Open access peer-reviewed chapter

Hydro/Hygrothermal Behavior of Plant Fibers and Its Influence on Bio-Composite Properties

Written By

Nouri Mustapha and Tahlaiti Mahfoud

Submitted: 22 October 2021 Reviewed: 10 January 2022 Published: 02 November 2022

DOI: 10.5772/intechopen.102580

From the Edited Volume

Natural Fiber

Edited by Han-Yong Jeon

Chapter metrics overview

105 Chapter Downloads

View Full Metrics

Abstract

Plant fibers have been shown to be highly sensitive to water molecules; this impacts the functionality of composites reinforced with these fibers, commonly known as bio-composites. This review aims to provide a comprehensive description of the behavior of plant fibers in the presence of water molecules in a liquid or gaseous state, as well as the different phenomena and mechanisms involved at the fiber scale and at the bio-composite scale via recent studies in this field. First, we will discuss the physical problem of sorption in polymers in a general way, and then we will focus on the case of plant fibers. Particular attention will be given to the adsorption kinetics of plant fibers and the models used to determine their diffusion parameters. In a second step, the effect of the incorporation of plant fibers in polymer matrices will be examined as well as the different factors influencing the diffusive behavior of bio-composites. In addition, the effect of hydro/hygrothermal aging on the mechanical properties of bio-composites will be discussed.

Keywords

  • plant fibers
  • bio-composite
  • hygrothermal behavior
  • aging
  • diffusion

1. Introduction

The depletion of natural resources, the pollution caused by industry, and the current environmental challenge have led researchers and industrialists to explore new ways of providing green and renewable resources, such as biomass. Among these promising natural resources in the field of composite materials are plant fibers, also known as cellulosic or lignocellulosic fibers. Indeed, the integration of these fibers into the field of composite materials as an alternative to synthetic fibers continues to attract much research and application interest. This is due to their neutrality with regard to CO2 emissions into the atmosphere, but also to the fact that these fibers have specific mechanical and thermal properties. In addition, these fibers have other advantages, such as their low cost, their biodegradability, and the fact that they are renewable.

However, the hydrophilic nature of these plant fibers and the diversity of their properties represent limitations for their use as reinforcement for composite materials. This is particularly true of hydrophobic matrices, which raises questions about the compatibility between the reinforcement and the matrix on the one hand, and the durability of the bio-composite on the other, especially when the latter is exposed to changing hygrothermal conditions.

This chapter will comprehensively address the hydro/hygrothermal behavior of plant fibers and their influence on the durability of bio-composites. First, we will discuss the physical problem of sorption in polymers in a general way, and then we will focus on the case of plant fibers. Particular attention will be given to the adsorption kinetics of plant fibers and the models used to determine their diffusion parameters. In a second step, the effect of the incorporation of plant fibers in polymer matrices will be examined as well as the different factors influencing the diffusive behavior of bio-composites. In addition, the effect of hydro/hygrothermal aging on the mechanical properties of bio-composites will be discussed.

Before starting this chapter, it is necessary to define some basic knowledge that are essential for understanding the physical phenomena that will be presented.

Advertisement

2. Generality on hydro/hygrothermal transfer

2.1 Terminology

2.1.1 Adsorption, absorption, and sorption

Absorption is a phenomenon of filling the pores of a solid with liquid water; it occurs by capillary condensation. Whereas adsorption is a process of attachment of water molecules, water vapor, to the surface of materials by intermolecular forces. These forces are of the same nature as those responsible for the imperfection of real gases and the condensation of vapors [1]. We, therefore, talk about the absorption of liquid water (hydrothermal aging) and the adsorption of water vapor (hygrothermal aging). As for sorption, it encompasses both phenomena.

2.1.2 Relationship between relative humidity and water activity

Under isothermal conditions, the water activity (aw) in a material is defined by the ratio of the water vapor pressure at its surface (pvs) to the pure water vapor pressure (pve). When the material is in equilibrium with its environment or there is no mass or heat transfer, pvs equals the water vapor pressure of air (pva). Therefore, the water activity of a material in equilibrium with its environment is equal to the relative humidity of the air (RHa).

aw=HRa=PvsPveT=PvaPveTE1

2.1.3 Condensation capillaire

This is the phenomenon where water vapor condenses into a liquid phase in a pore at an aw of less than 1, where aw = 1 represents the case of pure water.

2.1.4 Sorption isotherms

Under isothermal conditions, the curve representing the equilibrium moisture content (EMC) of a material as a function of aw is called the sorption isotherm if it was determined from the dry state of the material, and the desorption isotherm if the material was saturated at the initial state [2].

Isotherms are classified according to the new IUPAC classification [1] into six types (Figure 1(1)). The shape of the isotherm provides information about the pore size (macropores, mesopores, or micropores) and the types of adsorption (monolayer or multilayer). When the desorption process is different from the sorption process, a hysteresis loop is observed. These loops are also classified by IUPAC [1] into five categories (Figure 1(2)).

Figure 1.

IUPAC classification of physisorption isotherms (1) and hysteresis loops (2) [1].

2.1.5 Pore size

According to the IUPAC pore classification [1], macropores, mesopores, and micropores are, respectively, those pores whose width exceeds 50 nm, between 2 and 50 nm, and do not exceed 2 nm, respectively.

2.1.6 Aging

Aging is defined according to Berges [3] as an evolution of one or more properties of the material through a modification of the structure, the composition, or the morphology of the constituents. These changes can be temporary and dependent on the presence of the aging source (reversible aging), or permanent (irreversible aging) [4]. Furthermore, the ability of a material to resist irreversible aging is defined by its durability.

2.2 Sorption and diffusion of water molecules in polymers

Generally, polymers are penetrable by water molecules. The latter propagate progressively through the polymer macromolecular network when it is in direct contact with water molecules (gaseous or liquid) [5]. Over time, this leads to a weight gain that continues until the material reaches a saturation plateau [5]. This adsorbed amount is related to the total amount of hydrophilic sites, polymeric areas sensitive to receive water molecules, available in the polymer chains [6]. This also means that the chemical potentials of water in the polymer and in the ambient environment are equal [7]. The kinetics of water sorption and the amount of water adsorbed/absorbed depends on the polymer nature (hydrophobic or hydrophilic), the characteristics of the water (pH, deionized or salted water), and other thermodynamic parameters [4].

The most common method for evaluating water sorption processes in polymers is the recording of mass gain versus time data. The gravimetric curve represents the plot of this data; the mass of water absorbed/adsorbed (Mt) versus time (LF curve in Figure 2). This curve contains important information: the linear part of this curve informs about the penetration rate of water (the diffusion), while the saturation level presents the mass adsorbed/absorbed at infinity (M) [7].

Figure 2.

Typical gravimetric absorption curves of a fluid: (LF) Fick behavior, (A) pseudo-Fick behavior, (B) 2-step diffusion, (C) diffusion with mechanical damage, and (D) diffusion with chemical damage [8].

The diffusion behavior of polymers is often based on the Fick model. Nevertheless, some materials may exhibit non-Fick behavior in the presence of anomalies (curves A B, C, and D in Figure 2). The diffusion behavior can be identified mathematically using Eq. (2) [9]. If the parameter n is close to 0.5, this indicates that the diffusion of water, in this case, follows Fick’s law.

MtM=ktnE2

Where k, n represents the diffusion kinetic parameters.

In the case of bio-composites, we can see in the literature that the gravimetric curve of this kind of material shows a Fickian behavior at the macroscopic scale, see Section 4. However, at the microscopic scale, the problem is more complicated, water molecules can penetrate bio-composites by three different mechanisms: between polymer chains (matrix and fiber), by capillary action in micro-voids, and at the interface level (between fibers and the matrix) when chemical adhesion is absent.

2.3 Fick’s mechanism

Fick’s laws were established by the analogy between conductive heat transfer and mass transfer [10]. In an isotropic case, Fick’s second law is given by the equation below:

Ct=D2Cx2+2Cy2+2Cz2E3

Where: D is the diffusivity coefficient mm2/s, it is a scalar that defines the diffusion kinetics. C is the equilibrium water concentration.

