Open access peer-reviewed chapter

Extraction, Applications and Characterization of Plant Fibers

Written By

Richard Ntenga, Saidjo Saidjo, Annie Wakata, Pagore Djoda, Martin Tango and Etienne Mfoumou

Submitted: 23 November 2021 Reviewed: 07 February 2022 Published: 21 April 2022

DOI: 10.5772/intechopen.103093

From the Edited Volume

Natural Fiber

Edited by Han-Yong Jeon

Chapter metrics overview

784 Chapter Downloads

View Full Metrics

Abstract

During the second half of the twentieth century, industrial and scientific interests in plant fibers (PFs) have resulted in their resounding comeback as engineering materials. This chapter is concerned with the characterization of PF materials. Good knowledge of the properties of these materials is essential for safe design of the related structures. Bast fibers that are collected from the phloem surrounding the stem of certain dicotyledonous plants, for instance, are among the most used, owing to their higher tensile strength. However, for an optimum utilization of PFs, a relevant assessment of their physico-chemical and mechanical properties is very crucial. As it is now well established, PFs’ properties are largely influenced by their hierarchic composite microstructure and their viscoelastic behavior. This book chapter focuses on the presentation of various experimental approaches used to characterize the elastic and viscoelastic behaviors of plant fibers. Consideration of their blending in sheet form and relevant mechanical properties will also be of interest.

Keywords

  • plant fibers
  • mechanical testing
  • viscoelastic properties
  • cell wall
  • mechanical properties
  • inverse methods

1. Introduction

In the 1830s, the use of plant resources such as flax, hemp and others was widespread as their fibers were in high demand by the textile, the paper and sailing industries. These plants were grown over large areas for exploitation. However, with the progress of science and technology (loom, steam engine, development of cotton harvesting and processing technique and others), materials such as metals, ceramics, glass, polymers, stones and concrete were preferred to plant resources. In 1900, fiber plants experienced their lowest implantation in terms of surface area. Indeed, the rise of new materials has greatly contributed to the improvement of human living conditions through the construction of more robust and sophisticated habitats, the development of the automotive, railway, aeronautics, textile industries, etc. Subsequently, in a concern for economy, lightness and performance, the development of composite materials was born during the 1930s.

The industrial use of plant fibers began in the early twentieth century with the manufacturing of aircraft seats, fuel tanks or other electronic boxes in plant fibers reinforced polymer materials. The need for securing constructions or structures that are made up of these materials inevitably arose. From an engineering viewpoint, this is taken into consideration during the design, due to a good knowledge of the material characteristics. Plant fibers have specific properties that make them good candidate reinforcing materials for high-performance composites and other applications [1]. However, the mechanical properties of PFs vary considerably both within the same species and from one species to another. Humidity variation, for example, leads to shrinkage or swelling that changes mechanical properties [2]. Similarly, their thermal properties are by far very different from those of synthetic fibers.

Various studies also indicate that plant fibers exhibit, for example, a very complex anisotropic behavior [3, 4, 5]. This anisotropy must be accounted for if a reliable design is to be achieved. Close collaboration between scientific disciplines such as botany, chemistry, biochemistry, molecular structural biology, plant genetics, physics and mechanics allows each of them to make a constructive and complementary contribution. PFs must withstand stresses of all kinds when they are associated with their deriving plants. They are loaded when it comes to supporting the weight of the plant or when it comes to resisting the winds, storms and hurricanes so common in their environment. PFs are diverse, and can all be studied for their use as engineering materials, in order to take benefit of the particular advantages offered by each of them. Meanwhile, their mechanical, physical and chemical characterization can differ between members of the same species and from one species to another. They are most often in the form of bundles (technical fibers) comprising one to twenty elementary fibers. They have a complex hierarchical structure inducing anisotropy and, have great geometric and mechanical variability. Humidity variation, for example, leads to shrinkage or swelling and changes in mechanical properties.

The characterization of material generally involves so-called monotonous tests (tensile, compression, torsion, bending or a combination) according to the load’s direction (uniaxial or multiaxial), cyclic tests, hardness and resilience tests. Tensile test is undoubtedly the most common test applied to PFs [6, 7, 8, 9, 10, 11, 12] because it allows obtaining Young’s modulus, strength and elongation at break. Recent works show that PFs exhibit a delayed behavior over time and temperature [13, 14] highlighting their viscoelastic nature. A tensile test alone is therefore not sufficient to characterize these materials.

This chapter is structured in four sections. Following this introduction, Section 2 will give an overview of some essential applications, the supply chain and the techniques of separating fibers from their plant. In Section 3, we will describe the experimental characterization methods generally used to derive their structure morphology and their elastic, viscoelastic and thermomechanical properties. Section 4 is the conclusion.

Advertisement

2. Plant fibers

2.1 Common applications

The natural fiber derived from wood, sisal, hemp, coconut, cotton, kenaf, flax, jute, abaca, banana leaf fibers, bamboo, wheat straw or other fibrous material and the matrix can be a polymeric material. The key advantage of natural fibers and their composites over traditional materials is their biological and environmental durability as well as their superior biodegradability. Natural PFs are increasingly used in several fields of engineering applications because of their interesting properties [15]. Diverse abundance of natural fiber, shapes and forms is caused by their occurrence in different climatic zones, hence stimulating the interest and opportunities to conduct comprehensive studies for identifying new applications for the fibers in industry. Notably, they are gaining popularity due to their optimal use in reinforcement of bio-composite structures. These fibers are biodegradable, structurally sound and environmental friendly. However, a sound theoretical basis for modeling their structure and mechanical behavior has yet to be established. Thus, it will be a priority field of study that will challenge the scientists and researchers.

2.2 Fibers supply chain

The emerging trends and opportunities for natural fibers are broadening due to desirable attributes such as biodegradability, eco-friendly, sustainability and energy efficiency. Sustainability supply chain of natural fibers is assessed and rated based on the following criteria: water usage, CO2 emissions, cost, availability and any other impacts [16]. Moreover, in the fashion industry, businesses tend to identify the impacts of fibers on brands that contribute to the most impressive reduction in their impact on environmental footprint. Some of the preferred fibers include Linen, Tencel, Bamboo, Recycled Polyester, Recycled Wool, Cork, Organic Cotton and Hemp.

Perhaps the most important factor is the understanding of the entirety of the supply chain of natural fibers and the stages that contribute to having the biggest impacts. Consequently, a map of biodiversity quantitative impact indicators that help the companies determine where to focus their efforts in supply chain management to alleviate natural fiber environmental footprint was developed.

Nowadays, only 23% of companies take into account their environmental footprint when choosing their suppliers and between 40 and 60% of a company’s environmental footprint actually comes from its supply chain. Hence, in developing the natural fiber supply strategy, it is critical to understand the role of supply chain management and the associated impacts of environmental footprint. Network analysis, optimization of transhipment costs and decision analysis on optimal solutions to minimize both the supply chain cost and environmental footprint are essential toolkits in the advancement and promotion of natural fibers industry.

Moreover, over the last two decades, the trends in production of plant fibers have been declining due to popularity of synthetic fibers as well as adverse drought conditions. The fiber production plants spread across all continents of the globe. Table 1 illustrates the trends of different sources of fibers, production capacities and where they are produced.

Fiber sourceWorld production (103 tonnes)OriginCountry
Abaca70LeafMalaysia, Uganda, Philippines, Bolivia, Brazil
Bambou10,000StemAfrica, India, Brazil
Banana200StemAfrica, India, Brazil
BroomAbundantStem
Coir100FruitIndia, Sri Lanka, Philippines, Malaysia, Brazil
Cotton Lint18,500StemIndia, Europe, USA
Elephant GrassAbundantStemIndia, Africa
Flax810StemEurope
Hemp215StemYugoslavia, China
Jute2500StemIndia, Egypt, Guyana, Jamaica, Ghana, Malawi, Sudan, Tanzania, Brazil
Kenaf770StemIraq, Tanzania, Jamaica
LinseedAbundantFruitUSA
NettlesAbundantStemEurope
Oil Palm FruitAbundantFruitMalaysia, India, Brazil, Indonesia, Philippines
PalmyrahAbundantStemIndia
Ramie100StemHonduras, Mauritius
Roselli250StemBorneo, Guyana, Malaysia, Sri Lanka, Togo, Indonesia, Tanzania
Rice HuskAbundantFruit/grainIndia, Japan, Brazil, Others
Rice StrawAbundantStemIndia, Japan, Brazil, Others
Sisal380LeafEast Africa, Bahamas, Antigua, Kenya, Tanzania, India, Brazil
Sun Hemp70StemNigeria, Guyana, Siera Leone, India
Wheat StrawAbundantStemUSA, Brazil, India, Canada
Wood1,750,000StemAll Countries

Table 1.

