Robert, et al., Importance of agricultural and industrial waste in the field of nanocellulose and recent industrial developments of wood based nanocellulose: A review, ACS Sustain. Chem. Eng., 2018, 6(3), 2807–2828 CrossRef CAS.
  311. T. Sen and A. Paul, Confining concrete with sisal and jute FRP as alternatives for CFRP and GFRP, Int. J. Sustainable Built Environ., 2015, 4, 248–264 CrossRef.
  312. J. H. Lora and W. G. Glasser, Recent industrial applications of lignin: A sustainable alternative to nonrenewable matrials, J. Polym. Environ., 2002, 10(1), 39–48 CrossRef CAS.
  313. S. Biswas, S. Shahinur and M. Hasan, et al., Physical, mechanical and thermal properties of jute and bamboo fibre reinforced unidirectional epoxy composites, Procedia Eng., 2015, 105, 933–939 CrossRef CAS.
  314. M. Ramesh, K. Palanikumar and K. H. Reddy, Comparative evaluation on properties of hybrid glass fiber-sisal/jute reinforced epoxy composites, Procedia Eng., 2013, 51, 745–750 CrossRef CAS.
  315. M. Ramesh, K. Palanikumar and K. H. Reddy, Evaluation of mechanical and interfacial properties of sisal/jute/glass hybrid fiber reinforced polymer composites, Trans. Indian Inst. Met., 2016, 69, 1851–1859 CrossRef CAS.
  316. M. Ramesh, K. Palanikumar and K. H. Reddy, Impact behaviour analysis of sisal/jute and glass fiber reinforced hybrid composites, Adv. Mater. Res., 2014, 984–985, 266–272 Search PubMed.
  317. M. Ramesh, K. Palanikumar and K. H. Reddy, Influence of fiber orientation and fiber content on properties of sisal-jute-glass fiber-reinforced polyester composites, J. Appl. Polym. Sci., 2016, 133, 42968 CrossRef.
  318. S. K. Ramamoorthy, M. Skrifvars and A. Persson, A review of natural fibers used in biocomposites: plant, animal and regenerated cellulose fibers, Polym. Rev., 2015, 55, 107–162 CrossRef CAS.
  319. S. R. Ferreira, FdA. Silva and P. R. L. Lima, et al., Effect of hornification on the structure, tensile behavior and fiber matrix bond of sisal, jute and curaua fiber cement based composite systems, Constr. Build. Mater., 2017, 139, 551–561 CrossRef CAS.
  320. L. Mohammed, M. N. M. Ansari and G. Pua, et al., A review on natural fiber reinforced polymer composite and its applications, Int. J. Polym. Sci., 2015, 2015, 243947 Search PubMed.
  321. S. D. Salman, Effects of jute fibre content on the mechanical and dynamic mechanical properties of the composites in structural applications, Def. Technol., 2020, 16, 1098–1105 CrossRef.
  322. G. Kalusuraman, S. T. Kumaran and M. Aslan, et al., Use of waste copper slag filled jute fiber reinforced composites for effective erosion prevention, Measurement, 2019, 148, 106950 CrossRef.
  323. H. B. Nunez, P. J. H. Franco and D. E. R. Felix, et al., Surface modification and performance of jute fibers as reinforcement on polymer matrix: An overview, J. Nat. Fibers, 2019, 16(7), 944–960 CrossRef.
  324. A. V. Kiruthika and K. Veluraja, Physical properties of plant fibers (sisal, coir) and its composite material with tamarind seed gum as low-cost housing material, J. Nat. Fibers, 2017, 14, 801–813 CrossRef CAS.
  325. MdN. Khan, J. K. Roy and N. Akter, et al., Production and properties of short jute and short e-glass fiber reinforced polypropylene-based composites, Open J. Compos. Mater., 2012, 2(2), 40–47 CrossRef CAS.
  326. International Encyclopedia of Composites, ed. S. M. Lee, VCH, New York, 1990, vol. 4 Search PubMed.
  327. S. P. Kundu, S. Chakraborty and A. Roy, et al., Chemically modified jute fibre reinforced non-pressure (NP) concrete pipes with improved mechanical properties, Constr. Build. Mater., 2012, 37, 841–850 CrossRef.
  328. J. Ahmad, M. M. Arbili, A. Majdi, F. Althoey, A. Farouk Deifalla and C. Rahmawati, Performance of concrete reinforced with jute fibers (natural fibers): A review, J. Eng. Fibers Fabr., 2022, 17 DOI:10.1177/15589250221121871.
  329. K. S. Ahmed and S. Vijayarangan, Tensile, flexural and interlaminar shear properties of woven jute and jute-glass fabric reinforced polyester composites, J. Mater. Process. Technol., 2008, 207, 330–335 CrossRef CAS.
  330. E. T. N. Bisanda and M. P. Ansell, Properties of sisal-CNSL composites, J. Mater. Sci., 1992, 27, 1690–1700 CrossRef CAS.
  331. P. J. Roe and M. P. Ansell, Jute-reinforced polyester composites, J. Mater. Sci., 1985, 20, 4015–4020 CrossRef CAS.
  332. P. Feng, Y. Wu and T. Q. Liu, Non-uniform fiber-resin distributions of pultruded GFRP profiles, Composites, Part B, 2022, 231, 109543 CrossRef CAS.
  333. M. K. Sridhar, G. Basavarajappa and S. G. Kasturi, et al., Evaluation of jute as reinforcements in composites, Indian J. Text. Res., 1982, 79, 87–92 Search PubMed.
  334. H. Chandekar, V. Chaudhari and S. Waigaonkar, A review of jute fiber reinforced polymer composites, Mater. Today: Proc., 2020, 26, 2079–2082 CAS.
  335. M. Ramesh, L. Rajeshkumar and G. Sasikala, et al., A critical review on wood-based polymer composites: Processing, properties, and prospects, Polymers, 2022, 14(3), 589 CrossRef CAS PubMed.
  336. H. M. Akil, C. Santulli and F. Sarasini, et al., Environmental effects on the mechanical behaviour of pultruded jute/glass fibre-reinforced polyester hybrid composites, Compos. Sci. Technol., 2014, 94, 62–70 CrossRef CAS.
  337. H. M. Akil, L. W. Cheng and Z. A. Mohd Ishak, et al., Water absorption study on pultruded jute fibre reinforced unsaturated polyester composites, Compos. Sci. Technol., 2009, 69, 1942–1948 CrossRef CAS.
  338. M. H. Zamri, H. M. Akil and S. Safiee, et al., Predicting the coefficient of thermal expansion of pultruded composites with a natural fiber reinforcement, Mech. Compos. Mater., 2014, 50(5), 603–612 CrossRef CAS.
  339. K. Sever, The improvement of mechanical properties of jute fiber/LDPE composites by fiber surface treatment, J. Reinf. Plast. Compos., 2010, 29, 1921–1929 CrossRef CAS.
  340. S. N. Arju, A. M. Afsar and D. K. Das, et al., Role of reactive dye and chemicals on mechanical properties of jute fabrics polypropylene composites, Procedia Eng., 2014, 90, 199–205 CrossRef CAS.
  341. A. A. Kafi, K. Magniez and B. L. Fox, A surface-property relationship of atmospheric plasma treated jute composites, Compos. Sci. Technol., 2011, 71, 1692–1698 CrossRef CAS.
  342. M. R. Rahman, M. M. Huque and M. N. Islam, et al., Improvement of physico-mechanical properties of jute fibre reinforced polypropylene composites by post-treatment, Composites, Part A, 2008, 39, 1739–1747 CrossRef.
  343. M. R. Hossain, M. A. Islam and A. V. Vuurea, et al., Tensile behavior of environment friendly jute epoxy laminated composite, Procedia Eng., 2013, 56, 782–788 CrossRef CAS.
  344. M. J. Miah, M. A. Khan and R. A. Khan, Fabrication and characterization of jute fiber reinforced low density polyethylene based composites: Effects of chemical treatment, J. Sci. Res., 2011, 3(2), 249–259 CrossRef CAS.
  345. M. F. Hossen, S. Hamdan and M. R. Rahman, et al., Effect of clay content on the morphological, thermo-mechanical and chemical resistance properties of propionic anhydride treated jute fiber/polyethylene/nanoclay nanocomposites, Measurement, 2016, 90, 404–411 CrossRef.
  346. F. K. Liew, S. Hamdan and M. R. Rahman, et al., Thermomechanical properties of jute/bamboo cellulose composite and its hybrid composites: The effects of treatment and fiber loading, Adv. Mater. Sci. Eng., 2017, 2017, 8630749 Search PubMed.
  347. M. M. A. Sayeed, Comparison between tensile and damage analysis of hybrid jute/polypropylene needle punched nonwoven geotextiles produced from untreated and alkali treated jute fibers, J. Text. Inst., 2019, 110, 1800–1809 CrossRef CAS.
  348. N. Ranganathan, K. Oksman and S. K. Nayak, et al., Structure property relation of hybrid biocomposites based on jute, viscose and polypropylene: The effect of the fibre content and the length on the fracture toughness and the fatigue properties, Composites, Part A, 2016, 83, 169–175 CrossRef CAS.
  349. M. T. Zafar, S. N. Maiti and A. K. Ghosh, Effect of surface treatments of jute fibers on the microstructural and mechanical responses of poly(lactic acid)/jute fiber biocomposites, RSC Adv., 2016, 6, 73373–73382 RSC.
  350. V. Fiore, T. Scalici and F. Sarasini, et al., Salt-fog spray aging of jute-basalt reinforced hybrid structures: Flexural and low velocity impact response, Composites, Part B, 2017, 116, 99–112 CrossRef CAS.
  351. T. Sen and H. N. J. Reddy, Strengthening of RC beams in flexure using natural jute fibre textile reinforced composite system and its comparative study with CFRP and GFRP strengthening systems, Int. J. Sustainable Built Environ., 2013, 2, 41–55 CrossRef CAS.
  352. M. M. Kabir, H. Wang and T. Aravinthan, et al., Effects of natural fibre surface on composite properties: a review, Energy Environ. Sustain., 2011, 94–99 Search PubMed.
  353. B. W. Jo, S. Chakraborty and Y. S. Lee, Hydration study of the polymer modified jute fibre reinforced cement paste using analytical techniques, Constr. Build. Mater., 2015, 101, 166–173 CrossRef.
  354. S. Chakraborty, S. P. Kundu and A. Roy, et al., Improvement of the mechanical properties of jute fibre reinforced cement mortar: A statistical approach, Constr. Build. Mater., 2013, 38, 776–784 CrossRef.
  355. G. George, K. Joseph and E. R. Nagarajan, et al., Dielectric behaviour of PP/jute yarn commingled composites: Effect of fibre content, chemical treatments, temperature and moisture, Composites, Part A, 2013, 47, 12–21 CrossRef CAS.
  356. G. George, K. Joseph and E. R. Nagarajan, et al., Thermal, calorimetric and crystallisation behaviour of polypropylene/jute yarn bio-composites fabricated by commingling technique, Composites, Part A, 2013, 48, 110–120 CrossRef CAS.
  357. G. George, E. T. Jose and K. Jayanarayanan, et al., Novel bio-commingled composites based on jute/polypropylene yarns: Effect of chemical treatments on the mechanical properties, Composites, Part A, 2012, 43, 219–230 CrossRef.
  358. G. George, E. T. Jose and D. Akesson, et al., Viscoelastic behaviour of novel commingled biocomposites based on polypropylene/jute yarns, Composites, Part A, 2012, 43, 893–902 CrossRef CAS.
  359. X. Liu, Y. Cui and S. K. L. Lee, et al., Multiscale modeling of nano-SiO2 deposited on jute fibers via macroscopic evaluations and the interfacial interaction by molecular dynamics simulation, Compos. Sci. Technol., 2020, 188, 107987 CrossRef CAS.
  360. A. Ghosh, A. Ghosh and A. K. Bera, Bearing capacity of square footing on pond ash reinforced with jute-geotextile, Geotext. Geomembranes, 2005, 23, 144–173 CrossRef.
  361. B. K. Sarkar and D. Ray, Effect of the defect concentration on the impact fatigue ebdurance of untreated and alkali treated jute–vinylester composites under normal and liquid nitrogen atmosphere, Compos. Sci. Technol., 2004, 64, 2213–2219 CrossRef CAS.
  362. S. Mohanty, S. K. Verma and S. K. Nayak, Dynamic mechanical and thermal properties of MAPE treated jute/HDPE composites, Compos. Sci. Technol., 2006, 66, 538–547 CrossRef CAS.
  363. O. A. Khondker, U. S. Ishiaku and A. Nakai, et al., A novel processing technique for thermoplastic manufacturing of unidirectional composites reinforced with jute yarns, Composites, Part A, 2006, 37, 2274–2284 CrossRef.
