Next Article in Journal
Effects of the Face/Core Layer Ratio on the Mechanical Properties of 3D Printed Wood/Polylactic Acid (PLA) Green Biocomposite Panels with a Gyroid Core
Previous Article in Journal
Characterization of Microcrystalline Cellulose Isolated from Conocarpus Fiber
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Monomer Selection for In Situ Polymerization Infusion Manufacture of Natural-Fiber Reinforced Thermoplastic-Matrix Marine Composites

Faculty of Science and Engineering, School of Engineering, Computing & Mathematics, University of Plymouth, Drake Circus, Plymouth PL4 8AA, UK
*
Author to whom correspondence should be addressed.
Polymers 2020, 12(12), 2928; https://doi.org/10.3390/polym12122928
Submission received: 10 November 2020 / Revised: 28 November 2020 / Accepted: 30 November 2020 / Published: 7 December 2020
(This article belongs to the Section Polymer Composites and Nanocomposites)

Abstract

:
Awareness of environmental issues has led to increasing interest from composite researchers in using “greener” materials to replace synthetic fiber reinforcements and petrochemical polymer matrices. Natural fiber bio-based thermoplastic composites could be an appropriate choice with advantages including reducing environmental impacts, using renewable resources and being recyclable. The choice of polymer matrix will significantly affect the cost, manufacturing process, mechanical properties and durability of the composite system. The criteria for appropriate monomers are based on the processing temperature and viscosity, polymer mechanical properties, recyclability, etc. This review considers the selection of thermoplastic monomers suitable for in situ polymerization during resin, now monomer, infusion under flexible tooling (RIFT, now MIFT), with a primary focus on marine composite applications. Given the systems currently available, methyl methacrylate (MMA) may be the most suitable monomer, especially for marine composites. MMA has low process temperatures, a long open window for infusion, and low moisture absorption. However, end-of-life recovery may be limited to matrix depolymerization. Bio-based MMA is likely to become commercially available in a few years. Polylactide (PLA) is an alternative infusible monomer, but the relatively high processing temperature may require expensive consumable materials and could compromise natural fiber properties.

Graphical Abstract

1. Introduction

Fiber-reinforced composite materials have been widely used for marine applications (marine renewable energy devices, offshore oil/gas infrastructure, boat hulls, etc.) due to their superior resistance to biological and chemical attack in the harsh marine environment [1,2]. There is now an increasing trend to replace conventional synthetic fibers and petrochemical resins with renewable and recyclable materials for composite production [3]. Special attention to environmental issues should be taken for composites in marine applications due to the risk of microplastic contamination of the marine ecosystem [4,5]. Natural fibers (e.g., bamboo, coir, flax, hemp, jute, sisal, etc.) have attracted much interest as alternatives to conventional glass/carbon fibers in recent years [6]. There is limited literature on natural fiber-reinforced plastics (NFRP) composites in the marine environment, but the reviews by Moudood et al. [7], Al-Maharma and Al-Huniti [8] and Célino et al. [9] do consider the effects of water on NFRP. The advantages of natural fiber reinforcements include being renewable, biodegradable and low cost with high specific modulus and strength. However, the change from synthetic to natural fibers does incur a number of new considerations. There are over 100 review papers [10] addressing bast fibers and their composites published in the past 10 years, with 16 [11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26] identified with a 2020 publication date. Further, any reader seeking more information may consider our series of reviews that specifically cover fibers [27], organisms and enzymes [28], forensics [29], composites [30], modeling [31], interfaces [32], manufacture [33] and life cycle assessment (LCA) [34].
For the matrix in composites, thermoset resin composites become a problem at end-of-life, especially for large structures. Investigations into thermoplastics, as replacements for thermosets, are growing rapidly as these composites can be reprocessed, although fiber lengths are normally reduced during recycling. After pyrolysis, useable carbon fibers can be reclaimed from either matrix system, but the mechanical properties of glass fibers are severely reduced at high temperatures. Thomason [35] has suggested the glass fibers can be reprocessed to restore the properties, albeit with additional environmental burdens. Natural fiber composites at end-of-life have the possibility of incineration with energy recovery or disposal by composting with methane collection if the matrix is amenable to that option [36,37,38,39]. Bio-based or biodegradable thermoplastics are (potentially) available and could offer the opportunity for greater reduction of environmental burdens [40,41,42].
Thermoplastic composites have relatively high impact resistance [43,44,45]. Sutherland comprehensively reviewed the impact testing of marine composites [46,47,48]. In the impact testing of boat-building materials [49], thermoplastics demonstrated high specific impact strength compared with other materials due to the high level of plastic deformation. Otheguy [50] studied the impact behavior of a rigid inflatable boat manufactured with thermoplastic composites (glass fiber reinforced polypropylene) using a severe impact loading of 32 kJ (500 kg steel weight dropped from 6.5 m). The boat hull largely recovered its original shape with little chipping and no cracking at locations remote from the perforated region, which was significantly different from brittle thermosetting composites. An emergency repair method (fusion bonding repair) was successfully developed to maintain structural integrity and water tightness.
The composite manufacturing process, especially liquid composite molding (LCM), could make a significant contribution to the reduction of environmental impact. The resin transfer molding (RTM) approach uses a matched pair of mold tools to produce composite structures with high fiber volume fraction (high mechanical properties) [51,52,53,54,55,56,57]. RTM is viewed as an economical and efficient process with low styrene emission [58]. For larger structures, commercial manufacture of large structures uses (thermoset) resin infusion under flexible tooling (RIFT) [59,60,61,62,63], and the process is being adapted for (thermoplastic) monomer infusion under flexible tooling (MIFT). The hierarchy of processes is LCM > RTM > RIFT ≅ MIFT and the highest appropriate level in this list is used in the text below.
The production of thermoplastic matrix composites using the polymer in LCM is impractical due to the high melt viscosity. For LCM impregnation, in situ polymerization of a liquid monomer could be practical where the system has appropriate viscosity [64]. This review paper discusses the selection of thermoplastic monomers with the potential for producing natural fiber-reinforced marine composites via in situ polymerization infusion during MIFT. Key considerations are monomer initial viscosity, process temperature, fiber/matrix interfacial compatibility and bonding, availability of bio-sourced monomer, mechanical and thermal properties of the product, moisture uptake, environmental burdens and recyclability.