Fiber sources, country and annual production of plant fibers.

In 2018, world production of all apparel and textile fibers reached 110 million tons, with natural fiber production estimated at 32 million metric tons. Natural fibers accounted for 29% of the total world fiber production capacity, with most of annual yield variation linked to dry weather conditions. Moreover, the decline in the amounts of natural fibers in total fiber production in the last decade is due to the exponential growth in polyester production, whose demands were triggered by the fast-fashion apparel industry.

2.3 Plant fibers’ extraction methods

2.3.1 Methods commonly used

Cellulosic fibers originated from plants and trees such as cotton, flax, hemp, jute, ramie, kapok, coir and bamboo are termed natural PFs. Such fibers are derived from various parts of plants including leaves, stems (bast fibers), fruits and seeds. Because all natural PFs are made up of mainly cellulose, they are categorized as ‘natural cellulosic fibres’, which may consist of one plant cell or an aggregate of cells bounded together by non-cellulose materials. Major commercially used PFs include: seed fibers (cotton, coir, kapok), bast fibers (flax, hemp, ramie, bamboo, banana), leaf fibers (sisal, kenaf, pineapple, abaca). To date, bast fibers are produced and utilized to manufacture a wide array of traditional and novel products including ropes, nets, carpets, mats, brushes, mattresses, paper and board materials. Generally, PFs are classified into two groups, namely soft fibers and hard fibers. Soft fibers are obtained through labour-intensive processes. It involves the following steps: selection of plant and harvesting the plant, partial drying, pounding with stone mallet, scraped with devices similar to comb to clean the fibers, wash the fibers, dry in the sun and finally comb the fibers. Subsequently, the fibers are ready to be spun or twisted into thread or cord. Soft fibers are often used to make ropes, string, nets, bags, and hammocks.

Hard fibers are processed through successive phases of cutting, drying, cleaning, and soaking before they can be woven. They are strong and naturally flexible fibers, thus suitable and utilized to make furniture, birdcages, toys, baskets, and mats.

Figure 1(a) and (b) shows the matured flax plants grown under a controlled greenhouse environment and a setup of bench-scale trouph for water retting of flax stems [17].

Figure 1.

Greenhouse controlled experiments for flax plants [17]. (a) Matured flax plants in Phytotron and (b) matured flax stems undergoing water retting.

Historically, most plant fibers were extracted manually, supplemented by natural retting. Evidently, this process is tedious, time-consuming and the extracted quality of fibers depends on the skill of the labourer. Nowadays, these fibers are extracted by chemical, mechanical or biological methods.

Akubueze et al. [18], reviewed the chemical techniques employed to extract fibers from natural plants, which include alkali, acid and other reagents. The typical mechanical extraction methods involve the use of stripping the plant stem (typically known as Bacnis and Leonit processes). The latest mechanical extraction methods utilize the decortication process, whereby the plant stems are crushed between two drum rollers to obtain the fibers after removing the pulp. The use of decorticators increase fiber production by 20–25 times compared with the manual process. With biological processes, both consortium of microorganisms and enzymes are utilized to efficiently extract fibers from plant stems.

Overall, the mechanical extraction is incapable to remove the natural binding material (pectin) from the interspaces of the fibers within fiber bundle, chemical extraction is capable to remove the pectin within the fiber bundle but causes significant environmental pollution, whereas the biological extraction method provides increased fiber yield, with minimum detrimental effects to the environment.

According to the Centre for Learning and Teaching in Art and Design (CLTAD), bast fibers, for example, are generally obtained from the phloem, an inner skin of a plant. These fibers support the cells of the phloem to provide strength to the stem. During processing, the fibers need to be separated from both the interior (xylem) and exterior (epidermis) which is the outermost layer of cells. The processes for separating these fibers from plant stalks are known as retting and decortication. Bast fiber bundles are typically several feet long, composed of overlapping cellulose fibers and a cohesive gum (or pectin), which strengthens the stem of the plant. The processes with which the bast fibers are separated significantly influence the quality of fibers as there are many stages involved. Kumar et al. [19], reported that the processing of sustainable fiber starts with fiber extraction and yarn production followed by bleaching, dyeing, softening, printing and drying.

Moreover, the process that separates the fibers into smaller bundles and elementary fibers is known as retting. Fiber retting is a key process and is an important criterion that most industries value because it determines the ultimate properties of the fibers produced. Traditional retting methods include dew and water retting. Dew retting depends on ambient weather conditions, typically takes several weeks and hence the quality of fibers produced varies considerably. Similarly, water retting has been a primary method for low-cost production of bast fibers. The process involves submerging bast straws into water and then the decomposition of the pectic is effected by the activity of anaerobic microorganisms. The quality of retting is assessed by the weight, degumming rate and the fiber properties. The faster rate of weight loss is preferred, the degumming rate is evaluated as the percentage change in pectin content of phloem regions in the raw plant to those in water-retted plant, whereas the desired fiber properties include color, linear density and tensile strength. Ruan et al. [20], reported that water retting improved both whiteness and fineness as well as the mechanical properties of fibers.

Although water retting is capable to produce good quality fibers, the inherent long duration of 7–14 days and associated odor has made it less attractive. The retting period can be reduced to 100 h by using warm water (35°C), but high water consumption and unpleasant odor limit its use to some developing countries. Retting is the process by which pectin gets dissolved or softened from the fiber bundles and separates the fibers from stems through microbial activity. As such, a group of Clostridium microorganism is commonly known to play a significant role in the process by hydrolysing the pectin as it produces pectinase enzyme. These enzymes initially attack the cambium layer and then the other thin-walled cells in the cortex. This phenomenon takes place in most plant bast fibers as they have similar long filament structures, except those from cotton fibers which are single plant cells. As an example, for the retting process conducted in a bench-scale trouph under no-flow process water conditions, there were distinct features on how the fibers separate from bundles. Figure 2(a) and (b) show the scanning electron microscopy of the unretted and retted fibers of flax.

Figure 2.

A SEM shows the microstructure of flax fibers (a) before retting and (b) after the retting process [17].

Figure 3(a) shows that cellulosic fiber production accounted for 6% of the total in 2018, synthetic filament accounted for 45% and synthetic staple 20%. Similarly, Figure 3(b) depicts that cotton accounted for 81% of natural fiber production by weight in 2018, jute accounted for 7%, while coir and wool each accounted for 3%.

Figure 3.

World total fiber production and natural fiber production [21].

The synthetic fibers are dominated by polyester, which accounts for nearly 90% of world filament production and 70% of world synthetic staple production. The remaining synthetic fibers are composed mostly of nylon, acrylic and polypropylene.

Perhaps a key factor is to consider the role and contribution of human capital and household social economics. Employment statistics in natural fiber industries is difficult to estimate because households do not engage in consistent annual production. In Ref. [21] it is estimated that about 60 million households worldwide are engaged in natural fiber production, and hence the total employment, reflecting both full-time year-round employment and part-time or seasonal employment, is around 300 million, which represents about 4% of the world’s population.

Natural fibers possess superior advantages over synthetic fibers including widespread availability, low cost, low density, moderate strength modulus to weight ratio, high acoustic damping, low manufacturing energy consumption, low carbon footprint and biodegradability. Consequently, there are emerging concerted research initiatives that explore and promote the understanding of the characteristics of natural fibers [15, 22].

2.3.2 Other methods

As discussed in Section 2.3.1 above, dew and water retting are the most common processes for fiber retting. Plant fibers can also be extracted using chemical and enzymatic retting, which provide better control than dew and water retting. Unfortunately, chemical retting while effective in extraction of fibers, causes significant pollution challenges due to higher amount of chemicals utilized. For the chemical extraction methods, alkali and selected reagents have been employed. Alkali treatments promote the fibrillation, whereby the composite fiber bundle is degraded into smaller fibers. Sodium hydroxide (NaOH) is popularly used to reduce the fiber roughness, but also produces good quality fiber. Reagents such as sulfuric acid, hydrogen peroxide, protease and sodium citrate can also be used for chemical extraction [23].

Similarly, enzymatic retting is relatively expensive despite its shorter retting time, yet it produces acceptable fiber quality and is advantageous over other retting processes. In the enzymatic method, the selection of enzymes depends on the type of substrate, composition, size and lignin content. The most common enzymes utilized are cellulases and pectinases. Cellulase enzymes enhance the fiber smoothness by removing fibrils from the outer layer. As such, this results in reduction in the mechanical properties due to the damage caused in the fibers. Pectinases remove the inter-lamellar pectin, which is a natural adhesive compound between fibers.

