Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 7267 2023-06-13 20:41:17 |
2 format -1 word(s) 7266 2023-07-07 12:18:15 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Barreira-Pinto, R.; Carneiro, R.; Miranda, M.; Guedes, R.M. Polymer-Matrix Composites - Environmental Fatigue, Creep, Long-Term Durability. Encyclopedia. Available online: https://encyclopedia.pub/entry/45531 (accessed on 25 July 2024).
Barreira-Pinto R, Carneiro R, Miranda M, Guedes RM. Polymer-Matrix Composites - Environmental Fatigue, Creep, Long-Term Durability. Encyclopedia. Available at: https://encyclopedia.pub/entry/45531. Accessed July 25, 2024.
Barreira-Pinto, Rui, Rodrigo Carneiro, Mário Miranda, Rui Miranda Guedes. "Polymer-Matrix Composites - Environmental Fatigue, Creep, Long-Term Durability" Encyclopedia, https://encyclopedia.pub/entry/45531 (accessed July 25, 2024).
Barreira-Pinto, R., Carneiro, R., Miranda, M., & Guedes, R.M. (2023, June 13). Polymer-Matrix Composites - Environmental Fatigue, Creep, Long-Term Durability. In Encyclopedia. https://encyclopedia.pub/entry/45531
Barreira-Pinto, Rui, et al. "Polymer-Matrix Composites - Environmental Fatigue, Creep, Long-Term Durability." Encyclopedia. Web. 13 June, 2023.
Polymer-Matrix Composites - Environmental Fatigue, Creep, Long-Term Durability
Edit

Polymer-matrix composites are widely used in engineering applications. Yet, environmental factors impact their macroscale fatigue and creep performances significantly, owing to several mechanisms acting at the microstructure level. Seawater, due to a combination of high salinity and pressures, low temperature and biotic media present, also contributes to the acceleration of fatigue and creep damage. Similarly, other liquid corrosive agents penetrate into cracks induced by cyclic loading and cause dissolution of the resin and breakage of interfacial bonds. UV radiation either increases the crosslinking density or scissions chains, embrittling the surface layer of a given matrix. Temperature cycles close to the glass transition damage the fibre–matrix interface, promoting microcracking and hindering fatigue and creep performance. The microbial and enzymatic degradation of biopolymers is also studied, with the former responsible for metabolising specific matrices and changing their microstructure and/or chemical composition. 

polymer-matrix composites environmental fatigue creep degradation

1. Introduction

Polymer composites have long been a focus of intensive research, especially since they are, now more than ever, regarded as viable alternatives for many engineering applications. However, when alone, their properties, particularly mechanical strength and stiffness, do not usually completely satisfy the requirements of a given application, hence the need for reinforcements and the advent of polymer-matrix composites (PMCs). These reinforcements can be of multiple shapes, ranging from dispersed particles to fibres aligned along a preferential direction, effectively altering the behaviour of the material as a whole. They are all bonded together and held in place by a polymeric medium, i.e., a polymer matrix, responsible for transferring and distributing external loads in the composite and for protecting the reinforcements from the elements.
Despite their importance in withstanding mechanical loads, reinforcements barely account for any weathering and degradation protection, which falls entirely upon the matrix, nor do they dictate the fatigue and creep life of a composite.
Matrices can be either thermoset or thermoplastic, with the former including the vast majority of PMCs suited for structural applications, whereas the latter, whilst not so stiff and versatile, have seen a recent development driven by sustainability and recycling concerns. Additionally, renewable organic matrices, even if still thermoplastic, constitute on their own a class of materials with growing importance, predominantly in biological and medical applications, where they also experience fatigue and creep.
Accordingly, here verses on the influence of environmental effects on the fatigue and creep behaviour of a PMC, phenomena that, contrarily to the case of metals, are still not completely understood and are still a topic of dispute and prolific research among the scientific community. It aims at not only describing how several different environmental conditions interact with the micro-structure of a polymer but also at listing comprehensively the most important results found in the literature for typical polymer matrices.

2. Thermoset Polymers

Thermoset polymers are composed of chemically covalent bonds—cross-linked molecules in a network structure. On heating, they decompose—as opposed to softening—and do not undergo a liquid state. Once solidified by the cross-linking process, they cannot be reshaped. However, cross-linking results in great strength and rigidity owing to the restriction of the relative displacements of the macromolecules.

2.1. Epoxy

Epoxy resins are one of the major thermoset matrix materials, featuring two or more oxirane rings or epoxy groups as part of its chemical structure [1][2]. Fibre-reinforced epoxy (FRE) composites are extensively used in a broad spectrum of industrial applications such as aerospace, marine, and automotive, where durability, reliability and performance are required under hostile environments [3]. Hence, understanding its behaviour concerning fatigue and creep under various environmental conditions is pivotal in designing safe and dependable structures.

2.1.1. Water

The performance of FRE composites under moisture exposure is crucial, and it has been widely established that moisture has significant repercussions on the physical and chemical properties of such materials [3][4][5][6][7], owing to a large number of highly polar specific functional groups in the cured epoxy resin resulting in hydrophilic materials, i.e., with high water affinity [5]. Water uptake is correlated to mechanical performance losses mainly due to plasticising; however, hydrolysis reactions and water acting as a physical cross-link may occur as well [6][8]—refer to Section 2. The ramifications of such mechanisms are not confined to the static material properties and have been recorded to detrimentally affect the fatigue performance of epoxy-based composites. Tensile fatigue lives of glass/epoxy specimens were found to be reduced from 107 to 105 cycles when loaded at 65% of their ultimate tensile strength (UTS) immersed in distilled water at 25 C [9]. Hybrid glass-carbon/epoxy samples showed similar fatigue performances in wet and dry conditions of up to 45% UTS before a noticeable drop-off was registered [9]. Ageing of composite materials in aqueous environments also has prejudicial consequences on the mechanical performance: unidirectional carbon/epoxy laminates have been found to withstand more than 3 × 106 cycles at 80% of the ultimate flexural strength (UFS) when subjected to bending fatigue tests without preconditioning. Nevertheless, aged wet laminates—immersed in freshwater, seawater and seawater with 70 bar hydrostatic pressure at 50

2.1.2. Temperature

It is widely established that the mechanical properties and structural integrity of epoxy-based composites are significantly reduced [10][11] due to an immediate softening and rubber-like response at elevated temperatures—even only moderately above 𝑇𝑔[3]—restricting their use in high-temperature applications [12][13]. At temperatures considerably exceeding 𝑇𝑔, epoxy resins in general begin to decompose. The FRE composites initially undergo reversible mechanical changes with rising temperatures close to 𝑇𝑔, such as viscous softening [3]. As a consequence, the properties of both the polymeric resin and fibre–matrix interface can be adversely affected which is accompanied by a significant loss of mechanical performance [3]. This is evidenced in the fatigue life of 3D angle-interlock woven carbon/epoxy composites tested at 25 and 100 C, where a drop from 38% UTS to 30% UTS in the fatigue limit was observed with increasing temperature [14]—attributed to weakening in the interface bonding strength. Basalt/epoxy specimens were also recorded to have significant performance losses with increasing temperature concerning both fatigue and tensile properties—a linear decrease from −20 to 60 C with a reduction of 20% in UTS [15]. On the other hand, an increase in mechanical properties with decreasing temperature was also recorded. Mitsuhiro Okayasu and Yuki Tsuchiya [16] also reported an enhancement of the low-cycle fatigue performance by a factor of 1.5 on a unidirectional carbon/epoxy specimen at cryogenic temperatures. This factor decreases with the increasing number of cycles and at the later phase, the fatigue curve at −196 C is close to the one at room temperature. This could be attributed to a magnifying effect at low temperatures on chain movement inhibition due to high cross-link density as previously reported [17][18] and as stated by the authors.

2.1.3. Hygrothermal

Hygrothermal environments stem from the combined effects of water and temperature. In addition to the water uptake, a rise in temperature is capable of significantly accelerating the saturation of FRE due to a superior rate of diffusion. The enhanced chain mobility also exacerbates the drop in mechanical performance [3][19]. Irreversible chemical changes may also be induced in the material, i.e., that persist even after the composite is dried: a permanent performance loss of up to 17% has been reported on preconditioned glass/epoxy composites used in wind turbine blades—immersion in demineralised water at 50 C for 4800 h [20]. The combination of moisture and temperature also resulted in shorter fatigue lives by up to three orders of magnitude. The drop is fatigue performance is also evident on the fatigue tests of a wet epoxy resin—distilled water at 60 C for 2 months—for which the S-N curve was shifted by 20% whilst maintaining the same slope [21]. Hygrothermal ageing of glass/epoxy specimens in ambient, humid (40 C and 60% relative humidity, RH) and sub-zero specimens (−15 C) for 180 days resulted in a decrease from 106 cycles at 40% UTS to 5.6 × 105, 5.8 × 105 and 3.4×105 for each respective conditioning process [22]. Significant deterioration in fatigue behaviour was attributed to leaching of the glass fibres and hydrolysis of the epoxy resin by E. Vauthier et al. [23] on specimens aged in various temperature and humidity conditions. In addition, findings on the fatigue response of preconditioned natural FRE composites yielded similar outcomes. Immersion in distilled water at 70 C for 26 days was reported to induce noteworthy drops in the fatigue response of epoxy resins reinforced with flax fibres for the low-cycle range—for 20,000 cycles, the maximum stress is reduced from 196 to 148 MPa [24]. However, the obtained slopes of the fitted curves—9.4 for aged samples vs. 18.9 for unaged ones—point to an enhanced fatigue resistance for high numbers of cycles, seeing as both curves converge to 90 MPa for a lifetime of 5.3×106 cycles, and it has been proven that, for the material in question, the trend of the S-N curves is the same up to 108 cycles [24][25]. On non-woven flax/epoxy specimens severe degradation was also identified—at 50% loading level, the fatigue modulus loss increased to 70% for preconditioned samples in a water bath at 45 C from 25–30% for unaged ones [26].

2.1.4. Seawater

Through plasticisation and hydrolysis, seawater exposure in FRE composites can severely damage the thermophysical, chemical, and mechanical performance of the epoxy matrix [3][27]. Fatigue life predictions revealed that, for 107 cycles, basalt/epoxy composites aged in seawater for 6 weeks would undergo a reduction of 11% concerning the maximum applied stress [28]. Notched carbon/epoxy laminates—rectangular cross-sections with a central hole—suffered dire repercussions subsequent to immersion in natural and artificial seawater for 30 and 60 days in their low-cycle fatigue behaviour—for 104 cycles, the stress amplitude was 1.2 times higher for dry laminates [29]. Still, no discernible differences were observed in the fatigue behaviour regarding the immersion duration or the type of seawater and for high-cycle fatigue, the lifetime of the laminates was not affected—superior impact of holes and respective induced stress concentrations.