  364. A. S. M. Sayem, J. Haider and M. M. A. Sayeed, Development and characterization of multi layered jute fabric-reinforced HDPE composites, J. Compos. Mater., 2020, 54, 1831–1845 CrossRef CAS.
  365. M. I. Lautenschlager, L. Mayer and J. Gebauer, et al., Comparison of filler-dependent mechanical properties of jute fiber reinforced sheet and bulk molding compound, Compos. Struct., 2018, 203, 960–967 CrossRef.
  366. D. Plackett, T. L. Andersen and W. B. Pedersen, et al., Biodegradable composites based on L-polylactide and jute fibres, Compos. Sci. Technol., 2003, 63, 1287–1296 CrossRef CAS.
  367. V. Sridharan and N. Muthukrishnan, Optimization of machinability of polyester/modified jute fabric composite using grey relational analysis (GRA), Procedia Eng., 2013, 64, 1003–1012 CrossRef CAS.
  368. B. K. Goriparthi, K. N. S. Suman and N. M. Rao, Effect of fibre surface treatments on mechanical and abrasive wear performance of polylactide/jute composites, Composites, Part A, 2012, 43, 1800–1808 CrossRef CAS.
  369. A. Dong, F. Li and X. Fan, et al., Enzymatic modification of jute fabrics for enhancing the reinforcement in jute/PP composites, J. Thermoplast. Compos. Mater., 2018, 31, 483–499 CrossRef CAS.
  370. A. Dong, X. Fan and Q. Wang, et al., Enzymatic treatments to improve mechanical properties and surface hydrophobicity of jute fiber membranes, BioResources, 2016, 11(2), 3289–3302 CAS.
  371. L. Liu, J. Yu, L. Cheng and W. Qu, Mechanical properties of poly (butylene succinate) (PBS) biocomposites reinforced with surface modified jute fibre, Composites, Part A, 2009, 40, 669–674 CrossRef.
  372. N. Sultana, S. M. Z. Hossain and M. S. Alam, et al., An experimental investigation and modeling approach of response surface methodology coupled with crow search algorithm for optimizing the properties of jute fiber reinforced concrete, Constr. Build. Mater., 2020, 243, 118216 CrossRef.
  373. M. J. Islam, M. J. Rahman and T. Mieno, Safely functionalized carbon nanotube–coated jute fibers for advanced technology, Adv. Compos. Hybrid Mater., 2020, 3, 285–293 CrossRef CAS.
  374. T. F. Salem, S. Tirkes and A. O. Akar, et al., Enhancement of mechanical, thermal and water uptake performance of TPU/jute fiber green composites via chemical treatments on fiber surface, e-Polym., 2020, 20, 133–143 CrossRef CAS.
  375. N. Chand and U. K. Dwivedi, Effect of coupling agent on abrasive wear behaviour of chopped jute fibre-reinforced polypropylene composites, Wear, 2006, 261, 1057–1063 CrossRef CAS.
  376. K. Das, D. Ray and N. R. Bandyopadhyay, et al., Physico-mechanical properties of the jute micro/nanofibril reinforced starch/polyvinyl alcohol biocomposite films, Composites, Part B, 2011, 42, 376–381 CrossRef.
  377. M. Ramesh, R. Vimal and K. H. H. Subramaniyan, et al., Study of mechanical properties of jute-banana-glass fiber reinforced epoxy composites under various post curing temperature, Appl. Mech. Mater., 2015, 766–767, 211–215 Search PubMed.
  378. R. Vimal, K. H. H. Subramanian and C. Aswin, et al., Comparisonal study of succinylation and phthalicylation of jute fibres: study of mechanical properties of modified fibre reinforced epoxy composites, Mater. Today: Proc., 2015, 2, 2918–2927 CAS.
  379. O. H. Margoto, K. D. S. D. Prado and R. C. Mergulhao, et al., Mechanical and thermal characterization of jute fabric reinforced polypropylene composites: Effect of maleic anhydride, J. Nat. Fibers, 2022, 19(5), 1792–1804 CrossRef CAS.
  380. X. Cao, M. Huang and B. Ding, et al., Robust polyacrylonitrile nanofibrous membrane reinforced with jute cellulose nanowhiskers for water purification, Desalination, 2013, 316, 120–126 CrossRef CAS.
  381. M. Boopalan, M. Niranjanaa and M. J. Umapathy, Study on the mechanical properties and thermal properties of jute and banana fibre reinforced epoxy hybrid composites, Composites, Part B, 2013, 51, 54–57 CrossRef.
  382. T. T. L. Doan, H. Brodowsky and E. Mader, Jute fibre/epoxy composites: Surface properties and interfacial adhesion, Compos. Sci. Technol., 2012, 72, 1160–1166 CrossRef CAS.
  383. J. Gassan and V. S. Cutowski, Effect of corona discharge and UV treatment on the properties of jute-fibre epoxy composites, Compos. Sci. Technol., 2000, 60, 2857–2863 CrossRef CAS.
  384. A. C. Pereira, S. N. Monteiro and F. Sd Assis, et al., Charpy impact tenacity of epoxy matrix composites reinforced with aligned jute fibers, J. Mater. Res. Technol., 2017, 6, 312–316 CrossRef CAS.
  385. H. Abdellaoui, H. Bensalah and M. Raji, et al., Laminated epoxy biocomposites based on clay and jute fibers, J. Biol. Eng., 2017, 14, 379–389 Search PubMed.
  386. M. A. Pinto, V. B. Chalivendra and Y. K. Kim, et al., Improving the strength and service life of jute/epoxy laminar composites for structural applications, Compos. Struct., 2016, 156, 333–337 CrossRef.
  387. M. A. Pinto, V. B. Chalivendra and Y. K. Kim, et al., Effect of surface treatment and Z-axis reinforcement on the interlaminar fracture of jute/epoxy laminated composites, Eng. Fract. Mech., 2013, 114, 104–114 CrossRef.
  388. M. Rokbi, A. Khaldoune and M. R. Sanjay, et al., Effect of processing parameters on tensile properties of recycled polypropylene based composites reinforced with jute fabrics, Int. J. Lightweight Mater. Manuf., 2020, 3, 144–149 Search PubMed.
  389. T. B. Yallew, P. Kumar and I. Singh, Sliding wear properties of jute fabric reinforced polypropylene composites, Procedia Eng., 2014, 97, 402–411 CrossRef CAS.
  390. T. Berhanu, P. Kumar, and I. Singh, Mechanical behaviour of jute fibre reinforced polypropylene composites, in Proc. Of 5th International & 26th All India Manufacturing Technology, Design and Research Conference (AIMTDR 2014) December 12th–14th, IIT Guwahati, Assam, India, 2014 Search PubMed.
  391. K. G. Melese and I. Singh, Joining behavior of jute/sisal fibers based epoxy laminates using different joint configurations, J. Nat. Fibers, 2022, 19, 2053–2064 CrossRef CAS.
  392. D. Chandramohan and K. Marimuthu, Thrust force and torque in drilling the natural fiber reinforced polymer composite materials and evaluation of delamination factor for bone graft substitutes – a work of fiction approach, Int. J. Eng. Sci. Technol., 2010, 2, 6437–6451 Search PubMed.
  393. H. Toshihiko, X. Zhilan and Y. Yuqiu, et al., Tensile properties of bamboo, jute and kenaf mat-reinforced composite, Energy Procedia, 2014, 56, 72–79 CrossRef.
  394. B. Abdelhak, M. Noureddine, and M. Hacen, Improvement of the interfacial adhesion between fiber and matrix, Mechanic Mech. Eng., 2020, vol. 22, pp. 885–894 Search PubMed.
  395. J. Yang, Y. C. Ching and C. H. Chuah, Applications of lignocellulosic fibers and lignin in bioplastics: a review, Polymers, 2019, 11(5), 751 CrossRef CAS PubMed.
  396. A. C. H. Barreto, D. S. Rosa and P. B. A. Fechine, et al., Properties of sisal fibres treated by alkali solution and their application into cardanol-based biocomposites, Composites, Part A, 2011, 42, 492–500 CrossRef.
  397. M. Agrawal, R. Naik and S. Shetgar, et al., Surface treatment of jute fibre using eco-friendly method and its use in PP composites, Mater. Today: Proc., 2019, 18, 3268–3275 CAS.
  398. M. George, M. Chae and D. C. Bressler, Composite materials with bast fibres: Structural, technical and environmental benefits, Prog. Mater. Sci., 2016, 83, 1–23 CrossRef CAS.
  399. J. J. Kennad, K. Sankaranarayanasamy and C. S. Kumar, Chemical, biological, and nanoclay treatments for natural plant fiber-reinforced polymer composites: A review, Polym. Polym. Compos., 2020, 29, 1011–1038 CrossRef.
  400. X. Li, L. G. Tabil and S. Panigrahi, Chemical treatments of natural fibre for use in natural fibre reinforced composites: A review, J. Polym. Environ., 2007, 15, 25–33 CrossRef.
  401. S. Ochi, Mechanical properties of kenaf fibres and kenaf/PLA composites, Mech. Mater., 2008, 40(4–5), 446–452 CrossRef.
  402. C. Qin, N. Soykeabkaew and N. Xiuyuan, et al., The effect of fibre volume fraction and mercerization on the properties of all-cellulose composites, Carbohydr. Polym., 2008, 71, 458–467 CrossRef CAS.
  403. A. Arbelaiz, B. Fernandez and J. A. Ramos, et al., Thermal and crystallization studies of short flax fibre reinforced polypropylene matrix composites: effect of treatments, Thermochim. Acta, 2006, 440(2), 111–121 CrossRef CAS.
  404. R. D. T. Filho, F. A. Silva and E. M. R. Fairbairn, et al., Durability of compression molded sisal fiber reinforced mortar laminates, Constr. Build. Mater., 2009, 23(6), 2409–2420 CrossRef.
  405. J. Wei and C. Meyer, Degradation mechanisms of natural fiber in the matrix of cement composites, Cem. Concr. Res., 2015, 73, 1–16 CrossRef CAS.
  406. A. E. F. S. Almeida, G. H. D. Tonoli and S. F. Santos, et al., Improved durability of vegetable fiber reinforced cement composite subject to accelerated carbonation at early age, Cem. Concr. Compos., 2013, 42, 49–58 CrossRef CAS.
  407. R. Latif, S. Wakeel and N. Z. Khan, et al., Surface treatments of plant fibers and their effects on mechanical properties of fiber-reinforced composites: A review, J. Reinf. Plast. Compos., 2019, 38, 15–30 CrossRef CAS.
  408. P. Joseph, M. S. Rabello and L. H. Mattoso, et al., Environmental effects on the degradation behaviour of sisal fibre reinforced polypropylene composites, Compos. Sci. Technol., 2002, 62, 1357–1372 CrossRef CAS.
  409. F. Corrales, F. Vilaseca and M. Llop, et al., Chemical modification of jute fibres for the production of green-composites, J. Hazard. Mater., 2007, 144, 730–735 CrossRef CAS PubMed.
  410. K. Jayaraman, Manufacturing sisal–polypropylene composites with minimum fibre degradation, Compos. Sci. Technol., 2003, 63, 367–374 CrossRef CAS.
  411. X. Yuan, K. Jayaraman and D. Bhattacharyya, Plasma treatment of sisal fibres and its effects on tensile strength and interfacial bonding, J. Adhes. Sci. Technol., 2002, 16, 703–727 CrossRef CAS.
  412. J. Militky and A. Jabbar, Comparative evaluation of fibre treatments on the creep behaviour of jute/green epoxy composites, Composites, Part B, 2015, 80, 361–368 CrossRef CAS.
  413. Y. Seki, M. Sarikanat and K. Sever, et al., Effect of the low and radio frequency oxygen plasma treatment of jute fibre on mechanical properties of jute fibre/polyester composite, Fibres Polym., 2010, 11(8), 1159–1164 CrossRef CAS.
  414. M. Naebe, P. G. Cookson and J. Rippon, et al., Effects of plasma treatment of wool on the uptake of sulfonated dyes with different hydrophobic properties, Text. Res. J., 2010, 80(4), 312–324 CrossRef CAS.
  415. S. Garg, C. Hurren and A. Kaynak, Improvement of adhesion of conductive polypyrrole coating on wool and polyester fabrics using atmospheric plasma treatment, Synth. Met., 2007, 157(1), 41–47 CrossRef CAS.
  416. M. A. Khan, N. Haque and A. Al-Kafi, et al., Jute reinforced polymer composite by gamma radiation: effect of surface treatment with UV radiation, Polym.-Plast. Technol. Eng., 2006, 45(5), 607–613 CrossRef CAS.