Advertisement

3. Characterization of plant fibers

3.1 Fiber ultrastructure and morphology

The ultrastructure is about dimensions between the atomic and molecular domains. These are accessed using microscopes. Morphology and quantitative chemistry investigations on plant fibers can be achieved following various analytical techniques such as Fourier transform infrared spectroscopy (FTIR), high-performance liquid chromatography (HPLC) and thermogravimetric analysis (TGA), surface electron microscopy (SEM), atomic force microscopy (AFM) and transmission electron microscopy (TEM) [7, 22]. TEM, which uses the principle of electron diffraction leads to very high magnifications of about 5,000,000. Recent progress in instrumentation has made Raman microscopy an extraordinary analytical tool in biological and plant research [24]. The main advantage of confocal Raman microscopy (CRM) is its lateral spatial resolution and the fact that it provides not only chemical composition information but also structural information.

A plant fiber is a nanostructured, renewable, sustainable and biodegradable composite material (Figure 4) [25]. Its cell wall can be likened to a composite lamina, consisting of a few plies reinforced with fibrils. Each individual fiber is composed of a primary wall P and a secondary wall S, itself consisting of three layers S1, S2, S3. In the centre, there may be a cavity called lumen if the cell has not filled up completely during its development. Individual cells are interphased with the middle lamellae (ML) as presented in Figure 4. The S2 layer of the secondary wall represents about 80% of the section and governs the mechanical behavior of the fiber [26]. The middle lamella is a wall 0.5–2 μm thick that surrounds the fiber; it plays the role of matrix that maintains the cohesion of the fibers. It is mainly composed of hemicelluloses, pectin and lignin (about 70%) [27]. Figure 5(a)(d) show micrographs of the RC fiber [28] obtained on a Hitachi H-7650 TEM.

Figure 4.

Simplified structure of the wood cell wall as seen by Coté [25].

Figure 5.

TEM micrographs of the RC fiber (a) consecutive layers (16,400), (b) layer stacking (16,400), (c) warty sub-layer (7660) and (d) reinforcement by a small cell (10,900) [4].

The microfibrillar angle is defined as the angle that the microfibrils form with the longitudinal axis of the cell. These two parameters explain partially the difference in mechanical properties between different types of cortical fibers (Table 2) [15]. The microfibrillar angle has a major influence on the elastic properties of plant fibers. The weaker is this angle, the better are the properties for plant fibers to behave as a composite material, which presents better mechanical properties in the reinforcement direction [22, 33]. Xu and Liu [34] predicted that the cell wall elastic modulus of wood varies by a factor of 3 when microfibril angle changes from 40° to 10°.

FibersCrystallinity index (CrI)Microfibril angle (°)Cross-section area (mm2)Length of the cell (mm)Aspect ratio l/d
Coco45.01.203.335
Flax10.00.122.1687
Hemp6.20.0623960
Jute8.00.122.3110
Ramie7.50.031543500
RC [32]420.05–0.962>2000
Sisal56.6–66.220.01.102.2100

Table 2.

Structural parameters of some plant fibers [29, 30, 31].

The cellulose fibrils are oriented in a helix at an angle called micro-fibril angle, as shown in Figure 4. The microfibril angle in the S1 and S3 layers is greater than that of the S2 layer. It means that the fibrils in S1 and S3 layers are almost transversely oriented with respect to the fiber axis. According to the small microfibril angle in the S2 layer, its fibrils are oriented more parallel to the axis of the fiber [35]. In addition, for a given percentage of cellulose, the lower the microfibril angle, the higher the stiffness and strength of the fiber. The greater the microfibril angle, the greater the elongation at break [26]. Each microfibril can be considered as chains of cellulose crystals bound by amorphous zones [36].

The microfibril angle partly explains the elastic deformation of the plant fiber and therefore its elongation at break. Under relatively low tensile forces, a plant fiber undergoes a reversible deformation due to the progressive alignment of cellulose microfibrils with the fiber axis and an elasto-visco-plastic deformation of amorphous polymers. If the stress of the fiber is stronger, the deformation of it enters an irreversible phase that can continue until the rupture. A high microfibril angle implies a greater elastic deformation for a low tensile fiber stress. In addition, there is a negative correlation between the microfibril angle and the corresponding Young’s modulus (Figure 6) [37].

Figure 6.

Variation of the young modulus with the microfibril angle of a unit cell.

In order to estimate suitability of different fibers to engineering and other applications, it is necessary, among other things, to determine their mechanical properties in the longitudinal and transverse directions as well as the origin of the viscoelastic properties. Thus, we will present in the following paragraphs a state of the art on the main methods used to evaluate the elastic and viscoelastic properties of PFs. Various methods have been used to measure the angle of microfibrils in the S2 layer, which is generally considered a Z-helix. Nevertheless, some studies using cross-field pit punctuations such as those of Pysznski and Hejnowicz [38] on the tracheids of Norwegian Spruce show that in about 80% of the trees studied, the Z-shaped microfibrils have an angle of 10°–40° while in the remaining 20%, the angle is lower with variations in orientation. A complete list of the different microfibril angle measurement techniques with their advantages and disadvantages is given by Huang et al. [39]. Among these techniques, X-ray diffraction is fast, but it is impossible to measure the angle of a single fiber, because of the bundle, only an average of the angle on the X-rays affected cells can be determined. The results obtained by different methods are often contradictory. For example, the work of Herman et al. [40] on individual tracheids shows large variations in the microfibril angle within annual dark circles with a sharp decrease from spring cells to summer cells. While other studies by Lichtenegger et al. [41] using the SAXS (small-angle X-ray scattering) method, on the same cell type shows a higher microfibril angle in summer tracheids than in spring tracheids. Currently, it is necessary to understand where the differences in results obtained by the available measurement methods originate from and to find a method that gives safe and reproductive results. A technique was developed by Jang [42] which uses polarization confocal microscopy based on dichromic cell wall fluorescence when stained with specific fluorochromes showing a high affinity with cellulose. In this technique, sample preparation still needs to be addressed. In fact, very thin samples, only allow observation of fluorescence intensity in the S2 layer without interference with the other layers. A quick but reliable estimate of the Rhectophyllum Camerunense (RC) fiber [28] microfibrils angle was obtained on the SEM (following a microtome longitudinal section of the fiber coinciding with the S2 layer) and fluorescence micrographs.

3.2 Chemical constituents

The chemical composition of plant fibers depends largely on the particular needs of their stemming plant. However, cellulose, hemicellulose and lignin are the main constituents, and their content depends on the age, origin and extraction conditions of the fibers. Cellulose is the chemical constituent that contributes the most to the strength and stability of the plant cell wall and therefore of the fibers. The cellulose content of the fiber largely influences mechanical properties, the economic aspect and the production of the fiber. Fibers with a high cellulose content would be preferable for use in textiles, paper, composites and other fields of activity while those with a high hemicellulose content would be suitable for the production of ethanol and other fermentation products because hemicellulose is easy to hydrolyse in fermentable sugars. Thus, the value of plant fiber and its potential applications depends largely on its cellulose content. Let us say, however, that the value of a plant depends mainly on the quality of its fibers and their end-use and not on the cellulose content itself. As with all-natural products, mechanical and physical properties of natural fibers vary greatly. These properties depend on the chemical and structural composition which depends on the origin of extraction (from leaves, seeds or stems), the local environment where the plants grow, the age of the plants and the climate. The chemical composition, structure, defects and dimensions of the fiber cells are the main parameters that condition all properties of the fibers including mechanical properties [12]. With the exception of cotton, the constituents of plant fibers are cellulose, hemicellulose, lignin, pectin, waxes and water-soluble substances. The average chemical composition of some plant fibers is shown in Table 3.

FiberChemical content (%)
CelluloseHemicelluloseLigninPectin
Abaca63.219.65.1
Bamboo482319
Cotton835
Flax65–7010–162.92–4
Hemp6716.14
Jute55–6412–1812–330.2
Kenaf55–5918–206.8–84.5–5
Ramie68.613.10.6
RC68.21615.6
Sisal54–66127.30.8
TJ62.714.54.17.6
Wood83519–260

Table 3.

Chemical contents of some fibers.

The chemical bonds of the fibers can be determined with FTIR. Crystallographic properties can be analyzed with XRD. TGA, DTA and DSC are used to understand the thermal degradation behavior, the maximum degradation temperature of fibers. Pull-out tests applied to both raw and NaOH treated fibers aim for evaluation of the surface interaction of fibers with polymer matrices for composite materials applications.