2. Resin Infusion under Flexible Tooling (RIFT)

RIFT [59,60,61,62,63,65] has become the process of choice for large and complex composite structures. RIFT is a set of vacuum-driven processes using one solid mold tool and a flexible counter-face tool or membrane. The basic equipment required for RIFT is just a single mold tool, a vacuum pump and consumable materials (e.g., vacuum bag, flow media, peel ply, pipework). The single solid mold face reduces the tooling cost relative to RTM. The mold clamping forces for pressure-driven RTM increase with the area, and RIFT/MIFT soon becomes the only sensible LCM option for the production of large surface area composite structures, for example, boat hulls or wind turbine blades [63,66]. RIFT II denotes the process variant where the long-range flow of the liquid precursor (the uncured polymer matrix) into a flow medium/distribution mesh floods the surface of the laminate stack before through-thickness impregnation of the reinforcement. RIFT II is known by a variety of abbreviations with Seeman Composites resin infusion molding process (SCRIMPTM), or vacuum-assisted resin transfer molding (VARTM) [62], dominating terminology in the United States. In this paper, RIFT is used for production with thermosetting resins, and MIFT for the in situ polymerization process to produce thermoplastic matrix composites. The layout of the RIFT/MIFT is shown in Figure 1 below.
The key factors for a successful RIFT/MIFT process are (i) polymer rheology, (ii) reinforcement permeability, which quantifies resistance to resin flow [67], and (iii) processing temperature.

2.1. Polymer Rheology

Rheology is the science of flow and deformation of matter. For fluids, the key characteristic is viscosity: the ability of a fluid to resist a change in the arrangement of the molecules when under an applied strain or stress [1]. Note that a low viscosity indicates a high flow rate, or a high viscosity indicates a low flow rate. The polymer industry defines two forms of viscosity:
  • Dynamic viscosity (μ): the force required to overcome internal friction. The SI units are Pascal-seconds (Pa·s: identical to 1 kg·m−1·s−1), although the composites industry often uses centimeter–gram–second (CGS) units: centipoise (cP). There is a direct numerical equivalence between millipascal-seconds and centipoise (1 mPa·s = 1 cP).
  • Kinematic viscosity (η): the ratio of the viscous force to the inertial force where the latter is a function of the fluid density (ρ). The SI units are m2·s−1, although the parameter is often given in centistokes (CGS units: 1 centistoke is 1 mm2/s). Hence η = μ/ρ.
In LCM, the process is normally modeled using Darcy’s equation [68] where, for unidirectional flow in isotropic media, the volumetric flow rate (Q, units: m3/s) of a fluid in a saturated porous medium can be expressed as:
Q   =   K · A · Δ P μ · L
where K is a constant of proportionality known as the permeability (units: m2), A is the cross-section of the porous medium normal to the flow direction (units: m2), ΔP/L is the pressure gradient driving the flow (units: Pa/m) and μ is the dynamic fluid viscosity. For anisotropic media (composite systems), the equation should have a tensor form.
Rudd et al. [52] suggest that the most significant practical limitation on the suitability of a resin system for the LCM process is imposed by the viscosity. Resins with extremely low viscosity may be unsuitable for LCM processes as they may result in high porosity or gross voidage. Resins formulated for liquid composite molding processes typically have an initial viscosity of around 200 mPa·s. For curing thermoset resins or for in situ polymerization of thermoplastic monomers, the progress of the chemical reaction causes an increase in molecular size and a consequent increase in viscosity. At the low differential pressures used in infusion processes, the flow front becomes effectively stationary at a defined viscosity [62]. Becker [69] quotes an upper limit for viscosity in RTM of 800 mPa·s, while the non-injection point (NIP) has been defined as a viscosity of 1000 mPa·s [58].

2.2. Reinforcement Permeability

The permeability of the reinforcement also has a huge influence on the LCM processes. Equation (1) was originally derived for the flow of water through isotropic saturated rock to feed the drinking fountains in Dijon, France. Permeability is viewed as a unique characteristic of a saturated porous medium. The LCM process involves permeation through the multiscale pore spaces of the (an)isotropic reinforcement under two conditions: (i) unsaturated (wetting) then (ii) saturated (fully wetted).
Summerscales [70] and Park and Krawczak [71] have considered the differences in permeability for the same fabrics between the two permeation conditions. The reported difference between unsaturated and saturated permeabilities suggests that a change in surface energy occurs during wetting and dissipates some energy. Surface tension and contact angle measurements indicate that the difference in permeability arises from varying interactions at the microscopic level between the fiber and the fluid [72]. For conventional glass or carbon fibers, Kim et al. [73] and Diallo et al. [74] have reported that saturated permeabilities are always lower than unsaturated permeabilities. However, other authors have reported opposite results [67,75].
For LCM processes with natural fiber reinforcements, the permeability is more complicated. The absorption of the permeant fluid by natural fibers, and the consequent fiber swelling, lead to fiber diameter increase, which may be responsible for inconsistencies in permeability measurements for these reinforcements. Moreover, the choice of the liquid has a significant influence on the fiber-swelling ratio (wet apparent diameter/dry apparent diameter) [76,77,78,79], where “apparent” recognizes many natural fibers have non-circular cross-sections. An experimental study by Nguyen et al. [78] demonstrated that the permeability of flax fabrics showed a strong correlation to the liquid type dependent on liquid sorption into the natural fiber (swelling effect). Similar effects of liquid absorption, and consequent fiber swelling, on permeability, were also observed for jute fabrics. The average saturated permeability measured using diluted corn syrup was 23–70% smaller than when measured with motor oil [79]. Nguyen et al. [78,80] investigated the influence of liquid absorption and fiber swelling during RTM resin impregnation of flax fiber reinforcements and suggested a relationship between fiber swelling and permeability given by:
K = ( 1 f s w 2 · V f ) n + 1 A · ( f s w 2 · V f ) n
where K is the permeability, fSW is the fiber-swelling ratio, Vf is the fiber volume fraction, and A and n are empirically derived constants. The progress of the resin flow front may be delayed when swelling of the fibers behind the flow front constraints liquid moving forwards, or may have a favorable effect by forcing resin forwards. These opposing effects should be modeled using sink and source terms in the mass conservation equation. In consequence, the permeability value in the model may need to vary with exposure time and position in the preform. Nguyen used models with varying permeability (mass source/sink terms) to obtain better agreement with the experimental flow measurements than when using a constant permeability model. When the fiber volume fraction increased, the effect from the mass sink became greater. Since the high fiber volume fraction is a major advantage of RIFT (see Section 1), different resins could lead to vastly varying effects on the fiber swelling phenomenon (permeability of the reinforcement) during the RIFT impregnation process. In consequence, the influence of monomer selection on the permeability of natural fiber reinforcements needs to be considered but is beyond the scope of this paper.