The above equation could also be written in a case of radial diffusion in a solid cylinder of radius r as follows:

Ct=D1rCr+2Cr2E4

In the case where a uniform concentration is imposed on the borders, the simplified solution for long times (0<Mt/M<0.5) and short times (0.5<Mt/M<1) of the above relationship becomes, respectively, as follows [11]:

MtM=4πτ1/21π4τ12τ12.E5
MtM=1n=14n2expαn2τE6

With τ=Dt/r2 and αn the solutions of the Bessel equation of order n.

Equations (5) and (6) can be decomposed even more precisely according to the Mt/M intervals as follows [12]:

0<Mt/M<0.2:

MtM=4πτ1/2E7

0.2<Mt/M<0.5:

MtM=4πτ1/21π4τ12τ12.E8

0.5<Mt/M<0.7:

MtM=14expexp12τ12+expexp22τ22E9

0.7<Mt/M<1:

MtM=14expexp12τ12E10

With α1=2.40483etαn=5.52008.

For the case of a plane plate of thickness h with a uniform initial concentration on both surfaces, the approximate analytical solution of Fick’s law (Eq. (3)) for unidirectional diffusion along the thickness is expressed as follows [12]:

For the first half-absorption (Mt/M<0.5):

MtM=4hDtπE11

For the second half-absorption (Mt/M0.5):

ln1MtM=ln8π2π2Dth2E12

This is often used to illustrate the transfer of water vapor in continuous media and has been adopted by several authors to simulate the transfer of liquid water in materials assumed to be continuous at the macroscopic scale such as bio-composites and plant fibers, see Sections 3 and 4.

Advertisement

3. Hydro/hygrothermal aging of natural fibers

Plant fibers have been shown to be highly sensitive to water molecules, which impacts the functionality of bio-composites [13]. This phenomenon is often related to the morphology of these fibers, their cavity (lumen), and the free hydroxyl groups present on their surface (hydrophilic sites) [13, 14, 15, 16]. The components of plant fiber responsible for its hydrophilic character are mainly hemicellulose and pectin [7], although lignin and amorphous cellulose are also hydrophilic: lignin contains fewer -OH groups than the polysaccharides [17], while cellulose is less accessible. Therefore, a change in hygrothermal conditions could affect the degree of crystallinity of natural fibers, their stiffness, dimensional stability (swelling), and tensile strength [13, 17].

Table 1 shows the results of the saturation rates of the plant fibers after immersion in water. These results depend mainly on the nature of the plant species, the extraction method used, the application of the treatments, and the test protocol applied.

Furthermore, it appears that the hydrophilic character of natural fibers could be reduced by applying treatments, such as chemical and physical treatments [18, 19, 20, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48]. The reasons for these improvements are multiple and depend on the nature of the treatment used and its effect on the plant fibers. Chemical treatments aim to change the chemical composition of the fiber. For example, NaOH treatment removes some of the hydrophilic non-cellulosic materials such as lignin, hemicellulose, and pectin that cover the natural fiber interface. Also, free OH groups can react with the NaOH molecule forming fiber-cell-O-Na groups, which reduces the hydrophilic hydroxyl groups. The silane molecule, on the other hand, has an end containing alkoxy groups that can react with the hydroxyl groups of the plant fiber, giving it hydrophilic properties on its surface. The same mode of functioning for treatment with acetic acid, acrylic acid, and benzoyl [24, 46, 47]. Physical treatments such as Corona [31, 32, 33], Plasma [34, 35, 36], and thermal [18, 37, 38] mainly modify the structural and interfacial properties of the plant fibers without significantly changing their chemical composition.

The hydrophilic character of plant fibers strongly influences the hydro/hygrothermal properties (saturation rates, diffusion kinetics, and dimensional deformation) of bio-composites at the macroscopic scale, an increase in the fiber incorporation rate often leads to a change in these properties, see Section 4.

3.1 Sorption isotherm for plant fibers

The isotherm of cellulosic fibers generally has a sigmoidal shape with hysteresis loops between the adsorption and desorption curve [16, 17, 20, 25, 49, 50, 51, 52], in accordance with the type II isotherm according to the “International Union of Pure and Applied Chemistry (IUPAC)” classification [1]. Since plant fibers are much more hydrophilic than polymer matrices, bio-composites generally exhibit the same type of isothermal sorption curve.

The origin of the hysteresis phenomenon is not yet fully understood and is still a subject of debate in the literature. Considering fiber cell walls as micro-mesoporous materials, some authors have linked this to capillary condensation, which occurs at high RH, and which could also be present at low RH in the micropores [51]. However, it has been shown that capillary condensation could only occur for such material at very high RH [53]. Others have suggested that this may be related to the change of state of the amorphous components, notably hemicellulose and lignin [50, 52, 54]. It has been found that when adsorption occurs above the glass transition temperature of these polymers, also known as the softening point, hysteresis should be absent or minimized [50, 52]. Indeed, at room temperature, amorphous cell wall components of cellulosic fibers such as wood could soften around 65–75% RH [50]. Keating et al. [54] reported a loss of sorption hysteresis in artificial hemicellulose (galactomannan) films when the RH is above 75% at 25°C (Figure 3). Furthermore, Salmèn and Larsson [50] found that increasing the temperature and decreasing the crystallinity index of the fibers affects the presence of the hysteresis loop (Figure 4). They also noticed that the swelling effect could be involved when RH is below the softening temperature. A priori, when water molecules penetrate the matrix, nanopores could be created in the structure to receive these molecules and under desorption conditions, these nanopores could collapse [54]. In addition, Hill et al. [55] questioned the lignin content, whose magnitude of sorption hysteresis was greater when the content of this component was high.

Figure 3.

Sorption isotherm of a nanocomposite film and a guar film as a function of RH, showing hysteresis [54].

Figure 4.

Moisture adsorption and desorption curves as a function of RH for cellulose samples with different degrees of modification. Measurements were carried out at 25 and 65°C. The higher the degree of oxidation (ox), the lower the crystallinity [50].

In addition, several mathematical models exist in the literature to describe the isotherm of this kind of material, including theoretical, semi-theoretical, and empirical models [56]. The Guggenheim, Anderson, and de Boer (GAB), Hailwood Horrobin (H-H), and Generalized D’Arcy and Watt (GDW) models are the most widely used in the literature for plant fibers [20, 25, 55, 57, 58, 59, 60, 61, 62] (Table 2). These models also make physical sense for the attachment of water molecules at the pore scale and can be explained by an extension of Langmuir’s theory for multilayer adsorption. Indeed, each adsorption site can only adsorb one water molecule. These adsorbed molecules could subsequently be secondary adsorption sites for subsequent molecules. The GDW model assumes that some of these sites have the potential to become secondary adsorption sites. The H-H model, on the other hand, considers that water adsorbed by the cell wall can exist in two forms: multilayer water (Md) and monolayer water (Mh). However, while the monolayer in the GAB and DGW model is invariant, the monolayer described in the H-H model can change over the entire RH range.

FibersSaturation rate (%)References
Flax42.58Stamboulis et al. [18]
Alfa140.60Bessadok et al. [19]
Agave138Bessadok et al. [20]
Hemp62Saikia [21]
Hemp63Célino et al. [22]
Jute67.8
Lin62.5
Sisal60.6
Diss40Nouri et al. [23]

Table 1.

Saturation rates of the plant fibers after immersion in water.

ModèlesEquationParamètres
Guggenheim, Anderson and de Boer (GAB).EMCGAPaw=km0caw1kaw1kaw+kcawm0 the adsorption capacity in the first layer in contact with the solid, c, and k are parameters related to the sorption enthalpy.
Generalized D’Arcy and Watt (GDW).EMCGDWaw=km0aw1+kaw1k1ωaw1kawω represents the molecules adsorbed rate on primary sites which transform into secondary adsorption sites.
Hailwood and Horrobin (HH) model.EMCHHaw=18WK1K2aw1+K1K2aw+18WK2aw1+K2awW is the molecular the dry fiber weight per mole of sorption sites, K1 and K2 are constants.

Table 2.

Mathematical description of the GAB, GDW, and HH models.