  417. M. T. Hossain, M. S. Hossian and M. B. Uddin, et al., Preparation and characterization of sodium silicate-treated jute-cotton blended polymer-reinforced UPR-based composite: effect of γ-radiation, Adv. Compos. Hybrid Mater., 2021, 4, 257–264 CrossRef CAS.
  418. M. Strobel and C. S. Lyons, The role of low-molecular-weight oxidized materials in the adhesion properties of corona-treated polypropylene film, J. Adhes. Sci. Technol., 2003, 17(1), 15–23 CrossRef CAS.
  419. I. Sakata, M. Morita and N. Tsuruta, et al., Activation of wood surface by corona treatment to improve adhesive bonding, J. Appl. Polym. Sci., 1993, 49, 1251–1258 CrossRef CAS.
  420. M. N. Belgacem, P. Bataille and S. Sapieha, Effect of corona modification on the mechanical properties of polypropylene/cellulose composites, J. Appl. Polym. Sci., 1994, 53, 379–385 CrossRef CAS.
  421. S. Dong, S. Sapieha and H. P. Schreiber, Rheological properties of corona modified cellulose/polyethylene composites, Polym. Eng. Sci., 1992, 32(22), 1734–1739 CrossRef CAS.
  422. P. Bataille, M. Dufourd and S. Sapieha, Copolymerization of styrene on to cellulose activated by corona, Polym. Int., 1994, 24(4), 387–391 CrossRef.
  423. K. Tamargo-Martínez, A. Martínez-Alonso and M. A. Montes-Morán, et al., Effect of oxygen plasma treatment of PPTA and PBO fibres on the interfacial properties of single fibre/epoxy composites studied by Raman spectroscopy, Compos. Sci. Technol., 2011, 71(6), 784–790 CrossRef.
  424. M. Stepankova, J. Wiener and J. Dembický, Impact of laser thermal stress on cotton fabric, Cem. Concr. Compos., 2010, 18(3), 70–73 CAS.
  425. A. Baltazar-y-Jimenez, J. Juntaro and A. Bismarck, Effect of atmospheric air pressure plasma treatment on the thermal behaviour of natural fibres and dynamical mechanical properties of randomly-oriented short fibre composites, J. Biobased Mater. Bioenergy, 2008, 2(3), 264–272 CrossRef.
  426. R. Molina, P. Jovancic and F. Comelles, et al., Shrink-resistance and wetting properties of keratin fibres treated by glow discharge, J. Adhes. Sci. Technol., 2002, 16(11), 1469–1485 CrossRef CAS.
  427. H. Ma and C. W. Joo, Structure and mechanical properties of jute–polylactic acid biodegradable composites, J. Compos. Mater., 2010, 45, 1451–1460 Search PubMed.
  428. T. Yu, J. Ren and S. Li, et al., Effect of fibre surface-treatments on the properties of poly(lactic acid)/ramie composites, Composites, Part A, 2010, 41(4), 499–505 CrossRef.
  429. M. S. Huda, L. T. Drzal and A. K. Mohanty, et al., Effect of fibre surface-treatments on the properties of laminated biocomposites from poly (lactic acid) (PLA) and kenaf fibres, Compos. Sci. Technol., 2008, 68, 424–432 CrossRef CAS.
  430. M. A. Sawpan, K. L. Pickering and A. Fernyhough, Improvement of mechanical performance of industrial hemp fibre reinforced polylactide biocomposites, Composites, Part A, 2011, 42, 310–319 CrossRef.
  431. R. Csizmadia, G. Faludi and K. Renner, et al., PLA/wood biocomposites: improving composite strength by chemical treatment of fibres, Composites, Part A, 2013, 53, 46–53 CrossRef CAS.
  432. D. K. F. Anna, B. Aiswarya and B. Hong, et al., Effect of surface modification of jute fiber on the mechanical properties and durability of jute fiber-reinforced epoxy composites, Polym. Compos., 2018, 39, E2519–E2528 Search PubMed.
  433. L. Ammayappan, S. Chakraborty and I. Musthafa, et al., Standardization of a chemical modification protocol for jute fabric reinforcement, J. Nat. Fibers, 2022, 19(2), 562–574 CrossRef CAS.
  434. H. Wang, H. Memon and E. A. M. Hassan, et al., Effect of jute fiber modification on mechanical properties of jute fiber composite, Materials, 2019, 12, 1226 CrossRef CAS.
  435. J. O. Ighalo, C. A. Adeyanju and S. Ogunniyi, et al., An empirical review of the recent advances in treatment of natural fibers for reinforced plastic composites, Compos. Interfaces, 2021, 28, 925–960 CrossRef CAS.
  436. R. K. Samal, M. Mohanty and B. B. Panda, Effect of chemical modification on FTIR spectra; physical and chemical behavior of jute II, J. Polym. Mater., 1995, 12, 235–240 Search PubMed.
  437. A. K. Rana, A. Mandal and B. C. Mitra, et al., Short jute fibre-reinforced polypropylene composites: effect of compatibiliser, J. Appl. Polym. Sci., 1998, 69, 329–338 CrossRef CAS.
  438. M. H. Islam, M. R. Islam, M. Dulal, S. Afroj and N. Karim, The effect of surface treatments and graphene-based modifications on mechanical properties of natural jute fiber composites: A review, iScience, 2021, 25(1), 103597 CrossRef.
  439. P. Zadorecki and P. Flodin, Surface modification of cellulose fibres. I. Spectroscopic characterisation of surface-modified cellulose fibres and their copolymerisation with styrene, J. Appl. Polym. Sci., 1985, 30, 2419–2429 CrossRef CAS.
  440. P. Zadorecki and P. Flodin, Surface modification of cellulose fibres. II. The effect of cellulose fibre treatment on the performance of cellulose–polyester composites, J. Appl. Polym. Sci., 1985, 30, 3971–3983 CrossRef CAS.
  441. D. Ray, B. K. Sarkar and S. Das, et al., Dynamic mechanical and thermal analysis of vinylester–resin–matrix composites reinforced with untreated and alkali-treated jute fibres, Compos. Sci. Technol., 2002, 62, 911–917 CrossRef CAS.
  442. I. K. Varma, S. R. Ananthakrishnan and S. Krishnamoorthy, Effect of chemical treatment on thermal behavior of jute fibers, Text. Res. J., 1988, 58, 486–494 CrossRef CAS.
  443. M. K. Hossain, M. W. Dewan and M. Hosur, et al., Mechanical performances of surface modified jute fibre reinforced biopol nanophased green composites, Composites, Part B, 2011, 42, 1701–1707 CrossRef.
  444. D. Ray, B. K. Sarkar and R. K. Basak, et al., Study of thermal behaviour of alkali-treated jute fibres, J. Appl. Polym. Sci., 2002, 85, 2594–2599 CrossRef CAS.
  445. D. Ray, B. K. Sarkar and A. K. Rana, Fracture behavior of vinylester resin matrix composites reinforced with alkali-treated jute fibers, J. Appl. Polym. Sci., 2002, 85, 2588–2593 CrossRef CAS.
  446. D. Ray, B. K. Sarkar and A. K. Rana, et al., The mechanical properties of vinylester resin matrix composites reinforced with alkali-treated jute fibres, Composites, Part A, 2001, 32, 119–127 CrossRef CAS.
  447. L. Y. Mwaikambo and M. P. Ansell, Hemp fibre reinforced cashew nut shell liquid composites, Compos. Sci. Technol., 2003, 63, 1297–1305 CrossRef CAS.
  448. G. Rajesh and A. V. R. Prasad, Effect of fibre loading and successive alkali treatments on tensile properties of short jute fibre reinforced polypropylene composites, Adv. Mater. Manuf. Charact., 2013, 3(2), 528–532 Search PubMed.
  449. D. Ray, B. K. Sarkar and A. K. Rana, et al., Effect of alkali treated jute fibre on composite properties, Bull. Mater. Sci., 2001, 24(2), 129–135 CrossRef CAS.
  450. M. A. Gulbarga and S. B. Burli, Jute fiber-PP bio-composite: State of art, low investment, in-house and manual preparation of injection moldable bio-composite granules, Int. J. Sci. Res. Publ., 2013, 3(8), 1–6 Search PubMed.
  451. F. Sarker, P. Potluri and S. Afroj, et al., Ultrahigh performance of nano-engineered graphene-based natural jute fiber composites, ACS Appl. Mater. Interfaces, 2019, 11, 21166–21176 CrossRef CAS PubMed.
  452. S. J. Eichhorn and R. J. Young, Composite micromechanics of hemp fibres and epoxy resin microdroplets, Compos. Sci. Technol., 2004, 64, 762–772 CrossRef.
  453. M. A. Khan, F. Mina and L. T. Drzal, Influence of silane coupling agents of different functionalities on the performance of jute–polycarbonate composites, in 3rd International Wood and Natural Fibre Composite Symposium, September, 2000 Search PubMed.
  454. M. Z. Rong, M. Q. Zhang and Y. Liu, et al., Mechanical properties of sisal reinforced composites in response to water absorption, Polym. Polym. Compos., 2002, 10(6), 407–426 CrossRef CAS.
  455. A. Valadez-Gonzales, J. M. Cervantes-Uc and R. Olayo, et al., Chemical modification of henequen fibres with an organosilane coupling agent, Composites, Part B, 1999, 30, 321–331 CrossRef.
  456. J. Gassan and A. K. Bledzki, Effect of moisture content on the properties of silanized jute-epoxy composites, Polym. Compos., 1997, 18, 179–184 CrossRef CAS.
  457. J. Gassan and A. K. Bledzki, Effect of cyclic moisture absorption desorption on the mechanical properties of silanized jute–epoxy composites, Polym. Compos., 1999, 20(4), 604–611 CrossRef CAS.
  458. L. A. Pothan and S. Thomas, Polarity parameters and dynamic mechanical behaviour of chemically modified banana fiber reinforced polyester composites, Compos. Sci. Technol., 2003, 63, 1231–1240 CrossRef CAS.
  459. P. J. Herrera-Franco and A. Valadez-Gonzales, Mechanical properties of continuous natural fibre-reinforced polymer composites, Composites, Part A, 2004, 35, 339–345 CrossRef.
  460. S. A. Paul, D. Piasta and S. Spange, et al., Solvatochromic and electrokinetic studies of banana fibrils prepared from steam-exploded banana fibre, Biomacromol, 2008, 9, 1802–1810 CrossRef CAS.
  461. S. A. Paul, A. Boudenne and L. Ibos, et al., Effect of fibre loading and chemical treatments on thermophysical properties of banana fibre/polypropylene commingled composite materials, Composites, Part A, 2008, 39(9), 1582–1588 CrossRef.
  462. J. Gassan and A. K. Bledzki, Influence of fiber surface treatment on the creep behavior of jute fiber-reinforced polypropylene, J. Thermoplast. Compos. Mater., 1999, 12(5), 388–398 CrossRef CAS.
  463. A. C. Karmaker, A. Hoffmann and G. Hinrichsen, Influence of water uptake on the mechanical properties of jute fibre-reinforced polypropylene, J. Appl. Polym. Sci., 1994, 54, 1803–1807 CrossRef CAS.
  464. A. C. Karmaker and J. A. Youngquist, Injection molding of polypropylene reinforced with short jute fibers, J. Appl. Polym. Sci., 1996, 62(8), 1147–1151 CrossRef CAS.
  465. A. C. Kamaker, Effect of water absorption on dimensional stability and impact energy of jute fibre reinforced polypropylene, J. Mater. Sci. Lett., 1997, 16, 462–464 CrossRef.
  466. T. Paunikallio, J. Kasanen and M. Suvanto, et al., Influence of maleated polypropylene on mechanical properties of composite made of viscose fiber and polypropylene, J. Appl. Polym. Sci., 2003, 87, 1895–1900 CrossRef CAS.
  467. S. A. Paul, T. Reussmann and G. Mennig, et al., The role of interface modification on the mechanical properties of injection moulded composites from commingled polypropylene/banana granules, Compos. Interfaces, 2007, 14(7–9), 849–867 CrossRef CAS.
  468. S. N. Chattopadhyay, N. C. Pan and A. N. Roy, et al., Two step belaching of jute yarn and fabric using hydrogen peroxide and preacetic acid, J. Nat. Fibers, 2022, 19, 1159–1167 CrossRef CAS.
  469. M. A. Khan, M. M. Rahman and K. S. Akhunzada, Grafting of different monomers onto jute yarn by in situ UV-radiation method: Effect of additives, Polym.-Plast. Technol. Eng., 2002, 41(4), 677–689 CrossRef CAS.