2.3. Processing Temperature

The in situ polymerization of the thermoplastic matrix may require elevated temperature, so the processing temperature may be a crucial parameter in infusion processes, especially for natural fiber reinforcements. The successful implementation of MIFT could be compromised by inappropriate process consumables due to high process temperatures. Table 1 summarizes the maximum working temperatures of commercially available consumables. The potential for chemical interactions within the system should also be considered (e.g., does plasticization of pipework lead to collapse under vacuum at elevated temperatures?).
Using a thermoplastic monomer with a relatively high processing temperature could cause detrimental thermal degradation of natural fibers (e.g., flax, hemp) and their composites, especially during extended process times at high temperatures [81]. Gassan and Bledzki [82] studied the thermal degradation of flax fibers and found that temperatures below 170 °C had minimal effect on fiber properties, while above 170 °C, there was a significant reduction of tenacity and degree of polymerization. Chaishome et al. [83] reported that the thermal stability of flax fibers was significantly influenced by the thermal degradation of hemicellulose and pectin: during the production of a flax fiber thermoplastic composite, the mechanical properties could be improved by (i) reducing the hemicellulose and pectin content of flax fibers, (ii) decreasing the consolidation temperature and (iii) increasing the heating rate. Moreover, Chaishome and Rattanapaskorn [81] improved the thermal stability of flax fibers by alkaline treatment to remove hemicellulose and pectin from the flax fibers. For hemp fibers, Shahzad [84] found that the thermal degradation started in the range of 150–200 °C and prominently accelerated at around 250 °C. The thermal stability of hemp fiber was also analyzed by Ouajai and Shanks [85]: thermogravimetric analysis (TGA) and differential thermal analysis (DTA) results demonstrated that untreated fiber was less thermally stable than NaOH/Na2SO3 treated fiber, with degradation starting at 205 °C and 240 °C for each, respectively.

3. Monomers for Infusion

Molten thermoplastic polymers normally have viscosities far in excess of those used for LCM. However, in recent years, in situ polymerization to produce thermoplastic matrix composites has grown rapidly. Van Rijswijk and Bersee [92] reviewed in situ polymerization of most thermoplastics and classified the principal systems of potential use for MIFT into:
  • Ring-opening polymerization (ROP) in which cyclic molecules are opened into linear monomers or oligomers to produce high molecular weight polymers. The monomers are:
    Caprolactam (e.g., DSM fiber intermediates APA-6) to produce polyamide-6 (PA6);
    Laurolactam (e.g., EMS-Grivory APLC12) to produce polyamide-12 (PA12);
    Cyclic butylene terephthalate (CBT) oligomers (e.g., Cyclics Corporation) to produce polybutylene terephthalate (PBT) polyester;
    Cyclic bisphenol-A oligomers to produce polycarbonate;
    L-lactide to produce poly(L-lactide).
  • Vinyl polymerization where monomer unsaturation (double bonds) is opened to create free radicals which undergo an addition reaction to form long-chain polymers. The available monomer is:
    methyl methacrylate (MMA) (e.g., Arkema Elium® acrylic thermoplastic resin formulations specifically designed for RTM/MIFT manufacture of composite parts) to produce polymethyl methacrylate (PMMA).

3.1. Polyamides from Lactams

3.1.1. Polyamide-6 from Caprolactam

Anionic polymerization of lactams is the most developed method for in situ polymerization of thermoplastics via ROP [92,93]. The ε-caprolactam monomer (C6H11NO, melt temperature, Tm = 69 °C) undergoes ROP, which is usually conducted at 130–170 °C to produce high molecular weight polyamide-6 (PA6) [94]. Conversion up to 99.3% can be achieved in 3–60 min, depending on the type and amount of activator and catalyst added [92]. Pillay et al. [95] polymerized PA6 carbon fabric composite laminates during MIFT, with an ε-caprolactam monomer of low viscosity (~5 mPa·s at 100 °C) and the final polymer matrix demonstrated conversion of ~98%. Gong et al. [96] successfully produced self-reinforced composites with PA6 fiber and PA6 matrix using RTM via anionic polymerization of ε-caprolactam. Mechanical properties of samples produced at four temperatures (140–200 °C step 20 °C) were investigated. Both tensile and flexural strength peaked when processing at 160 °C. The conversion of monomer and the void fraction of composite were all >93% and <2.5%, respectively. A comprehensive study on the vacuum infusion of PA6 composites has been undertaken [97,98,99]. Using hexamethylene diisocyanate as an activator and caprolactam magnesium bromide as a catalyst, the viscosity remained low (<100 mPa·s) for most of the injection time with a processing window (viscosity < 1000 mPa·s) up to 20 min. Final conversion was in the range of 93–97% at processing temperatures between 130 and 180 °C [97]. The infusion process was optimized by considering both polymerization temperature [98] and the choice of activator/initiator [99]. The feasibility of using vacuum infused PA6 thermoplastic composites for MW-size wind turbine blades were reported. A 10% reduction in material cost was expected relative to the epoxy counterparts [100].
Bio-based ε-caprolactam is potentially available:
  • In 2014, Genomatica (San Diego, CA, USA) announced an intention to develop enzymatic pathways to produce hexamethylenediamine, adipic acid and caprolactam [101];
  • In early 2018, Genomatica and Aquafil (Trento, Italy) announced a partnership to commercialize a Genomatica process for caprolactam derived from renewable feedstocks [102,103];
  • Lee et al. (2019) reviewed renewable routes to obtain monomeric precursors for PA66 and PA6 from food waste [104].