3.2 Water adsorption kinetics of plant fibers

The diffusive behavior of plant fibers are often based on the Fick model [10, 12, 18, 21, 22, 25, 57, 63]. However, it is sometimes described by non-Fick diffusion models. Célino et al. [22] used different classical diffusion models to define this phenomenon for plant fibers: the Fick model, the two-stage Fick model developed by Loh et al. [64], and the Langmuir model developed by Kibler et al. [65]. They concluded that the Fick model better represents the kinetics of water vapor adsorption in the studied fibers. However, the Langmuir model seems to fit better in the case of liquid water absorption during the immersion process (Figure 5(1)). In addition, Saikia [21] observed during his work on hemp, okra, and betel nut fibers that absorption behaved in two stages. The first stage took place very quickly and obeyed Fick’s law of diffusion. The second absorption step represents a non-Fickian diffusion (Figure 5(2)).

Figure 5.

(1) Immersion absorption curve for different plant fibers [22]. (2) Water content absorbed by: (a) hemp, (b) okra, and (c) betel nut as a function of the square root of time at different temperatures [21].

On the other hand, the analytical solution of Fick’s law in the case of plant fibers are approximate, the sorption is generally subdivided into two or even four zones, each of which is defined by its own law of behavior (see Eqs. (7)(10)). Therefore, some authors propose a single diffusion coefficient to describe the diffusive behavior within plant fibers [18, 63], but others suggest two different diffusion coefficients [12, 25, 57]: D1 for the first half-sorption and D2 for the second half-sorption. Gouanvé et al. [12] found during their work that the D1 and D2 values of flax fibers are similar throughout the aw range studied. However, Alix et al. [25] observed a distinct behavior on the same type of fiber, the value of D2 was found to be significantly larger than D1 over the whole range of aw studied. According to the authors, this dissimilarity is caused by the heterogeneity of the fibers where D2 should be more representative of water diffusion in the core of the fiber while D1 characterizes diffusion through the surface. The same findings were raised by Bessadok et al. [57] on Agave fibers and Nouri et al. [23] on Diss fibers.

In addition, a multitude of results on the diffusion coefficients of plant fibers have been found in the literature (Table 3), this is due mainly to the adopted test protocol, especially for the case of immersion in water, and to the assumed hypothesis (the diffusion occurs from the cross-section and the latter is assumed to be circular). Célino et al. [22] studied the sorption mechanism of different plant fibers (flax, hemp, sisal, and jute) under different conditions: immersion in water and conditioned at 80% RH. The results showed, for each type of fiber, a significant difference of the order of 10−2 between the diffusion coefficients of each case studied. For example, flax fibers showed a diffusion coefficient in the case of water immersion of 5.9 e−06 mm2/s and 2.00 e−04 mm2/s under 80% RH. In the case of relative humidity conditions, similar results were also obtained by Rouidier [10] on flax fibers. However, Gouavné et al. [12] noted a difference of more than three decades under the same conditions. Furthermore, in the case of water immersion, Stamboulis et al. [18] noted a D value of the same order of magnitude as those found by Célino et al. [22] and Rouidier [10] under relative humidity for flax fibers. On the other hand, Nouri et al. [23] studied the effect of different treatments on the diffusive behavior of Diss fibers under different conditions. The results showed diffusion coefficients of the same order of magnitude for the different case studies. As an example, fibers treated with 5% NaOH showed a D1 when immersed in water and conditioned under a RH = 68% of 4.30 × 10−8 mm2/s and 1.6 × 10−8 mm2/s, respectively.

FibersConditionDiffusion modelD (mm2/s)References
Jute51% RHFick3.38 e−07Mannan and Talukder [63]
FlaxWater immersionFick1.93 e−04Stamboulis et al. [18]
FlaxUnder a wide range of RHFick6.50 e−07Gouavné et al. [12]
Agave57%FickD1 = 7.27 e−03
D2 = 3.29 e−02
Bessadok et al. [57]
75%FickD1 = 4.55 e−03
D2 = 2.24 e−03
87%FickD1 = 7.64 e−03
D2 = 2.51 e−03
HempWater immersion at 300 KFick5.20 e−04Saikia [21]
Flax33% RHFick2.06 e−04Roudier [10]
50% RHFick2.79 e−04
75% RHFick3.69 e−04
HempWater immersionLangmuir5.6 e−06Célino et al. [22]
JuteLangmuir5.9 e−06
FlaxLangmuir6.8 e−06
SisalLangmuir9.1 e−06
Hemp80% RHFick2.27 e−04
JuteFick4.02 e−04
FlaxFick2.00 e−04
SisalFick1.17 e−04
DissWater immersionFickD1 = 7.90 e−08
D2 = 5.6 e−07
Nouri et al. [23]

Table 3.

Diffusion parameters of plant fibers under different aging conditions.

In addition, most analytical approaches consider the cross-sectional shape of vegetal fiber bundles as circular. However, this hypothesis is not necessarily applicable to all vegetal fibers as shown by the observations made on the different vegetal fibers [66, 67, 68, 69]. Assuming that the diffusion occurs from the cross-section, the morphology of the cross-section can influence the results obtained when calculating the diffusion coefficient. Recently, Nouri et al. [70] have proposed a coupled (experimental-numerical) approach to improve the modeling of water diffusion through Diss considering two fibers geometries, one often suggested in the literature (circular) and the other one revealed by microscopy (ellipsoidal). The modeling was applied for untreated and treated Diss fibers to determine the diffusion coefficient. Moreover, a single diffusion coefficient was enough to describe diffusion behavior in the case of Diss fibers, contrary to what is practiced with analytical approaches. It was demonstrated that for the same fiber cross-section area, faster diffusion will occur on high perimeter fibers. This means that when the fibers are represented by an ellipsoidal section instead of a circular one for the same cross-sectional area, the diffusion coefficient is less important. Therefore, a significant decrease of 1.5–2.4 in the diffusion coefficient was observed for the fibers studied.

Concerning the evolution of diffusion coefficients as a function of the evolution of relative humidity, Gouanvé et al. [12] observed an increase in D1 and D2 of flax fibers when aw is lower than 0.50; above this value, a decrease was observed for these coefficients (Figure 6a). These findings were explained, according to the authors, by the dual-mode of sorption, Langmuir, and Henry, in the first part of the isotherm as well as the braking effect of swelling at high aw. The same findings were raised by Bessadok et al. [57] (Figure 6b). According to them, this indicates a water sorption mechanism according to Park’s model: part of the water is adsorbed on specific sites (low mobility of the fixed water molecules) and the rest is dissolved according to Henry’s process (higher mobility of the dissolved molecules) and then the subsequent formation of aggregates at high aw (low mobility of the aggregates). Alix et al. [25] observed similar behavior for flax fibers with a decrease in diffusion coefficients when aw is less than 0.1. Above this value, D1, D2 follow the behavior observed by Gouanvé et al. [12] (Figure 6c). This early decrease was explained by the fact that water molecules interact with polar fiber groups leading to hydrogen bonds that increase the cohesion between cellulose chains, thus reducing the mobility of water. In contrast to the previous authors, Nouri et al. [23] found a decrease in D (D1, D2) with increasing aw (Figure 6d). On the other hand, Roudier [10] found that as the relative humidity increases, the diffusion coefficient for flax fibers increase. According to him, this relationship is described by a linear law (Figure 7). It should be noted that the fibers studied here were dried before each conditioning, contrary to the other works which are based on the results of isotherm adsorption.

Figure 6.

Evolution of the water diffusion coefficients (D1 and D2) as a function of aw for: (a) flax fibers [12], (b) agave fibers [57], (c) flax fibers [25], and (d) Diss fibers [23].

Figure 7.

Linear evolution of the water diffusion coefficient for flax fibers as a function of relative humidity [10].

Furthermore, treatments also have an important impact on the hygrothermal properties of plant fibers. Mannan et al. [63] found that after various treatments on jute fibers, the diffusion coefficients decreased during delignification, bleaching, and soap washing, suggesting that moisture was absorbed in the amorphous region of the fibers. Stamboulis et al. [18] studied the effect of the heat treatment, Durbalin, on the hygrothermal properties of flax fibers. The raw fibers always showed a higher diffusion coefficient than the treated fibers, independent of the RH studied. Bessadok et al. [57] observed a decrease in the diffusion coefficient of Agave fibers after their chemical treatment (acrylic acid, styrene, maleic anhydride, and acetylation). On the same line, Nouri et al. [23] noticed a decrease in the diffusion coefficient of Diss fibers for the water absorption case after the different treatments carried out (thermal, acetic acid, NaOH, and silane treatments).