  470. M. M. Hassan, M. R. Islam and M. A. Khan, Improvement of physicomechanical properties of jute yarn by photografting with 3-(trimethoxysilyl) propylmethacrylate, J. Adhes. Sci. Technol., 2003, 17(5), 737–750 CrossRef CAS.
  471. C. Chuai, K. Almdal and L. Poulsen, et al., Conifer fibres as reinforcing materials for polypropylene-based composites, J. Appl. Polym. Sci., 2001, 80, 2833–2841 CrossRef CAS.
  472. G. Cantero, A. Arbelaiz, R. Llano-Ponte and I. Mondragon, Effects of fibre treatment on wettability and mechanical behaviour of flax/polypropylene composites, Compos. Sci. Technol., 2003, 63, 1247–1254 CrossRef CAS.
  473. R. K. Samal, H. S. Samantaray and R. N. Samal, Graft copolymerization with a new class of acidic peroxo salt IV. Grafting of acrylamide onto jute fiber using potassium monopersulphate: Catalyzed by Fe(II), Polym. J., 1986, 18, 471–478 CrossRef CAS.
  474. G. S. Deshmukh, Advancement in hemp fibre polymer composites: a comprehensive review, J. Polym. Eng., 2022, 42, 575–598 CrossRef CAS.
  475. A. K. Mohanty, S. Patnaik and B. C. Singh, Graft copolymerization of acrylonitrile onto acetylated jute fibers, J. Appl. Polym. Sci., 1989, 37, 1171–1181 CrossRef CAS.
  476. H. M. D. Rahman, J. M. D. Alam and I. H. M. D. Mondal, Modification of Jute Fiber with Vinyl Acetate and Methyl Vinyl Ketone for Textile Applications, Adv. Res. Text. Eng., 2023, 8(1), 1082 Search PubMed.
  477. S. S. Tripathy, S. Jena and S. B. Misra, et al., A study on graft copolymerization of methyl methacrylate onto jute fiber, J. Appl. Polym. Sci., 1985, 30, 1399–1406 CrossRef CAS.
  478. X. Liu, Y. Cui and S. Hao, et al., Influence of depositing nano-SiO2 particles on the surface microstructure and properties of jute fibers via in situ synthesis, Composites, Part A, 2018, 109, 368–375 CrossRef CAS.
  479. K. C. M. Nair, S. Thomas and G. Groeninckx, Thermal and dynamic mechanical analysis of polystyrene composites reinforced with short sisal fibers, Compos. Sci. Technol., 2001, 61, 2519–2529 CrossRef.
  480. C. S. R. Freire, A. J. D. Silvestre and C. P. Neto, et al., Composites based on acylated cellulose fibers and low-density polyethylene: effect of the fiber content, degree of substitution and fatty acid chain length on final properties, Compos. Sci. Technol., 2008, 68, 3358–3364 CrossRef CAS.
  481. J. George, S. S. Bhagawan and S. Thomas, Improved interactions in chemically modified pineapple leaf fibre reinforced polyethylene composites, Compos. Interfaces, 1998, 5(3), 201–223 CrossRef CAS.
  482. L. A. Pothan, J. George and S. Thomas, Effect of fibre surface treatments on the fibre–matrix interaction in banana fibre reinforced polyester composites, Compos. Interfaces, 2002, 9(4), 335–353 CrossRef CAS.
  483. P. V. Joseph, K. Joseph and S. Thomas, Short sisal fiber reinforced polypropylene composites: the role of interface modification on ultimate properties, Compos. Interfaces, 2002, 9(2), 171–205 CrossRef CAS.
  484. B. C. Mitra, R. K. Basak and M. Sarkar, Studies on jute-reinforced composites, its limitations, and some solutions through chemical modifications of fibres, J. Appl. Polym. Sci., 1998, 67(6), 1093–1100 CrossRef CAS.
  485. T. Paunikallio, M. Suvanto and T. T. Pakkanen, Composition, tensile properties and dispersion of polypropylene composites reinforced with viscose fibers, J. Appl. Polym. Sci., 2004, 91, 2676–2684 CrossRef CAS.
  486. H. Anuar, A. Zuraida and J. G. Kovacs, et al., Improvement of mechanical properties of injection molded polylactic acid kenaf fibre biocomposite, J. Thermoplast. Compos. Mater., 2012, 25, 153–164 CrossRef CAS.
  487. J. I. P. Singh, S. Singh and V. Dhawan, Effect of alkali treatment on mechanical properties of jute fiber-reinforced partially biodegradable green composites using epoxy resin matrix, Polym. Polym. Compos., 2019, 28(6), 388–397 CrossRef.
  488. K. P. Mieck, A. Nechwatal and C. Knobelsdorf, Faser-Matrix-Haftung in Kunststoffverbunden aus Thermoplastischer Matrix und Flachs: 1. Die Ausrüstung mit Silanen, Angew. Makromol. Chem., 1995, 224, 73–88 CrossRef CAS.
  489. A. K. Mohanty, M. A. Khan and G. Hinrichsen, Influence of chemical surface modification on the properties of biodegradable jute fabric–polyester amide composites, Composites, Part A, 2000, 31, 143–150 CrossRef.
  490. D. Verma and K. L. Goh, Effect of mercerization/alkali surface treatment of natural fibres and their utilization in polymer composites: Mechanical and morphological studies, J. Compos. Sci., 2021, 5, 175 CrossRef CAS.
  491. H.-P. Fink, E. Walenta and J. Kunze, The structure of natural cellulosic fibres-Part 2. The supermolecular structure of bast fibres and their changes by mercerization as revealed by X-Ray diffraction and 13C-NMR-spectroscopy, Papier, 1999, 53(9), 534–542 CAS.
  492. E. Dinand, M. Vignon and H. Chanzy, et al., Mercerization of primary wall cellulose and its implication for the conversion of cellulose I→cellulose II, Cellulose, 2002, 9, 7–18 CrossRef CAS.
  493. A. K. Bledzki, K. Specht, H.-P. Fink, et al., Abschlussbericht des FNRVerbundvorhabens, Höher belastbare, leichte Fahrzeuginnenbauteil, Gülzow, Germany, December 2000 Search PubMed.
  494. P. V. Joseph, K. Joseph and S. Thomas, et al., The thermal and crystallisation studies of short sisal fibre reinforced polypropylene composites, Composites, Part A, 2003, 34(3), 253–266 CrossRef.
  495. D. Feng, D. F. Caulfield and A. R. Sanadi, Effect of compatibilizer on the structure–property relationships of kenaf-fiber/polypropylene composites, Polym. Compos., 2001, 22(4), 506–517 CrossRef CAS.
  496. A. R. Sanadi and D. F. Caulfield, Transcrystalline interphases in natural fibre–PP composites: effect of coupling agent, Compos. Interfaces, 2000, 7(1), 31–43 CrossRef CAS.
  497. W. Qiu, F. Zhang and T. Endo, et al., Preparation and characteristics of composites of high-crystalline cellulose with polypropylene: effects of maleated polypropylene and cellulose content, J. Appl. Polym. Sci., 2003, 87, 337–345 CrossRef CAS.
  498. A. R. Sanadi, D. Feng, and D. F. Caulfield, Highly filled lignocellulosic reinforced thermoplastics: effect of interphase modification, in Proceedings of the 18th Riso International Symposium on Materials Science: Polymeric Composites – Expanding the Limits, 1997, pp. 465–470 Search PubMed.
  499. K. L. Fung, R. K. Y. Li and S. C. Tjong, Interface modification on the properties of sisal fiber-reinforced polypropylene composites, J. Appl. Polym. Sci., 2002, 85, 169–176 CrossRef CAS.
  500. P. R. Hornsby, E. Hinrichsen and K. Tarverdi, Preparation and properties of polypropylene composites reinforced with wheat and flax straw fibres, J. Mater. Sci., 1997, 32, 1009–1015 CrossRef CAS.
  501. K. Van de Velde and P. Kiekens, Influence of fibre and matrix modifications on mechanical and physical properties of flax fibre reinforced poly (propylene), Macromol. Mater. Eng., 2001, 286(4), 237–242 CrossRef CAS.
  502. R. Karnani, M. Krishnan and R. Narayan, Biofibre-reinforced polypropylene composites, Polym. Eng. Sci., 1997, 37(2), 476–483 CrossRef CAS.
  503. Z. Y. Sun, H. S. Han and G. C. Dai, Mechanical properties of injection-molded natural fibre-reinforced polypropylene composites: formulation and compounding processes, J. Reinf. Plast. Compos., 2010, 29, 637–650 CrossRef CAS.
  504. H. D. Rozman, K. W. Tan and R. N. Kumar, et al., The effect of lignin as a compatibilizer on the physical properties of coconut fiber-polypropylene composites, Eur. Polym. J., 2000, 36, 1483–1494 CrossRef CAS.
  505. W. Thielemans and R. Wool, Butyrated kraft lignin as compatibilizing agent for natural fiber reinforced thermoset composites, Composites, Part A, 2004, 35, 327–338 CrossRef.
  506. M. Avella, L. Casale and R. Dellerba, et al., Broom fibers as reinforcing materials for polypropylene-based composites, J. Appl. Polym. Sci., 1998, 68, 1077–1089 CrossRef CAS.
  507. S. Takase and N. Shiraishi, Studies on composites from wood and polypropylenes II, J. Appl. Polym. Sci., 1989, 37, 645–659 CrossRef CAS.
  508. M. Kazayawoko, J. J. Balatinecz and L. M. Matuana, Surface modification and adhesion mechanisms in wood fibre–polypropylene composites, J. Mater. Sci., 1999, 34, 6189–6199 CrossRef CAS.
  509. M. N. C. Ichazo, J. Gonzalez and R. Perera, et al., Polypropylene/wood flour composites: treatments and properties, Compos. Struct., 2001, 54, 207–214 CrossRef.
  510. A. Paul, K. Joseph and S. Thomas, Effect of surface treatments on the electrical properties of low-density polyethylene composites reinforced with short sisal fibers, Compos. Sci. Technol., 1997, 57, 67–79 CrossRef CAS.
  511. Y. Karaduman and O. L. Gokcan, Effect of enzymatic pretreatment on the mechanical properties of jute fiber-reinforced polyester composites, J. Compos. Mater., 2013, 47(10), 1293–1302 CrossRef.
  512. Y. Li, H. Sun and Y. Zhang, et al., The three-dimensional heterostructure synthesis of ZnO/cellulosic fibers and its application for rubber composites, Compos. Sci. Technol., 2019, 177, 10–17 CrossRef CAS.
  513. S. Zhang, F. Zhang and Y. Pan, et al., Multiwall-carbon-nanotube/cellulose composite fibers with enhanced mechanical and electrical properties by cellulose grafting, RSC Adv., 2018, 8, 5678–5684 RSC.
  514. S. Ovali and E. Sancak, Invetigation on mechanical properties of jute fiber reinforced low density polyethylene composites, J. Nat. Fibers, 2022, 19, 3109–3126 CrossRef CAS.
  515. Y. Karaduman, M. M. A. Sayeed and L. Onal, et al., Viscoelastic properties of surface modified jute fibre/polypropylene nonwoven composites, Composites, Part B, 2014, 67, 111–118 CrossRef CAS.
  516. J. Tusnim, N. S. Jenifar and M. Hasan, Effect of chemical treatment of jute fiber on thermo-mechanical properties of jute and sheep wool fiber reinforced hybrid polypropylene composites, J. Thermoplast. Compos. Mater., 2022, 35, 1981–1993 CrossRef CAS.
  517. L. A. Pothan, S. Thomas and N. R. Neelakantan, Short banana fibre reinforced polyester composites: mechanical, failure and aging characteristics, J. Reinf. Plast. Compos., 1997, 16, 744–765 CrossRef CAS.
  518. S. Kalia, K. Thakur, A. Celli, M. A. Kiechel and C. L. Schauer, Surface modification of plant fibers using environment friendly methods for their application in polymer composites, textile industry and antimicrobial activities: A review, J. Environ. Chem. Eng., 2013, 1, 97–112 CrossRef CAS.
  519. V. Sadrmanesh and Y. Chen, Bast fibres: structure, processing, properties, and applications, Int. Mater. Rev., 2019, 64, 38–406 Search PubMed.
  520. J. A. Khan and M. A. Khan, The use of jute fibers as reinforcements in composites, Biofiber Reinforcements in Composite Materials, 2015, pp. 3–34 Search PubMed.
  521. P. K. Pal and S. R. Ranganathan, Jute plastics composites for the building industry, Pop. Plast., 1986, 31, 22–24 CAS.
  522. R. Davidson, and N. L. Hancox, Proc. Int. Symposium on Biocomposites and Blends Based on Jute and Allied Fibres, Calcutta, 1994, p. 83 Search PubMed.