2.1.5. Liquid Chemical Agents and UV Radiation

FRE composites are often employed in outdoor applications where the material is subjected to corrosive medium and ultraviolet (UV) radiation in addition to moisture and temperature [3]. Concerning the former, alkaline solutions have been shown to severely diminish the fatigue life of pultruded carbon/epoxy specimens following a prolonged immersion for a 2-year period. Fatigue life predictions resulting from several methods, namely, S-N curves, Whitney’s method and Sendeckyj’s method, identified a reduction in the maximum allowable stresses for 2 million cycles from 69.28 to 60.75% UTS, 73.39 to 68.55% UTS and 65.00 to 54.41% UTS for each respective method [30]. E-glass/epoxy specimens aged in aqueous solutions of NAOH having concentrations of 5%, 15% and 25% for 1, 2, 3 and 4 months suffered a reduction of 21.36 and 62.38% in fatigue strength and life [31]. The same procedure applied to an HCl aqueous solution resulted in more severe losses of 41.95% and 97.19% [31]. On the other hand, the energy associated with the wavelengths—290 to 400 nm—of the solar radiation incident on the surface of the earth is comparable to the breakdown energies of covalent bonds of epoxy and consequently can result in discolourations, chain scission-induced microscopic degradation and reduction in mechanical properties [3][32]. The exposure of tri-axial carbon/epoxy laminates 313-bulb UV for 750 h had major consequences on fatigue life for stress ranges smaller than 500.0 MPa with a major decrease from above 105 cycles to below 104 cycles [33]. Exposure to 𝛾-rays corresponding to doses of 20, 100 and 200 kGy also detrimentally impacted the fatigue performance of glass/epoxy specimens, particularly at low-cycle fatigue (𝑁<104) where a gradual decrease in the maximum stress was identified with increasing irradiation doses. This also translated into fewer life cycles for a given stress value with increasing doses. Nonetheless, no drops in the fatigue life were detected at high-cycle fatigue stages (𝑁>104) [34].

2.2. Vinyl Ester

Vinyl ester resins originate from an addition reaction between and epoxy resin and an unsaturated carboxylic acid such as acrylic or methacrylic and boast great mechanical performance comparable to epoxy whilst simultaneously offering superb processability similar to a polyester resin [2].

2.2.1. Hygrothermal

The combined effect of moisture uptake and elevated temperatures has already been established to degrade the fatigue life of fibre-reinforced epoxy polymers. This is extended to vinyl ester resins, with immersion in water at 40 C for periods ranging from 30 to 600 days being proven to significantly deteriorate the fatigue performance of glass/vinyl ester laminates [35]. Still, while the laminates showed a greater loss in fatigue life at the superior stress levels—70%—the vinyl ester resin samples showed more degradation at 30% of the ultimate stress. The findings revealed that for brief exposure intervals, the fatigue behaviour for composite materials is independent of the moisture content. However, the impact becomes more apparent with prolonged exposure.

2.2.2. Seawater

Exposure to aqueous environments has been shown to severely impact the fatigue performance of glass/vinyl ester composite pipes [36]. The fatigue strength of corroded specimens due to salt water spray was found to be decreased from a stress amplitude of 24 MPa for 106 cycles to 15 MPa after exposure for 300 h. Immersion in oxygenated seawater also displayed detrimental consequences to fatigue performance with the effect being more prevalent with increasing exposure periods. This deterioration effect was particularly obvious in the low cycle zone, where specimens exposed for 10,000 h saw a reduction in fatigue life by a factor of 6. Exposure to 100% RU of up to 3000 and 10,000 h elicited a slight performance loss, with the repercussions being exacerbated in the low cycle region where the fatigue strength and life were respectively reduced by ∼2 MPa and a factor of 2. The impact of exposure to both fresh and saltwater on the fatigue durability of the glass fibre-reinforced/vinyl ester composite was also characterised by McBagonluri et al. [37]. Both environmental circumstances have been observed to result in a significant decrease in fatigue strength. The authors come to the conclusion that water intrusion in both freshwater and saltwater conditions causes the load-bearing fibres to gradually deteriorate, which leads to fatigue failure. The flexural fatigue life of aged—exposure to water and salt solutions containing mass fractions of either 5% NaCl or 10% NaCl for up to 6570 h—pultruded glass/vinyl ester composites was recorded to be severely reduced from beyond 107 cycles to within 106 to 107 cycles at 30% of the flexural strength of the specimens [38].
Similar results were reported for carbon fibre-reinforced vinyl ester resins, with fatigue lives of fully saturated woven carbon/vinyl ester—immersed in synthetic seawater for a period of 140 days—evidenced a decrease up to ∼62% across all strain ranges with this effect being more prevalent for higher strain ranges—low-cycle fatigue—where the average measured reduction is between 37% at 0.46% strain range and 90% at 0.6% strain range [39]. Fatigue lives of carbon/vinyl ester laminates tested immersed in seawater induced a staggering 71% degradation in the number of cycles to failure when compared to dry specimens, whereas simply preconditioning—immersion in seawater at 40 C for 6 months—resulted in a drop of 30% [40].

2.2.3. Liquid Chemical Agents

Under liquid chemically active environments, PMCs are expected to corrode and to suffer losses in their fatigue response. This is evident in the striking 58% decrease in the slope of the S-N curves for unaged basalt/vinyl ester tendons sprayed with artificial acid rain—pH of 2.5. This was attributed to the strong corrosive properties of the acid that, with penetration, induced chemical corrosion in the constituents of the PMC in addition to the internal stresses originating from fluid ingress [41].

2.3. Polyester

Polyester macromolecules are attained from the reaction between a diacid or dianhydride with a dihydroxy compound. Numerous types of sail and motor boats, fishing boats, and naval vessels extensively employ polyester resin-based fibre-reinforced composites in the construction of their hulls [2]. Hence, understanding their environmental fatigue in such an environment is of the utmost importance.

2.3.1. Water

Apart from seawater, moisture uptake has already been established to detrimentally impact the fatigue life of fibre-reinforced epoxy polymers. This is extended to polyester resins, wherein tensile fatigue lives of polyester-based GFRP laminates were recorded to decrease to one-ninth of the dry samples’ ones for a stress level of 40% UTS following a moisture uptake of 0.38% with respect to the initial weight of the material [42].

2.3.2. Seawater

The effects of seawater on polyester-reinforced polymer composites have also been well established [43][44]. Reductions in fatigue performance were recorded and established to be exacerbated with increasing immersion times for randomly oriented short glass fibre-reinforced polyester composites immersed in natural seawater for 90, 180 and 270 days [43]. Furthermore, fatigue lifetimes under a simulated marine environment for preconditioned—immersion for 30 days in a 3.5 wt% NaOH solution—non-crimp wound glass fibre-reinforced polyester composites were recorded to be significantly shortened. Still, the same failure mechanisms were detected for both dry specimens tested in air and preconditioned ones tested in artificial seawater—the S-N curves exhibited the same slope [44].

3. Thermoplastic Polymers

Thermoplastic matrix composites have long been used for several different varied and ordinary applications; however, there has been an increase in investment and development of technologies for their use in more demanding scenarios, such as the aerospace industry. They boast higher matrix toughness and impact resistance, faster and more flexible production cycles as well as a better recyclability, crucial considering a more sustainable industry. However, their needs for high-temperature processing, poor fibre impregnation and high sensibility to heat come as disadvantages, deeming their implementation difficult [45][46][47]. Contrarily to thermosets, thermoplastic polymers may present a crystalline phase, whose degree may vary accordingly to the cooling rate from the melting temperature [12]. Biron [46] divided a wide range of thermoplastics into three different types of usage: commodities, technical thermoplastics and speciality thermoplastics. In this section, a thermoplastic matrix from each type (PP, PA and PEEK) is presented and, for each polymer, its resistance to environmental factors reviewed from the available data in the literature.

3.1. Polypropylene

Polypropylene is a partially crystalline thermoplastic used for a wide range of applications (automotive, appliances, furniture, etc.) [48], featuring relatively good mechanical properties at ambient temperature as well as good dimensional stability, flame resistance and fatigue resistance [48][49]. Furthermore, PP has been appointed as a very promising matrix for the use of natural reinforcements [49].

3.1.1. Water

As previously mentioned, water absorption by polymer-matrix composites leads to mechanical and chemical changes that pose a problem to the life of several components. Polypropylene matrix composites have been documented to have their fatigue life reduced with an increase in moisture content [50][51]. Ayman et al. [51] conducted tests on wet short glass fibre-reinforced PP samples and concluded that there is a reduction of fatigue life when exposed to moisture.
This phenomenon is once again explained by the influence of water on the interfacial strength between fibre and matrix. The effect is reduced by increasing the reinforcement content; nevertheless, the fibres are also subject to corrosion, which leads to premature fracture with increasing water content [51].

3.1.2. Temperature

Similarly to other PMCs, PP matrix composites face a reduction of fatigue and creep life with an increase in temperature [52][53]. Viña et al. performed several fatigue tests, at different temperatures, on sheets of glass fibre-reinforced polypropylene.

3.1.3. Hygrothermal

Moisture absorption and temperature are related to each other more often than not. When exposed to higher temperatures, water diffusion is easier and enhances the moisture degradation capability. The influence of temperature on the moisture absorption of a hemp-reinforced PP  [110,111], at 80 C, the moisture uptake can be up to 3600% faster than when subject to ambient temperature. Naturally, the moisture uptake rates are highly dependent on the fibre content and nature [52][54][55], especially for natural fibres (hydrophilic by nature) [56].

3.1.4. Seawater

Polypropylene matrix composites are being considered for use in civil construction and others under marine environments. Being so, it is necessary to assess how the salinity affects the behaviour of such materials [57][58]. Dong et al. [57] have tested sheets of glass fibre-reinforced polypropylene underwater with different saline concentrations for a duration of 6 months. It was concluded that, under seawater, the water uptake increases for the same interval of time compared to distilled water. The corrosion of glass fibres was also evident, with fatigue and creep resistance naturally degraded by the influence of these mechanisms [57].
Even though the same effect was registered by other authors such as Najafi et al. [58] with wood–polypropylene samples, there is still limited research regarding the behaviour of polypropylene composites under saltwater environments, mainly due to the high dependence of the response on different fibre types and contents.

3.1.5. UV Radiation

The effect of UV radiation on the fatigue life of different polypropylene composites has been documented by Joseph et al. [59] in their research. An important reduction of the mechanical properties of PP composites when exposed to UV radiation for long periods was observed, which occurs mainly due to chain scission (photo-oxidation) and to the appearance of cracks due to chemi-crystallisation.
Nasri et al. [60] have also studied the influence of UV radiation on the mechanical properties of PP composites. Using specimens of flax or pine fibre-reinforced PP, the response to a flexural test after different exposure times was carried out. It was concluded that the longer the exposure time, the more the mechanical properties are degraded, both in stiffness and strength.

3.2. Polyamide

PA is a thermoplastic used as a matrix mainly for short-fibre-reinforced composites in a wide range of applications. The low density, good mechanical properties, heat stability and the wide range of processing techniques, including additive manufacture [61], make this material highly desirable for a vast number of engineering solutions [48][62]. Among the different polyamides available, the most used are PA6 and PA6,6 belonging to the aliphatic polyamide family, also known as nylons [48][63].

3.2.1. Water

It is widely established that polyamides present a significant hygroscopic behaviour [48][64]; therefore, some research has been conducted in order to evaluate the mechanical response of PA composites under aqueous environments. Barbouchi et al. [65] carried fatigue tests on short glass fibre-reinforced PA6,6 samples on atmospheres with different % RH. Besides registering a steep decrease in the glass transition temperature 𝑇𝑔, there was an important reduction of both stiffness and strength. Consequently, the fatigue behaviour was diminished by the water intake; however, the fatigue endurance limit was observed to be the same for both dry and wet samples. The same substantial reduction of mechanical properties was observed by several other authors [66][67][68]. Water uptake, up to 8.5 wt%, decreasing with the amount of fibre content, was found to have a significant effect on the fatigue of the PA composite samples. These changes are explained by the plasticising effect of water which acts mainly on the amorphous phase, predominant in the PA microstructure [67].