3.1.2. Polyamide-12 from Laurolactam

Polyamide-12 (PA12) is produced by ROP of ω-laurolactam (C12H23NO, Tm = 154 °C), with similar activators and initiators as for PA6. Unlike PA6, polymerization of PA12 must be performed above the melting temperature of the final polymer (175 °C) to avoid polymerization inhibition due to premature crystallization [105]. Therefore, the processing temperatures are usually between 180 and 240 °C and a cooling process is required to solidify the final polymer [92]. Zingraff et al. [106] investigated the RTM process for anionic polymerization of PA12 (with carbon fabrics). The initial viscosity for the molten monomer was 23 mPa·s at 180 °C and the polymerization was completed at 190 °C for 1 h. Under optimal flow conditions, the average void content of the composite plates could reduce from 15% to less than 1%. Reactive PA12 is currently marketed by EMS Chemie, Switzerland [92]. Laurolactam is normally derived from petrochemicals by butadiene trimerization to cyclododecatriene (CDT) [107]. To the best of the authors’ knowledge, no bio-based laurolactam has been identified.

3.1.3. Polyamide-6/12 from Lactams

A mixture of ε-caprolactam and ω-laurolactam may be easier to process than either monomer in isolation. Rusu and Rusu [108] synthesized copolyamides with 0, 10, 30, 40 or 50% ω-laurolactam in the monomer feed with one formulation mixed at 110 °C for one minute under a nitrogen atmosphere. The mixture was then polymerized at 160 °C for 30 min. The copolymer is less likely to crystallize than the homopolymers and will consequently have reduced mechanical properties (lower density, so fewer bonds/m3 to react stress) and lower environmental resistance (due to greater free volume).

3.2. Polybutylene Terephthalate from Cyclic Butylene Terephthalate

Many researchers have reported LCM production of thermoplastic polybutylene terephthalate (PBT) composite from cyclic butylene terephthalate (CBT) oligomers. Commercial CBT was usually used in these studies, with the polymerization temperature ranging from 190–260 °C [64,109,110,111,112,113]. Commercial CBT is normally a mixture of oligomers between two and seven repeat units, i.e., from the dimer to the heptamer [111,114]. The melting point of an oligomer mixture is usually lower than the pure, discrete cyclic oligomer. CBT softens at 140 °C and melts completely at 160–190 °C, whereas the pure cyclic dimer melts at 196 °C [115] and the pure cyclic tetramer melts at 248 °C [116]. The initial melt viscosity of the oligomer mixture is 150 mPa·s at 150 °C and falls to 20–30 mPa·s at 190 °C [92,115,117], whereas the dynamic viscosity of conventional PBT was reported to be ~1000 Pa·s at 250 °C [118]. Parton and Verpoest found the final conversions of CBT oligomers were between 92 and 98% when polymerized at 190 °C for 30 min.
Pang et al. [119] and Abt and Sanchez-Soto [120] reviewed the polymerization of CBT oligomers as the precursor material for thermoplastic polyesters and their composites. The properties of the final product are highly dependent on the processing temperature. Isothermal polymerization below the Tm of PBT (∼225 °C) produces PBT with a high degree of crystallinity, which results in a brittle matrix [93,120]. Toughening of PBT may be required to improve the fracture toughness of PBT composites, e.g., by adding polycaprolactone [111,121]. CBT was a product of Cyclics Corporation (Schwarzheide, Germany). No bio-based CBT has been identified to date.

3.3. Polycarbonate from Cyclic Bisphenol A Oligomer

Thermoplastic polycarbonate (PC) can be polymerized from macrocyclic bisphenol A (BPA, C15H16O2, Tm ~200 °C) through ring-opening metathesis polymerization (ROMP). High molecular weight (Mw) polycarbonate (up to 300,000) with the conversion of over 99% can be achieved by polymerizing cyclic oligomers at 300 °C for 30 min, in the presence of various basic catalysts [122].
Salem et al. successfully demonstrated that polycarbonate matrix composites (with glass fibers) could be produced by RTM [123]. At 250 °C processing temperature, molten cyclic BPA oligomer had a viscosity at ~1 Pa·s, which is comparable with the upper limit for viscosity in RTM discussed in Section 2.1. The polymerization was conducted at 300 °C to decrease the reaction time without degrading the polymer. The final BPA polycarbonate matrix showed a high polymer content of over 95% and a molecular weight of up to 50,000. Some matrix voids were observed in the composites structure, but these can be reduced by an additional consolidation step in a hot-press. To the best of the authors’ knowledge, no bio-based BPA has been found.

3.4. Poly(L-lactide) from L-Lactide

Polylactic acid or polylactide (PLA) is a synthetic thermoplastic polymer produced by direct condensation polymerization of lactic acid (C3H6O3) or by ROP of the lactide dimer (C6H8O4) [124]. The monomer is produced from the fermentation of 100% natural and renewable agricultural resources, such as corn [125]. The in situ polymerization of PLA based composites has attracted many studies, especially for particle-reinforced composites, due to the more homogeneous particle dispersion which could be achieved [126]. Zhuang et al. [127] reported the production of TiO2/PLA nanocomposite via in situ polymerization of stereoisomeric L-lactide at 150 °C for 24 h with magnetic stirring. Similarly, a graphene/PLA composite was made using the same method at 170 °C for 4 h [128].
Louisy et al. [129] reported the production of glass/poly(L-lactide) composites by RTM via in situ ROP of L-lactide with tin octoate Sn(Oct)2 as the catalyst. The Tm of L-lactide is ~110 °C, so to achieve low viscosity before molding, the monomer and catalyst were mixed in a 120 °C pot (maximum conversion being only 38% in this condition). The polymerization in the mold was conducted at 185 °C, and the resulting poly(L-lactide) matrices exhibited conversions and molar masses up to 99% and 78,000 g·mol−1, respectively.