In 1838, Anselm Payen proposed that cell walls of many plant cells be made of the same substance to which he gave the name cellulose. Cellulose is a natural polymer whose molecule, formed by long chains, consists of units of D-anhydroglucopyranoses (formula: (C6 H10 O5)n) linked by β-(1,4)-glycosidic bonds in position C1 and C4 (Figure 7). It represents the most abundant biological molecule on our planet. It is present in plants, algae, bacteria and some animals.

Figure 7.

Cellulose molecule.

Cellulose is the major constituent of wood and is the major constituent of cotton and other textile fibers such as flax, hemp, jute and ramie. Its degree of polymerization varies according to the plant species. It can be 14,000 for native cellulose, but the insulation and purification procedures reduce it very sharply by about 2500. Cellulose contributes to the strength and rigidity of the fiber thanks to its strongly oriented chains. These macromolecular chains can be arranged, either regularly, in crystalline regions, or randomly in amorphous regions. Mechanical properties of natural fibers depend on their type of cellulose, as each type has its own cellular geometry. If cellulose is a prime structural constituent for the vast majority of plant cell walls, then hemicellulose with lignin acts as binding materials. Properties depend on the fiber cell geometry of each type of cellulose and its degree of polymerization.

Hemicelluloses represent the second most abundant constituent of plant fiber. Hemicelluloses are polysaccharides found in lignocelluloses alongside cellulose and pectin. Hemicelluloses, unlike cellulose, are composed of several sugars that form short chains with ramifications. The sugars present can be divided into different groups: pentoses (xylose, arabinose), hexoses (glucose, mannose, galactose), hexo-uronic acids (glucuronic acid and methyl-glucuronic acid) and l-deoxyhexoses (rhamnose and fucose). Hemicelluloses are, by definition, water-soluble polysaccharides that can be extracted from the plant cell walls using alkaline solutions. They are the most hydrophilic biopolymers in the cell wall that promote moisture absorption. In their natural state, they have a degree of polymerization that varies from 200 to 300, and their structure depends on the plant species. The best-studied class of hemicelluloses are xyloglucans. They have a bridging role between cellulose microfibrils in order to strengthen the cell wall by interaction with cellulose and, in some cell walls, with lignin. They consist of a glucose chain and short side chains of xylose, galactose and fructose.

Lignin together with cellulose and hemicelluloses is part of the wood industry. Its proportion in wood varies between 15 and 30% [43]. Lignin or ‘lignins’ are three-dimensional polymers from the radical polymerization of three phenylpropenoic alcohols: coniferryl alcohol, sinapyl alcohol and p-coumaryl alcohol [44]. Lignin contributes to the rigidity of cell walls, and thus to the erect port of terrestrial higher plants. Lignin also offers a protective barrier against the microbial attack of plants. Indeed, due to its chemical nature, lignin is very resistant to various chemical agents and biological degradation. To sum up, lignin polymers make the cell wall rigid and impermeable, allowing the transport of water and nutrients through the vascular system by protecting plants from microbial invasion. Lignin is totally amorphous and hydrophobic. It is not hydrolysed by acids, but hot soluble in soda, easily oxidized and also condensable with phenol.

Pectins are polymers of acidic polysaccharides, composed of a main chain of uronic acid bound in 1–4. Regularly, rhamnose molecules are interspersed between these monomers by bonds 1–2 and 1–4. Some of these rhamnose units carry side chains composed of neutral oses among which galactose and arabinose are the most abundant. The type of bond between uronic acid and rhamnose molecules forms elbows. The pectin macromolecule appears like a zigzag. This arrangement contributes to its special properties and provides some flexibility to plants. Pectins are extracted from the fiber by a chemical method either by boiling water or by ethylene diamine tetraacetic acid.

3.3 Density

Different methods can be used including solid pycnometers or gas pycnometers [45, 46, 47]. The choice of gases (helium for example) or immersion liquids such as toluene, ethanol and xylene is decisive for quality results [46, 47]. Fibers must be dried for at least 72 h in a desiccator containing silica (previously regenerated). Fibers are then cut into lengths of 5–15 mm and then introduced into the pycnometer which is eventually placed in the desiccator for at least 24 h. Before carrying out the hydrostatic weighing with the immersion liquids, the vortex agitation of fibers to evacuate the microbubbles between needs to be done. Significant degassing could occur at this stage and provides information on the porosity rate of the fibers [28].

3.4 Mechanical characterization

3.4.1 General considerations

In general, PFs are suitable for reinforcing plastics (thermosets and thermoplastics) and textiles manufacturing thanks to their relatively high strength and low density. The tensile strength and the modulus of elasticity of PFs are very important characteristics for the use of fibers as reinforcements in composite and textile materials. However, the tensile test data for most fibers in service have yet to be studied, as the data found in the literature are scattered and often unreliable. In fact, methods used for the characterization are not identical. Table 4 shows the tensile mechanical properties of some plant fibers compared to synthetic fibers [48]. The properties of the fibers and their structure depend on several factors such as the origin, variety, conditions of growth and harvesting of fibers associated with the treatments, the location in the stem, the presence or absence of a lumen, measurement techniques that vary greatly from one research team to another. These factors can make a difference for the same type of fiber and influence test results.

FiberDensity (g/cm3)Diametre (μm)Length (mm)Tensile strength (MPa)Young modulus (GPa)Elongation at break (%)Moisture content (%)
Abaca1.510–30 (20)4.6–5.2 (4.9)430–813 (621.5)31.1–33.6 (32.35)2.914
Bamboo0.6–1.1 (0.85)25–88 (56.5)1.5–4 (2.75)270–862 (566)17–89 (53)1.3–8 (4.65)11–17 (14)
Banana1.3512–30 (21)0.4–0.9 (0.65)529–914 (721.5)27–32 (29.5)5–6 (5.5)10–11 (10.5)
Coir1.27–30 (18.5)0.3–3 (1.65)175615–25 (20)10
Cotton1.2112–35 (23.5)15–56 (35.5)287–597 (442)6–10 (8)2–10 (6)33–34 (33.5)
Flax1.385–38 (21.5)10–65 (37.5)343–1035 (689)50–70 (60)1.2–3 (2.1)7
Hemp1.4710–51 (30.5)5–55 (30)580–1110 (845)30–60 (45)1.6–4.5 (3.05)8
Jute1.235–25 (15)0.8–6 (3.4)187–773 (480)20–55 (37.5)1.5–3.1 (2.3)12
Kenaf1.212–36 (24)1.4–11 (6.2)295–930 (612.5)22–60 (41)2.7–6.9 (4.8)) 6.2–12 (9.1)
Pineapple1.58–41 (24.5)3–8 (5.5)170–1627 (898.5)60–82 (71)1–3 (2)14
Ramie1.4418–80 (49)40–250 (145)400–938 (669)61.4–128 (94.7)2–4 (3)12–17 (14.5)
RC0.9470–350 (120)450–1500 (557.1)5.8 (±3.5)27.5
Sisal1.27–47 (27)0.8–8 (4.4)507–855 (681)9–22 (15.5)1.9–3 (2.45)11
TJ(1.398)40–90 ()(404.0)(32.3)(1.8)

Table 4.

Mechanical properties of some selected plant fibers versus synthetic fibers [48].

Selection of plant fiber implies a prior study of its mechanical properties, chemical resistance, dimensional stability, separation process, etc. It is worth recalling that linear cellulosic macromolecules are linked by hydrogen bonds and are closely associated with hemicelluloses and lignin, which confer stiffness to fiber. One of the issues of natural fibers is the scattered information and the differences in mechanical properties reported. Likewise, the lack of standards for both producers and users of these materials regarding methods to collect, process, post-process and characterize plant fibers underlines the complexity in the selection.

3.4.2 Quasi-static tensile test

Quasi-static tensile test is the method commonly used in the literature for the characterization of the mechanical properties of plant fibers in the longitudinal direction. This type of characterization presents challenges linked to the assembly and to the single nature of the fiber. In addition, the geometry of the plant fiber makes it often difficult to conduct the tests. Therefore, evaluation of the mean diameter along the fiber using a microscope is necessary for the performance of the test. The single fiber is mounted on a paper frame and a drop of glue is used to stick the fibers. The role of this paper frame is to facilitate the handling and alignment of the fiber on the jaws of the experimental device as shown in Figure 8 [32].

Figure 8.

Tensile test and gripping tab specimens for plant fibers.

The large dispersion of the mechanical properties of the plant fibers observed (Figure 9) is mostly related to the test conditions. The research work by Ntenga et al. [14] focused on the choice of the stress speed and the gage length, in order to keep the deformation in the elastic domain and reduce this dispersion during the tests. The machine cross-head speed of 1 mm/min and the gage length of 10 mm were found to cause less dispersion of the mechanical properties in a tensile test.