The relationship between the diffusion properties of plant fibers and bio-composites is little studied in the literature due to the lack of reliable experimental protocols for plant fibers (for the case of immersion in water) on the one hand, and the complexity of the problem on the other hand. In the next section, we give more explanation about the effect of integrating plant fibers as reinforcement to the polymer matrix on the diffusive behavior of bio-composites.

Advertisement

4. Hydro/hygrothermal aging of bio-composites

In contrast to plant fibers, water sensitivity is not as noticeable in most polymer matrices, especially petroleum-based matrices. In this section, we investigate the effect of incorporating plant fibers on the hydro/hygrothermal behavior of bio-composites.

4.1 Diffusion behavior of bio-composites

We can see in the literature that most of the plant fiber/PP bio-composites show a Fickian diffusion: wood powder/polypropylene (PP) [71], jute fiber/PP [72, 73], date palm fiber/PP [74], hemp fiber/P and bagasse fiber/PP [75], Kenaf fiber/PP [76], wood fiber/PP [77], wood fiber/high density polyethylene (HDPE) [78, 79], flax fiber/Elium and flax fiber/epoxy [7], flax/epoxy fiber [3], short jute fiber/PLA [80], Areca fine fibers, and Calotropis gigantea fiber/phenol formaldehyde [81].

For the case of a flat plate of thickness h with infinite dimensions, the approximate analytical solution of Fick’s law for unidirectional diffusion through the thickness is expressed by Eqs. (11) and (12). This is of course to determine the homogenized properties of the bio-composite at the macroscopic scale (effective properties). Furthermore, in the case of finite dimensions, the direction of diffusion can be perturbed by so-called edge effects; the diffusion is not only unidirectional along with the thickness but also occurs in other directions. Chilali [7] studied the influence of the width (w)/thickness (h) ratio on the water diffusion coefficient of Lin fiber/epoxy and Lin fiber/Elium composites. The results showed that when this ratio is equal to 60 (180 × 180 × 3 mm3) the edge effect is no longer observable and diffusion occurs along with the thickness similar to the case of the samples (20 × 20 × 3 mm3) sealed on the edges. Therefore, and in addition to what was explained in Section 2.3 for the plane plate case, Shen and Springer [82] proposed a correction to the diffusion coefficient (Dc) in the case of finite sample dimensions for a homogeneous material:

Dc=D1+hl+hw2E13

This correction is often applied in the literature on bio-composites to avoid the edge effect [72, 83, 84, 85].

On the other hand, the interpretation of the diffusion process of composites by analytical approaches often does not reveal all the secrets of this phenomenon, especially when such a complex material is studied.

For a better understanding, the use of numerical approaches using the finite element method is frequently reported in the literature. We can distinguish two types of modeling of the diffusive behavior of conventional composites in the literature: those that consider the composite as a homogeneous material defined by its effective properties [86, 87, 88], others that represent it in a more realistic way by taking into account the fibers [80, 89, 90, 91, 92]. These models can be made in 3D or 2D. In addition, the use of Fick’s law is often observed [80, 86, 87, 88, 89, 90, 91], although we can sometimes see the use of non-Fick’s laws such as Langmuir’s [89, 92, 93]. However, there is little work on bio-composites. Chilali [7] modeled the diffusive behavior of bidirectional flax fiber/epoxy and flax fiber/Elium bio-composites by a 2D biphasic model using the 2D Fick model. Berges [3] reproduced the same methodology for unidirectional flax/epoxy fiber composites. Jiang et al. [80] modeled the 3D diffusive behavior of short jute/PLA composites based on microstructure identification by X-ray tomography. Nouri [94] proposed a 2D numerical model for Diss/PP bio-composites considering three phases: fibers, matrix, and interface. This was done based on microscopic pictures of the bio-composite microstructure. This was achieved based on microscopic pictures of the bio-composite microstructure. The difficulty in this kind of modeling is the non-adaptation of the values determined at the fiber scale, notably the diffusion coefficient as explained in Section 3, to the value that should be implemented in the model. Generally, the authors need to recalculate the fiber diffusion coefficient by an inverse method.

4.2 Factors influencing the hydro/hygrothermal behavior of bio-composites

In general, water molecules penetrate bio-composites by three different mechanisms: between polymer chains, by capillary action in micro-voids, and in the interfaces between fibers and the matrix [81]. Therefore, the process of water diffusion through bio-composites is influenced, mainly, by two types of factors, namely internal factors (related to the bio-composite structure and the nature of its phases) and external factors (relative humidity and temperature).

The literature has shown that the fiber content influences the hydro/hygrothermal properties of bio-composites, Aloa et al. [28] studied the water absorption behavior of hemp fiber/PLA bio-composites and found that water absorption and swelling increased with the addition of hemp fibers to the PLA matrix. The same results were observed by Reza and Krishna [95], Sanjeevi et al. [81], and Mishra and Verma [96] for water absorption. The latter authors explained this, in their study of wood flour/PP composites, by the increase in free OH groups with increasing wood flour; this makes the bio-composite more hydrophilic. However, the diffusion coefficient does not seem to be impacted by increasing the loading rate, and the bio-composites show a lower diffusion coefficient than PP (a slowing wood flour effect). In contrast, results obtained by Law and Ishak [33] show that the diffusion coefficient of Kenaf/PP fiber composites increases with increasing loading rate (an accelerating Kenaf fiber effect). The authors explained this by the ability of the matrix to surround most of the fibers at low loading rates acting as a barrier to water diffusion. This barrier effect of the matrix decreases with increasing loading rate, leading to accelerated diffusivity. Similar results were reported by Joseph et al. [97] on Sisal/PP fiber composites.

On the other hand, a better fiber/matrix adhesion can lead to a reduction in the number of hydrophilic sites in the bio-composite and consequently influence its diffusive behavior. Beg and Pickering [98] found a decrease in the diffusion coefficient of kraft fiber/PP composites with a loading rate of 40% after the addition of 4% by weight MA-g-PP (coupling agent), see Figure 8. This decrease has also been observed by many researchers [76, 99]. However, the work of Mishra and Verma [96] has shown the opposite.

Figure 8.

Evolution of moisture content of composites (kraft fiber/PP) as a function of water immersion time during hygrothermal aging at temperatures of 30, 50, and 70°C (dotted line: with 4 wt.% MAPP; solid line: without MAPP) [98].

On the other hand, Sanjevvi et al. [81] found that fiber size had a significant influence on the adsorption rate of Areca fine fibers (AFFs) and C. gigantea fiber/phenol formaldehyde bio-composites whose long-fiber composite showed the highest water uptake of about 20% when the volume loading rate was 25%. In addition, Pérez-Fonseca et al. [100] found that the size of the fibers (short: 150–212 mm and long: 300–425 mm) did not have a significant effect on the water absorption kinetics of Agave/PP and Pine Sawdust/PP composites. Nevertheless, the composites containing Agave fibers had about three times higher absorption than the composites loaded with pine sawdust at the same loading rate and the diffusivity was about two times higher (Table 4).

BiocompositesConditionsD (mm2/s)References
Kraft fiber (40%)/PPWater immersion at 30°C2.90 e−7Beg and Pickering [98]
Water immersion at 50°C5.70 e−7
Kraft fiber (40%)/PP + 4% MAPPWater immersion at 30°C2.50 e−7Beg and Pickering [98, 101]
Agave fiber (30%)/PPWater immersion7.10 e−7Pérez-Fonseca et al. [100]
Pine sawdust (30%)/PP4.10 e−7
Wood flour (30%)/PPWater immersion at 27°C5.70 e−4Mishra and Verma [96]
Kenaf fibers (30%)/PPWater immersion1.30 e−6Law et al. [76, 99]
Kenaf fibers (30%)/PP + 5% MAPP8.70 e−7
Kenaf fibers (25%)/PP + 1% MAPPWater immersion9.40 e−6Tajvidi et al. [102]
Kenaf fibers (50%)/PP + 2% MAPP1.30 e−4
Wood flour (25%)/PP + 1% MAPP1.60 e−5
Wood flour (50%)/PP + 2% MAPP1.30 e−5
Sisal fibers (30%)/PPWater immersion3.62 e−8Joseph et al. [97]
Diss fibers (20%)/PPWater immersion4.30 e−7Nouri and Tahlaiti [94]

Table 4.

Diffusion coefficients of different biocomposites reported in the literature.