3.5. Polymethyl Methacrylate from Methyl Methacrylate (Acrylic) Monomer

The methyl methacrylate (MMA, C5H8O2) monomer can be converted by vinyl (addition) polymerization into polymethyl methacrylate (PMMA), often simply called acrylic [92,93]. Arkema (Paris, France) Elium® acrylic resin is claimed as the first thermoplastic resin compatible with RTM/infusion manufacture producing composite structures with mechanical properties similar to thermosets [130]. The viscosity and processing temperature of Elium® resin are as low as 100 mPa·s and 20 °C (ambient), respectively [131]. Elium® resin is widely researched for thermoplastic composites via RTM/infusion, with use increasing due to the desirable characteristics above.
Raponi et al. [132] have analyzed the thermal, rheological, and dielectric characteristics during polymerization of Elium® liquid thermoplastic monomer for infusion manufacturing of composite materials. A three-step thermal cycle comprising isothermal 25 °C for 25 min, then 80 °C for 30 min, and a final isothermal 110 °C for 120 min is strongly recommended for high/full polymerization of the resin. Bhudolia et al. [133] investigated the Mode I fracture toughness of carbon fiber-Elium® composites and found excellent performance (72% increase) relative to carbon fiber-epoxy composites. Obande et al. [134] compared the mechanical and thermomechanical characteristics between Elium® and epoxy glass fiber composites. The acrylic composites showed equivalent modulus, 33% higher tensile strength and a 19% increase in Mode I fracture toughness against their counterparts with epoxy resin. They also proposed optimal infusion parameters to manufacture carbon fiber-Elium® composites with high fiber volume fraction (up to 60%) and low void content (<1%) [135]. Some researchers have successfully produced natural fiber-Elium® composites and tested their mechanical properties [136,137]. In 2016, the CANOE technical platform successfully produced a 9-m thermoplastic composite boat using MIFT of Elium® at room temperature [138]. A number of commercial organizations have developed proprietary processes for manufacturing bio-based MMA. Major players in the global synthetic and bio-based MMA market include Arkema Group, Asahi-Kasei, BASF SE, Dow Chemicals, Evonik and Mitsubishi Rayon [139].
Besides the thermoplastic polymers discussed above, which have already been studied for the infusion process, some other polymers, for example, thermoplastic polyurethane (TPU), poly-(aryl) ether ketone (PEK), polyethylene terephthalate (PET) and polyphthalamide (PPA), may also be candidates for producing thermoplastic composites via infusion due to their relatively low monomer viscosities. The key parameters of the infusion process for each polymer are summarized in Table 2.

4. Properties of Thermoplastic Polymers

The major mechanical and thermal properties of the candidate thermoplastic polymers are summarized in Table 3 and directly influence the application of thermoplastic composites in the marine environment. As mentioned above, the viscosity and processing temperature of the monomer determine the feasibility of the infusion process and are a significant factor in monomer selection. The selection criteria are discussed in detail in Section 5 below.

5. Monomer Selection Criteria

The discussion above leads to a number of criteria for the selection of the monomer/resin for in situ polymerization manufacture of natural fiber thermoplastic composites for marine applications:
  • Essential criteria
    • The viscosity of the monomer must be <1000 mPa·s (NIP) to enable the infusion process.
    • The processing temperature must be <200 °C to minimize the thermal degradation of the natural fibers and to reduce the cost of consumables.
    • Tg of the cured matrix should be above the maximum use temperature to minimize the creep effect in highly stressed applications.
    • Low water sensitivity is needed to maintain proper mechanical and thermal properties in marine environments.
  • Desirable criteria
    • Monomer/resin should be bio-based or have potential bio-based sources available.
    • The open window for infusion should be relatively long to enable the production of a large-scale demonstrator/product with 3D geometry in the future.
    • The cost of the monomer/resin should be relatively low.
    • Low embodied energy and other environmental burdens of the product across the entire life cycle, and recyclable at end-of-life.
The assessment of the respective monomers is summarized in Table 4.

5.1. Essential Criteria

5.1.1. Viscosity

As mentioned in Section 2.1, the upper limit of the viscosity for monomer infusion is 800 mPa·s; thus, TPU and PPA were not suitable as the thermoplastic matrix from the MIFT method. In addition, although the monomers of PA6 and PA12 possess extremely low viscosity, they may still be applicable by delayed infusion (in order to increase the viscosity) after adding the catalyst.

5.1.2. Process Temperature

For PA12, PBT, PC, TPU, PEK, PET and PPA, it is clear that their processing temperatures are too high, which will result in thermal degradation of natural fibers. Pretreatment of natural fibers may help to maintain their properties above 180 °C [81,83,84,85], but will surely increase the environmental burdens in composite production. In addition, as discussed in Section 2.3, the high processing temperature will also increase the cost of the consumables. These factors eliminate these polymers from further consideration in the current context. The processing temperatures of PA6, PLA and PMMA are relatively low, which should lead to reduced process energy, lower cost of consumables, and no requirement for fiber pretreatment.

5.1.3. Glass Transition Temperature

For applications where the polymer will be exposed to high stress, it is essential that the glass transition temperature (Tg) exceeds the maximum use temperature as the material will transform from elastic/brittle to viscoelastic/tough at this temperature and become susceptible to creep deformation under sustained load. The data in Table 4 indicate that TPU and PBT do not meet the set criteria for ambient temperature use.