Figure 9.

Tensile stress/strain curves for the four cross-head speeds of gage length 10 mm [14].

3.4.3 Nano-indentation test

Nanoindentation is a technique used to characterize the longitudinal and transverse mechanical properties of fibers at the cell wall scale. Commonly measured properties are Young’s modulus and material hardness. In the literature, nanoindentation tests have been carried out to access both transverse and longitudinal mechanical properties on wood fibers [31, 49] and recently on flax fibers [50]. According to Cisse [51], nanoindentation only gives access to local behavior of the fiber, and the identification of mechanical properties requires knowledge and use of a behavior model. The testing technique consists of applying a force to the indenter and taking the area of the indentation, in order to determine the Young’s modulus and the hardness of the material (Figure 10(a) and (b)).

Figure 10.

(a) Nano indentation experimental device and (b) indentor impression Berkovich [52].

A typical set of nanoindentation tests results [53] is shown in Figure 11.

Figure 11.

Transverse modulus of plant fibers obtained in nano indentation.

Differences in transverse and longitudinal modulus noted between the fibers can be explained not only by the differences in micro-fibrillary angles but also by the rate of cellulose that varies between fibers. Hemp and sisal in particular have a cellulose content of around 60%, while that of flax is over 75%; however, the mechanical properties of cellulose are much superior to those of lignin, hemicelluloses and pectins, other constituents of natural fibers [50].

3.4.4 Dynamic mechanical analysis

A large amount of work exists in the field of vibration-based non-destructive testing (NDT) including an extensive survey of over 300 papers by Kong et al. [54]. Indeed, the vibration-based technique has been a very active area of research for many years, however, has always dealt with rigid bodies. As an extension of the use of this technique, the purpose of this section is to present the applicability of the low-frequency vibration-based technique towards estimation of dynamic Young’s modulus of natural fiber-based materials, initially having no bending stiffness. This technique enhances the applicability of non-contact acoustic non-destructive testing to the estimation of dynamic characteristics of thin materials, where the current standard method [55] is not applicable.

Let us consider a thin rectangular specimen having a length b, a width a, a thickness h and a density ρ. Figure 12 shows the specimen configuration in a Cartesian coordinate system at equilibrium (i) and vibrating in flexural mode (ii) and (iii).

Figure 12.

Specimen configuration (i): undeformed, and vibrating at (ii): fundamental frequency, (iii): second frequency in flexural mode.

The specimen, considered as a membrane, initially has no bending stiffness. It is then slightly stretched in the y-direction, in order to make it possible to vibrate transversally (i.e. in the z-direction). The tensile force F is assumed to remain constant during small vibrations in the y-z plane.

In general, for a specimen having intrinsic elasticity, the equation of motion is expressed as follows:

2ξt2c22ξx2+2ξty2+d22ξx2+2ξty22ξ=pxytρhE1

where ξ is the displacement normal to the plane (x,y) which coincides with the equilibrium position of the membrane, c is the velocity of propagation of bending waves, p is the external pressure on the surface of the membrane. c and d are defined as follow:

c=Tρhandd2=Eh212ρ1ν2E2

where E, T and ν are the elastic modulus, tensile force per unit length of the edge, and Poisson’s ratio.

The frequency equation with the fixed-fixed boundary condition shown in Figure 12 above was derived in Mfoumou et al. [56] to obtain the frequency of vibration ωmn of each bending mode as follows:

ωmn=cπma2+πnb21+d22c2πma2+πnb2,m,n=0,1,2,E3

where m and n are the mode numbers.

For a plant fiber-based material considered as a membrane; therefore, no account of intrinsic elasticity is taken so that Eq. (3) is simplified, and the normal frequencies equation is expressed as:

ωmn=cπma2+πnb2,m,n=0,1,2,E4

The Young’s modulus can therefore be determined using the flexural resonance method by monitoring normal modes of vibration. These modes for an oscillating system are special solutions where all the parts of the system are oscillating with the same frequency. At these modes, considering only bending modes in the length direction (m = 0), the relationship between the frequency in hertz and the state of strain was established as follows [56]:

f0n2=E.n24.ρ.b2.εE5

thus, enabling extraction of the constant E from experimentally measured normal mode frequencies and corresponding strains.

3.4.5 Creep test

Both creep experiment and relaxation experiment are two techniques commonly used to characterize the delayed behavior of ‘conventional’ materials. A creep test consists of imposing an almost instantaneous stress load on the plant fiber and maintaining it constantly over time and then proceeding to a discharge. The resulting deformation under the action of the load is creep, and that under the action of discharge is recovery. In general, the creep responses can be broken down into three stages depending on the strain rate as shown in the following Figure 13. The first stage in which creep occurs at a decreasing rate is called primary creep; the second step, commonly called secondary creep, is carried out at a relatively constant speed; and the third stage, tertiary creep, occurs at an increasing rate and terminates with material fracture.

Figure 13.

Creep/recovery test of an elementary hemp fiber under a constant environment [53].

The creep test was successfully carried out on an elementary hemp fiber and the results allowed it possible to highlight the viscoelastic nature of the plant fiber [51]. Figure 13 shows the creep test results obtained.

3.4.6 Relaxation test

3.4.6.1 The context

When a constant strain is applied to a material for a long period, cross-links or the primary bonds that form between molecules start breaking with time and spontaneously lose their bonding capability. High level of strain or long period is the main reason for intermolecular bond breakage, thus creating stress decay over time, called stress relaxation. The rate of bond breakage influences the rate of stress relaxation. Other factors control the rate of bond breakdown, such as stress on the bond, chemical interference, molecular chain mobility which allows molecular chains to move out from their position. The behavior of stress relaxation in plant fibers is also influenced by temperature, humidity, and strain levels. The stress relaxation tests are therefore mainly performed with different ranges of temperature, humidity and strain levels. The time taken to reach the end of relaxation is called relaxation time. From other studies, it is reported that at higher temperature relaxation time becomes shorter, while at lower temperature it becomes longer but the shape of relaxation does not change with temperature [57]; moreover, the variation of strain level affects the stress relaxation [58]. The literature also reports the sensitivity of this class of material to loading-directionality, and ductile and brittle phenomena [59].

3.4.6.2 Stress relaxation measurement

During structural design, the properties of the material must be considered. Elastic Modulus is one of the most important material properties describing the stiffness of the material. When a force is applied to an object, modulus of elasticity or elastic modulus gives the mathematical description of the object’s tendency to be deformed elastically.

In orthotropic materials such as wood-based natural fibers, the strain quickly increases linearly with the stress, then exhibit a nonlinear behavior when the strain exceeds the proportional limits. When the stress relaxation tests are conducted for a very small deformation, the viscoelasticity of the material can be considered linear. During stress relaxation test, the material relieves stress over time as well as the elastic modulus of material Et also decreases with time at a constant temperature. According to linear viscoelastic material [60], the elastic modulus relaxation can then be defined as:

Et=atsoE6

where, so is the constant strain and at is stress of material as a function of time. Indeed, elastic modulus relaxation is the relaxation of modulus of elasticity of material.

3.4.6.3 Sample, experiments, and results

A rectangular strip of specimen is placed between the clamps of the tensile test machine (see Figure 8), and it is slightly loaded within its elastic region. The specimen is tested in uniaxial stress-state at a strain rate of 1 mm/mm with 0.4% strain changes. The elongation is kept constant at 0.4% strain level (1 mm extension) for 5400 s and time, stress, and strain are recorded.

Experiments were carried out for paperboard (PPR) without crack and PPR with crack. Five specimens were tested for each case and each experiment continued for 5400 s (1.5 h) with 1 mm extension. The reason for taking 1 mm extension was to keep the deformation within the elastic region.

The stress relaxation of each specimen was monitored and analyzed at constant.

elongation. The load, stress and time data for constant strain were obtained from the experiments. From the testing of five specimens in each case, we have plotted stress versus time curves. The plotted stress relaxation of PPR without and with the presence of a side crack is presented in Figure 14.

Figure 14.

Stress relaxation of paperboard with and without crack.

Figure 15 show the stress relaxation behavior of PPR at different strain levels (two different extension levels, 1 mm and 0.5 mm).

Figure 15.

Stress relaxation of paperboard for 1 mm and 0.5 mm extension.