  523. A. N. Shah and S. C. Lakkad, Mechanical properties of jute-reinforced plastics, Fibre Sci. Technol., 1981, 15, 41–46 CrossRef.
  524. R. A. Clark and M. P. Ansell, Jute and glass fibre hybrid laminates, J. Mater. Sci., 1986, 21, 269–276 CrossRef.
  525. C. Pavithran, K. Gopakumar and S. V. Prasad, et al., Copper coating on coir fibres, J. Mater. Sci., 1981, 16, 1548–1556 CrossRef CAS.
  526. M. K. Sridhar, G. Basavarappa and G. S. Kasturi, et al., Mechanical Properties of Jute/Polyester Composites, Indian J. Technol., 1984, 22, 213–215 CAS.
  527. I. K. Varma, S. R. Ananthakrishnan and S. Krishnamoorthy, Composites of glass/modified jute fabric and unsaturated polyester resin, Composites, 1989, 20, 383–388 CrossRef CAS.
  528. S. Das, Jute Composite and its Applications, International workshop IJSG, 2009 Search PubMed.
  529. Y. H. Celik, E. Kilickap and A. I. Kilickap, An experimental study on milling of natural fiber (jute)-reinforced polymer composite, J. Compos. Mater., 2019, 53, 3127–3137 CrossRef CAS.
  530. V. S. Chinta, P. R. Reddy, K. E. Prasad, et al., Investigation of fracture parameters of jute/glass reinforced hybrid composite and analysis by using FEA, Emerging Trends in Mechanical Engineering, Lecture Notes in Mechanical Engineering, ed. L. Vijayaraghavan, et al.,  DOI:10.1007/978-981-32-9931-3_22.
  531. P. Krishnasamy, G. Rajamurugan and M. Thirumurugan, Dynamic mechanical characteristics of jute fiber and 304 wire mesh reinforced epoxy composite, J. Ind. Text., 2021, 51, 540–588 CrossRef CAS.
  532. A. Kumar and A. Srivastava, Preparation and mechanical properties of jute fiber reinforced epoxy composites, Ind. Eng. Manag., 2017, 6(4), 234 Search PubMed.
  533. J. I. P. Singh, S. Singh and V. Dhawan, Influence of fiber volume fraction and curing temperature on mechanical properties of jute/PLA green composites, Polym. Polym. Compos., 2020, 28, 273–284 CrossRef CAS.
  534. S. Nahar, R. A. Khan and K. Dey, et al., Comparative studies of mechanical and interfacial properties between jute and bamboo fiber-reinforced polypropylene-based composites, J. Thermoplast. Compos. Mater., 2012, 25(1), 15–32 CrossRef.
  535. F. A. Mirza, S. M. Rasel, A. M. Afsar, et al., Injection molding and mechanical properties evaluation of short jute fiber polypropylene reinforced composites. Natural Filler and Fibre Composites: Development and Characterisation, WIT Transactions on State of the Art in Science and Engineering, 2014,  DOI:10.2495/978-1-78466-147-2/006.
  536. V. Prasad, A. Joy and G. Venkatachalam, et al., Finite element analysis of jute and banana fibre reinforced hybrid polymer matrix composite and optimization of design parameters using ANOVA technique, Procedia Eng., 2014, 97, 1116–1125 CrossRef CAS.
  537. A. Chatterjee, S. Kumar and H. Singh, Tensile strength and thermal behavior of jute fibre reinforced polypropylene laminate composite, Compos. Commun., 2020, 22, 100483 CrossRef.
  538. M. Sheng, W. Yimin, Z. Anding, et al., Study on jute fiber reinforced polypropylene (PP) composite, Composite Technologies for 2020 Proceedings of the Fourth Asian–Australasian Conference on Composite Materials (ACCM 4), 2004, pp. 52–56 Search PubMed.
  539. A. A. Singh and S. Palsule, Jute fiber reinforced chemically functionalized polypropylene self-compatibilizing composites by Palsule process, J. Compos. Mater., 2016, 50(9), 1199–1212 CrossRef CAS.
  540. K. Deepak, S. V. P. Vattikuti and B. Venkatesh, Experimental investigation of jute fibre reinforced nano clay composite, Prog. Mater. Sci., 2015, 10, 238–242 CAS.
  541. P. Kumar, M. Tiwari and M. E. Makhata, et al., Effect of rate of loading on jute fibre-reinforced polymer composite, Trans. Indian Inst. Met., 2020, 73, 1573–1577 CrossRef CAS.
  542. I. M. D. Rosa, C. Santulli and F. Sarasini, et al., Post-impact damage characterization of hybrid configurations of jute/glass polyester laminates using acoustic emission and IR thermography, Compos. Sci. Technol., 2009, 69, 1142–1150 CrossRef.
  543. K. Sever, M. Sarikanat and Y. Seki, et al., Surface treatments of jute fabric: The influence of surface characteristics on jute fabrics and mechanical properties of jute/polyester composites, Ind. Crops Prod., 2012, 35(1), 22–30 CrossRef CAS.
  544. D. Ray, N. R. Bose and A. K. Mohanty, et al., Modification of the dynamic damping behaviour of jute/vinylester composites with latex interlayer, Composites, Part B, 2007, 38, 380–385 CrossRef.
  545. S. Shahinur, M. Hasan and Q. Ahsan, Physical and mechanical properties of chemically treated jute fiber reinforced MAHgPP green composites, Appl. Mech. Mater., 2017, 860, 134–139 Search PubMed.
  546. A. Jabbar, J. Militky and J. Wiener, et al., Nanocellulose coated woven jute/green epoxy composites: Characterization of mechanical and dynamic mechanical behavior, Compos. Struct., 2017, 161, 340–349 CrossRef.
  547. Z. F. Albahash and M. N. M. Ansari, Investigation on energy absorption of natural and hybrid fiber under axial static crushing, Compos. Sci. Technol., 2017, 151, 52–61 CrossRef CAS.
  548. S. B. R. Devireddy and S. Biswas, Thermo-physical properties of short banana–jute fiber reinforced epoxy based hybrid composites, Proc. Inst. Mech. Eng., Part L, 2018, 238, 939–951 Search PubMed.
  549. V. Mishra and S. Biswas, Evaluation of three body abrasive wear behavior of bidirectional jute fiber reinforced epoxy composites, J. Ind. Text., 2015, 44(5), 781–797 CrossRef.
  550. P. T. R. Swain and S. Biswas, Influence of fiber surface treatments on physico-mechanical behaviour of jute/epoxy composites impregnated with aluminium oxide filler, J. Compos. Mater., 2017, 51, 3909–3922 CrossRef CAS.
  551. M. K. Gupta, Investigations on jute fibre-reinforced polyester composites: Effect of alkali treatment and poly(lactic acid) coating, J. Ind. Text., 2020, 49, 923–942 CrossRef CAS.
  552. M. K. Singh and S. Zafar, Effect of layering sequence on mechanical properties of woven kenaf/jute fabric hybrid laminated microwave-processed composites, J. Ind. Text., 2022, 51, 2731S–2752S CrossRef CAS.
  553. H. E. Balcioglu, Fracture behaviors of SiC particle filled and jute fiber reinforced natural composites, J. Nat. Fibers, 2022, 19, 2338–2355 CrossRef CAS.
  554. S. S. Chestee, P. Poddar and T. K. Sheel, et al., Short jute fiber reinforced polypropylene composites: Effect of nonhalogenated fire retardants, Adv. Chem., 2017, 2017, 1049513 Search PubMed.
  555. S. Sengupta, Development of jute fabric for jute-polyester bio-composite considering structure-property relationship, J. Nat. Fibers, 2022, 19, 1864–1878 CrossRef CAS.
  556. J. M. Ferreira, C. Capela and J. Manaia, et al., Mechanical properties of woven mat jute/epoxy composites, Mater. Res., 2016, 19(3), 702–710 CrossRef CAS.
  557. R. Hu and J. Lim, Fabrication and mechanical properties of completely biodegradable hemp fibre reinforced polylactic acid composites, J. Compos. Mater., 2007, 41, 1655–1669 CrossRef CAS.
  558. X. Peng, M. Fan and J. Hartley, et al., Properties of natural fibre composites made by pultrusion process, J. Compos. Mater., 2011, 46, 237–246 CrossRef.
  559. H. M. Akil, I. M. De Rosa and C. Santulli, et al., Flexural behaviour of pultruded jute/glass and kenaf/glass hybrid composites monitored using acoustic emission, Mater. Sci. Eng., A, 2010, 527, 2942–2950 CrossRef.
  560. M. H. Zamri, H. M. Akil and A. A. Bakar, et al., Effect of water absorption on pultruded jute/glass fibre-reinforced unsaturated polyester hybrid composites, J. Compos. Mater., 2012, 46, 51–61 CrossRef CAS.
  561. N. Nosbi, H. M. Akil and Z. A. M. Ishak, et al., Degradation of compressive properties of pultruded kenaf fibre reinforced composites after immersion in various solutions, Mater. Des., 2010, 31, 4960–4964 CrossRef CAS.
  562. W. M. Raymond, Handbook of Pultrusion Technology, Chapman and Hall, New York, 1985 Search PubMed.
  563. W. D. Callister Jr, Materials Science, Engineering: an Introduction, John Wiley & Sons, Inc., 6th edn, 2003 Search PubMed.
  564. M. A. Gunning, L. M. Geever and J. A. Killion, et al., The effect of processing conditions for polylactic acid based fibre composites via twin screw extrusion, J. Reinf. Plast. Compos., 2014, 33, 648–662 CrossRef.
  565. S. Shahinur, and M. Hasan, Jute/coir/banana fiber reinforced bio-composites: critical review of design, fabrication, properties and applications, Encycl Renew Sustain Mater,  DOI:10.1016/B978-0-12-803581-8.10987-7.
  566. Y. Yang, M. Murakami and H. Hamada, Molding method, thermal and mechanical properties of jute/PLA injection molding, J. Polym. Environ., 2012, 20, 1124–1133 CrossRef CAS.
  567. M. A. Pinto, V. B. Chalivendra and Y. K. Kim, et al., Evaluation of surface treatment and fabrication methods for jute fibre/epoxy laminar composites, Polym. Compos., 2014, 35, 310–317 CrossRef CAS.
  568. H. F. M. de Queiroz, M. D. Banea and D. K. K. Cavalcanti, Adhesively bonded joints of jute, glass and hybrid jute/glass fibre-reinforced polymer composites for automotive industry, Appl. Adhes. Sci., 2021, 9, 2 CrossRef CAS.
  569. D. Klosterman, R. Chartoff, G. Graves, N. Osborne and B. Priore, Interfacial characteristics of composites fabricated by laminated object manufacturing, Composites, Part A, 1998, 28, 1165–1174 CrossRef.
  570. A. Bourmaud, D. U. Shah and J. Beaugrand, et al., Property changes in plant fibres during the processing of bio-based composites, Ind. Crops Prod., 2020, 154, 112705 CrossRef CAS.
  571. T. Jirawattanasomkul, S. Likitlersuang and N. Wuttiwannasak, et al., Effects of heat treatment on mechanical properties of jute fiber-reinforced polymer composites for concrete confinement, J. Mater. Civ. Eng., 2020, 32(12) DOI:10.1061/(ASCE)MT.1943-5533.0003456.
  572. S. Sathees Kumar, Effect of natural fiber loading on mechanical properties and thermal characteristics of hybrid polyester composites for industrial and construction fields, Fibers Polym., 2020, 21, 1508–1514 CrossRef CAS.
  573. S. Mishra, J. B. Naik and Y. P. Patil, The compatibilising effect of maleic anhydride on swelling and mechanical properties of plant-fibre-reinforced novolac composites, Compos. Sci. Technol., 2000, 60, 1729–1735 CrossRef CAS.
  574. J. Dehury, Processing & characterization of jute/glass fiber reinforced epoxy based hybrid composites, M. Tech. thesis, National Institute of Technology Rourkela, India, 2013 Search PubMed.
  575. S. Singha and V. K. Thakur, Mechanical properties of natural fiber reinforced polymer composites, Bull. Mater. Sci., 2008, 31, 791–799 CrossRef.
  576. A. N. Fraga, E. Frulloni and O. Osa, et al., Relationship between water absorption and dielectric behaviour of natural fibre composite materials, Polym. Test., 2006, 25(2), 181–187 CrossRef CAS.
  577. N. M. Mehta and P. H. Parsania, Fabrication and evaluation of some mechanical and electrical properties of jute-biomass based hybrid composites, J. Appl. Polym. Sci., 2006, 100, 1754–1758 CrossRef CAS.