3.2.2. Temperature

Due to the high presence of an amorphous phase in some PA composites, mainly due to the high cooling rate verified during manufacturing, the mechanical response is expected to be highly affected by an increase in temperature. Kawai et al. [69] evaluated the fatigue response of short carbon fibre-reinforced PA6 samples at different temperatures (room temperature, 50 C, 70 C). It was concluded that the fatigue strength for the same number of cycles decreases with an increase in temperature. The results are on par with the substantial stiffness and strength reduction (approx. 50%) that were verified under higher temperatures. Monte et al. [70] documented a similar level of reduction in strength and fatigue resistance with higher temperatures, adding that the characterisation of tensile-compressive fatigue was very difficult due to instability problems, since at T = 130 C, the samples were immensely compliant. Hysteric heating was also verified by previous research to further decrease the fatigue performance of PA composites [70][71].

3.2.3. Hygrothermal

In the previous section, both temperature and water uptake were shown to be detrimental to the properties of PA matrix composites. In order to establish a relation between the two variables, Sang et al. [72] ran different tests to evaluate the moisture uptake with the temperature. Moisture uptake increased in a Fickian way reaching the plateau at the maximum moisture uptake. The same results were obtained by Lei et al. [73], who also documented an important degree of swelling and reduction of mechanical properties.
The use of PA matrix composites in hot and humid environments is then discouraged since the high content of the amorphous phase is heavily influenced by environmental conditions.

3.2.4. Seawater

In order to evaluate the effect of salt concentration in the moisture uptake or degradation of PA matrix composites, Haadar et al. [74] immersed short glass fibre-reinforced PA in different environments, distilled water and saline solution. The water uptake was observed to evolve in a similar way for both conditions until near the saturation level, where the presence of sodium chloride slightly decreased the water diffusivity. The impact on the mechanical properties was similar, with NaCl-immersed specimens retaining a slightly higher percentage of their dry properties than the ones immersed in distilled water. Therefore, despite inducing a small change, fatigue and creep behaviour between distilled and saltwater environments are not significantly different.

3.2.5. UV Radiation

The assessment of the response of polyamide composites to UV radiation is a necessity for natural light-exposed engineering applications. Markovičová et al. [75] conducted some experiments to evaluate the UV radiation effect on glass fibre-reinforced PA, concluding that, up to 750 h of exposure, there was a moderate loss of tensile strength and flexural stiffness; nonetheless, there was recovery after 1000 h, which was attributed to partial cross-linking of the matrix. Furthermore, Pinpathomrat et al. [76] observed that unreinforced PA6 had a significant loss of approximately 85% of tensile strength after being exposed for 7 days, yet the same behaviour was not obtained for the reinforced counterpart [77]. A clear influence of UV exposure on the fatigue and creep behaviour of PA matrix composites requires more research.

3.3. Polyetheretherketone

PEEK is a partially crystalline polymer belonging to the polyaryletherketone (PAEK) family, constituted by high-performance thermoplastics [12][78]. Its stability at high temperatures (300 C), compatibility with high-strength fibres, high strength, stiffness and good resistance to chemicals and radiation make it especially attractive for more demanding applications such as the aerospace and medical industries [12][48][78].

3.3.1. Water

Determining the behaviour of PEEK polymers with different water contents becomes a necessity when a designed component is made to last in any environment. After three weeks of immersion, a weight change on the uptake equilibrium on boiling water of 0.45% was recorded [79]. At 65% RH, equilibrium was attained only after 1000 h, with the specimen exhibiting a weight change of only 0.08%. After performing the fatigue test analysis, it was concluded that the effect of the moisture absorption had no relevant effects on the fatigue performance of the composite (expected behaviours due to the low moisture absorbance). This was further verified by other studies [80], which also appointed a low moisture uptake as the main reason behind the maintenance of the fatigue performance.

3.3.2. Temperature

The use of PEEK polymer-matrix composites for high performance requires the assessment of its behaviour under different temperatures. Studies have reported that an increase in temperature leads to a degradation of both longitudinal stiffness and strength (especially near the glass transition temperature) of the samples, mainly attributed to the movement of chains in the amorphous parts [81][82]. Consequently, there is a reduction of the fatigue life of the composite.
The same behaviour was also documented by Kawai et al. [82], particularly on the effect of temperature on the off-axis loading of unidirectional AS4/PEEK samples. The failure surfaces showed tearing of the matrix, probably due to the higher shear stresses acting along the off-axis direction. The effect of temperature on creep was also studied [83], and a rise in the tensile creep compliance with higher levels of temperature was registered, showing excellent creep resistance at room temperature everywhere but near the glass transition temperature (above 130 C), where creep compliance rises sharply [84][85]. Both the fatigue and the creep-temperature responses of PEEK composites are highly dependent on the matrix, mainly due to its viscoelastic behaviour at high temperatures [83][84][85].

3.3.3. Hygrothermal

Exposure to higher temperatures leads to a higher moisture intake, with the mechanical properties such as flexural strength experiencing a slight reduction, up to 15%, under 80 C at 85% RH [86]. Several other studies [79][87][88] concluded that the mechanical properties of PEEK composites subject to different temperatures and moisture contents do not degrade in a substantial way, not being, however, totally negligible.

3.3.4. Seawater

Boinard et al. [89] exposed samples of PEEK to different combinations of temperature and salt content (distilled water and brine) and concluded that neither the absorption nor the diffusion process are affected by the presence of sodium chloride. Schambron et al. [90] obtained the same results in their research on carbon-fibre/PEEK bone plates. Mechanical and fatigue tests after exposure to different ageing environments with several NaCl concentration levels also showed that the degradation of mechanical properties is hardly affected by immersion in water or saline environments at room temperature and that the fatigue life is not altered.

3.3.5. UV Radiation

PEEK matrices have shown little to no reaction, in mechanical properties, to UV exposure. Nakamura et al. [91] evaluated the mechanical properties of PEEK sheets after being exposed to UV radiation, and recorded no significant reduction in tensile strength. A slight increase resulting from the embrittlement has been registered due to the UV-induced crosslinking [92]. This increase in strength is inversely proportional to the static tensile stress applied to the specimens while being exposed. Niu et al. [93] registered a decrease of 7.5% of stiffness after 1560 h of alternating cycles between 8 h of UV irradiation of 1.55 W/m2 at 70 C and 4 h of dark condensation at 50 C, having identified fibre/matrix debonding near the surface. This reduction in stiffness may slightly compromise the fatigue life of PEEK composites, especially with higher degrees of crystallinity. Regardless, for a significant reduction of mechanical properties to happen, very demanding and specific conditions are needed, showing that PEEK has good resistance to UV radiation for the majority of applications, much due to its stable aromatic structure, which makes it completely inert to a wide range of chemical environments [94][95].

4. Biopolymers

As researchers thrive to reduce today’s society’s environmental footprint, adopting more and more sustainable practices, the polymer industry has concomitantly reacted with the introduction of the so-called biopolymers, also referred to as biodegradable or bio-based/biogenic [96]. Despite being interchangeably used, they entail different meanings: bio-based/biogenic points to the origin of the polymer or the raw material from which it was produced, specifying materials, generally, macromolecular polymers, that derive exclusively from renewable sources, whereas biodegradable depends only on the chemical structure of a polymer and refers to the recycling options available at the end of its service life, meaning they can be fully degraded by microorganisms or enzymes under ambient conditions (i.e., outside composting plants) within a reasonable time length resulting in nontoxic products. These two terms are, thus, independent of each other and should not be confused, e.g., polymers based on fossil resources can still be biodegradable, and bio-based polymers are not necessarily biodegradable [96][97].
Researchers focus on some of the most representative bio-based and biodegradable polymer matrices, namely starch—and its derivatives, particularly TPS, PLA and PHAs—the most commonly used being PHB. On one hand, starch and PLA are, by far, the most productive biodegradable plastics, accounting for 38.4% and 25.0% of the total biodegradable plastic capacity, according to European Bioplastics Data from 2019 [98][99], with PLA totalling 449,000 tons per year, its applications ranging from packaging to medical devices. On the other hand, PHAs, despite their small market, high production costs and complex extraction processes, are regarded as the most promising biodegradable polymer—with a market worth of USD 57 million in 2019 almost doubling to USD 98 million in 2021 (projected), being not only biodegradable and naturally produced, but also biocompatible with many organisms [98].
Additionally, to the best of the authors’ knowledge, there are very few works focusing on the environmental fatigue and creep behaviours of these polymers—let alone their composites—which comes as no surprise, considering that the shift of focus towards these materials happened only recently, and the experimental methods are yet to mature.

4.1. Polylactic Acid

PLA is a polyester obtained from agroresources that is chemically synthesised using monomers from the fermentation of plant-derived carbohydrates [97][100][101], which is nearly amorphous in its pure form. Due to its low glass transition temperature (50–80 C), in several applications, it is often reinforced with fibres—often organic, such as kenaf or sisal—or particles. As mentioned, very few studies have been performed regarding PLA mechanical, let alone environmental, fatigue behaviour [102][103][104][105], most of which focus on the properties of additive manufacturing specimens [106][107][108][109][110][111][112][113][114][115][116] and some in PLA composites [117][118] and creep behaviour [105][119].

4.1.1. Water

PLA boasts a low water and oxygen acceptability level [120][121], with degradation occurring due to hydrolysis of ester linkages followed by microbial breakdown [122][123]. This preferentially affects the amorphous regions of the polymer by erosion—only after a long time do the crystalline regions undergo hydrolysis—and the degree of crystallinity of PLA samples has been found to increase with degradation time [124]. Additionally, the pH value was coupled with the degradation rate, and the degradation products themselves contributed to altering the pH of the medium; more specifically, he decomposition rates increased with increasing pH in the range 3.4–7, yet all samples assessed, both semicrystalline and amorphous, attained the same pH value after about 20 days of hydrolysis [124].
D-PLA, despite being hydrolysable, is considered non-degradable, contrarily to the degradable L-PLA [97][125][126][127]. Several studies have reported a negligible (or even no) mass loss when immersed in sterile water [100][128], even immersed in a laboratory aquatic medium over a period of 112 days (constant light, 30 C) [97].
Regarding composites, moisture absorption has been correlated with the volume of fibres and voids in PLA/hemp composites [129], and although in the neat polymer a certain time is required for the water to reach the inner core, composite samples are quickly filled with water, leading to homogeneous hydrolysis [130]. Likewise, PLA/TiO2 nanocomposites were tested at (37 ± 1) C and pH = 7.4 over more than one year, and it was found that the TiO2 particles accelerated the degradation kinetics, with the hydrolysis starting at the interface between fillers and matrix [131]. It has been found for PLA/PEG gels that the loading frequency does not affect the degradation rate for blends with a low cross-link density, whereas a higher frequency is linked to faster degradation rates in highly cross-linked polymers [132].