Furthermore, as the ambient temperature increases, the activity of the polymer molecules increases, which accelerates the diffusion of water molecules within the bio-composites. In the same line, Beg and Pickering [98] observed that increasing temperature increases the water uptake kinetics of bio-composites (Figure 8).

Table 4 summarizes the diffusion coefficient values of different bio-composites reported in the literature.

4.3 Hydro/hygrothermal aging

Hydro/hygrothermal aging can be qualified as a physicochemical degradation. Indeed, in the short term, the composite undergoes plasticization due to the infiltration of water molecules between the polymer chains. This type of aging is mostly reversible when the material returns to its initial state. In addition, plant fibers are known to have an affinity for water, unlike most matrices. This causes differential swelling which can also lead to micro-cracks at the fiber/matrix interphase. These cracks can propagate further through a succession of adsorption/desorption cycles (Figure 9). This damage is irreversible and affects the performance of bio-composites. In addition, chemical degradation of the matrix and even of the fibers can occur; this is called hydrolysis. The presence of temperature can aggravate these phenomena [103].

Figure 9.

Degradation mechanism during hygrothermal aging [13]: (a) fibre swells and matrix microcrack at the interface (b) water infiltration into microcracks and the interface (c) dissolution of some fibre component (d) fibres-matrix debonding.

Law and Ishak [99] found a plasticization effect on Kenaf/PP composites after saturation, with a decrease of between 14 and 35% in all tensile and flexural mechanical properties for the composites at 40% loading. After drying the samples again, a partial recovery of their initial mechanical properties (in flexion and in traction) was noted.

Beg and Pickering [98] observed degradation in the tensile mechanical properties of kraft fiber (40%)/PP composites, with and without 4% MAPP, aged for 238 days at different temperatures of 30, 50, and 70°C. The property retention capacity was lower with increasing temperature. The authors associate these results with the degradation of the fibers and/or the fiber/matrix interface (Figure 10).

Figure 10.

Effects of hygrothermal aging on the fracture surface SEMs of composites (kraft fiber (40%)/PP) of: (a) unaged composites and (b) composites aged at 70C for 238 days [98].

Freund [104] carried out a study of cyclic hygrothermal aging at 80°C on Lin/Elium composites during his thesis. Each cycle contains two phases: a saturation phase at 80% relative humidity and a drying phase at 10% relative humidity. It was found that after each cycle the composites lost mass and mechanical properties with a decrease in stiffness of about 50% after five cycles and a decrease in tensile stress from 110 to 30 MPa (Figure 11). These degradations also had an effect on the adsorption kinetics: the more the material undergoes aging cycles, the faster its adsorption becomes.

Figure 11.

Evolution of Young’s modulus and maximum stress of composites flax/Elium as a function of humidity [104].

The author summarized this change in behavior into two reasons: swelling of the composite, especially the fibers, and degradation of the fibers. These findings were justified by a change in the failure mechanism of the composite due to a degradation of the fiber/matrix interface (Figure 12), and the potential degradation of the flax fibers were justified by a decrease in their crystallinity index from 77.2 to 38.4%.

Figure 12.

Fracture facies of flax/Elium composite before and after aging [104].

On the other hand, Berges [3] studied the cyclic aging of composites at 70°C with two humidity conditions 90% RH and 15% RH for adsorption and desorption, respectively. For each cycle, a total duration of 4 days was chosen, that is, 2 days for each half cycle. Tensile tests carried out after four and nine cycles revealed a reproducible behavior for tensile modulus with a decrease in elongation at break and stress at the break with the number of cycles undergone. This reduction was related to fiber/matrix decohesion and resin damage.

Wang and Petru [105] compared the results of the mechanical properties of unidirectional (30% by volume)/epoxy flax fiber composites undergoing natural aging outdoors and artificial aging by immersion in water at 60°C. It was concluded that 60, 120, and 180 days of natural aging correspond to 1.1, 37.2, and 167.9 h of artificial aging, respectively. Comparisons were made regarding flexural strength and flexural modulus.

Advertisement

5. Conclusion

Plant fibers have been shown to be highly sensitive to water molecules, which impacts the functionality of bio-composites. This phenomenon is often related to the morphology of these fibers, their cavity (lumen), and the free hydroxyl groups present on their surface. It appears that the hydrophilic character of natural fibers could be reduced by applying treatments, such as chemical and physical treatments.

The isotherm of cellulosic fibers generally has a sigmoidal shape with hysteresis loops between the adsorption and desorption curve in accordance with the type II isotherm. The Guggenheim, Anderson, and de Boer (GAB), Hailwood Horrobin (H-H), and Generalized D’Arcy and Watt (GDW) models are the most widely used in the literature to describe the isotherm of plant fibers.

The diffusive behavior of plant fibers are often based on the Fick model. However, it is sometimes described by non-Fick diffusion models. The same conclusions have been reported for bio-composites. On the other hand, the process of water diffusion through bio-composites is influenced, mainly, by two types of factors, namely internal factors (related to the bio-composite structure and the nature of its phases) and external factors (relative humidity and temperature).

Hydro/hygrothermal aging can cause irreversible damage to the bio-composite due to the different nature of the matrix and the plant fibers. This is caused by a differential swelling between these components.

Advertisement

Acknowledgments

This research work has been conducted with the financial support of FEDER—Region Pays de la Loire in the framework of the CIPTAP R&D project.