3.4.6.4 Formulation of relaxation

The data obtained from the stress relaxation experiments are decreasing type of data with function of time and this type of data can be fitted to the poly-exponential function of the following form:

yt=i=1NaiebitE7

where, ai and bi are the unknown parameters. Several methods have been developed to estimate these parameters. The most available methods are graphical method, regression-difference equation method, method of partial sums, Fourier Transform method, Foss’s method [61] for a sum of two exponentials. However, their uses are limited. For example, the graphical method is not suited where there are consistent fluctuations, regression-difference method and method of partial sums are only appropriate for equally spaced data and Fourier Transform method is suitable for exponentially spaced observation. On the other hand, the use of Foss method is broader than any other method, and even not equally and exponentially spaced data can be treated using this method [61].

3.4.6.5 Model parameters extraction using experimental data

The parameters of a set of mechanical models can be calculated from experimental data. MATLAB, for example, can be used to extract the parameters from the data. To analyze the suitability of the mechanical model with the experimental stress relaxation, Maxwell Model, Two-unit Maxwell Model, Modified Two-unit Maxwell Model, Standard linear solid model are constructed and then compared with the experimental relaxation. Analytical description of these models is given in [62].

In Ref. [56] we have chosen Foss method to develop curve fitting for all models and then compared with the experimental relaxation. Whereas in Ref. [15] we used the Zapas-Phillips method. The best-fitted model with the experimental data was then selected to analysis all experimental data and mathematically stress relaxation equations were derived.

To predict the stress relaxation behavior of natural fibers, we derived the mathematical equations for PPR with and without presence of crack. These equations were derived by the Modified Two-unit Maxwell model which suits best with the experimental result. Though we carried out our experimental tests with five specimens for each kind of test and among them three specimen-data were taken into consideration, but here we will construct the stress relaxation equation for only one specimen for each case.

Below the comparison, diagrams between experimental relaxation data and the Modified Two-unit Maxwell are shown in Figures 16 and 17. The stress relaxation equation for each case is derived using Modified Two-unit Maxwell model.

Figure 16.

Stress relaxation of paperboard—curve fitting.

Figure 17.

Stress relaxation of paperboard with crack–curve fitting.

3.4.7 Inverse characterization

Suitability of materials inverse characterization, destructive or non-destructive, is widely investigated [52, 63, 64]. Furtado et al. [65] used an ultrasound shear wave viscoelastography method to determine the viscoelastic complex shear modulus of macroscopically homogeneous tissues. Ilczyszyn et al. [66] performed the mechanical characterization of flax fibers using an inverse optimization simplex method.

The aim here is to use macro-micro approaches to achieve an efficient estimation of the fiber properties. In fact, homogenization laws of the micromechanics of the elastic/viscoelastic behavior of composite materials provide relationships of the properties of these materials in terms of their constituents’ properties. For an orthotropic material, the knowledge of its off-axes elastic modules in a set of θ directions leads to the calculation of the fibers’ elastic constants for instance. Analytical relationships of elastic constants that account for the orientation of the fibers can be found in the literature. A presentation of an inverse method based on the composite cylinder assembly proposed by Hashin [67] to characterize the anisotropy of plant fibers was discussed in Ref. [64] for a transversely isotropic 2-phase composite. In this case, the matrix and fiber phases are assumed isotropic and transversely isotropic respectively. Using a stress field in cylindrical coordinates and applying Hooke’s law gives the following results:

xyz stand for off-axis Cartesian coordinates (reference coordinate system).

123 represent material axes of a unidirectional composite.

For a tensile test in the x-direction, the Young modulus is expressed as:

1Exx=1E11cos4θ+14G23+14K23+ν122E11sin4θ+1G122ν12E11sin2θcos2θE8

There are five independent properties to be determined E11G23G12K23ν12.

Analytical expressions of the five properties in terms of fiber and matrix phase properties and the volume fractions are given by:

E11=Ezfϕf+EzmϕmνfνmΔEzfνrzfErf+νfEzf2νmϕmE9
ν12¯=21νm2EmνfνmΔ+νmE10
k23=EmϕfI1I4I2I321νmϕfI4L1I2L2+1+νmϕfI4L1I2L2I4I1+I2I3νmϕmI1L2I3L1E11
G12¯Gm=Gmϕm+Grzfϕf+1Gmϕf+1+GrzfϕmE12
G23¯Gm=1+ϕfGmGfGm+km+73Gm2km+83GmGrzfϕmE13

with

Gf=Erf21+νfGm=Em21+νmkm=Em312νmE14
Δ=νfErf1Eθf+νfνrzfErf+νfEzfϕmϕfνmEmϕmϕf1Emϕf+1ϕf+2νm2Em;E15
I1=ϕmνrzfErf+νfEzf+2νmEmϕf;I2=ϕmEzf+νfEmL1=νrzfErf+νfEzf;E16
I3=νfErfνfEθfϕm1Emνmϕf1;E17
I4=νfEzfϕm+νmEmϕfL2=νfErf1Eθf+1Emνm1E18

Eqs. (8)(13) are then solved for Erf,Ezf,νrzf,νf,GfandGrzf using a multiparameter optimisation algorithm.

Advertisement

4. Conclusion

There are evolving global challenges on the utilization of non-renewable resources in the manufacturing industry and increasingly stringent environmental legislation. Both consumers and regulatory agencies are thriving for products that reduce dependency on fossil fuels and thus, are more environmentally friendly. As such, this paves for an opportunity to embrace the use of natural fibers in products and composites leading to significant growth of biobased economy, which the present chapter intends to stimulate.

The field of study of plant fibers that can be industrially exploited remains open. In this chapter, a particular emphasis has been put on their production, in particular on the methods that are generally used to separate them from their originating plants. To date, the question of improving the quality of the extracted fiber has been satisfactorily answered, particularly as regards the possibility of combining several methods when necessary. Some other questions still require research. These include, among others, growing conditions for seed multiplication and fiber production, harvesting methods, optimisation of fiber separation, the molecular basis for improving fiber decortication and performance. The knowledge gained from this work could be used to design new varieties of fibers, tailored for specific industrial applications. Similarly, the recourse to proteomics [68, 69], to isolate genes involved in the biosynthesis of cell wall lignin and hemicellulose in tobacco. Variations in these constituents can affect the fiber quality and cellulose availability. This could then lead to a new orientation on molecular selection research as well as genetic modifications studies to improve the quality of plant fibers.

Morphology and surface behavior of plant fibers are studied using various techniques such as XRD, FTIR, SEM, AFM, TEM and thermogravimetric analysis that helps in understanding the nature of natural fibers.

In terms of the mechanical behavior of plant fibers, important milestones have been achieved to highlight the influence of the chemical composition and structural parameters of the plant wall on their tensile properties. The microstructure of plant fibers is very complex, precisely when it comes to defining generalizable geometric and analytical models that describe it. As mentioned above, improving the mechanical properties of fibers may require the introduction of new types of fibers. And we could mention in this regard the ongoing research on spinning with solvents [70, 71], to obtain fibers of greater strength and low scattered properties. Understanding how fiber morphology affects the properties of composite materials is essential. More precisely, it is important for the selection of new fibers and for the cultivation of fibrous plants genetically selected. This would help to predict their potential for reinforcement in other materials to achieve desired properties.

Investigation of the viscoelastic properties of plant fibers has also been outlined. A variety of dynamic modulus measurement methods exists including ultrasonic wave propagation and the flexural resonance method presented here, for which normal modes of vibration are monitored. Stress relaxation tests are to be carried out to retrieve stress over time as well as the elastic modulus of the fiber material. A mathematical method for extracting the relaxation modulus from relaxation experimental data has to be proposed to this end. Proper selection of the testing vibrational mode and machine cross-head speed (during relaxation) appear important in the suggested methods in order to avoid dispersive results. The Young’s modulus that is obtained from the dynamic behavior of the specimen should, therefore, reflects the frequency dependence of the material.

Advertisement

Acknowledgments

The authors wish to acknowledge the Director of the University Institute of Technology of the University Ngaoundéré, Prof. Mohammadou Bouba Adji, for providing research facilities within the department of mechanical engineering.