4.1.2. Temperature

PLA has a glass transition temperature close to 60 C, below which the migration of moisture and microbial agents into the material is hampered by the increased stiffness and reduced mobility of the chains, slowing down its degradation rate [129]. Contrarily, temperatures above 50 C have been found to accelerate the abiotic hydrolysis of PLA [133][134][135][136], even if its biodegradation kinetics are generally considered slow, [137]—for a more comprehensive review, see Karamanlioglu et al. [138]. Temperature has been determined to be the main factor affecting biodegradation in microorganism-rich environments such as compost, soil and seawater with faster degradation rates registered therein than in their sterile counterparts [135], also accounting for the differences between the distinct environments—soil and aqueous media [97][134][136][139]. Samples tested at temperatures from 25 to 50 C exhibited degradation effects ranging from no damage after one year to a mass loss of more than 40% just after 5 weeks [135][136], or, similarly, a decrease of 55% in tensile strength over 11 months [140][141].
Focusing on composites, hydrophilic fillers, e.g., starch, kenaf fibres or wood–flour, have been reported to increase the thermal decomposition rate under composting conditions for as long as 90 days, e.g., from 60% to 80% when the starch content increased from 10% to 40%, the same occurring when the 𝑇𝑔is decreased by blending with other polymers [142][143][144]. In addition, in PLA/KBF and PLA/flax biocomposites, a larger loading was correlated with an accelerated decomposition, due to the increased micropore surface area which, in turn, increases the transporting channels inside the composite used by humidity and microorganisms [143].

4.1.3. Seawater

PLA has been found to degrade very slowly in seawater, due to the lack of effective microorganisms that only allows for a (slow) abiotic hydrolysis to occur [100]. The high salinity in seawater compared to distilled water delays the diffusion process and the degradation rate [98][145]. Even after 1 year, no visible degradation was reported for the most common type of PLA [98][128][145], and a weight loss of just 2.5% over 600 days was registered in a simulated marine environment [97][146].
Natural and static laboratory conditions have been compared, and the differences have been found negligible after a period of 10 weeks at 25 C for various degrees of crystallinity (apart from the higher weight loss associated with mechanical forces) with a slight increase in stiffness and strength during the first weeks [147][148]. Specimens tested in a French harbour for 6 months exhibited only a reduction below 10% in tensile strength with laboratory tests concluding that seawater temperature significantly affects the mechanical properties [149].
However, even at moderate temperatures (around 40 C), the tensile stiffness of PLA matrices in PLA/flax biocomposites has remained constant, with the stiffness and strength of the composite diminishing [150], behaviour attributed to wet ageing of the interfacial adhesion. For most organic reinforcements, the water uptake increases with the fibre content [151], leading, after long exposure periods, to the appearance of cracks in the fibre and in the matrix [150][152].
Finally, samples immersed in a phosphate-buffered saline solution have been tested and are shown to keep a strength above 200 MPa, from an initial bending strength of 280 MPa, for up to 25 weeks [130][153].

4.1.4. UV Radiation

UV radiation has been shown to reduce the toughness of PLA/PHB samples due to surface modification, as it promotes solvent-induced crystallisation [120]. In a different study, samples exposed to natural sunlight for 90 days, despite exhibiting no discernible cracks or defects, suffered a 15–25% decrease in molecular mass and an increase in storage modulus [129][154]. Mass losses, changes in crystalline structures and degradation of thermal and mechanical properties, particularly Young’s modulus, were also noticeable under both laboratory and natural UV irradiation [154].

4.1.5. Microorganisms

Microorganisms act in different environments, and throughout this section, researchers focus on natural soil and compost biotic media. Regarding the former, although PLA is considered naturally biodegradable, it hardly degrades in a natural environment, due to a lack of microorganisms capable of decomposing it [97][98][139][155][156][157], e.g., a weight loss of 5% after 180 days with no visible changes has been recorded by [97][158]. In compost, however, it takes about 6–9 months for its complete decomposition, especially since the temperature therein sits close to its 𝑇𝑔 (58–65 vs. 60 C, respectively), promoting hydrolysis [98][135][136][159]. At 65 C, samples have been recorded to suffer a mass loss of 34% after just 7 weeks [97][123], with further studies confirming the trend [100][135][160][161][162]. Moreover, the temperature has been shown to be the key factor affecting the polymer and the microbial communities present; for instance, in [163][164], the recorded degradation increased from 43% to 75% over roughly the same period when the temperature was increased from 37 C to 55 C.
Degradation happens in two steps: (1) hydrolysis—first chemical, as already described, and then enzymatic—to reduce the molecular weight followed by (2) metabolisation of the chain monomers by the microorganisms present [100][101][129]. As in the case of water uptake, microbial communities target predominantly the amorphous regions, since the polymer chains are more flexible there, increasing the crystallinity of the attacked regions and, thus, embrittling the polymer matrix [100][135][136][165].
Finally, in PLA matrix composites, the presence of the fillers usually favours moisture uptake and, hence, the decomposition by microorganisms [129][156], which, besides affecting the total mass as in [152], may also impact the overall mechanical, thermal and surface properties. Alterations in the microstructure, promoting embrittlement, and, with them, an increase in the glass transition temperature and crystallinity degree have been reported in the literature [129][156].

4.2. Polyhydroxyalkanoates

PHA—their most important representative being poly(hydroxybutyrate) (PHB)—are biobased and biodegradable polymers produced by soil bacteria, typically stiff and brittle, with low permeability to water and oxygen [97][100]. Taking a closer look at PHB, its microstructure consists of alternating layers of crystalline and amorphous domains, the latter being preferentially targeted by hydrolytic degradation.

4.2.1. Water

PHB is known to degrade in water, with the tensile strength decreasing significantly in the initial stage of degradation and slowing down as decomposition evolves and the rate at which decomposition occurs depending on the water inorganic composition and temperature [152][166]. Mass losses of 7% and 35% were recorded after 180 and 358 days of water exposure, respectively [167], and a value of 8.5% was measured for simulated conditions over one year [100][128].

4.2.2. Temperature

Temperature is an important factor when assessing the biodegradability of PHB composites [97], e.g., samples at 40 C exhibited a weight loss of 0.64% per day [168], whilst a 50% weight loss was registered for specimens kept at temperatures of up to 30 C during 50 days.

4.2.3. Seawater

PHB is totally biodegradable in seawater, experiencing, much unlike PLA, fast hydrolysis [98]; several studies have been reviewed in [98]. Surface erosion was generally considered the main initiator of degradation, promoted by the high availability of microorganisms capable of degrading these matrices—even if still slower than in compost, after which other mechanisms would act in the polymer matrix. Different studies have been performed with rather distinct outcomes: weight loss of 7% after 1 year at 25 C [128] against 42% after just 160 days at 29 C or even 60% after 35 days in [98][100]. One common observation, nevertheless, is that the degradation of the polymer chains is uniform and independent of the chain length [100]. Furthermore, after 140-day submergence in a marine environment, the degradation levels of both the amorphous and the crystalline phases were reportedly the same [169][170]. Finally, PHA fibres, albeit not exhibiting signs of hydrolysis even after 120 days exposed to a saline solution, contrasting against PHB/PHV ones, have been shown to affect the degradability of the polymer matrix itself due to their water uptake [171].

4.2.4. Microorganisms

Comparatively to seawater, these polymers degrade much faster in soil and compost [98] according to multiple studies [129][172][173][174][175][176][177]. The involved kinetics depend strongly on temperature, as it influences the biotic activity, leading to distinct rates over a period of 200 days, namely 0.05%/day at 15 C, 0.12%/day at 28 C and 0.45%/day at 40 C, as reported by [100]. Likewise, a mass loss of 98% after one year has been measured by [100]. These results have been replicated for composting conditions [100], with results ranging from a mass loss of 6% for temperatures of 6–32 C over 150 days all the way to complete degradation at 58 C after 28 days. Degradation occurs preferentially in the amorphous regions, being accelerated by applied strains (both static, creep-inducing, and cyclic), inducing a change in semi-crystalline domains and increasing the amorphous fraction [178]. Finally, regarding composites, the incorporation of TiO2 particles has been reported to slow the degradation under sediment burial [175].

4.3. Thermoplastic-Starch

TPS is a biobased and biodegradable polymer(-matrix) boasting an amorphous structure that is synthesised by adding natural plasticisers to starch obtained from agricultural resources—such as potatoes, corn and rice, all rich in polysaccharides—leading to a thermoplastic polymer [97][100]. Typically, TPS-based composites exhibit low mechanical strength and impact resistance, being typically brittle and affected by water uptake [179].

4.3.1. Water

Starch alone possesses many polar groups that react forming hydrogen bonds, leading to high water absorption [100]. This, nonetheless, can be offset by different additives, both in the neat polymer and as reinforcements in composites [97]. Looking specifically at composites, cellulose fibres have been recorded to enhance both mechanical and thermal properties, whilst decreasing the water uptake [180][181].

4.3.2. Seawater

TPS has proven to be degradable in seawater, albeit slowly, with several blends and composites registering weight loss and deterioration of tensile properties almost exclusively in the starch fraction [182][183], as occurred in, e.g., PHB/TPS blends immersed in coastal water in Puerto Rico over a period of 1 year or PE/TPS blends exposed in laboratory tests simulating harbour conditions in the Baltic Sea over 20 months [182][184][185]. Enzymatic hydrolysis has been reported to be the main contributing mechanism, which, in composites, is favoured by the increased water uptake associated with most reinforcements, and by high levels of strain that augment the surface area, thus helping the microbial communities in seawater to infiltrate the composite. Moreover, as expected, the rate of degradation depends on the microorganisms present: in river water (highest number of bacteria), a weight loss of 32% was reported, decreasing to 3.3% in a marine environment, with studies indicating rates as high as 2% day1[98].

4.3.3. Microorganisms

Starch and TPS composites are known to degrade quickly under the influence of biotic factors, with studies performed both in soil burial and composting conditions [186][187][188][189][190]. Regarding the former, values of 100% and 72.6% mass loss have been reported in the literature for a period of 30 days at 30 C [191] and 25 C [192], respectively, whereas for the latter, an unexpected longer time was measured in several studies [193][194], taking up to 84 days at 60 C to fully decompose. A low starch content is, additionally, prone to slowing fatigue crack propagation and diminishing the incidence of microbial attack, with a high starch content promoting faster degradation and creep kinetics due to an overall increase in the penetration of water, light, oxygen and microbial attack. Contrasting results have, nevertheless, been found in the literature, e.g., with studies reporting a long time for complete degradation—as long as 12 years for some TPS/PE blends under composting conditions [97][123].