References

  1. 1. Thommes M, Kaneko K, Neimark AV, Olivier JP, Rodriguez-Reinoso F, Rouquerol J, et al. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure and Applied Chemistry. 2015;87:1051-1069. DOI: 10.1515/pac-2014-1117
  2. 2. Sun L-M, Meunier F, Brodu N, Manero M-H. Adsorption—Aspects théoriques. 2016. p. 24
  3. 3. Berges M. Mécanismes de dégradation sous sollicitations hydrothermomécaniques de biocomposites et renforts en fibres végétales: application au développement de mobiliers urbains ultralégers et mobiles [thesis]. Bourgogne Franche-Comté; 2018
  4. 4. Testoni GA. In situ long-term durability analysis of biocomposites in the marine environment. n.d. p. 235
  5. 5. Regazzi A. Contribution à l’étude du vieillissement couplé thermo-hydro-mécanique de biocomposite PLA/lin [PhD thesis]. Ecole Nationale Supérieure des Mines de Paris; 2013
  6. 6. Fraga AN, Alvarez VA, Vazquez A, de la Osa O. Relationship between dynamic mechanical properties and water absorption of unsaturated polyester and vinyl ester glass fiber composites. Journal of Composite Materials. 2003;37:1553-1574. DOI: 10.1177/0021998303029421
  7. 7. Chilali A. Étude expérimentale et modélisation de la durabilité des biocomposites à fibres de lin [thesis]. Reims; 2017
  8. 8. Weitsman YJ, Elahi M. Effects of fluids on the deformation, strength and durability of polymeric composites—An overview. Mechanics of Time Dependent Materials. 2000;4:107-126. DOI: 10.1023/A:1009838128526
  9. 9. Assarar M, Scida D, El Mahi A, Poilâne C, Ayad R. Influence of water ageing on mechanical properties and damage events of two reinforced composite materials: Flax-fibres and glass–fibres. Materials & Design. 2011;32:788-795. DOI: 10.1016/j.matdes.2010.07.024
  10. 10. Roudier A. Analyse multi-échelle du comportement hygro-mécanique des fibres de lin. Génie Mécanique. Université Blaise Pascal—Clermont II; 2012
  11. 11. Crank J. The Mathematics of Diffusion. 2nd ed. Oxford, England: Clarendon Press; 1975
  12. 12. Gouanvé F, Marais S, Bessadok A, Langevin D, Métayer M. Kinetics of water sorption in flax and PET fibers. European Polymer Journal. 2007;43:586-598. DOI: 10.1016/j.eurpolymj.2006.10.023
  13. 13. Azwa ZN, Yousif BF, Manalo AC, Karunasena W. A review on the degradability of polymeric composites based on natural fibres. Materials and Design. 2013;47:424-442. DOI: 10.1016/j.matdes.2012.11.025
  14. 14. Guo X, Wu Y. Characterizing molecular structure of water adsorbed by cellulose nanofiber film using in situ micro-FTIR spectroscopy. Journal of Wood Chemistry and Technology. 2017;37:383-392. DOI: 10.1080/02773813.2017.1306078
  15. 15. Célino A, Gonçalves O, Jacquemin F, Fréour S. Utilisation de la spectrométrie infrarouge pour une quantification rapide du taux d’humidité dans des fibres végétales. Revue des Composites et des Materiaux Avances. 2014;24:81-95. DOI: 10.3166/rcma.24.81-95
  16. 16. Guo X, Liu L, Wu J, Fan J, Wu Y. Qualitatively and quantitatively characterizing water adsorption of a cellulose nanofiber film using micro-FTIR spectroscopy. RSC Advances. 2018;8:4214-4220. DOI: 10.1039/C7RA09894D
  17. 17. Himmel S, Mai C. Effects of acetylation and formalization on the dynamic water vapor sorption behavior of wood. Holzforschung. 2014;69:633-643. DOI: 10.1515/hf-2014-0161
  18. 18. Stamboulis A, Baillie CA, Peijs T. Effects of environmental conditions on mechanical and physical properties of flax fibers. Composites Part A: Applied Science and Manufacturing. 2001;32:1105-1115
  19. 19. Bessadok A, Marais S, Gouanvé F, Colasse L, Zimmerlin I, Roudesli S, et al. Effect of chemical treatments of Alfa (Stipa tenacissima) fibres on water-sorption properties. Composites Science and Technology. 2007;67:685-697. DOI: 10.1016/j.compscitech.2006.04.013
  20. 20. Bessadok A, Marais S, Roudesli S, Lixon C, Métayer M. Influence of chemical modifications on water-sorption and mechanical properties of Agave fibres. Composites Part A: Applied Science and Manufacturing. 2008;39:29-45. DOI: 10.1016/j.compositesa.2007.09.007
  21. 21. Saikia D. Studies of water absorption behavior of plant fibers at different temperatures. International Journal of Thermophysics. 2010;31:1020-1026. DOI: 10.1007/s10765-010-0774-0
  22. 22. Célino A, Fréour S, Jacquemin F, Casari P. Characterization and modeling of the moisture diffusion behavior of natural fibers. Journal of Applied Polymer Science. 2013;130:297-306. DOI: 10.1002/app.39148
  23. 23. Nouri M, Tahlaiti M, Grondin F, Belarbi R. The influence of chemical and thermal treatments on the Diss fiber hygroscopic behaviors. Journal of Natural Fibers. 2020:1-14. DOI: 10.1080/15440478.2020.1848733
  24. 24. Kabir MM, Wang H, Lau KT, Cardona F. Chemical treatments on plant-based natural fibre reinforced polymer composites: An overview. Composites Part B: Engineering. 2012;43:2883-2892. DOI: 10.1016/j.compositesb.2012.04.053
  25. 25. Alix S, Philippe E, Bessadok A, Lebrun L, Morvan C, Marais S. Effect of chemical treatments on water sorption and mechanical properties of flax fibres. Bioresource Technology. 2009;100:4742-4749. DOI: 10.1016/j.biortech.2009.04.067
  26. 26. Kalia S, Kaith BS, Kaur I. Pretreatments of natural fibers and their application as reinforcing material in polymer composites—A review. Polymer Engineering & Science. 2009;49:1253-1272. DOI: 10.1002/pen.21328
  27. 27. Haque R, Saxena M, Shit SC, Asokan P. Fibre-matrix adhesion and properties evaluation of sisal polymer composite. Fibers and Polymers. 2015;16:146-152. DOI: 10.1007/s12221-015-0146-2
  28. 28. Alao PF, Marrot L, Kallakas H, Just A, Poltimäe T, Kers J. Effect of hemp fiber surface treatment on the moisture/water resistance and reaction to fire of reinforced PLA composites. Materials. 2021;14:4332. DOI: 10.3390/ma14154332
  29. 29. Balaji AN, Nagarajan KJ. Characterization of alkali treated and untreated new cellulosic fiber from Saharan Aloe vera cactus leaves. Carbohydrate Polymers. 2017;174:200-208. DOI: 10.1016/j.carbpol.2017.06.065
  30. 30. Asim M, Jawaid M, Abdan K, Ishak MR. Effect of alkali and silane treatments on mechanical and fibre-matrix bond strength of Kenaf and pineapple leaf fibres. Journal of Bionic Engineering. 2016;13:426-435. DOI: 10.1016/S1672-6529(16)60315-3
  31. 31. Gassan J, Gutowski VS. Effects of corona discharge and UV treatment on the properties of jute-fibre epoxy composites. Composites Science and Technology. 2000;60:2857-2863
  32. 32. Pizzi A, Kueny R, Lecoanet F, Massetau B, Carpentier D, Krebs A, et al. High resin content natural matrix–natural fibre biocomposites. Industrial Crops and Products. 2009;30:235-240. DOI: 10.1016/j.indcrop.2009.03.013
  33. 33. Ragoubi M, Bienaimé D, Molina S, George B, Merlin A. Impact of corona treated hemp fibres onto mechanical properties of polypropylene composites made thereof. Industrial Crops and Products. 2010;31:344-349. DOI: 10.1016/j.indcrop.2009.12.004
  34. 34. Sinha E, Panigrahi S. Effect of plasma treatment on structure, wettability of jute fiber and flexural strength of its composite. Journal of Composite Materials. 2009;43:1791-1802. DOI: 10.1177/0021998309338078
  35. 35. Graupner N, Albrecht K, Hegemann D, Müssig J. Plasma modification of man-made cellulose fibers (Lyocell) for improved fiber/matrix adhesion in poly(lactic acid) composites. Journal of Applied Polymer Science. 2013;128:4378-4386. DOI: 10.1002/app.38663
  36. 36. Marais S, Gouanvé F, Bonnesoeur A, Grenet J, Poncin-Epaillard F, Morvan C, et al. Unsaturated polyester composites reinforced with flax fibers: Effect of cold plasma and autoclave treatments on mechanical and permeation properties. Composites Part A: Applied Science and Manufacturing. 2005;36:975-986. DOI: 10.1016/j.compositesa.2004.11.008
  37. 37. Ariawan D, Salim MS, Taib RM, Thirmizir MZA, Ishak ZAM. Interfacial characterisation and mechanical properties of heat treated non-woven Kenaf fibre and its reinforced composites. Composite Interfaces. 2018;25:187-203. DOI: 10.1080/09276440.2017.1354562