References

  1. 1. Thomas S, Paul SA, Pothan LA, Deepa B. Cellulose Fibers: Bio- and Nano-Polymer Composites. Springer-Verlag-Berlin, Heidelberg. 2011. DOI: 10.1007/978-3-642-17370-7. ISBN: 9783642173691
  2. 2. Persson K. Micromechanical Modelling of Wood and Fibre Properties, Lund University. Department of Mechanics and Materials: Lund, Sweden; 2000
  3. 3. Navi P, Heger F. Comportement Thermo-hydromécanique du bois: Applications Technologiques et dans les Structures. Lausanne, Suisse: Presses Polytecniques et Universitaires Romandes; 2005
  4. 4. Ntenga R, Béakou A, Atangana Atéba J, Ayina Ohandja L. Estimation of the elastic anisotropy of sisal fibres by an inverse method. Journal of Materials Science. 2008;43:6206-6213. DOI: 10.1007/s10853-008-2925-2
  5. 5. Cichocki FR, Thomason JL. Thermoelastic anisotropy of a natural fiber. Composites Science and Technology. 2002;62:669-678. DOI: 10.1016/S0266-3538(02)00011-8
  6. 6. Goudenhooft C, Siniscalco D, Arnould O, Bourmaud A, Sire O, Gorshkova T, et al. Investigation of the mechanical properties of flax cell walls during plant development: The relation between performance and cell wall structure. Fibers. 2018;6:1-9. DOI: 10.3390/fib6010006
  7. 7. Hulle A, Kadole P, Katkar P. Agave Americana leaf fibers. Fibers. 2015;3:64-75. DOI: 10.3390/fib3010064
  8. 8. Yu Y, Jiang Z, Fei B, Wang G, Wang H. An improved microtensile technique for mechanical characterization of short plant fibers: A case study on bamboo fibers. Journal of Materials Science. 2011;46:739-746. DOI: 10.1007/s10853-010-4806-8
  9. 9. Réquilé S, Goudenhooft C, Bourmaud A, Le Duigou A, Baley C. Exploring the link between flexural behaviour of hemp and flax stems and fibre stiffness. Industrial Crops and Products. 2018;113:179-186. DOI: 10.1016/J.INDCROP.2018.01.035
  10. 10. Duval A, Bourmaud A, Augier L, Baley C. Influence of the sampling area of the stem on the mechanical properties of hemp fibers. Materials Letters. 2011;65:797-800. DOI: 10.1016/j.matlet.2010.11.053
  11. 11. Krishna KV, Kanny K. The effect of treatment on kenaf fiber using green approach and their reinforced epoxy composites. Composites. Part B, Engineering. 2016;104. DOI: 10.1016/j.compositesb.2016.08.010
  12. 12. Kompella MK, Lambros J. Micromechanical characterization of cellulose fibers. Polymer Testing. 2002;21:523-530. DOI: 10.1016/S0142-9418(01)00119-2
  13. 13. Pacheco JEL, Bavastri CA, Pereira JT. Viscoelastic relaxation modulus characterization using Prony series. Latin American Journal of Solids and Structures. 2015;12(2):420-445. DOI: 10.1590/1679-78251412
  14. 14. Ntenga R, Saïdjo S, Beda T, Béakou A. Estimation of the effects of the cross-head speed and temperature on the mechanical strength of kenaf bast fibers using Weibull and Monte-Carlo statistics. Fibers. 2019;7:89. DOI: 10.3390/fib7100089
  15. 15. Saïdjo S, Ntenga R, Mfoumou E, Beda T, Beakou A. Predicting the relaxation modulus for the study of the delayed behaviour of kenaf fibres in stress relaxation. Fibres & Textiles in Eastern Europe. 2021;29:19-25. DOI: 10.5604/01.3001.0014.7783
  16. 16. Akhil New Normal: Overview of Sustainable Supply Chains and Preferred Fibers. Available from: https://greenstory.ca/blog/new-normal-sustainable-supply-chains-preferred-fibers/ [Accessed: November 5, 2021]
  17. 17. Tango M, Cook K, Njagi R, Browne R, Mfoumou E, Ghosh S, et al. Developing Bio-Economy of Flax Fibers in Atlantic Canada. Darmouth, Canada: Acadia university. 2018
  18. 18. Akubueze EU, Ezeanyanaso CS, Muniru S, Affo G, Igwe CC. Extraction & production of agro—Sack from banana (Musa sapientum) & plantain (Musa Paradisiaca L) fibres for packaging agricultural produce. International Journal of Agriculture and Crop Sciences. 2015;9:9-14
  19. 19. Paramasivam SK, Panneerselvam D, Sundaram D. Extraction, characterization and enzymatic degumming of banana fiber extraction, characterization and enzymatic degumming of Banana. Journal of Natural Fibers. 2020;00:1-10. DOI: 10.1080/15440478.2020.1764456
  20. 20. Ruan P, Raghavan V, Gariepy Y, Du J. Characterization of flax water retting of different durations in laboratory condition and evaluation of its fiber properties. BioResources. 2015;10:3553-3563. DOI: 10.15376/biores.10.2.3553-3563
  21. 21. Townsend T. Natural Fibres and the World Economy. DNFI/FAO. 2019. Available from: dnfi.org/coir/natural-fibres-and-the-world-economy-july-2019_18043. (file accessed on 7th nov 2021)
  22. 22. Béakou A, Ntenga R. Structure, morphology and mechanical properties of Rhectophyllum camerunense (RC) plant fiber. Part II: Computational homogenization of the anisotropic elastic properties. Computational Materials Science. Mans, France: Thesis of the University of Mans. 2011;50:1550-1558. DOI: 10.1016/j.commatsci.2010.12.013. Available from: https://tel.archives-ouvertes.fr/tel-00793179
  23. 23. Mumthas ACSI, Wickramasinghe GLD, Gunasekera USW. Effect of physical, chemical and biological extraction methods on the physical behaviour of banana pseudo-stem fibres: Based on fibres extracted from five common Sri Lankan cultivars. Journal of Engineered Fibers and Fabrics. 2019;14:1-15. DOI: 10.1177/1558925019865697
  24. 24. Cabrales L, Abidi N, Manciu F. Characterization of developing cotton fibers by confocal raman microscopy. Fibers. 2014;2:285-294. DOI: 10.3390/fib2040285
  25. 25. Harrington JJ, Booker R, Astley RJ. Modelling the elastic properties of softwood—Part I: The cell-wall lamellae. Holz als Roh-und Werkstoff. 1995;56:37-41
  26. 26. Baley C. Renforcement des polymeres par des fibres vegetales, elements cles. Revue des Composites et des Matériaux Avancés. 2007;16:1-8
  27. 27. Karray M. Caractérisation des propriétés mécaniques et électriques des matériaux composites par méthode holographique numérique 3D et analyse diélectrique. Mans, France: Thesis of the University of Mans. 2012. Available from: https://tel.archives-ouvertes.fr/tel-00793179
  28. 28. Béakou A, Ntenga R, Lepetit J, Atéba JA, Ayina LO. Physico-chemical and microstructural characterization of “Rhectophyllum camerunense” plant fiber. Composites. Part A, Applied Science and Manufacturing. 2008;39:67-74. DOI: 10.1016/j.compositesa.2007.09.002
  29. 29. Huang Y, Fei B, Wei P, Zhao C. Mechanical properties of bamboo fiber cell walls during the culm development by nanoindentation. Industrial Crops and Products. 2016;92:102-108. DOI: 10.1016/j.indcrop.2016.07.037
  30. 30. Mukherjee PS, Satyanarayana KG. Structure and properties of some vegetable fibres. Part 2. Pineapple fibre (Anannus Comosus). Journal of Materials Science. 1986;21:51-56
  31. 31. Tze WTY, Wang S, Rials TG, Pharr GM, Kelley SS. Nanoindentation of wood cell walls: Continuous stiffness and hardness measurements. Composites. Part A, Applied Science and Manufacturing. 2007;38:945-953. DOI: 10.1016/j.compositesa.2006.06.018
  32. 32. ASTM International ASTM D3379-89 Standard Test Method for Tensile Strength and Young’s Modulus for High Modulus Single Filament Fibers. USA: ASTM. 1989
  33. 33. Charlet K, Gaillard-Martinie B, Béakou A. Comportement mécanique et modélisation numérique de la lamelle mitoyenne de lin. Annales de Chimie—Science des Matériaux. 2012;37:341-350. DOI: 10.3166/RCMA.22.341-350
  34. 34. Xu P, Liu H. Models of microfibril elastic modulus parallel to the cell axis (Wood Science and Technology (2004) 38 (363-374)). Wood Science and Technology. 2004;38:577. DOI: 10.1007/s00226-004-0251-7
  35. 35. Ho THI, T.N. Etude de l’influence de la température et de l’humidité sur les propriétés mécaniques en traction des fibres de chanvre et de coco. Université du Québec; 2008. DOI: 10.1016/j.cell.2009.01.043. ISBN:9781493918621. ISSN:10974172. PMID: 20056882