References

  1. Chawla, K.K. Composite Materials: Science and Engineering; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2012.
  2. Ratna, D. Handbook of Thermoset Resins; ISmithers: Shawbury, UK, 2009.
  3. Najafi, M.; Eslami-Farsani, R.; Saeedi, A.; Ebrahimnezhad-Khaljiri, H. The Effect of Environmental Conditions on the Synthetic Fiber-Reinforced Epoxy Composites. In Handbook of Epoxy/Fiber Composites; Springer: Berlin/Heidelberg, Germany, 2022; pp. 1–51.
  4. Pérez-Pacheco, E.; Cauich-Cupul, J.; Valadez-González, A.; Herrera-Franco, P. Effect of moisture absorption on the mechanical behavior of carbon fiber/epoxy matrix composites. J. Mater. Sci. 2013, 48, 1873–1882.
  5. Lee, M.C.; Peppas, N.A. Water transport in epoxy resins. Prog. Polym. Sci. 1993, 18, 947–961.
  6. Apicella, A.; Nicolais, L. Effect of water on the properties of epoxy matrix and composite. In Epoxy Resins and Composites I; Springer: Berlin/Heidelberg, Germany, 1985; pp. 69–77.
  7. Roy, S. Moisture-induced degradation. In Long-Term Durability of Polymeric Matrix Composites; Springer: Berlin/Heidelberg, Germany, 2012; pp. 181–236.
  8. Li, L.; Zhang, S.; Chen, Y.; Liu, M.; Ding, Y.; Luo, X.; Pu, Z.; Zhou, W.; Li, S. Water transportation in epoxy resin. Chem. Mater. 2005, 17, 839–845.
  9. Shan, Y.; Liao, K. Environmental fatigue behavior and life prediction of unidirectional glass–carbon/epoxy hybrid composites. Int. J. Fatigue 2002, 24, 847–859.
  10. Taylor, R.; Bogaard, R. Thermophysical Properties of Epoxy and Phenolic Composites; ACS Publications: Washington, DC, USA, 1995.
  11. Gupta, V.; Drzal, L.; Lee, C.C.; Rich, M. The temperature-dependence of some mechanical properties of a cured epoxy resin system. Polym. Eng. Sci. 1985, 25, 812–823.
  12. Mouritz, A.P. Introduction to Aerospace Materials; Elsevier: Amsterdam, The Netherlands, 2012.
  13. Irving, P.E.; Soutis, C. Polymer Composites in the Aerospace Industry; Woodhead Publishing: Sawston, UK, 2019.
  14. Li, D.s.; Dang, M.g.; Jiang, L. Elevated temperature effect on tension fatigue behavior and failure mechanism of carbon/epoxy 3D angle-interlock woven composites. Compos. Struct. 2021, 268, 113897.
  15. Zhao, X.; Wang, X.; Wu, Z.; Keller, T.; Vassilopoulos, A.P. Temperature effect on fatigue behavior of basalt fiber-reinforced polymer composites. Polym. Compos. 2019, 40, 2273–2283.
  16. Okayasu, M.; Tsuchiya, Y. Mechanical and fatigue properties of long carbon fiber reinforced plastics at low temperature. J. Sci. Adv. Mater. Devices 2019, 4, 577–583.
  17. Nettles, A.T.; Biss, E.J. Low Temperature Mechanical Testing of Carbon-Fiber/Epoxy-Resin Composite Materials; Technical Report; NASA: Washington, DC, USA, 1996.
  18. Callaghan, M. Use of resin composites for cryogenic tankage. Cryogenics 1991, 31, 282–287.
  19. Choi, S.; Douglas, E.P. Complex hygrothermal effects on the glass transition of an epoxy-amine thermoset. ACS Appl. Mater. Interfaces 2010, 2, 934–941.
  20. Rocha, I.B.C.M.; Raijmaekers, S.; Nijssen, R.; Van der Meer, F.; Sluys, L.J. Hygrothermal ageing behaviour of a glass/epoxy composite used in wind turbine blades. Compos. Struct. 2017, 174, 110–122.
  21. Krauklis, A.E.; Gagani, A.I.; Echtermeyer, A.T. Hygrothermal aging of amine epoxy: Reversible static and fatigue properties. Open Eng. 2018, 8, 447–454.
  22. Padmaraj, N.; Vijaya, K.M.; Dayananda, P. Experimental investigation on fatigue behaviour of glass/epoxy quasi-isotropic laminate composites under different ageing conditions. Int. J. Fatigue 2021, 143, 105992.
  23. Vauthier, E.; Abry, J.; Bailliez, T.; Chateauminois, A. Interactions between hygrothermal ageing and fatigue damage in unidirectional glass/epoxy composites. Compos. Sci. Technol. 1998, 58, 687–692.
  24. Jeannin, T.; Berges, M.; Gabrion, X.; Léger, R.; Person, V.; Corn, S.; Piezel, B.; Ienny, P.; Fontaine, S.; Placet, V. Influence of hydrothermal ageing on the fatigue behaviour of a unidirectional flax-epoxy laminate. Compos. Part B Eng. 2019, 174, 107056.
  25. Jeannin, T.; Gabrion, X.; Ramasso, E.; Placet, V. About the fatigue endurance of unidirectional flax-epoxy composite laminates. Compos. Part B Eng. 2019, 165, 690–701.
  26. Habibi, M.; Laperrière, L.; Hassanabadi, H.M. Effect of moisture absorption and temperature on quasi-static and fatigue behavior of nonwoven flax epoxy composite. Compos. Part B Eng. 2019, 166, 31–40.
  27. Mourad, A.H.I.; Abdel-Magid, B.M.; El-Maaddawy, T.; Grami, M.E. Effect of seawater and warm environment on glass/epoxy and glass/polyurethane composites. Appl. Compos. Mater. 2010, 17, 557–573.
  28. Wang, X.; Zhao, X.; Wu, Z. Fatigue degradation and life prediction of basalt fiber-reinforced polymer composites after saltwater corrosion. Mater. Des. 2019, 163, 107529.
  29. Branco, R.; Reis, P.N.; Neto, M.A.; Costa, J.D.; Amaro, A.M. Seawater Effect on Fatigue Behaviour of Notched Carbon/Epoxy Laminates. Appl. Sci. 2021, 11, 11939.
  30. Wu, P.; Xu, L.; Luo, J.; Zhang, X.; Bian, W. Influences of long-term immersion of water and alkaline solution on the fatigue performances of unidirectional pultruded CFRP plate. Constr. Build. Mater. 2019, 205, 344–356.
  31. Balcıoğlu, H.E.; Sakin, R.; Gün, H. The effect of different environmental condition on flexural strength and fatigue behavior of e-glass/epoxy composites. Iran. J. Sci. Technol. Trans. Mech. Eng. 2021, 45, 165–180.
  32. Kumar, B.G.; Singh, R.P.; Nakamura, T. Degradation of carbon fiber-reinforced epoxy composites by ultraviolet radiation and condensation. J. Compos. Mater. 2002, 36, 2713–2733.
  33. Mosallam, A.; Xin, H.; He, S.; Agwa, A.A.; Adanur, S.; Salama, M.A. Thermal cycling and ultraviolet radiation effects on fatigue performance of triaxial CFRP laminates for bridge applications. J. Compos. Mater. 2022, 56, 279–294.
  34. Zheng, L.F.; Wang, L.N.; Wang, Z.Z.; Wang, L. Effects of γ-ray irradiation on the fatigue strength, thermal conductivities and thermal stabilities of the glass fibres/epoxy resins composites. Acta Metall. Sin. Engl. Lett. 2018, 31, 105–112.
  35. Acosta, F.A.; Roman, R.; Pando, M.A.; Godoy, L.A. Fatigue Strength of Composite Materials Considering Hygrothermal Degradation. Mec. Comput. 2016, 34, 41–60.
  36. Khan, Z. Degradation of fatigue resistance of filament wound glass fibre reinforced/vinyl ester pipes exposed to aqueous environments. Plast. Rubber Compos. 2011, 40, 397–401.
  37. McBagonluri, F.; Garcia, K.; Hayes, M.; Verghese, K.; Lesko, J. Characterization of fatigue and combined environment on durability performance of glass/vinyl ester composite for infrastructure applications. Int. J. Fatigue 2000, 22, 53–64.
  38. Liao, K.; Schultheisz, C.R.; Hunston, D.L. Long-term environmental fatigue of pultruded glass-fiber-reinforced composites under flexural loading. Int. J. Fatigue 1999, 21, 485–495.
  39. Prabhakar, P.; Garcia, R.; Imam, M.A.; Damodaran, V. Flexural fatigue life of woven carbon/vinyl ester composites under sea water saturation. Int. J. Fatigue 2020, 137, 105641.
  40. Siriruk, A.; Penumadu, D. Degradation in fatigue behavior of carbon fiber–vinyl ester based composites due to sea environment. Compos. Part B Eng. 2014, 61, 94–98.
  41. Liu, X.; Li, F.; Wang, X. Synergistic effect of acidic environmental exposure and fatigue loads on FRP tendons. Constr. Build. Mater. 2022, 314, 125584.
  42. Ferdous, W.; Manalo, A.; Yu, P.; Salih, C.; Abousnina, R.; Heyer, T.; Schubel, P. Tensile fatigue behavior of polyester and vinyl ester based GFRP laminates—A comparative evaluation. Polymers 2021, 13, 386.
  43. Djeghader, D.; Redjel, B. Fatigue resistance of randomly oriented short glass fiber reinforced polyester composite materials immersed in seawater environment. Mech. Ind. 2017, 18, 604.
  44. Altunsaray, E.; Neser, G.; Erbil, C.; Gürsel, K.; Ünsalan, D. Environmental fatigue behavior of non-crimp, E-glass fiber reinforced polyester composites for marine applications. Mater. Werkst. 2012, 43, 1053–1058.
  45. Van Rijswijk, K.; Bersee, H. Reactive processing of textile fiber-reinforced thermoplastic composites—An overview. Compos. Part A Appl. Sci. Manuf. 2007, 38, 666–681.
  46. Biron, M. Chapter 6—Thermoplastic Composites. In Thermoplastics and Thermoplastic Composites, 3rd ed.; Biron, M., Ed.; Plastics Design Library, William Andrew Publishing: Norwich, NY, USA, 2018; pp. 821–882.
  47. Grigore, M.E. Methods of Recycling, Properties and Applications of Recycled Thermoplastic Polymers. Recycling 2017, 2, 24.
  48. Da Silva, L. Materiais De Construção; Publindústria: Porto, Portugal, 2013.
  49. Shubhra, Q.T.; Alam, A.; Quaiyyum, M. Mechanical properties of polypropylene composites: A review. J. Thermoplast. Compos. Mater. 2013, 26, 362–391.
  50. Kanny, K.; Mohan, T. Surface treatment of sisal fiber composites for improved moisture and fatigue properties. Compos. Interfaces 2013, 20, 783–797.
  51. Abdelhaleem, A.M.; Megahed, M.; Saber, D. Fatigue behavior of pure polypropylene and recycled polypropylene reinforced with short glass fiber. J. Compos. Mater. 2018, 52, 1633–1640.
  52. Bledzki, A.K.; Faruk, O. Creep and impact properties of wood fibre–polypropylene composites: Influence of temperature and moisture content. Compos. Sci. Technol. 2004, 64, 693–700.
  53. Viña, J.; Argüelles, A.; Canteli, A.F. Influence of Temperature on the Fatigue Behaviour of Glass Fibre Reinforced Polypropylene. Strain 2011, 47, 222–226.