5.1.4. Moisture Content and Depression of Mechanical and Thermal Properties

Moisture absorption is another important parameter as it has a great influence on composite properties. Davies [153] investigated the influence of seawater aging on the mechanical properties of carbon fiber-Elium® composites. Elium® was indicated to have a lower sensitivity to seawater than epoxy. The Elium® polymer saturated in seawater at 60 °C showed a ~20% decrease in tensile strength, comparing to the virgin polymer. The fatigue performance of glass fiber-Elium® composites is comparable to that of glass fiber-epoxy at similar fiber content [154]. However, there are various formulations of Elium® available, and the full performance data set is not yet available. Ishak and Lim [155] studied the effect of moisture absorption on the tensile properties of pure PBT polymer and short glass fiber reinforced PBT. Results showed that at 100% RH, the tensile strength of unreinforced and reinforced PBT decreased 96% and 68%, respectively, while at 81.2% RH, the moisture influence on tensile strength was insignificant. The water absorption effect on the mechanical properties of carbon fiber reinforced PC was studied by Tanaka et al. [156]. Water was absorbed in both the PC resin and the fiber/matrix interface. For the PC resin, increasing water absorption time leads to no change in tensile modulus but a decrease in the tensile strength. For the carbon fiber/PC composite, the decrease showed only when water absorption time reached 400 h. Prabhakaran et al. [157] compared the properties of PA6 at dry and 50% RH conditions. Results demonstrated that the tensile modulus and yield strength were reduced by 65% (3.4 GPa to 1.2 GPa) and 50% (90 MPa to 45 MPa), respectively.
In the context of composites intended for marine use, the hydrophilic nature of the polymer is an important consideration. Colin and Verdu [158] observed three kinds of chemical groups:
  • non-hydrophilic groups ~generally absorb less than 0.1 w/o (weight percent) of water,
  • moderately hydrophilic groups ~generally absorb less than 3 w/o water, and
  • strongly hydrophilic groups ~saturated state generally limited to values <10 w/o water.
Although the moisture absorption of the polymer cannot directly reveal that of the composite material (other parameters, e.g., the fiber reinforcement, matrix void content and the fiber/matrix interface also have important influences), it can still provide a general idea of the rate and saturation moisture absorption of composite materials. In particular, natural fibers absorb large amounts of moisture [153], so special attention should be paid to the influence of water absorption in the composite reinforcement.
Hydrophobic polymers may not wet and bond to natural hydrophilic fibers leading to a weak fiber-matrix interface [6]. This may be a competing mechanism in the monomer/polymer selection. In a strongly hydrophilic polymer, high moisture absorption is likely to reduce the properties of the composite. In a hydrophobic polymer, moisture may weaken an already poor fiber-matrix interface, and in consequence, reduce the mechanical properties of the composite. A moderately hydrophilic polymer may be more suitable for natural fiber composites in marine environments. The achievement of optimal composite properties depends on a good fiber-to-matrix bond. This issue, and the associated coupling agents or modifiers (especially maleic anhydride polypropylene copolymer) technology, have been reviewed elsewhere for natural fiber composites [159,160,161,162].
Wright [163] plotted the fall in Tg for (thermosetting) epoxy resins (from five separate published papers) as a function of moisture content and found “as a rough rule-of-thumb” there was a drop of 20 °C for each 1% of water pick-up (data available up to 7% moisture content). Assuming similar effects of water absorption across a range of polymers, high water absorption leading to a significant reduction of Tg could be detrimental in many marine applications. For saturated PMMA at 1.92% water pick-up, the Tg was depressed by ~20 °C [164]. For PLA microspheres [165], Tg was reduced from 52 °C (~0.3% H2O) to 37 °C (3.5% H2O), implying a need for the cautious design of PLA matrix composites to be used in humid tropical environments. When exposed to solar heating, the surface temperature can be significantly above ambient temperature [166]. The response is dependent on the chemical nature of the dye/pigment, and the heating may be reduced by choosing low solar absorbance materials.
Based on the evidence above, PA6 resin is not suitable for marine composites due to the high moisture absorption (up to 11% [93]), which leads to a significant mechanical and thermal property reduction of its composites. Further, the already low Tg of PA6 falls below the ambient temperature at around 4% moisture uptake making the polymer matrix susceptible to creep in wet conditions [167].
Prabhakaran et al. [93] reported a decision-making methodology (multiple attribute decision making) in resin selection for a vacuum infused glass fiber reinforced thermoplastic wind turbine blade. They considered the viscosity, processing temperature, cost and availability of the candidate resin systems. A scoring method to evaluate, the ranking of the resin indicated: PA6 > PBT > PMMA > PPA > PA12 > PET > PC > TPU > PEK. They concluded that PA6 was the best matrix material for the wind turbine blade. Based on the additional criteria for stressed components to be used in the marine environment, PLA and PMMA are selected as the only sensible matrix materials for the matrix of natural fiber-reinforced marine composites manufactured by MIFT. It is clear that the monomer selection is significantly influenced by many aspects, e.g., the type of fiber reinforcement and the application environment.

5.2. Desirable Criteria

5.2.1. Bio-Based Monomer

PLA monomer is produced from 100% bio-based and renewable resources (Section 3.4). Although PLA meets most of the specific requirements above, it currently requires process temperatures close to the degradation of onset temperatures for lignocellulosic materials. Furthermore, the biodegradability of PLA may lead to low durability in the marine environment [153]. A study showed that the biodegradation ratios of PLA bag and bottle packaging were ~8.4% and ~5.7%, respectively, after 365 days in the ocean water [168]. A similar degradation phenomenon of PLA in a simulated marine environment was also observed by Pelegrini et al. [169], who found the degradation of PLA composite was facilitated by natural buriti fiber reinforcement. This disadvantage of PLA resin requires further consideration.
PMMA resin does not show the drawbacks above; it almost perfectly meets the essential criteria. Among the four thermoplastic polymers discussed here, PMMA possesses the highest Tg (107 °C), minimizing the potential for creep. The cost of Elium® MMA monomer is higher [93]; nevertheless, the low processing temperature could significantly reduce the cost of consumables and energy for infusion. Bio-based MMA monomers are not yet available, but they are under development as mentioned in Section 3.5, and commercial bio-based infusion MMA is likely to become available soon. Furthermore, Elium® resin is weldable [131].
Soroudi and I Jakubowicz [170] have reviewed the available recycling methods, quality and costs for bio-based plastics. Álvarez-Chávez et al. [171] provided insight into the health and environmental impacts of bio-based plastics, with their analysis indicating that no commercially available or developed bio-based plastic was fully sustainable. A fundamental challenge in thermomechanical reprocessing of bio-based materials is the inherent degradation that may change the polymeric structure and hence affect the performance of the recycle. Yates and Barlow [172] reviewed LCA for bio-based polymers (focusing on PLA, PHA, and starch polymers) and found that, while reductions in non-renewable energy use (NREU) and global warming potential (GWP) can be achieved, the reported impacts from other environmental burdens were higher making it difficult to decide which materials had the lowest detriment to the environment.
An alternative disposal route for bio-based materials (natural fibers and plant-based resins) is biological waste treatment, which Hermann et al. [36] divide into aerobic composting or anaerobic digestion and ambient (≤35 °C) or elevated (50–60 °C) temperature processes. A biodegradable material is expected to reach a defined extent of degradation by biological activity under specific environmental conditions within a given time under standard test conditions [173]. Krzan et al. [174] have recently reviewed the standards and certification appropriate to environmentally degradable plastics. The biodegradation of polymeric materials under controlled composting conditions is the subject of a number of standard methods, including ASTM D5338 [175], ASTM D6400 [176], ASTM D6868 [177], EN 13432 [178] or ISO 14852 [179]. ASTM D6691 [180], ASTM D7473 [181], ASTM D7991 [182] and BS ISO 22766 [183] address test methods for measuring plastics (bio-)degradability in marine environments. However, Zumstein et al. [184] have eloquently argued that experimental observations of the carbon from degraded polymers (such as visual disappearance of plastic, plastic mass loss, a decrease in the plastic’s tensile strength, shortening of the average polymer chain length, or microbial growth) are all ill-suited to assess plastic biodegradation as the remnant material may remain a hazard until completely integrated back into the biological cycle.
Any chosen bio-based monomer should have low environmental burdens, environmental toxicology, embodied energy and CO2 emissions when compared to the synthetic equivalent.