  38. 38. Rong MZ, Zhang MQ, Liu Y, Yang GC, Zeng HM. The effect of fiber treatment on the mechanical properties of unidirectional sisal-reinforced epoxy composites. Composites Science and Technology. 2001;61:1437-1447. DOI: 10.1016/S0266-3538(01)00046-X
  39. 39. Fook LT, Yatim JM. An experimental study on the effect of alkali treatment on properties of Kenaf fiber for reinforced concrete elements. International Journal of Research in Engineering and Technology. 2015;4:37-40. DOI: 10.15623/ijret.2015.0408007s
  40. 40. Chandrasekar M, Ishak MR, Sapuan SM, Leman Z, Jawaid M. A review on the characterisation of natural fibres and their composites after alkali treatment and water absorption. Plastics, Rubber and Composites. 2017;46:119-136. DOI: 10.1080/14658011.2017.1298550
  41. 41. Huang Z, Ge H, Yin J, Liu F. Effects of fiber loading and chemical treatments on properties of sisal fiber-reinforced sheet molding compounds. Journal of Composite Materials. 2017;51:3175-3185. DOI: 10.1177/0021998316683440
  42. 42. Faruk O, Bledzki AK, Fink H-P, Sain M. Biocomposites reinforced with natural fibers: 2000–2010. Progress in Polymer Science. 2012;37:1552-1596. DOI: 10.1016/j.progpolymsci.2012.04.003
  43. 43. Bledzki AK, Mamun AA, Lucka-Gabor M, Gutowski VS. The effects of acetylation on properties of flax fibre and its polypropylene composites. Express Polymer Letters. 2008;2:413-422. DOI: 10.3144/expresspolymlett.2008.50
  44. 44. Bledzki A. Composites reinforced with cellulose based fibres. Progress in Polymer Science. 1999;24:221-274. DOI: 10.1016/S0079-6700(98)00018-5
  45. 45. Vishnu Vardhini K, Murugan R, Surjit R. Effect of alkali and enzymatic treatments of banana fibre on properties of banana/polypropylene composites. Journal of Industrial Textiles. 2018;47:1849-1864. DOI: 10.1177/1528083717714479
  46. 46. Li X, Tabil LG, Panigrahi S. Chemical treatments of natural fiber for use in natural fiber-reinforced composites: A review. Journal of Polymers and the Environment. 2007;15:25-33. DOI: 10.1007/s10924-006-0042-3
  47. 47. Motaung TE, Linganiso LZ, Kumar R, Anandjiwala RD. Agave and sisal fibre inforced polyfrurfuryl alcohol composite. Journal of Thermoplastic Composite Materials. 2017;30:1323-1343. DOI: 10.1177/0892705716632858
  48. 48. Bledzki AK, Reihmane S, Gassan J. Properties and modification methods for vegetable fibers for natural fiber composites. Journal of Applied Polymer Science. 1996;59:1329-1336. DOI: 10.1002/(SICI)1097-4628(19960222)59:8<1329::AID-APP17>3.0.CO;2-0
  49. 49. Placet V, Day A, Beaugrand J. The influence of unintended field retting on the physicochemical and mechanical properties of industrial hemp bast fibres. Journal of Materials Science. 2017;52:5759-5777. DOI: 10.1007/s10853-017-0811-5
  50. 50. Salmén L, Larsson PA. On the origin of sorption hysteresis in cellulosic materials. Carbohydrate Polymers. 2018;182:15-20. DOI: 10.1016/j.carbpol.2017.11.005
  51. 51. Shi J, Avramidis S. Water sorption hysteresis in wood: I review and experimental patterns—Geometric characteristics of scanning curves. Holzforschung. 2017;71:307-316. DOI: 10.1515/hf-2016-0120
  52. 52. Fredriksson M, Thybring EE. Scanning or desorption isotherms? Characterising sorption hysteresis of wood. Cellulose. 2018;25:4477-4485. DOI: 10.1007/s10570-018-1898-9
  53. 53. Thygesen L, Thybring E, Hoffmeyer P. Water sorption in wood and modified wood at high values of relative humidity. Part I: Results for untreated, acetylated, and furfurylated Norway spruce. Holzforschung. 2010;64:315-323. DOI: 10.1515/HF.2010.044
  54. 54. Keating BA, Hill CAS, Sun D, English R, Davies P, McCue C. The water vapor sorption behavior of a galactomannan cellulose nanocomposite film analyzed using parallel exponential kinetics and the Kelvin-Voigt viscoelastic model. Journal of Applied Polymer Science. 2013;129:2352-2359. DOI: 10.1002/app.39132
  55. 55. Hill CAS, Norton A, Newman G. The water vapor sorption behavior of natural fibers. Journal of Applied Polymer Science. 2009;112:1524-1537. DOI: 10.1002/app.29725
  56. 56. Trabelsi A, Belarbi R, Turcry P, Aït-Mokhtar A. Characterization of water vapor desorption variability and its effects on drying simulation for an in-situ high performances concrete. Construction and Building Materials. 2011;141:140–151
  57. 57. Bessadok A, Langevin D, Gouanvé F, Chappey C, Roudesli S, Marais S. Study of water sorption on modified Agave fibres. Carbohydrate Polymers. 2009;76:74-85. DOI: 10.1016/j.carbpol.2008.09.033
  58. 58. Marzec A, Lewicki PP. Antiplasticization of cereal-based products by water. Part I. Extruded flat bread. Journal of Food Engineering. 2006;73:1-8. DOI: 10.1016/j.jfoodeng.2004.12.002
  59. 59. Koko CA, Diomande M, Kouame BK, Assidjo EN. Détermination expérimentale et modélisation des isothermes d’adsorption d’eau des amandes d’Irvingia gabonensis de la région du Haut-Sassandra (Côte d’Ivoire). n.d. p. 17
  60. 60. Bui R, Labat M, Aubert J-E. Comparison of the saturated salt solution and the dynamic vapor sorption techniques based on the measured sorption isotherm of barley straw. Construction and Building Materials. 2017;141:140-151. DOI: 10.1016/j.conbuildmat.2017.03.005
  61. 61. Patcharawijit A, Choodum N, Yamsaengsung R. Effects of superheated steam treatment on moisture adsorption and mechanical properties of pre-dried rubberwood. Drying Technology. 2019;37:1647-1655. DOI: 10.1080/07373937.2018.1529037
  62. 62. Furmaniak S, Terzyk AP, Gauden PA, Rychlicki G. Applicability of the generalised D’Arcy and Watt model to description of water sorption on pineapple and other foodstuffs. Journal of Food Engineering. 2007;79:718-723. DOI: 10.1016/j.jfoodeng.2006.02.036
  63. 63. Mannan KM, Talukder MAI. Characterization of raw, delignified and bleached jute fibres by study of absorption of moisture and some mechanical properties. Polymer. 1997;38:2493-2500. DOI: 10.1016/S0032-3861(96)00804-X
  64. 64. Loh WK, Crocombe AD, Abdel Wahab MM, Ashcroft IA. Modelling anomalous moisture uptake, swelling and thermal characteristics of a rubber toughened epoxy adhesive. International Journal of Adhesion and Adhesives. 2005;25:1-12. DOI: 10.1016/j.ijadhadh.2004.02.002
  65. 65. Carter HG, Kibler KG. Langmuir-type model for anomalous moisture diffusion in composite resins. Journal of Composite Materials. 1978;12:118-131. DOI: 10.1177/002199837801200201
  66. 66. Alves Fidelis ME, Pereira TVC, Gomes O da FM, de Andrade Silva F, Toledo Filho RD. The effect of fiber morphology on the tensile strength of natural fibers. Journal of Materials Research and Technology. 2013;2:149-157. DOI: 10.1016/j.jmrt.2013.02.003
  67. 67. Ferreira SR, Pepe M, Martinelli E, de Andrade Silva F, Toledo Filho RD. Influence of natural fibers characteristics on the interface mechanics with cement based matrices. Composites Part B: Engineering. 2018;140:183-196. DOI: 10.1016/j.compositesb.2017.12.016
  68. 68. Newman RH, Le Guen MJ, Battley MA, Carpenter JEP. Failure mechanisms in composites reinforced with unidirectional Phormium leaf fibre. Composites Part A: Applied Science and Manufacturing. 2010;41:353-359. DOI: 10.1016/j.compositesa.2009.11.001
  69. 69. Nouri M, Griballah I, Tahlaiti M, Grondin F, Beaugrand J. Plant extraction and physicochemical characterizations of untreated and pretreated Diss fibers (Ampelodesmos mauritanicus). Journal of Natural Fibers. 2019:1-11. DOI: 10.1080/15440478.2019.1687062
  70. 70. Nouri M, Tahlaiti M, Meroufel A, Grondin F. Enhanced modeling of water diffusion in natural fibers: Application to Diss fibers. Journal of Natural Fibers. 2021:1-10. DOI: 10.1080/15440478.2021.1982817
  71. 71. Wang W, Guo X, Zhao D, Liu L, Zhang R, Yu J. Water absorption and hygrothermal aging behavior of wood-polypropylene composites. Polymers. 2020;12:782. DOI: 10.3390/polym12040782
  72. 72. Chandekar H, Chaudhari V, Waigaonkar S, Mascarenhas A. Effect of chemical treatment on mechanical properties and water diffusion characteristics of jute-polypropylene composites. Polymer Composites. 2019; 41:1447-1461. DOI: 10.1002/pc.25468