  54. Lin, Q.; Zhou, X.; Dai, G. Effect of hydrothermal environment on moisture absorption and mechanical properties of wood flour–filled polypropylene composites. J. Appl. Polym. Sci. 2002, 85, 2824–2832.
  55. Panthapulakkal, S.; Sain, M. Studies on the Water Absorption Properties of Short Hemp—Glass Fiber Hybrid Polypropylene Composites. J. Compos. Mater. 2007, 41, 1871–1883.
  56. Saheb, D.N.; Jog, J.P. Natural fiber polymer composites: A review. Adv. Polym. Technol. 1999, 18, 351–363.
  57. Dong, S.; Zhou, P.; Guo, R.; Li, C.; Xian, G. Durability study of glass fiber reinforced polypropylene sheet under simulated seawater sea sand concrete environment. J. Mater. Res. Technol. 2022, 20, 1079–1092.
  58. Kazemi, S.; Younesi-Kordkheili, H. Effect of sea water on water absorption and flexural properties of wood-polypropylene composites. Holz Roh-Werkstoff 2011, 69, 553–556.
  59. Joseph, P.; Rabello, M.S.; Mattoso, L.; Joseph, K.; Thomas, S. Environmental effects on the degradation behaviour of sisal fibre reinforced polypropylene composites. Compos. Sci. Technol. 2002, 62, 1357–1372.
  60. Nasri, K.; Toubal, L.; Loranger, É.; Koffi, D. Influence of UV irradiation on mechanical properties and drop-weight impact performance of polypropylene biocomposites reinforced with short flax and pine fibers. Compos. Part C Open Access 2022, 9, 100296.
  61. Dul, S.; Fambri, L.; Pegoretti, A. High-Performance Polyamide/Carbon Fiber Composites for Fused Filament Fabrication: Mechanical and Functional Performances. J. Mater. Eng. Perform. 2021, 30, 5066–5085.
  62. Kausar, A. Advances in Carbon Fiber Reinforced Polyamide-Based Composite Materials. Adv. Mater. Sci. 2019, 19, 67–82.
  63. Botelho, E.; Figiel, Ł.; Rezende, M.; Lauke, B. Mechanical behavior of carbon fiber reinforced polyamide composites. Compos. Sci. Technol. 2003, 63, 1843–1855.
  64. De Moura, M.S.F.; de Morais, A.; de Magalhães, A. Materiais Compósitos: Materiais, Fabrico e Comportamento Mecânico; Edições Técnicas, Publindústria: Porto, Portugal, 2005.
  65. Barbouchi, S.; Bellenger, V.; Tcharkhtchi, A.; Castaing, P.; Jollivet, T. Effect of water on the fatigue behaviour of a pa66/glass fibers composite material. J. Mater. Sci. 2007, 42, 2181–2188.
  66. Hassan, A.; Abd Rahman, N.M.M.; Yahya, R. Moisture Absorption Effect on Thermal, Dynamic Mechanical and Mechanical Properties of Injection-Molded Short Glass-Fiber/Polyamide 6,6 Composites. Fibers Polym. 2012, 13, 899–906.
  67. Mortazavian, S.; Fatemi, A.; Khosrovaneh, A. Effect of Water Absorption on Tensile and Fatigue Behaviors of Two Short Glass Fiber Reinforced Thermoplastics. SAE Int. J. Mater. Manuf. 2015, 8, 435–443.
  68. Chaichanawong, J.; Thongchuea, C.; Areerat, S. Effect of moisture on the mechanical properties of glass fiber reinforced polyamide composites. Adv. Powder Technol. 2016, 27, 898–902.
  69. Kawai, M.; Takeuchi, H.; Taketa, I.; Tsuchiya, A. Effects of temperature and stress ratio on fatigue life of injection molded short carbon fiber-reinforced polyamide composite. Compos. Part A Appl. Sci. Manuf. 2017, 98, 9–24.
  70. De Monte, M.; Moosbrugger, E.; Quaresimin, M. Influence of temperature and thickness on the off-axis behaviour of short glass fibre-reinforced polyamide 6.6–cyclic loading. Compos. Part A Appl. Sci. Manuf. 2010, 41, 1368–1379.
  71. Lin, S.H.; Ma, C.C.M.; Tai, N.H. Long fiber reinforced polyamide and polycarbonate composites. II: Fatigue behavior and morphological property. J. Vinyl Addit. Technol. 1996, 2, 80–86.
  72. Sang, L.; Wang, C.; Wang, Y.; Hou, W. Effects of hydrothermal aging on moisture absorption and property prediction of short carbon fiber reinforced polyamide 6 composites. Compos. Part B Eng. 2018, 153, 306–314.
  73. Lei, Y.; Zhang, T.; Zhang, J.; Zhang, B. Dimensional stability and mechanical performance evolution of continuous carbon fiber reinforced polyamide 6 composites under hygrothermal environment. J. Mater. Res. Technol. 2021, 13, 2126–2137.
  74. Haddar, N.; Ksouri, I.; Kallel, T.; Mnif, N. Effect of hygrothermal ageing on the monotonic and cyclic loading of glass fiber reinforced polyamide. Polym. Compos. 2014, 35, 501–508.
  75. Markovičová, L.; Zatkalíková, V.; Garbacz, T. Effect of UV Radiation to Change the Properties of the Composite PA+GF. World Acad. Sci. Eng. Technol. Int. J. Chem. Mol. Nucl. Mater. Metall. Eng. 2015, 9, 1224–1227.
  76. Pinpathomrat, B.; Yamada, K.; Yokoyama, A. The effect of UV irradiation on polyamide 6/carbon-fiber composites based on three-dimensional printing. SN Appl. Sci. 2020, 2, 1–9.
  77. Pillay, S.; Vaidya, U.K.; Janowski, G.M. Effects of moisture and UV exposure on liquid molded carbon fabric reinforced nylon 6 composite laminates. Compos. Sci. Technol. 2009, 69, 839–846.
  78. Kurtz, S.M. Chapter 1—An Overview of PEEK Biomaterials. In PEEK Biomaterials Handbook, 2nd ed.; Kurtz, S.M., Ed.; Plastics Design Library, William Andrew Publishing: Norwich, NY, USA, 2019; pp. 3–9.
  79. Dickson, R.; Jones, C.; Harris, B.; Leach, D.; Moore, D. The environmental fatigue behaviour of carbon fibre-reinforced polyether ether ketone. J. Mater. Sci. 1985, 20, 60–70.
  80. Selzer, R.; Friedrich, K. Mechanical properties and failure behaviour of carbon fibre-reinforced polymer composites under the influence of moisture. Compos. A Part Appl. Sci. Manuf. 1997, 28, 595–604.
  81. Jen, M.H.R.; Tseng, Y.C.; Kung, H.K.; Huang, J. Fatigue response of APC-2 composite laminates at elevated temperatures. Compos. Part B Eng. 2008, 39, 1142–1146.
  82. Kawai, M.; Yajima, S.; Hachinohe, A.; Kawase, Y. High-temperature off-axis fatigue behaviour of unidirectional carbon-fibre-reinforced composites with different resin matrices. Compos. Sci. Technol. 2001, 61, 1285–1302.
  83. Maksimov, R.D.; Kubát, J. Time and temperature dependent deformation of poly(ether ether ketone) (PEEK). Mech. Compos. Mater. 1997, 33, 517–525.
  84. Xiao, X.; Hiel, C.; Cardon, A. Characterization and modeling of nonlinear viscoelastic response of PEEK resin and PEEK composites. Compos. Eng. 1994, 4, 681–702.
  85. Katouzian, M.; Bruller, O.S.; Horoschenkoff, A. On the Effect of Temperature on the Creep Behavior of Neat and Carbon Fiber Reinforced PEEK and Epoxy Resin. J. Compos. Mater. 1995, 29, 372–387.
  86. Ma, C.C.M.; Lee, C.L.; Chang, M.J.; Tai, N.H. Hygrothermal behavior of carbon fiber-reinforced poly(ether ether ketone) and poly(phenylene sulfide) composites. I. Polym. Compos. 1992, 13, 448–453.
  87. Ma, C.C.M.; Yur, S.W. Environmental effects on the water absorption and mechanical properties of carbon fiber reinforced PPS and PEEK composites. Part II. Polym. Eng. Sci. 1991, 31, 34–39.
  88. Bismarck, A.; Hofmeier, M.; Dörner, G. Effect of hot water immersion on the performance of carbon-reinforced unidirectional poly(ether ether ketone) (PEEK) composites: Stress rupture under end-loaded bending. Compos. Part A Appl. Sci. Manuf. 2007, 38, 407–426.
  89. Boinard, E.; Pethrick, R.; MacFarlane, C. The influence of thermal history on the dynamic mechanical and dielectric studies of polyetheretherketone exposed to water and brine. Polymer 2000, 41, 1063–1076.
  90. Schambron, T.; Lowe, A.; McGregor, H.V. Effects of environmental ageing on the static and cyclic bending properties of braided carbon fibre/PEEK bone plates. Compos. Part B Eng. 2008, 39, 1216–1220.
  91. Nakamura, H.; Nakamura, T.; Noguchi, T.; Imagawa, K. Photodegradation of PEEK sheets under tensile stress. Polym. Degrad. Stab. 2006, 91, 740–746.
  92. Batista, N.; Rezende, M.; Botelho, E.C. The Influence of Crystallinity on the Weather Resistance of CF/PEEK Composites. Appl. Compos. Mater. 2021, 28, 235–246.
  93. Niu, Y.F.; Yang, Y.; Li, T.Y.; Yao, J.W. Effects of UV irradiation and condensation on poly(ether-ether-ketone)/carbon fiber composites from nano- to macro-scale. High Perform. Polym. 2018, 30, 230–238.
  94. Kurtz, S.M.; Devine, J.N. PEEK biomaterials in trauma, orthopedic, and spinal implants. Biomaterials 2007, 28, 4845–4869.
  95. Kurtz, S.M. Chapter 6—Chemical and Radiation Stability of PEEK. In PEEK Biomaterials Handbook; Kurtz, S.M., Ed.; Plastics Design Library, William Andrew Publishing: Oxford, UK, 2012; pp. 75–79.
  96. Koltzenburg, S.; Maskos, M.; Nuyken, O. Polymer Chemistry; Springer: Berlin/Heidelberg, Germany, 2017.
  97. Lambert, S.; Wagner, M. Environmental performance of bio-based and biodegradable plastics: The road ahead. Chem. Soc. Rev. 2017, 46, 6855–6871.
  98. Wang, G.X.; Huang, D.; Ji, J.H.; Völker, C.; Wurm, F.R. Seawater-Degradable Polymers—Fighting the Marine Plastic Pollution. Adv. Sci. 2021, 8, 2001121.
  99. European-Bioplastics. Bioplastics Market Data 2019. Available online: https://www.european-bioplastics.org/wp-content/uploads/2019/11/Report_Bioplastics-Market-Data_2019_short_version.pdf (accessed on 15 December 2022).
  100. Kliem, S.; Kreutzbruck, M.; Bonten, C. Review on the biological degradation of polymers in various environments. Materials 2020, 13, 4586.
  101. Kolstad, J.J.; Vink, E.T.; De Wilde, B.; Debeer, L. Assessment of anaerobic degradation of Ingeo™ polylactides under accelerated landfill conditions. Polym. Degrad. Stab. 2012, 97, 1131–1141.
  102. Rizvi, S.H.A.; Che, J.; Mehboob, A.; Zaheer, U.; Chang, S.H. Experimental study on magnesium wire–polylactic acid biodegradable composite implants under in vitro material degradation and fatigue loading conditions. Compos. Struct. 2021, 272, 114267.