5.2.2. Open Window for Infusion

In the production of a PLA matrix composite by in situ polymerization during the RTM process, Louisy et al. [129] found that the working time could be up to 3 h by using L-lactide monomer. According to the technical data sheet, Elium® 188 XO (a grade of Elium® resin designed for natural fibers) has a processing open window of up to 60 min, a low peak exothermic temperature during polymerization. The processing open windows for both PMMA and PLA are significantly longer than the ~20 min open window of PA6 reported in [97]. The longer processing time will facilitate the production of large composite structures via resin infusion.

Latent Catalysts or Hardeners

Catalysts or hardeners with no activity under normal (ambient) conditions, which can be activated by an external stimulus (heat, ultraviolet, etc.) to initiate chemical reactions, are known as “latent” catalysts or hardeners, respectively. The catalysis is delayed until a single activation event triggers multiple subsequent catalytic reaction sites. The three principal initiation mechanisms are thermal-, photochemical- and mechano-catalysis. Mechano-catalysis is most likely to be used in the development of autonomous self-healing materials [185,186,187].
Subject to the identification of appropriate monomer for infusion of thermoplastic matrix composites, there could be scope for extending the open window (keeping the viscosity low until mold fill) of the system through the use of a latent catalyst or harder. Although Elium® resin shows an open window up to 60 min, this could still be a significant advantage for the infusion of large-scale composite structure with 3D geometry.

5.2.3. Cost

The cost of raw materials is very dependent on feedstock costs. The quantity purchased, the negotiating power of the purchaser, delivery costs, the presence of multiple competitive suppliers and the technology readiness level of the system (costs fall with scale-up of production facilities). Comparing the costs for the two remaining monomers from suppliers on Alibaba [188], L-lactide shows a higher price (€2600/ton) than MMA (€1700/ton). The cost for MMA is similar to epoxy (€1500/ton) and high compared to unsaturated polyester (€1060/ton). The additional costs of the recyclable thermoplastics may be mitigated by environmental charges imposed on thermosets at end-of-life.

5.2.4. Recyclability

A main advantage of thermoplastics compared to thermosets is the potential for recyclability of thermoplastic composites as studied by many researchers [1,50,189]. The composites produced by ring-opening of monomers/oligomers (e.g., PLA) and vinyl polymerization (e.g., PMMA) should be processable by granulation and re-melting to produce short fiber composites of lower duty than the original continuous fiber-reinforced composites.
As one of the main advantages of Elium® resin, Arkema offers two types of recycling processes on its website: (i) mechanical recycling and (ii) reactive recycling [190]. In mechanical recycling, the Elium® composite structures are ground down to granules; these fragments can be used to make new composite parts after being heated. In reactive recycling, the granules are heat-depolymerized into initial raw monomer at a higher temperature, which can be utilized for totally new composites. Meanwhile, the remaining carbon or glass fibers can be reused.
Cousins et al. [191] comprehensively discussed four techniques for recycling glass fiber Elium® composites, (i) thermal decomposition, (ii) mechanical grinding, (iii) thermoforming and (iv) dissolution. Thermal decomposition aims to decompose (and lose) the Elium® resin and recycle the glass fiber reinforcement. Mechanical grinding is a mature technology in composite recycling and only requires low energy (0.29 MJ/kg composite). Ground materials made into samples by injection molding demonstrated higher tensile modulus and strength compared with virgin composites with short glass fibers. In terms of thermoforming, a curved spar cap was successfully straightened at 120 °C and 5.4 kPa for 8 h. Furthermore, a prototype skateboard was constructed with thin sheets planed from the straightened composite components. Glass fiber Elium® composites can also be recycled via dissolution into their constituent parts. Although the volatile solvent was required, both fiber reinforcements and matrix materials (~90%) can be recycled for an energy cost of 4.0 MJ/kg composite. Further optimization is required to identify the most suitable recycling methods for natural fiber Elium® composites.
A significant issue that could arise with a PMMA matrix is the performance when exposed to fire. At elevated temperatures, the polymer depolymerizes to produce a flammable monomer. On the night of 2 August 1973 at Douglas on the Isle of Man, Oroglas acrylic glazing was implicated in the rapid-fire spread through the Summerland leisure center. Fifty people died, and 80 were seriously injured [192].
In terms of the PLA resin composites, Le Duigou et al. [193] investigated the effect of recycling (by repeated injection molding) on the tensile properties of short flax fiber reinforced PLA composites. Results showed that the composite maintained tensile modulus even after six injection cycles. It would be interesting to investigate the feasibility of PLA fiber-PLA composites (although PLA fiber is a semi-synthetic fiber) in the marine environment, as it could offer great advantages in composite recycling.
Indicative environmental data are given in Table 5. LCA will be undertaken in the InterReg SeaBioComp project when (a) the process conditions for optimum composite production are known, (b) raw materials manufacturers have a more mature technology with enhanced efficiencies, (c) economies of scaled-up production are realized, and (d) bio-based precursors are likely to have entered the market.