  578. R. A. Pethrick and D. Hayward, Real time dielectric relaxation studies of dynamic polymeric systems, Prog. Polym. Sci., 2002, 27, 1983–2017 CrossRef CAS.
  579. N. Chand and D. Jain, Effect of sisal fibre orientation on electrical properties of sisal fibre reinforced epoxy composites, Composites, Part A, 2005, 36, 594–602 CrossRef.
  580. E. Boinard and R. A. Pethrick, The influence of thermal history on the dynamic mechanical and dielectric studies of polyetheretherketone exposed to water and brine, Polymer, 2000, 41, 1063–1076 CrossRef CAS.
  581. M. N. A. M. Taib, and N. M. Julkapli, Dimensional stability of natural fiber-based and hybrid composites, in Mechanical and Physical Testing of Biocomposites, Fibre-Reinforced Composites and Hybrid Composites,  DOI:10.1016/B978-0-08-102292-4.00004-7.
  582. V. Tserki, N. E. Zafeiropoulos and F. Simon, et al., A study of the effect of acetylation and propionylation surface treatments on natural fibres, Composites, Part A, 2005, 36, 1110–1118 CrossRef.
  583. Y. Li, Y. W. Mai and L. Ye, Effects of fibre surface treatment on fracture mechanical properties of sisal–fibre composites, Compos. Interfaces, 2005, 12, 141–163 CrossRef CAS.
  584. M. Y. Khalid, Z. U. Arif and M. F. Sheikh, et al., Mechanical characterization of glass and jute fiber-based hybrid composites fabricated through compression molding technique, Int. J. Material Form., 2021, 14, 1085–1095 CrossRef.
  585. G. Kalaprasad, K. Joseph and S. Thomas, Influence of short glass fiber addition on the mechanical properties of sisal reinforced low density polyethylene composites, J. Compos. Mater., 1997, 31, 509–527 CrossRef CAS.
  586. C. M. Praveen Kumar, R. B. Ashok, M. Kumar and C. P. Roopa, Natural nano-fillers materials for the bio-composites: A review, J. Indian Chem. Soc., 2022, 99, 100715 CrossRef.
  587. N. Graupner, F. Sarasini and J. Mussig, Ductile viscose fibres and stiff basalt fibres for composite applications-An overview and the potential of hybridization, Composites, Part B, 2020, 194, 108041 CrossRef CAS.
  588. M. A. Aziz, P. Paramasivam and S. L. Lee, Prospects for natural fibre reinforced concretes in construction, Int. J. Cem. Compos. Lightweight Concr., 1981, 3, 123–132 CrossRef.
  589. R. Jarabo, E. Fuente and M. C. Monte, et al., Use of cellulose fibres from hemp core in fibre-cement production, on flocculation, retention, drainage and product properties effect, Ind. Crops Prod., 2012, 39, 89–96 CrossRef CAS.
  590. M. Ganesan and G. Nallathambi, Alkali-treated coir fibre-pith composite for waste water treatment, J. Ind. Text., 2022, 51, 749S–768S CrossRef CAS.
  591. M. Z. Rong, M. Q. Zhang and Y. Liu, et al., The effect of fibre treatment on the mechanical properties of unidirectional sisal-reinforced epoxy composites, Compos. Sci. Technol., 2001, 61, 1437–1447 CrossRef CAS.
  592. S. Das, A. K. Singha and A. Chaudhuri, et al., Lengthwise jute fibre properties variation and its effect on jute–polyester composite, J. Text. Inst., 2019, 110, 1695–1702 CrossRef CAS.
  593. V. Mishra and S. Biswas, Physical and mechanical properties of bi-directional jute fibre epoxy composites, Procedia Eng., 2013, 51, 561–566 CrossRef CAS.
  594. A. A. Shaikh and S. A. Channiwala, Experimental and analytical investigation of jute polyester composite for long continuous fibre reinforcement, J. Reinf. Plast. Compos., 2006, 25, 863–873 CrossRef CAS.
  595. Y. Karaduman and L. Onal, Dynamic mechanical and thermal properties of enzyme-treated jute/polyester composites, J. Compos. Mater., 2013, 47, 2361–2370 CrossRef.
  596. H. Katogi, Y. Shimamura, T. Tohgo, et al., Fatigue behavior of unidirectional jute spun yarn, in Proceedings of 18th Int Conf Compos Mater, 2011 Search PubMed.
  597. D. U. Shah, P. J. Schubel and M. J. Clifford, et al., Fatigue life evaluation of aligned plant fibre composites through S-N curves and constant-life diagrams, Compos. Sci. Technol., 2013, 74, 139–149 CrossRef CAS.
  598. M. Jacob, S. Thomas and K. T. Varughese, Mechanical properties of sisal/oil palm hybrid fibre reinforced natural rubber composites, Compos. Sci. Technol., 2004, 64, 955–965 CrossRef CAS.
  599. N. Sombatsompop and K. Chaochanchaikul, Effect of moisture content on mechanical properties, thermal and structural stability and extrudate texture of poly(vinyl chloride)/wood sawdust composites, Polym. Int., 2004, 53, 1210–1218 CrossRef CAS.
  600. H. N. Dhakal, Z. Y. Zhang and N. Bennett, Influence of fibre treatment and glass fibre hybridisation on thermal degradation and surface energy characteristics of hemp/unsaturated polyester composites, Composites, Part B, 2012, 43, 2757–2761 CrossRef CAS.
  601. U. Z. Haydar, A. H. Khan and A. Hossain, et al., Mechanical and electrical properties of jute fabrics reinforced polyethylene/polypropylene composites: Role of gamma radiation, Polym.-Plast. Technol. Eng., 2009, 48, 760–766 CrossRef.
  602. M. A. Khan, R. A. Khan and G. Haydaruzzaman, et al., Study on the physico-mechanical properties of starch-treated jute yarn-reinforced polypropylene composites: effect of gamma radiation, Polym.-Plast. Technol. Eng., 2009, 48, 542–548 CrossRef CAS.
  603. M. Ramakrishna, V. Kumar and N. S. Yuvraj, Recent development in natural fiber reinforced polypropylene composites, J. Reinf. Plast. Compos., 2009, 28, 1169–1189 CrossRef.
  604. H. S. Yang, H. J. Kim and B. J. Lee, et al., Water absorption behavior and mechanical properties of lignocellulosic filler-polyolefin bio-composites, Compos. Struct., 2006, 72, 429–437 CrossRef.
  605. M. S. Jamil, I. Ahmed and I. Abdullah, Effects of rice husk filler on the mechanical and thermal properties of liquid natural rubber compatibilized high-density polyethylene/natural rubber blends, J. Polym. Res., 2006, 13, 315–321 CrossRef CAS.
  606. X. Li, S. Panigrahi and L. G. Tabil, A study on flax fiber-reinforced polyethylene biocomposites, Am. Soc. Agric. Biol. Eng., 2009, 25(4), 525–531 CAS.
  607. J. P. Lopez, S. Boufi and N. E. El Mansouri, et al., PP composites based on mechanical pulp, deinked newspaper and jute strands: A comparative study, Composites, Part B, 2012, 43, 3453–3461 CrossRef CAS.
  608. B. Singh, M. Gupta and A. Verma, The durability of jute fibre-reinforced phenolic composites, Compos. Sci. Technol., 2000, 60, 581–589 CrossRef CAS.
  609. M. El Messiry, S. El-Tarfawy and R. El Deeb, Study pultruded jute fabric effect on the cementitious thin composites mechanical properties with low fiber volume fraction, Alexandria Eng. J., 2017, 56, 415–421 CrossRef.
  610. A. K. Bledzki and A. Jaszkiewicz, Mechanical performance of biocomposites based on PLA and PHBV reinforced with natural fibres–a comparative study to PP, Compos. Sci. Technol., 2010, 70, 1687–1696 CrossRef CAS.
  611. S. Bruzaud and A. Bourmaud, Thermal degradation and (nano) mechanical behaviour of layered silicate reinforced poly(3-hydroxybutyrate-co-3-hydroxyvalerate) nanocomposites, Polym. Test., 2007, 26, 652–659 CrossRef CAS.
  612. M. J. John and R. D. Anandjiwala, Chemical modification of flax reinforced polypropylene composites, Composites, Part A, 2009, 40(4), 442–448 CrossRef.
  613. S. Chakraborty, S. P. Kundu and A. Roy, et al., Polymer modified jute fibre as reinforcing agent controlling the physical and mechanical characteristics of cement mortar, Constr. Build. Mater., 2013, 49, 214–222 CrossRef CAS.
  614. T. Lindstrom, and L. Wagberg, An overview of some possibilities to modify fibre surfaces for tailoring composite interfaces, in Proceedings of the 23rd International Symposium on Materials Science, Denmark, 2002, p. 35 Search PubMed.
  615. W. Webo, M. Maringa and L. Masu, The combined effect of mercerisation, silane treatment and acid hydrolysis on the mechanical properties of sisal fibre/epoxy resin composites, MRS Adv., 2020, 5, 1225–1233 CrossRef CAS.
  616. E. K. Gamstedt and R. Talreja, Fatigue damage mechanisms in unidirectional carbon-fibre-reinforced plastics, J. Mater. Sci., 1999, 34, 2535–2546 CrossRef CAS.
  617. S. A. Hitchen, S. L. Ogin and P. A. Smith, Effect of fibre length on fatigue of short carbon fibre/epoxy composite, Compos, 1995, 26(4), 303–310 CrossRef CAS.
  618. B. P. Jang, W. Kowbel and B. Z. Jang, Impact behaviour and repeated impact testing of polymer composites, Compos. Sci. Technol., 1992, 44, 107–118 CrossRef CAS.
  619. B. P. Jang, C. T. Huang and C. Y. Hsieh, et al., Repeated impact failure of continuous fibre reinforced thermoplastic and thermoset composites, J. Compos. Mater., 1991, 25, 1171–1203 CrossRef CAS.
  620. M. Quaresimin, Multiaxial fatigue testing of composites: from the pioneers to future directions, Strain, 2015, 51, 16–29 CrossRef.
  621. M. Quaresimin, L. Susmel and R. Talreja, Fatigue behaviour and life assessment of composite laminates under multiaxial loadings, Int. J. Fatigue, 2010, 32, 2–16 CrossRef CAS.
  622. S. Amijima, T. Fujii and M. Hamaguchi, Static and fatigue tests of a woven glass fabric composite under biaxial tension-torsion loading, Compos, 1991, 22, 281–289 CrossRef CAS.
  623. A. Bentur, and S. Mindess, Fibre Reinforced Cementitious Composites, CRC Press, USA and Canada, 2nd edn, 2006 Search PubMed.
  624. P. N. Balaguru, and S. P. Shah, Fibre-reinforced Cement Composites, McGraw Hill Inc., UK, 1992 Search PubMed.
  625. S. I. Haruna, H. Zhu, Y. E. Ibrahim, J. Shao, M. Adamu and A. I. B. Farouk, Experimental and statistical analysis of U-shaped polyurethane-based polymer concrete under static and impact loads as a repair material, Buildings, 2022, 12(11), 1986 CrossRef.
  626. R. Boujmal, H. Essabir and S. Nekhlaoui, et al., Composite from polypropylene and henna fiber: Structural, mechanical and thermal properties, J. Biobased Mater. Bioenergy, 2014, 8, 246–252 CrossRef CAS.
  627. J. Raghavan and M. Meshii, Creep rupture of polymer composites, Compos. Sci. Technol., 1997, 57(4), 375–388 CrossRef CAS.
  628. B. A. Acha, M. M. Reboredo and N. E. Marcovich, Creep and dynamic mechanical behavior of PPejute composites: effect of the interfacial adhesion, Composites, Part A, 2007, 38(6), 1507–1516 CrossRef.
  629. A. K. Bledzki and O. Faruk, Creep and impact properties of wood fibre polypropylene composites: influence of temperature and moisture content, Compos. Sci. Technol., 2004, 64(5), 693–700 CrossRef CAS.
  630. N. E. Marcovich and M. A. Villar, Thermal and mechanical characterization of linear low-density polyethylene/wood flour composites, J. Appl. Polym. Sci., 2003, 90(10), 2775–2784 CrossRef CAS.
  631. M. A. Hidalgo-Salazar, J. H. Mina and P. J. Herrera-Franco, The effect of interfacial adhesion on the creep behaviour of LDPE-Al-Fique composite materials, Composites, Part B, 2013, 55, 345–351 CrossRef CAS.
  632. Y. Xu, Q. Wu and Y. Lei, et al., Creep behavior of bagasse fibre reinforced polymer composites, Bioresour. Technol., 2010, 101(9), 3280–3286 CrossRef CAS PubMed.