  73. 73. Ameer MH, Nawab Y, Ali Z, Imad A, Ahmad S. Development and characterization of jute/polypropylene composite by using comingled nonwoven structures. Journal of the Textile Institute. 2019;110:1652-1659. DOI: 10.1080/00405000.2019.1612502
  74. 74. Boukettaya S, Alawar A, Almaskari F, Ben Daly H, Abdala A, Chatti S. Modeling of water diffusion mechanism in polypropylene/date palm fiber composite materials. Journal of Composite Materials. 2018;52:2651-2659. DOI: 10.1177/0021998317752228
  75. 75. Kord B, Hosseinihashemi SK, Modirzare M. Influence of fungal infection on the long-term water absorption and morphological behavior of bagasse fiber/polypropylene composites at different exposure times. Science and Engineering of Composite Materials. 2013;20:351-357. DOI: 10.1515/secm-2013-0005
  76. 76. Lee JM, Mohd Ishak ZA, Mat Taib R, Law TT, Ahmad Thirmizir MZ. Mechanical, thermal and water absorption properties of kenaf-fiber-based polypropylene and poly(butylene succinate) composites. Journal of Polymers and the Environment. 2013;21:293-302. DOI: 10.1007/s10924-012-0516-4
  77. 77. Hosseinihashemi SK, Arwinfar F, Najafi A, Nemli G, Ayrilmis N. Long-term water absorption behavior of thermoplastic composites produced with thermally treated wood. Measurement. 2016;86:202-208. DOI: 10.1016/j.measurement.2016.02.058
  78. 78. Gozdecki C, Wilczyn’ski A. Effects of wood particle size and test specimen size on mechanical and water resistance properties of injected wood–high density polyethylene composite. Wood and Fiber Science. 2015;47:365-374
  79. 79. Mrad H, Alix S, Migneault S, Koubaa A, Perré P. Numerical and experimental assessment of water absorption of wood-polymer composites. Measurement. 2018;115:197-203. DOI: 10.1016/j.measurement.2017.10.011
  80. 80. Jiang N, Li Y, Li D, Yu T, Li Y, Xu J, et al. 3D finite element modeling of water diffusion behavior of jute/PLA composite based on X-ray computed tomography. Composites Science and Technology. 2020;199:108313. DOI: 10.1016/j.compscitech.2020.108313
  81. 81. Sanjeevi S, Shanmugam V, Kumar S, Ganesan V, Sas G, Johnson DJ, et al. Effects of water absorption on the mechanical properties of hybrid natural fibre/phenol formaldehyde composites. Scientific Reports. 2021;11:13385. DOI: 10.1038/s41598-021-92457-9
  82. 82. Shen C-H, Springer GS. Moisture absorption and desorption of composite materials. Journal of Composite Materials. 1976;10:2-20. DOI: 10.1177/002199837601000101
  83. 83. Rangaraj SV, Smith LV. Effects of moisture on the durability of a wood/thermoplastic composite. Journal of Thermoplastic Composite Materials. 2000;13:140-161. DOI: 10.1177/089270570001300204
  84. 84. Cheour K, Assarar M, Scida D, Ayad R, Gong X-L. Effect of water ageing on the mechanical and damping properties of flax-fibre reinforced composite materials. Composite Structures. 2016;152:259-266. DOI: 10.1016/j.compstruct.2016.05.045
  85. 85. Sahu P, Gupta MK. Lowering in water absorption capacity and mechanical degradation of sisal/epoxy composite by sodium bicarbonate treatment and PLA coating. Polymer Composites. 2020;41:668-681. DOI: 10.1002/pc.25397
  86. 86. Kumosa L, Benedikt B, Armentrout D, Kumosa M. Moisture absorption properties of unidirectional glass/polymer composites used in composite (non-ceramic) insulators. Composites Part A: Applied Science and Manufacturing. 2004;35:1049-1063. DOI: 10.1016/j.compositesa.2004.03.008
  87. 87. Jiang X, Kolstein H, Bijlaard FSK. Moisture diffusion in glass–fiber-reinforced polymer composite bridge under hot/wet environment. Composites Part B: Engineering. 2013;45:407-416. DOI: 10.1016/j.compositesb.2012.04.067
  88. 88. Bhatti IS, Yan Z. Comparative hygrothermal ageing analysis of glass and carbon fiber polymer composites: Numerical and experimental study. In: 2015 12th International Bhurban Conference on Applied Sciences and Technology (IBCAST). 2015. pp. 28-33. DOI: 10.1109/IBCAST.2015.7058474
  89. 89. Joliff Y, Belec L, Chailan JF. Modified water diffusion kinetics in an unidirectional glass/fibre composite due to the interphase area: Experimental, analytical and numerical approach. Composite Structures. 2013;97:296-303. DOI: 10.1016/j.compstruct.2012.09.044
  90. 90. Joliff Y, Belec L, Heman MB, Chailan JF. Experimental, analytical and numerical study of water diffusion in unidirectional composite materials—Interphase impact. Computational Materials Science. 2012;64:141-145. DOI: 10.1016/j.commatsci.2012.05.029
  91. 91. Jain D, Mukherjee A. Three-dimensional hygromechanical analysis of fibre polymer composites: Effect of boundary conditions. Composites Part B: Engineering. 2016;90:203-211. DOI: 10.1016/j.compositesb.2015.12.011
  92. 92. Peret T, Clement A, Freour S, Jacquemin F. Numerical transient hygro-elastic analyses of reinforced Fickian and non-Fickian polymers. Composite Structures. 2014;116:395-403. DOI: 10.1016/j.compstruct.2014.05.026
  93. 93. Bond DA. Moisture diffusion in a fiber-reinforced composite: Part I—Non-Fickian transport and the effect of fiber spatial distribution. Journal of Composite Materials. 2005;39:2113-2141. DOI: 10.1177/0021998305052030
  94. 94. Nouri M, Tahlaiti M. A dual-scale numerical model for the diffusive behaviour prediction of biocomposites based on randomly oriented fibres. Construction Technologies and Architecture. 2022;1:584-587. DOI: 10.4028/www.scientific.net/CTA.1.584
  95. 95. Masoodi R, Pillai KM. A study on moisture absorption and swelling in bio-based jute-epoxy composites. Journal of Reinforced Plastics and Composites. 2012;31:285-294. DOI: 10.1177/0731684411434654
  96. 96. Mishra S, Verma J. Effect of compatibilizers on water absorption kinetics of polypropylene/wood flour foamed composites. Journal of Applied Polymer Science. 2006;101:2530-2537. DOI: 10.1002/app.23916
  97. 97. Joseph PV, Rabello MS, Mattoso LHC, Joseph K, Thomas S. Environmental effects on the degradation behaviour of sisal fibre reinforced polypropylene composites. Composites Science and Technology. 2002;62:1357-1372. DOI: 10.1016/S0266-3538(02)00080-5
  98. 98. Beg MDH, Pickering KL. Mechanical performance of Kraft fibre reinforced polypropylene composites: Influence of fibre length, fibre beating and hygrothermal ageing. Composites Part A: Applied Science and Manufacturing. 2008;39:1748-1755. DOI: 10.1016/j.compositesa.2008.08.003
  99. 99. Law TT, Ishak ZAM. Water absorption and dimensional stability of short Kenaf fiber-filled polypropylene composites treated with maleated polypropylene. Journal of Applied Polymer Science. 2011;120:563-572. DOI: 10.1002/app.33184
  100. 100. Pérez-Fonseca AA, Robledo-Ortíz JR, Moscoso-Sánchez FJ, Rodrigue D, González-Núñez R. Injection molded self-hybrid composites based on polypropylene and natural fibers. Polymer Composites. 2014;35:1798-1806. DOI: 10.1002/pc.22834
  101. 101. Beg MDH, Pickering KL. Reprocessing of wood fibre reinforced polypropylene composites. Part II: Hygrothermal ageing and its effects. Composites Part A: Applied Science and Manufacturing. 2008;39:1565-1571. DOI: 10.1016/j.compositesa.2008.06.002
  102. 102. Tajvidi M, Najafi SK, Moteei N. Long-term water uptake behavior of natural fiber/polypropylene composites. Journal of Applied Polymer Science. 2006;99:2199-2203. DOI: 10.1002/app.21892
  103. 103. Bergeret A, Krawczak P. Liaison renfort/matrice—Définition et caractérisation. Ref: TIP100WEB—“Plastiques et composites”. 2006. Available from: https://www.techniques-ingenieur.fr/base-documentaire/materiaux-th11/caracterisation-et-proprietes-d-usage-des-composites-42144210/liaison-renfort-matrice-am5305/ [Accessed: 19 July 2020]
  104. 104. Freund L. Étude du vieillissement hygrothermique des composites renforcés de fibres naturelles: Approche expérimentale et modélisation. n.d. p. 165
  105. 105. Wang X, Petrů M. Degradation of bending properties of flax fiber reinforced polymer after natural aging and accelerated aging. Construction and Building Materials. 2020;240:117909. DOI: 10.1016/j.conbuildmat.2019.117909

Written By

Nouri Mustapha and Tahlaiti Mahfoud

Submitted: 22 October 2021 Reviewed: 10 January 2022 Published: 02 November 2022