  103. Zhang, J.; Hirschberg, V.; Rodrigue, D. Mechanical fatigue of biodegradable polymers: A study on polylactic acid (PLA), polybutylene succinate (PBS) and polybutylene adipate terephthalate (PBAT). Int. J. Fatigue 2022, 159, 106798.
  104. Algarni, M. Fatigue Behavior of PLA Material and the Effects of Mean Stress and Notch: Experiments and Modeling. Procedia Struct. Integr. 2022, 37, 676–683.
  105. Liber-Kneć, A.; Kuźniar, P.; Kuciel, S. Accelerated fatigue testing of biodegradable composites with flax fibers. J. Polym. Environ. 2015, 23, 400–406.
  106. Azadi, M.; Dadashi, A. Experimental fatigue dataset for additive-manufactured 3D-printed Polylactic acid biomaterials under fully reversed rotating-bending bending loadings. Data Brief 2022, 41, 107846.
  107. Bagheri, A.; Aghareb Parast, M.S.; Kami, A.; Azadi, M.; Asghari, V. Fatigue testing on rotary friction-welded joints between solid ABS and 3D-printed PLA and ABS. Eur. J. Mech.—A/Solids 2022, 96, 104713.
  108. Mayén, J.; Del Carmen Gallegos-Melgar, A.; Pereyra, I.; Poblano-Salas, C.A.; Hernández-Hernández, M.; Betancourt-Cantera, J.; Mercado-Lemus, V.; Del Angel Monroy, M. Descriptive and inferential study of hardness, fatigue life, and crack propagation on PLA 3D-printed parts. Mater. Today Commun. 2022, 32, 103948.
  109. Azadi, M.; Dadashi, A.; Dezianian, S.; Kianifar, M.; Torkaman, S.; Chiyani, M. High-cycle bending fatigue properties of additive-manufactured ABS and PLA polymers fabricated by fused deposition modeling 3D-printing. Forces Mech. 2021, 3, 100016.
  110. Ezeh, O.; Susmel, L. On the notch fatigue strength of additively manufactured polylactide (PLA). Int. J. Fatigue 2020, 136, 105583.
  111. Jerez-Mesa, R.; Travieso-Rodriguez, J.; Llumà-Fuentes, J.; Gomez-Gras, G.; Puig, D. Fatigue lifespan study of PLA parts obtained by additive manufacturing. Procedia Manuf. 2017, 13, 872–879.
  112. Ezeh, O.; Susmel, L. On the fatigue strength of 3D-printed polylactide (PLA). Procedia Struct. Integr. 2018, 9, 29–36.
  113. Ezeh, O.; Susmel, L. Fatigue strength of additively manufactured polylactide (PLA): Effect of raster angle and non-zero mean stresses. Int. J. Fatigue 2019, 126, 319–326.
  114. Senatov, F.; Niaza, K.; Stepashkin, A.; Kaloshkin, S. Low-cycle fatigue behavior of 3d-printed PLA-based porous scaffolds. Compos. Part B Eng. 2016, 97, 193–200.
  115. Gomez-Gras, G.; Jerez-Mesa, R.; Travieso-Rodriguez, J.A.; Lluma-Fuentes, J. Fatigue performance of fused filament fabrication PLA specimens. Mater. Des. 2018, 140, 278–285.
  116. Ezeh, O.; Susmel, L. Fatigue behaviour of additively manufactured polylactide (PLA). Procedia Struct. Integr. 2018, 13, 728–734.
  117. Shahar, F.S.; Hameed Sultan, M.T.; Safri, S.N.A.; Jawaid, M.; Abu Talib, A.R.; Basri, A.A.; Md Shah, A.U. Fatigue and impact properties of 3D printed PLA reinforced with kenaf particles. J. Mater. Res. Technol. 2022, 16, 461–470.
  118. Travieso-Rodriguez, J.A.; Zandi, M.D.; Jerez-Mesa, R.; Lluma-Fuentes, J. Fatigue behavior of PLA-wood composite manufactured by fused filament fabrication. J. Mater. Res. Technol. 2020, 9, 8507–8516.
  119. Soares, J.S.; Moore, J.E., Jr.; Rajagopal, K.R. Constitutive framework for biodegradable polymers with applications to biodegradable stents. Asaio J. 2008, 54, 295–301.
  120. Rasal, R.M.; Hirt, D.E. Toughness decrease of PLA-PHBHHx blend films upon surface-confined photopolymerization. J. Biomed. Mater. Res. Part A Off. J. Soc. Biomater. Jpn. Soc. Biomater. Aust. Soc. Biomater. Korean Soc. Biomater. 2009, 88, 1079–1086.
  121. Hiljanen-Vainio, M.; Varpomaa, P.; Seppälä, J.; Törmälä, P. Modification of poly (L-lactides) by blending: Mechanical and hydrolytic behavior. Macromol. Chem. Phys. 1996, 197, 1503–1523.
  122. Auras, R.A.; Lim, L.T.; Selke, S.E.; Tsuji, H. Poly (Lactic Acid): Synthesis, Structures, Properties, Processing, and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2011.
  123. Mulbry, W.; Reeves, J.B.; Millner, P. Use of mid-and near-infrared spectroscopy to track degradation of bio-based eating utensils during composting. Bioresour. Technol. 2012, 109, 93–97.
  124. Gorrasi, G.; Pantani, R. Effect of PLA grades and morphologies on hydrolytic degradation at composting temperature: Assessment of structural modification and kinetic parameters. Polym. Degrad. Stab. 2013, 98, 1006–1014.
  125. Shogren, R.; Doane, W.; Garlotta, D.; Lawton, J.; Willett, J. Biodegradation of starch/polylactic acid/poly (hydroxyester-ether) composite bars in soil. Polym. Degrad. Stab. 2003, 79, 405–411.
  126. Kim, E.Y.; Lee, J.K.; Lee, W.K. Hydrolytic kinetics of langmuir monolayers of enantiomeric poly (lactide)s. Curr. Appl. Phys. 2006, 6, 735–738.
  127. Tokiwa, Y.; Calabia, B.P. Biodegradability and biodegradation of poly (lactide). Appl. Microbiol. Biotechnol. 2006, 72, 244–251.
  128. Bagheri, A.R.; Laforsch, C.; Greiner, A.; Agarwal, S. Fate of so-called biodegradable polymers in seawater and freshwater. Glob. Challenges 2017, 1, 1700048.
  129. Brebu, M. Environmental degradation of plastic composites with natural fillers—A review. Polymers 2020, 12, 166.
  130. Mano, J.F.; Sousa, R.A.; Boesel, L.F.; Neves, N.M.; Reis, R.L. Bioinert, biodegradable and injectable polymeric matrix composites for hard tissue replacement: State of the art and recent developments. Compos. Sci. Technol. 2004, 64, 789–817.
  131. Luo, Y.B.; Wang, X.L.; Wang, Y.Z. Effect of TiO2 nanoparticles on the long-term hydrolytic degradation behavior of PLA. Polym. Degrad. Stab. 2012, 97, 721–728.
  132. Nicodemus, G.; Shiplet, K.; Kaltz, S.; Bryant, S. Dynamic compressive loading influences degradation behavior of PEG-PLA hydrogels. Biotechnol. Bioeng. 2009, 102, 948–959.
  133. Tokiwa, Y.; Calabia, B.P.; Ugwu, C.U.; Aiba, S. Biodegradability of plastics. Int. J. Mol. Sci. 2009, 10, 3722–3742.
  134. Kawai, F. The biochemistry and molecular biology of xenobiotic polymer degradation by microorganisms. Biosci. Biotechnol. Biochem. 2010, 74, 1743–1759.
  135. Karamanlioglu, M.; Robson, G.D. The influence of biotic and abiotic factors on the rate of degradation of poly (lactic) acid (PLA) coupons buried in compost and soil. Polym. Degrad. Stab. 2013, 98, 2063–2071.
  136. Pranamuda, H.; Tokiwa, Y.; Tanaka, H. Polylactide degradation by an Amycolatopsis sp. Appl. Environ. Microbiol. 1997, 63, 1637–1640.
  137. Hottle, T.A.; Agüero, M.L.; Bilec, M.M.; Landis, A.E. Alkaline amendment for the enhancement of compost degradation for polylactic acid biopolymer products. Compost. Sci. Util. 2016, 24, 159–173.
  138. Karamanlioglu, M.; Preziosi, R.; Robson, G.D. Abiotic and biotic environmental degradation of the bioplastic polymer poly (lactic acid): A review. Polym. Degrad. Stab. 2017, 137, 122–130.
  139. Ohkita, T.; Lee, S.H. Thermal degradation and biodegradability of poly (lactic acid)/corn starch biocomposites. J. Appl. Polym. Sci. 2006, 100, 3009–3017.
  140. Rudnik, E.; Briassoulis, D. Degradation behaviour of poly (lactic acid) films and fibres in soil under Mediterranean field conditions and laboratory simulations testing. Ind. Crop. Prod. 2011, 33, 648–658.
  141. Ho, K.L.G.; Pometto, A.L.; Gadea-Rivas, A.; Briceño, J.A.; Rojas, A. Degradation of polylactic acid (PLA) plastic in Costa Rican soil and Iowa state university compost rows. J. Environ. Polym. Degrad. 1999, 7, 173–177.
  142. Petinakis, E.; Liu, X.; Yu, L.; Way, C.; Sangwan, P.; Dean, K.; Bateman, S.; Edward, G. Biodegradation and thermal decomposition of poly (lactic acid)-based materials reinforced by hydrophilic fillers. Polym. Degrad. Stab. 2010, 95, 1704–1707.
  143. Wu, C.S. Characterizing Biodegradation of PLA and PLA-g-AA/Starch Films Using a Phosphate-Solubilizing Bacillus Species. Macromol. Biosci. 2008, 8, 560–567.
  144. Othman, M.; Nor Azowa, I.; Ruzaidi, C.M.; Mohd Nazarudin, Z.; Halim, Z. Biodegradability analysis of KBF reinforced poly (lactic acid) biocomposites. In Advanced Materials Research; Trans Tech Publ: Stafa-Zurich, Switzerland, 2012; Volume 576, pp. 434–437.
  145. Gexia, W.; Dan, H.; Wei, Z.; Junhui, J. Degradation performance of typical biodegradable polyesters in seawater. J. Funct. Polym. 2020, 33, 492–499.
  146. Pelegrini, K.; Donazzolo, I.; Brambilla, V.; Coulon Grisa, A.M.; Piazza, D.; Zattera, A.J.; Brandalise, R.N. Degradation of PLA and PLA in composites with triacetin and buriti fiber after 600 days in a simulated marine environment. J. Appl. Polym. Sci. 2016, 133.
  147. Tsuji, H.; Suzuyoshi, K. Environmental degradation of biodegradable polyesters 2. Poly (ε-caprolactone), poly , and poly (L-lactide) films in natural dynamic seawater. Polym. Degrad. Stab. 2002, 75, 357–365.
  148. Tsuji, H.; Suzuyoshi, K. Environmental degradation of biodegradable polyesters 1. Poly (ε-caprolactone), poly , and poly (L-lactide) films in controlled static seawater. Polym. Degrad. Stab. 2002, 75, 347–355.
  149. Deroiné, M.; Le Duigou, A.; Corre, Y.M.; Le Gac, P.Y.; Davies, P.; César, G.; Bruzaud, S. Accelerated ageing of polylactide in aqueous environments: Comparative study between distilled water and seawater. Polym. Degrad. Stab. 2014, 108, 319–329.
  150. Le Duigou, A.; Davies, P.; Baley, C. Seawater ageing of flax/poly (lactic acid) biocomposites. Polym. Degrad. Stab. 2009, 94, 1151–1162.