  633. A. J. Nunez, N. E. Marcovich and M. I. Aranguren, Short-term and long-term creep of polypropylene–woodflour composites, Polym. Eng. Sci., 2004, 44(8), 1594–1603 CrossRef CAS.
  634. G. E. Dieter, Mechanical Metallurgy, McGraw-Hill, Singapore, 1988, pp. 348–368 Search PubMed.
  635. J. Kim, C. Baillie and J. Poh, et al., Fracture toughness of CFRP with modified epoxy resin matrices, Compos. Sci. Technol., 1992, 43, 283–297 CrossRef CAS.
  636. R. C. Petersen, J. E. Lemons and M. S. McCracken, Fracture toughness micromechanics by energy methods with a photocure fibre-reinforced composite, Polym. Compos., 2007, 28(3), 311–322 CrossRef CAS PubMed.
  637. H. Zarei, M. Kroger and H. Albertsen, An experimental and numerical crashworthiness investigation of thermoplastic composite crash boxes, Compos. Struct., 2008, 85(3), 245–257 CrossRef.
  638. S. Agrawal, K. K. Singh and P. Sarkar, Impact damage on fibre-reinforced polymer matrix composite – a review, J. Compos. Mater., 2013, 48(3), 317–332 CrossRef.
  639. K. Oksman and C. Clemons, Mechanical properties and morphology of impact modified polypropylene – wood flour composites, J. Appl. Polym. Sci., 1998, 67, 1503–1513 CrossRef CAS.
  640. M. Bera, R. Alagirusamy and A. Das, A study on interfacial properties of jute–PP composites, J. Reinf. Plast. Compos., 2010, 29(20), 3155–3161 CrossRef CAS.
  641. R. B. Adusumalli, M. Reifferscheid and H. K. Weber, et al., Shear strength of the lyocell fibre/polymer matrix interface evaluated with the microbond technique, J. Compos. Mater., 2010, 46(3), 359–367 CrossRef.
  642. B. Wang, K. Li, C. Zhang, B. Hu, Y. Huang, T. Wang and Q. Lu, Influence of cellulose ultrastructure on the catalytic pyrolysis for selective production of levoglucosenone, Ind. Crops Prod., 2023, 192, 116072 CrossRef CAS.
  643. M. E. Eugenio, M. Ruiz-Montoya, R. Martín-Sampedro, D. Ibarra and M. J. Díaz, Influence of cellulose characteristics on pyrolysis suitability, Processes, 2021, 9(9), 1584 CrossRef CAS.
  644. C. Li, Y. Sun and D. Dong, et al., Co-pyrolysis of cellulose/lignin and sawdust: Influence of secondary condensation of the volatiles on characteristics of biochar, Energy, 2021, 226, 120442 CrossRef CAS.
  645. E. Sinha and S. K. Rout, Influence of fibre surface treatment on structural, thermal and mechanical properties of jute and its composite, Bull. Mater. Sci., 2009, 32(1), 65–76 CrossRef CAS.
  646. C. Vajrasthira, T. Amornsakchai and B. Limcharoen, Fibre–matrix interactions in aramid short fibre reinforced thermoplastic polyurethane composites, J. Appl. Polym. Sci., 2003, 87, 1059–1067 CrossRef CAS.
  647. J. Rout, S. S. Tripathy and M. Mishra, et al., The influence of fibre surface modification on mechanical properties of coir-polyester composites, Polym. Compos., 2002, 22, 468–472 CrossRef.
  648. S. Mohanty, S. K. Verma and S. S. Tripathy, et al., Effect of surface treatment on the interface of HDPE-jute composites, Int. J. Plast. Technol., 2003, 6, 75–81 Search PubMed.
  649. D. Bikiaris, P. Matzinos and J. Prinos, et al., Use of silanes & copolymers as adhesion promoters in glass fibre/polyethylene composites, J. Appl. Polym. Sci., 2001, 80, 2877–2888 CrossRef CAS.
  650. M. Botev, H. Betchev and D. Bikiaris, et al., Mechanical properties & viscoelastic behavior of basalt fibre reinforced polypropylene, J. Appl. Polym. Sci., 1999, 74, 523–531 CrossRef CAS.
  651. E. M. Woo and J. C. Seferis, Viscoelastic characterization of high performance epoxy matrix composites, Polym. Compos., 1991, 12(4), 273–280 CrossRef CAS.
  652. O. Beaudoin, A. Bergeret and J. C. Quantin, et al., Factors influencing viscoelastic properties of poly (butylene terephthalate) reinforced short glass fibres, Polym. Compos., 2002, 23(4), 87–94 CrossRef CAS.
  653. M. A. L. Manchado, J. Biagiotti and J. M. Kenny, Comparative study of the effects of different fibres on the processing and properties of ternary composites based on PP-EPDM blends, Polym. Compos., 2002, 23(5), 779–789 CrossRef.
  654. M. Ashida and T. Noguchi, Dynamic moduli for short fibre-CR composites, J. Appl. Polym. Sci., 1984, 29, 661–670 CrossRef CAS.
  655. K. SudarshanRao, Dynamic mechanical behavior of unfilled and graphite filled carbon epoxy composites, IOP Conf. Ser.: Mater. Sci. Eng., 2021, 1126, 012033 CAS.
  656. J. Kubat, M. Rigdahl and M. Welander, Characterizations of interfacial interactions in high density polyethylene filled with glass spheres using dynamic mechanical analysis, J. Appl. Polym. Sci., 1990, 39, 1527–1539 CrossRef CAS.
  657. M. S. Sohn, K. S. Kim and S. H. Hong, et al., Dynamic mechanical properties of particle reinforced EPDM composites, J. Appl. Polym. Sci., 2003, 87, 1595–1601 CrossRef CAS.
  658. S. Mohanty, S. K. Verma and S. K. Nayak, et al., Influence of fiber treatment on the performance of sisal–polypropylene composites, J. Appl. Polym. Sci., 2004, 94, 1336–1345 CrossRef CAS.
  659. E. Alfthan, A. de Ruvo and W. Brown, Glass transition temperature of oligosachcharides, J. Reinf. Plast. Compos., 1993, 12, 139–155 CrossRef.
  660. S. Joseph, M. S. Sreekala and S. Thomas, Viscoelastic properties of oil palm fibre reinforced phenol formaldehyde composites, Int. J. Plast. Technol., 2002, 5(1), 28–35 Search PubMed.
  661. T. Aurich and G. Mennig, Characterization of injection moulded flax fibre reinforced polypropylene, Int. J. Plast. Technol., 2002, 5(1), 9–14 Search PubMed.
  662. A. K. Rana, B. C. Mitra and A. N. Banerjee, Short jute fibre-reinforced polypropylene composites: dynamic mechanical study, J. Appl. Polym. Sci., 1999, 71, 531–539 CrossRef CAS.
  663. P. Ghosh, N. R. Bose and B. C. Mitra, et al., Dynamic mechanical analysis of FRP composites based on different fibre reinforcements & epoxy resin as the matrix material, J. Appl. Polym. Sci., 1997, 62, 2467–2472 CrossRef.
  664. K. N. E. Verghese, R. E. Jensen and J. J. Lesko, et al., Effects of molecular relaxation behavior on sized carbon fibre–vinylester matrix composites, Polymer, 2001, 42, 1633–1645 CrossRef CAS.
  665. N. S. Hon David, and N. Shiraishi, Wood & Cellulose Chemistry, Marcel Dekker Inc., New York & Basel, 1991 Search PubMed.
  666. L. A. Pothan, Z. Oommen and S. Thomas, Dynamic mechanical analysis of banana fibre reinforced polyester composites, Compos. Sci. Technol., 2003, 63(2), 283–293 CrossRef CAS.
  667. A. Afaghi-Khatibi and Y.-W. Mai, Characterisation of fibre/matrix interfacial degradation under cyclic fatigue loading using dynamic mechanical analysis, Composites, Part A, 2002, 33(11), 1585–1592 CrossRef.
  668. I. Tzanakis, M. Hadfield and B. Thomas, et al., Future perspectives of sustainable tribology, Renewable Sustainable Energy Rev., 2012, 16, 4126–4140 CrossRef CAS.
  669. G. Fang, Z. Zhaozhu and Z. Huijuan, et al., Tribological behavior of spun kevlar fabric composites filled with fluorinated compounds, Tribol. Int., 2010, 43, 1466–1471 CrossRef.
  670. P. K. Bajpai, I. Singh and J. Madaan, Tribological behavior of natural fibre reinforced PLA composites, Wear, 2013, 297, 829–840 CrossRef CAS.
  671. J. Wu and X. H. Cheng, The tribological properties of kevlar pulp reinforced epoxy composites under dry sliding and water lubricated condition, Wear, 2006, 261, 1293–1297 CrossRef CAS.
  672. P. L. Menezes, P. K. Rohatgi and M. R. Lovell, Studies on the tribological behavior of natural fibre reinforced polymer composites, Green Tribology, Green Energy and Technology, Springer, Verlag, Berlin Heidelberg, 2012,  DOI:10.1007/978-3-642-23681-5_12.
  673. M. Punyapriya, Statistical analysis for the abrasive wear behavior of bagasse fibre reinforced polymer composite, Int. J. Appl. Res. Mech. Eng., 2012, 2, 154–158 Search PubMed.
  674. B. Witold, K. Vera and V. Domagoj, et al., Tribology of polymers and polymer-based composites, J. Mater. Educ., 2010, 32, 273–290 Search PubMed.
  675. A. Patnaik, A. Satapathy and S. Biswas, Investigations on three-body abrasive wear and mechanical properties of particulate filled glass epoxy composites, Malays. Polym. J., 2010, 5, 37–48 Search PubMed.
  676. P. Rajasekhar, G. Ganesan and C. Senthilkumar, Studies on tribological behavior of polyamide filled jute fibre-nano-ZnO hybrid composites, Procedia Eng., 2014, 97, 2099–2109 CrossRef CAS.
  677. M. Fahim, and N. Chand, Tribology of FRP Composites, Allied Publishers, Delhi, 2000, p. 47 Search PubMed.
  678. E. Lackey, G. V. James, K. Inamdar, and B. Hancock, Composites 2004 Convention and Trade Show American Composites Manufacturers Association, Tampa, Florida, USA, 2004, pp. 1–9 Search PubMed.
  679. M. S. Sreekala, M. G. Kumaran and S. Thomas, Water sorption in oil palm fibre reinforced phenol formaldehyde composites, Composites, Part A, 2002, 33, 763–777 CrossRef.
  680. M. S. Sreekala and S. Thomas, Effect of fibre surface modification on water-sorption characteristics of oil palm fibres, Compos. Sci. Technol., 2003, 63, 861–869 CrossRef CAS.
  681. A. Valadez-Gonzalez, J. M. Cervantes-Uc and R. Olayo, et al., Effect of fibre surface treatment on the fibre–matrix bond strength of natural fibre reinforced composites, Composites, Part B, 1999, 30, 309–320 CrossRef.
  682. G. Mehta, L. T. Drzal and A. K. Mohanty, et al., Effect of fibre surface treatment on the properties of biocomposites from nonwoven industrial hemp fibre mats and unsaturated polyester resin, J. Appl. Polym. Sci., 2006, 99, 1055–1068 CrossRef CAS.
  683. A. Stamboulis, C. A. Baillie and S. K. Garkhail, et al., Environmental durability of flax fibres and their composites based on polypropylene matrix, Appl. Compos. Mater., 2000, 7, 273–294 CrossRef CAS.
  684. M. M. Thwe and K. Liao, Effects of environmental aging on the mechanical properties of bamboo–glass fibre reinforced polymer matrix hybrid composites, Composites, Part A, 2002, 33, 43–52 CrossRef.
  685. L. A. Pothan and S. Thomas, Effect of hybridization and chemical modification on the water-absorption behavior of banana fibre-reinforced polyester composites, J. Appl. Polym. Sci., 2004, 91, 3856–3865 CrossRef CAS.
  686. K. Jarukumjorn and N. Suppakarn, Effect of glass fibre hybridization on properties of sisal fibre–polypropylene composites, Composites, Part B, 2009, 40, 623–627 CrossRef.
  687. G. S. Springer, Environmental Effects on Composite Materials, Tecnomic, Lancaster (PA), 1988 Search PubMed.
  688. A. Saaidia, A. Belaadi and M. Boumaaza, et al., Effect of water absorption on the behavior of jute and sisal fiber biocomposites at different lengths: ANN and RSM modeling, J. Nat. Fibers, 2023, 20, 2140326 CrossRef.
  689. A. Espert, F. Vilaplana and S. Karlsson, Comparison of water absorption in natural cellulosic fibres from wood and one-year crops in polypropylene composites and its influence on their mechanical properties, Composites, Part A, 2004, 35, 1267–1276 CrossRef.