  36. 36. Siqueira G, Bras J, Dufresne A. Cellulosic bionanocomposites: A review of preparation, properties and applications. Polymers (Basel). 2010;2:728-765. DOI: 10.3390/polym2040728
  37. 37. Nenonene AY. Elaboration et caractérisation mécanique de panneaux de particules de tige de kénaf et de bioadhésifs à base de colle d’os, de tannin ou de mucilage. 2009; pp. 217. Available from: http://oatao.univ-toulouse.fr/7834/1/nenonene
  38. 38. Pyszinski W, Hejnowicz Z. Is the orientation of the fibrilar helix in the main layer of the cell walls constant or variable wthin the tree. Acta Societatis Botanicorum Poloniae. 1972;41:27-38
  39. 39. Huang C, Kutscha NP, Leaf GJ, Megraw RA. Microfibril Angle in Wood. New York, Basel: Buterfield; 1998. pp. 177-205
  40. 40. Herman M, Dutilleul P, Avella-Shaw T. Growth rate effects on intra ring and inter-ring trajectories of microfibril angle in Norway spruce (Picea Abies). IAWA Journal. 1999;20:3-21
  41. 41. Lichtenegger H, Reiterer A, Stanzl Tschegg SE, Fratzel P. Variation of cellulose microfibril angles in soft woods and hard woods: A possible strategy of mechanical optimization. Journal of Structural Biology. 1999;128:257-269
  42. 42. Jang HF. Measurement of fibril angle in wood fibres with polarisation confocal microscopy. Journal of Pulp and Paper Science. 1998;24:224-230
  43. 43. El Hadji BL. Nouveaux matériaux composites thermoformables à base de fibres de cellulose. Institut National Polytechnique de Grenoble; 2008. Available from: https://tel.archives-ouvertes.fr/tel-00268828
  44. 44. Sedan D. Etude des interactions physico-chimiques aux interfaces fibres de chanvre/ciment. Influence sur les propriétés mécaniques du composite. Université de Limoge; 2007. Available from: http://aurore.unilim.fr/ori-oai-search/notice/view/unilim-ori-23351
  45. 45. ASTM B923-02 Standard Test Method for Metal Powder Skeletal Density by Helium or Nitrogen Pycnometry. USA: ASTM. 2008
  46. 46. Amiri A, Triplett Z, Moreira A, Brezinka N, Alcock M, Ulven CA. Standard density measurement method development for flax fiber. Industrial Crops and Products. 2017;96:196-202. DOI: 10.1016/j.indcrop.2016.11.060
  47. 47. Le Gall M, Davies P, Martin N, Baley C. Recommended flax fibre density values for composite property predictions. Industrial Crops and Products. 2018;114:52-58. DOI: 10.1016/j.indcrop.2018.01.065
  48. 48. Peças P, Carvalho H, Salman H, Leite M. Natural fibre composites and their applications: A review. Fibers. 2018;2:1-20. DOI: 10.3390/jcs2040066
  49. 49. Gindl W, Schöberl T. The significance of the elastic modulus of wood cell walls obtained from nanoindentation measurements. Composites. Part A, Applied Science and Manufacturing. 2004;35:1345-1349. DOI: 10.1016/j.compositesa.2004.04.002
  50. 50. Bourmaud A, Keryvin V, Charleux L. Approche du comportement viscoélastique d ’ une fibre de lin par analyse du recouvrement d ’ empreintes de nanoindentation [Investigation of the viscoelastic behavior of a flax fiber by analysing the nanoindentation imprint recovery]. 2011. pp. 1-8. Availabel from: https://hal.archives-ouvertes.fr/hal-00597508
  51. 51. Cisse O. Characterization of the hygro-mechanical behavior of elementary bast fibers from hemp [PhD thesis]. University of Franche-Comté; 2014
  52. 52. Scida D, Bourmaud A, Baley C. Influence of the scattering of flax fibres properties on flax/epoxy woven ply stiffness. Materials and Design. Courbevoie, France: AFM, Maison de la Mécanique. 2017;122:136-145. DOI: 10.1016/j.matdes.2017.02.094
  53. 53. Bourmaud A, Baley C. Rigidity analysis of polypropylene/vegetal fibre composites after recycling. Polymer Degradation and Stability. 2009;94:297-305. DOI: 10.1016/j.polymdegradstab. 2008.12.010
  54. 54. Kong X, Cai CS, Hu J. The state-of-the-art on framework of vibration-based structural damage identification for decision making. Applied Sciences. 2017;7:497. DOI: 10.3390/app7050497
  55. 55. ASTM International. ASTM C1548-02 Standard Test Method for Dynamic Young’s Modulus, Shear Modulus, and Poisson’s Ratio of Refractory Materials by Impulse Excitation of Vibration. USA: ASTM. 2020
  56. 56. Mfoumou E, Hedberg C. Static versus low frequency dynamic elastic modulus measurement of thin films. Electronic Journal “Technical Acoustics”. 2006;17:1-16
  57. 57. Vahedi N, Wu C, Vassilopoulos AP, Keller T. Thermomechanical characterization of a balsa-wood-veneer structural sandwich core material at elevated temperatures. Construction and Building Materials. 2020;230:117037. DOI: 10.1016/J.CONBUILDMAT.2019.117037
  58. 58. Ornaghi HL, Almeida JHS, Monticeli FM, Neves RM. Stress relaxation, creep, and recovery of carbon fiber non-crimp fabric composites. Composites Part C: Open Access. 2020;3:100051. DOI: 10.1016/J.JCOMC.2020.100051
  59. 59. Benvenuti E, Orlando N, Gebhardt C, Kaliske M. An orthotropic multi-surface damage-plasticity FE-formulation for wood: Part I—Constitutive model. Computers and Structures. 2020;240:106350. DOI: 10.1016/J.COMPSTRUC.2020.106350
  60. 60. Malloy RA. Prototyping and experimental stress analysis. In: Plastic Part Design for Injection Molding—An Introduction. Munich: Prototyping and Experimental Stress Analysis; 1994. pp. 285-339. Available from: http://www.hanser-elibrary.com/doi/book/10.3139/9783446433748
  61. 61. Foss SD. A method of exponential curve fitting by numerical integration. Biometrics. 1970;26:815. DOI: 10.2307/2528726
  62. 62. Xu Q, Engquist B. A mathematical model for fitting and predicting relaxation modulus and simulating viscoelastic responses. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. Proc. R. Soc. A 474: 20170540. 2018;474. DOI: 10.1098/RSPA.2017.0540
  63. 63. Kazemirad S, Bernard S, Hybois S, Tang A, Cloutier G. Ultrasound shear wave viscoelastography: Model-independent quantification of the complex shear modulus. IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control. 2016;63:1399-1408. DOI: 10.1109/TUFFC.2016.2583785
  64. 64. Ntenga R, Béakou A, Atangana Atéba J, Ayina Ohandja L. Estimation of the elastic anisotropy of sisal fibres by an inverse method. Journal of Materials Science. 2008;43:67–74. DOI: 10.1007/s10853-008-2925-2
  65. 65. Furtado SCR, Araujo A, Silva A. Inverse characterization of vegetable fibre-reinforced composites exposed to environmental degradation. Compos. Struct. 2018;189:529-544. DOI: 10.1016/J.COMPSTRUCT.2018.01.092
  66. 66. Ilczyszyn F, Cherouat A, Montay G. Nouvelle approche pour la caractérisation mécanique des fibres naturelles. In: Proceedings of the 20ème Congrès Français de Mécanique. France: AFM, Maison de la Mécanique, Courbevoie. 2011. pp. 1-6. Available from: http://hdl.handle.net/2024/46776
  67. 67. Hashin Z. On the elastic behaviour of fibre-reinforced materials of arbitrary transverse phase geometry. Journal of the Mechanics and Physics of Solids. 1965;13:119-134
  68. 68. Das PP, Lin Q, Wong SM. Comparative proteomics of tobacco mosaic virus-infected Nicotiana tabacum plants identified major host proteins involved in photosystems and plant defence. Journal of Proteomics. 2019;194:191-199. DOI: 10.1016/J.JPROT.2018.11.018
  69. 69. Tan BC, Lim YS, Lau SE. Proteomics in commercial crops: An overview. Journal of Proteomics. 2017;169:176-188. DOI: 10.1016/j.jprot.2017.05.018
  70. 70. Klemm D, Heublein B, Fink HP, Bohn A. Cellulose: Fascinating biopolymer and sustainable raw material. Angewandte Chemie International Edition. 2005;44:3358-3393. DOI: 10.1002/ANIE.200460587
  71. 71. Heinze T. Cellulose: Structure and properties. Advances in Polymer Science. 2015;271:1-52. DOI: 10.1007/12_2015_319

Written By

Richard Ntenga, Saidjo Saidjo, Annie Wakata, Pagore Djoda, Martin Tango and Etienne Mfoumou

Submitted: 23 November 2021 Reviewed: 07 February 2022 Published: 21 April 2022