  151. Bledzki, A.; Jaszkiewicz, A.; Murr, M.; Sperber, V.; Lützendgrf, R.; Reußmann, T. Processing techniques for natural-and wood-fibre composites. In Properties and Performance of Natural-Fibre Composites; Elsevier: Amsterdam, The Netherlands, 2008; pp. 163–192.
  152. Faruk, O.; Bledzki, A.K.; Fink, H.P.; Sain, M. Biocomposites reinforced with natural fibers: 2000–2010. Prog. Polym. Sci. 2012, 37, 1552–1596.
  153. Shikinami, Y.; Okuno, M. Bioresorbable devices made of forged composites of hydroxyapatite (HA) particles and poly L-lactide (PLLA): Part I. Basic characteristics. Biomaterials 1999, 20, 859–877.
  154. Mosnáčková, K.; Šlosár, M.; Kollár, J.; Janigová, I.; Šišková, A.; Chmela, Š.; Sikorska, W.; Perd’ochová, D.; Gálisová, I.; Alexy, P.; et al. Ageing of plasticized poly (lactic acid)/poly (3-hydroxybutyrate)/carbon black mulching films during one season of sweet pepper production. Eur. Polym. J. 2019, 114, 81–89.
  155. Suyama, T.; Tokiwa, Y.; Ouichanpagdee, P.; Kanagawa, T.; Kamagata, Y. Phylogenetic affiliation of soil bacteria that degrade aliphatic polyesters available commercially as biodegradable plastics. Appl. Environ. Microbiol. 1998, 64, 5008–5011.
  156. Vasile, C.; Pamfil, D.; Râpă, M.; Darie-Niţă, R.N.; Mitelut, A.C.; Popa, E.E.; Popescu, P.A.; Draghici, M.C.; Popa, M.E. Study of the soil burial degradation of some PLA/CS biocomposites. Compos. Part B Eng. 2018, 142, 251–262.
  157. Urayama, H.; Kanamori, T.; Kimura, Y. Properties and biodegradability of polymer blends of poly (l-lactide) s with different optical purity of the lactate units. Macromol. Mater. Eng. 2002, 287, 116–121.
  158. Song, J.; Murphy, R.; Narayan, R.; Davies, G. Biodegradable and compostable alternatives to conventional plastics. Philos. Trans. R. Soc. B Biol. Sci. 2009, 364, 2127–2139.
  159. Laycock, B.; Nikolić, M.; Colwell, J.M.; Gauthier, E.; Halley, P.; Bottle, S.; George, G. Lifetime prediction of biodegradable polymers. Prog. Polym. Sci. 2017, 71, 144–189.
  160. Rudeekit, Y.; Numnoi, J.; Tajan, M.; Chaiwutthinan, P.; Leejarkpai, T. Determining biodegradability of polylactic acid under different environments. J. Met. Mater. Miner 2008, 18, 83–87.
  161. Sikorska, W.; Musiol, M.; Nowak, B.; Pajak, J.; Labuzek, S.; Kowalczuk, M.; Adamus, G. Degradability of polylactide and its blend with poly in industrial composting and compost extract. Int. Biodeterior. Biodegrad. 2015, 101, 32–41.
  162. Spiridon, I.; Darie, R.N.; Kangas, H. Influence of fiber modifications on PLA/fiber composites. Behavior to accelerated weathering. Compos. Part B Eng. 2016, 92, 19–27.
  163. Yagi, H.; Ninomiya, F.; Funabashi, M.; Kunioka, M. Mesophilic anaerobic biodegradation test and analysis of eubacteria and archaea involved in anaerobic biodegradation of four specified biodegradable polyesters. Polym. Degrad. Stab. 2014, 110, 278–283.
  164. Yagi, H.; Ninomiya, F.; Funabashi, M.; Kunioka, M. Thermophilic anaerobic biodegradation test and analysis of eubacteria involved in anaerobic biodegradation of four specified biodegradable polyesters. Polym. Degrad. Stab. 2013, 98, 1182–1187.
  165. Luzi, F.; Fortunati, E.; Puglia, D.; Petrucci, R.; Kenny, J.; Torre, L. Study of disintegrability in compost and enzymatic degradation of PLA and PLA nanocomposites reinforced with cellulose nanocrystals extracted from Posidonia Oceanica. Polym. Degrad. Stab. 2015, 121, 105–115.
  166. Voinova, O.; Gladyshev, M.; Volova, T.G. Comparative study of PHA degradation in natural reservoirs having various types of ecosystems. In Macromolecular Symposia; Wiley Online Library: Weinheim, Germany, 2008; Volume 269, pp. 34–37.
  167. Mergaert, J.; Wouters, A.; Swings, J.; Anderson, C. In situ biodegradation of poly (3-hydroxybutyrate) and poly (3-hydroxybutyrate-co-3-hydroxyvalerate) in natural waters. Can. J. Microbiol. 1995, 41, 154–159.
  168. Mergaert, J.; Webb, A.; Anderson, C.; Wouters, A.; Swings, J. Microbial degradation of poly (3-hydroxybutyrate) and poly (3-hydroxybutyrate-co-3-hydroxyvalerate) in soils. Appl. Environ. Microbiol. 1993, 59, 3233–3238.
  169. Volova, T.; Boyandin, A.; Vasil’ev, A.; Karpov, V.; Kozhevnikov, I.; Prudnikova, S.; Rudnev, V.; Xuån, B.B.; Dũng, V.V.; Gitel’Zon, I. Biodegradation of polyhydroxyalkanoates (PHAs) in the South China Sea and identification of PHA-degrading bacteria. Microbiology 2011, 80, 252–260.
  170. Reis, K.; Pereira, J.; Smith, A.; Carvalho, C.; Wellner, N.; Yakimets, I. Characterization of polyhydroxybutyrate-hydroxyvalerate (PHB-HV)/maize starch blend films. J. Food Eng. 2008, 89, 361–369.
  171. Mohammadi, M.; Ghaffari-Moghaddam, M. Recovery and extraction of polyhydroxyalkanoates (PHAs). Polyhydroxyalkanoate (PHA) Based Blends Compos. Nanocompos. 2014, 30, 47.
  172. Boyandin, A.N.; Prudnikova, S.V.; Karpov, V.A.; Ivonin, V.N.; Đỗ, N.L.; Nguyễn, T.H.; Lê, T.M.H.; Filichev, N.L.; Levin, A.L.; Filipenko, M.L.; et al. Microbial degradation of polyhydroxyalkanoates in tropical soils. Int. Biodeterior. Biodegrad. 2013, 83, 77–84.
  173. Saad, G.R.; Khalil, T.M.; Sabaa, M.W. Photo-and bio-degradation of poly (ester-urethane) s films based on poly and poly (ε-Caprolactone) blocks. J. Polym. Res. 2010, 17, 33–42.
  174. Sadi, R.K.; Fechine, G.J.; Demarquette, N.R. Photodegradation of poly (3-hydroxybutyrate). Polym. Degrad. Stab. 2010, 95, 2318–2327.
  175. Sridewi, N.; Bhubalan, K.; Sudesh, K. Degradation of commercially important polyhydroxyalkanoates in tropical mangrove ecosystem. Polym. Degrad. Stab. 2006, 91, 2931–2940.
  176. Savenkova, L.; Gercberga, Z.; Nikolaeva, V.; Dzene, A.; Bibers, I.; Kalnin, M. Mechanical properties and biodegradation characteristics of PHB-based films. Process Biochem. 2000, 35, 573–579.
  177. Mergaert, J.; Anderson, C.; Wouters, A.; Swings, J.; Kersters, K. Biodegradation of polyhydroxyalkanoates. FEMS Microbiol. Rev. 1992, 9, 317–321.
  178. Ruka, D.R.; Sangwan, P.; Garvey, C.J.; Simon, G.P.; Dean, K.M. Biodegradability of poly 3-hydroxybutyrate/bacterial cellulose composites under aerobic conditions, measured via evolution of carbon dioxide and spectroscopic and diffraction methods. Environ. Sci. Technol. 2015, 49, 9979–9986.
  179. Kunioka, M.; Kawaguchi, Y.; Doi, Y. Production of biodegradable copolyesters of 3-hydroxybutyrate and 4-hydroxybutyrate by Alcaligenes eutrophus. Appl. Microbiol. Biotechnol. 1989, 30, 569–573.
  180. Wattanakornsiri, A.; Pachana, K.; Kaewpirom, S.; Sawangwong, P.; Migliaresi, C. Green composites of thermoplastic corn starch and recycled paper cellulose fibers. Songklanakarin J. Sci. Technol. 2011, 33, 461–467.
  181. Jha, K.; Kataria, R.; Verma, J.; Pradhan, S. Potential biodegradable matrices and fiber treatment for green composites: A review. AIMS Mater. Sci. 2019, 6, 119–138.
  182. Guzman-Sielicka, A.; Janik, H.; Sielicki, P. Degradation of Polycaprolactone Modified with TPS or CaCO3 in Biotic/Abiotic Seawater. J. Polym. Environ. 2012, 20, 353–360.
  183. Guzman, A.M.; Janik, H.Z.; Mastalerz, M.; Kosakowska, A.M. Pilot study of the influence of thermoplastic starch based polymer packaging material on the growth of diatom population in sea water environment. Pol. J. Chem. Technol. 2011, 13, 57–61.
  184. Iman, S.; Gordon, S.; Shogren, R.; Tosteson, T.R.; Govind, N.S.; Green, R.V. Degradation of Strchpoly (β-Hydoxybutyrate-co-β-Hydroxyvalerate) bioplastic in tropical coastal waters. Appl. Environ. Microbiol 1999, 65, 431–437.
  185. Rutkowska, M.; Heimowska, A.; Krasowska, K.; Janik, H. Biodegradability of polyethylene starch blends in sea water. Pol. J. Environ. Stud. 2002, 11, 267–272.
  186. Vaverková, M.; Toman, F.; Adamcová, D.; Kotovicová, J. Study of the biodegrability of degradable/biodegradable plastic material in a controlled composting environment. Ecol. Chem. Eng. 2012, 19, 347.
  187. Torres, F.; Troncoso, O.; Torres, C.; Díaz, D.; Amaya, E. Biodegradability and mechanical properties of starch films from Andean crops. Int. J. Biol. Macromol. 2011, 48, 603–606.
  188. Bastioli, C.; Cerutti, A.; Guanella, I.; Romano, G.; Tosin, M. Physical state and biodegradation behavior of starch-polycaprolactone systems. J. Environ. Polym. Degrad. 1995, 3, 81–95.
  189. Di Franco, C.; Cyras, V.P.; Busalmen, J.P.; Ruseckaite, R.A.; Vázquez, A. Degradation of polycaprolactone/starch blends and composites with sisal fibre. Polym. Degrad. Stab. 2004, 86, 95–103.
  190. Mohee, R.; Unmar, G. Determining biodegradability of plastic materials under controlled and natural composting environments. Waste Manag. 2007, 27, 1486–1493.
  191. Bootklad, M.; Kaewtatip, K. Biodegradation of thermoplastic starch/eggshell powder composites. Carbohydr. Polym. 2013, 97, 315–320.
  192. Zain, A.H.M.; Ab Wahab, M.K.; Ismail, H. Biodegradation behaviour of thermoplastic starch: The roles of carboxylic acids on cassava starch. J. Polym. Environ. 2018, 26, 691–700.
  193. Vikman, M.; Itävaara, M.; Poutanen, K. Measurement of the biodegradation of starch-based materials by enzymatic methods and composting. J. Environ. Polym. Degrad. 1995, 3, 23–29.
  194. Vikman, M.; Itävaara, M.; Poutanen, K. Biodegradation of starch-based materials. J. Macromol. Sci. Part A Pure Appl. Chem. 1995, 32, 863–866.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , ,
View Times: 706
Revisions: 2 times (View History)
Update Date: 07 Jul 2023
1000/1000
Video Production Service