  690. J. Zhou and J. P. Lucas, The effects of a water environment on anomalous absorption behavior in graphite/epoxy composites, Compos. Sci. Technol., 1995, 53, 57–64 CrossRef CAS.
  691. ASTM-D570-98: Standard Test Method for Water Absorption of Plastics, American Society for Testing and Materials, 2005 Search PubMed.
  692. Y. J. Weitsman, Effects of fluids on polymeric composites – a review, in Comprehensive Composite Materials, ed. A. Kelly, and C. Zweben, Elsevier, 2000, vol. 2, pp. 369–401 Search PubMed.
  693. H. B. Daly, H. B. Brahim and N. Hfaied, et al., Investigation of water absorption in pultruded composites containing fillers and low profile additives, Polym. Compos., 2007, 28, 355–364 CrossRef.
  694. H. J. Russell, Environmental Effects on Engineered Materials, Marcel Dekker, Inc., 2001 Search PubMed.
  695. M. Rizal, A. Z. Mubarak and A. Razali, et al., Free vibration characteristics of jute fibre reinforced composite for the determination of material properties: Numerical and experimental studies, AIP Conf. Proc., 2019, 2187, 050020 CrossRef CAS.
  696. Z. Mahboob, Y. Chemisky, F. Meraghni and H. Bougherara, Mesoscale modelling of tensile response and damage evolution in natural fibre reinforced laminates, Composites, Part B, 2017, 119, 168–183 CrossRef CAS.
  697. J. Andersons, J. Modniks and E. Sparnins, Modeling the nonlinear deformation of flax-fiber-reinforced polymer matrix laminates in active loading, J. Reinf. Plast. Compos., 2015, 34(3), 248–256 CrossRef CAS.
  698. S. M. Panamoottil, R. Das and K. Jayaraman, Towards a multiscale model for flax composites from behaviour of fibre and fibre/polymer interface, J. Compos. Mater., 2016, 51(6), 859–873 CrossRef.
  699. C. Poilane, Z. E. Cherif, F. Richard, A. Vivet, B. Ben Doudou and J. Chen, Polymer reinforced by flax fibres as a viscoelastoplastic material, Compos. Struct., 2014, 112, 100–112 CrossRef.
  700. J. Sliseris, L. Yan and B. Kasal, Numerical modelling of flax short fibre reinforced and flax fibre fabric reinforced polymer composites, Composites, Part B, 2016, 89, 143–154 CrossRef CAS.
  701. M. R. Mansor, M. S. Sapuan, E. S. Zainudin, et al., Life cycle assessment of natural fiber polymer composites, Agricultural Biomass Based Potential Materials, ed. K. R. Hakeem, et al.,  DOI:10.1007/978-3-319-13847-3_6.
  702. D. I. Nagy, I. V. Rowe and R. Kahhat, et al., Life cycle assessment of bagasse fiber reinforced biocomposites, Sci. Total Environ., 2020, 720, 137586 CrossRef PubMed.
  703. A. D. L. Rosa, G. Recca and J. Summerscales, et al., Bio-based versus traditional polymer composites. A life cycle assessment perspective, J. Cleaner Prod., 2014, 74, 135–144 CrossRef.
  704. I. K. Kookos, A. Koutinas and A. Vlysidis, Life cycle assessment of bioprocessing schemes for ploy(3-hydroxybutyrate) production using soybean oil and sucrose as carbon sources, Resour., Conserv. Recycl., 2019, 141, 317–328 CrossRef.
  705. C. Alves, A. P. S. Dias and A. C. Diogo, et al., Eco-composite: the effects of the jute fibre treatments on the mechanical and environmental performance of the composite materials, J. Compos. Mater., 2010, 45(5), 573–589 CrossRef.
  706. S. D. Pandita, X. Yuan and M. A. Manan, et al., Evaluation of jute/glass hybrid composite sandwich: Water resistance, impact properties and life cycle assessment, J. Reinf. Plast. Compos., 2014, 33(1), 14–25 CrossRef.
  707. Y. S. Song, J. R. Youn and T. G. Gutowski, Life cycle energy analysis of fibre reinforced composites, Composites, Part A, 2009, 40, 1257–1265 CrossRef.
  708. Y. Liu, R. Tang and J. Yu, et al., Investigation of interfacial structure of coupling agent treated fillers by fourier transform infrared spectroscopy and attenuated total reflection-FTIR spectroscopy, Polym. Compos., 2002, 23(1), 28–33 CrossRef CAS.
  709. P. Ganan, R. Zulunga and A. Restrepo, et al., Plantain fibre bundles isolated from colombian agro-industrial restudies, Bioresour. Technol., 2008, 99, 486–491 CrossRef CAS PubMed.
  710. M. M. Haque, M. Hasan and M. S. Islam, et al., Physico–mechanical properties of chemically treated palm and coir fibre reinforced polypropylene composites, Bioresour. Technol., 2009, 100, 4903–4906 CrossRef CAS PubMed.
  711. A. I. Vogel, Text Book of Practical Organic Chemistry, New Delhi, Orient Longmans, 3rd edn, 1985, p. 1039 Search PubMed.
  712. M. Bass and V. Freger, Facile evaluation of coating thickness on membranes using ATR-FTIR, J. Membr. Sci., 2015, 492, 348–354 CrossRef CAS.
  713. H. Wu, C. Silva and Y. Yu, et al., Hydrophobic functionalization of jute fabrics by enzymatic-assisted grafting of vinyl copolymers, New J. Chem., 2017, 41, 3773–3780 RSC.
  714. M. Troedec, D. Sedan and C. Peyratout, et al., Influence of various chemical treatments on the composition and structure of hemp fibres, Composites, Part A, 2008, 39, 514–522 CrossRef.
  715. J. Szopa, M. W. Kwiatkowska and A. Kulma, et al., Chemical composition and molecular structure of fibres from transgenic flax producing polyhydroxybutyrate, and mechanical properties and platelet aggregation of composite materials containing these fibres, Compos. Sci. Technol., 2009, 69, 2438–2446 CrossRef CAS.
  716. A. Devis, and D. Sims, Weathering of Plastics, Applied Science Publishers, UK, 1983 Search PubMed.
  717. T. L. Jose, A. Joseph and M. Skrifvars, et al., Thermal and crystallisation behaviour of cotton-polypropylene commingled composite systems, Polym. Compos., 2010, 31(8), 1487–1494 CrossRef.
  718. T. B. Yallew, P. Kumar and I. Singh, A study about hole making in woven jute fabric reinforced polymer composites, Proc. Inst. Mech. Eng., Part L, 2016, 230, 888–898 CAS.
  719. F. Chegdani and M. E. Mansori, New multiscale approach for machining analysis of natural fiber reinforced bio-composites, J. Manuf. Sci. Eng. Trans. ASME, 2019, 141, 1–24 Search PubMed.
  720. B. V. Kavad, A. B. Pandey and M. V. Tadavi, et al., A review paper on effects of drilling on glass fiber reinforced plastic, Proc. Technol., 2014, 14, 457–464 CrossRef.
  721. M. M. A. Nassar, R. Arunachalam and K. I. Alzebdeh, Machinability of natural fiber reinforced composites: a review, Int. J. Adv. Manuf. Technol., 2017, 88, 2985–3004 CrossRef.
  722. R. Vinayagamoorthy and N. Rajeswari, Analysis of cutting forces during milling of natural fibered composites using fuzzy logic, Int. J. Compos. Mater. Manuf., 2012, 2, 15–21 Search PubMed.
  723. K. Balasubramanian, M. T. H. Sultan and F. Cardona, et al., Machining analysis of natural fiber reinforced composites using fuzzy logic, IOP Conf. Ser.: Mater. Sci. Eng., 2016, 152, 012051 Search PubMed.
  724. B. R. Sankar, P. Umamaheswarrao and V. Srinivasalu, et al., Optimization of the milling process on jute polyester composite using taguchi based grey relational analysis coupled with principle component analysis, Mater. Today: Proc., 2015, 2, 2522–2531 CAS.
  725. B. V. Ramnath, S. Sharavanan and J. Jayakrishnan, Optimization of process parameters in drilling of fibre hybrid composites using Taguchi and grey relational analysis, IOP Conf. Ser.: Mater. Sci. Eng., 2017, 183, 012003 Search PubMed.
  726. Y. H. Celik and M. S. Alp, Determination of milling performance of jute and flax fiber reinforced composites, J. Nat. Fibers, 2022, 19, 782–796 CrossRef CAS.
  727. V. Fiore, Natural fibres and their composites, Polymers, 2020, 12, 2380 CrossRef CAS PubMed.
  728. F. Namvar, M. Jawaid and P. M. Tahir, et al., Potential use of plant fibres and their composites for biomedical applications, BioResources, 2014, 9(3), 5688–5706 CrossRef.
  729. A. K. Mohanty, S. Vivekanandhan and J.-M. Pin, et al., Composites from renewable and sustainable resources: Challenges and innovations, Science, 2018, 362, 536–542 CrossRef CAS PubMed.
  730. Y. Y. Khine and M. H. Stenzel, Surface modified cellulose nanomaterials: a source of non-spherical nanoparticles for drug delivery, Mater. Horiz., 2020, 7, 1727–1758 RSC.
  731. S. Sengupta, P. Ghosh and I. Mustafa, Properties of poly-vinyl alcohol bonded jute (Corchorus olitorius) nonwoven fabric and its performance as disposable carry bag, J. Nat. Fibers, 2022, 19, 2034–2052 CrossRef CAS.
  732. B. Ravi Kumar, C. H. N. Srimannarayana and K. A. Krishnan, et al., Experimental evolution on comparative mechanical properties of jute-flax fibre reinforced composite structures, Struct. Eng., Mech., 2020, 74(4), 515–520 Search PubMed.
  733. A. S. Herrmann, J. Nickel and U. Riedel, Construction materials based upon biologically renewable resources – from components to finished parts, Polym. Degrad. Stab., 1998, 59, 251–261 CrossRef CAS.
  734. M. A. Semsarzadesh, Fiber matrix interactions in jute reinforced polyester resin, Polym. Compos., 1986, 7, 23–25 CrossRef.
  735. R. Sun, J. Fang and A. Goodwin, et al., Fractionation and characterization of polysaccharides from abaca fibre, Carbohydr. Polym., 1998, 37, 351–359 CrossRef CAS.
  736. K. G. Satyanarayana, K. Sukumaran and P. S. Mukherjee, et al., Natural fiber–polymer composite, Cem. Compos., 1990, 12, 117–136 CrossRef CAS.
  737. S. Hattalli, A. Benaboura and A. Castellan, Adding value to Alfa grass (Stipa tenacissima L.) soda lignin as phenolic resins, Polym. Degrad. Stab., 2002, 75, 259–264 CrossRef.
  738. V. Vilay, M. Mariatti and R. M. Taib, et al., Effect of fibre surface treatment and fibre loading on the properties of bagasse fibre reinforced unsaturated polyester composites, Compos. Sci. Technol., 2008, 68, 631–638 CrossRef CAS.
  739. S. Sreenivasulu and A. C. Reddy, Mechanical properties evaluation of bamboo fibre reinforced composite materials, Int. J. Engine Res., 2014, 3(1), 187–194 Search PubMed.
  740. W. Hoareau, W. Trindade and B. Siegmund, et al., Sugar cane bagasse and curaua lignins oxidized by chlorine dioxide and reacted with furfuryl alcohol: Characterization and stability, Polym. Degrad. Stab., 2004, 86, 567–576 CrossRef CAS.
  741. T. P. Sathishkumar and P. Navaneethakrishnan, Tensile and flexural properties of snake grass natural fiber reinforced isophthallic polyester composites, Compos. Sci. Technol., 2012, 72, 1183–1190 CrossRef CAS.
  742. L. Bacci, S. Baronti and S. Predieri, et al., Fibre yield and quality of fibre nettle (Urtica dioica L.) cultivated in Italy, Ind. Crops Prod., 2009, 29, 480–484 CrossRef.
  743. R. Swamy, G. M. Kumar and Y. Vrushabhendrappa, et al., Study of areca reinforced phenol formaldehyde composites, J. Reinf. Plast. Compos., 2004, 23, 1373–1382 CrossRef CAS.
  744. J. D. Almeida, R. Aquino and S. Monteiro, Tensile mechanical properties, morphological aspects and chemical characterization of piassava (Attalea funifera) fibres, Composites, Part A, 2006, 37, 1473–1479 CrossRef.
  745. M.-p. Ho, H. Wang and J.-H. Lee, et al., Critical factors on manufacturing processes of natural fiber composites, Composites, Part B, 2011, 43(8), 3549–3562 CrossRef.

This journal is © The Royal Society of Chemistry 2024