6. Current Challenges and Future Perspectives

It is challenging to select a suitable monomer for the manufacture of natural-fiber thermoplastic composites via MIFT, as only a limited set of monomers can meet the basic criteria. PMMA does not have an available bio-based monomer yet, while PLA requires high processing temperature (high energy consumption and cost of consumables) as well as its relatively low durability. Furthermore, the biggest challenge may be a sustainable and cost-effective approach for the end-of-life thermoplastic composites dependent on the polymer matrix. Suitable systems using thermoplastics should be much higher in the recycling hierarchy than thermoset resins.
In spite of these challenges, there is an expectation that the production of natural-fiber thermoplastic composites via MIFT for marine applications (especially for large structures) is a trend that will grow rapidly, with its superior renewable and recyclable characteristics as well as low manufacturing cost. These developments will be expedited by:
  • The bio-based MMA monomer being produced on at an industrial scale;
  • The modification of lactide monomer resin to reduce processing temperature and enhance the durability of PLA;
  • Modified, or new, monomers/polymers to meet the criteria (e.g., reduced moisture absorption of infusible polyamides) in order to introduce more candidates;
  • The use of copolymer systems as the matrix.

7. Conclusions

The use of natural fibers as the reinforcement for “sustainable composites” calls into question the use of a thermoset resin matrix. Processing of thermoplastic matrix composites normally requires a temperature that will damage the fibers. In consequence, there is increasing interest in situ polymerization during MIFT. The monomer selection for marine applications has been discussed in this review. The parameters considered were (i) monomer viscosity, (ii) processing temperature, (iii) moisture absorption, (iv) mechanical properties, (v) bio-based availability, (vi) process open window, (vii) cost, and (viii) recyclability. It was concluded that a commercially available acrylic resin best fits the criteria for monomer selection as the thermoplastic matrix. The commercial resin is not currently bio-based but is expected to become available in the near future. Special attention should be given to the recycling of the composites to minimize the cost/energy and environmental impact; LCA is needed to confirm the sustainability. Moreover, latent catalysts may be developed to further improve the open window for the infusion. Finally, PLA resin could be an alternative, although the relatively high processing temperature and potential durability issues may limit its use.

Author Contributions

Conceptualization, Y.Q., J.S.; writing—original draft preparation, Y.Q., J.S.; writing—review and editing, Y.Q., J.S., J.G.-J., M.M., R.P.; supervision, J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was conducted within the SeaBioComp project, which has received funding from Interreg 2 Seas Mers Zeeën program 2014–2020 co-funded by the European Regional Development Fund under subsidy contract No. 2S06-006.
Polymers 12 02928 i001

Acknowledgments

The authors would like to thank Peter Davies (Ifremer), Chung-Hae Park (IMT Lille-Douai) and Gertjan Vancoillie (CenTexBel) for useful comments and insight provided.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Graham-Jones, J.; Summerscales, J. Marine Applications of Advanced Fibre-Reinforced Composites; Woodhead Publishing: Sawston, UK, 2015. [Google Scholar]
  2. Pemberton, R.; Summerscales, J.; Graham-Jones, J. Marine Composites: Design and Performance; Woodhead Publishing: Sawston, UK, 2019. [Google Scholar]
  3. Satyanarayana, K.G.; Arizaga, G.G.; Wypych, F. Biodegradable composites based on lignocellulosic fibers—An overview. Prog. Polym. Sci. 2009, 34, 982–1021. [Google Scholar] [CrossRef]
  4. Andrady, A.L. Microplastics in the marine environment. Mar. Pollut. Bull. 2011, 62, 1596–1605. [Google Scholar] [CrossRef] [PubMed]
  5. Wright, S.L.; Thompson, R.C.; Galloway, T.S. The physical impacts of microplastics on marine organisms: A review. Environ. Pollut. 2013, 178, 483–492. [Google Scholar] [CrossRef] [PubMed]
  6. Pickering, K.L.; Efendy, M.A.; Le, T.M. A review of recent developments in natural fibre composites and their mechanical performance. Compos. Part A Appl. Sci. Manuf. 2016, 83, 98–112. [Google Scholar] [CrossRef] [Green Version]
  7. Moudood, A.; Rahman, A.; Öchsner, A.; Islam, M.; Moudood, A. Flax fiber and its composites: An overview of water and moisture absorption impact on their performance. J. Reinf. Plast. Compos. 2018, 38, 323–339. [Google Scholar] [CrossRef] [Green Version]
  8. Al-Maharma, A.Y.; Al-Huniti, N. Critical Review of the Parameters Affecting the Effectiveness of Moisture Absorption Treatments Used for Natural Composites. J. Compos. Sci. 2019, 3, 27. [Google Scholar] [CrossRef] [Green Version]
  9. Célino, A.; Fréour, S.; Jacquemin, F.; Casari, P. The hygroscopic behavior of plant fibers: A review. Front. Chem. 2014, 1, 43. [Google Scholar] [CrossRef] [Green Version]
  10. Summerscales, J. Virtual Book on Bast Fibres and Their Composites. Available online: https://ecm-academics.plymouth.ac.uk/jsummerscales/mats347/bast_book.htm (accessed on 19 November 2020).
  11. Baley, C.; Gomina, M.; Breard, J.; Bourmaud, A.; Davies, P. Variability of mechanical properties of flax fibres for composite reinforcement. A review. Ind. Crops Prod. 2020, 145, 111984. [Google Scholar] [CrossRef]
  12. Brutch, N.B.; Matvienko, I.; Porokhovinova, E.A.; Pavlov, A.V.; Nôžková, J.; Koshkin, V. Effect of Photoperiod on Linum usitatissimum L. Characters. J. Nat. Fibers 2019, 17, 1345–1354. [Google Scholar] [CrossRef]
  13. Lee, C.H.; Abdan, K.; Lee, S.H.; Liu, M. A Comprehensive Review on Bast Fiber Retting Process for Optimal Performance in Fibers Reinforced Polymer Composites. Adv. Mater. Sci. Eng. 2020, 13. [Google Scholar] [CrossRef]
  14. Syduzzaman, M.; Al Faruque, A.; Bilisik, K.; Naebe, M. Plant-Based Natural Fibre Reinforced Composites: A Review on Fabrication, Properties and Applications. Coatings 2020, 10, 973. [Google Scholar] [CrossRef]
  15. Li, M.; Pu, Y.; Thomas, V.M.; Yoo, C.G.; Ozcan, S.; Deng, Y.; Nelson, K.; Ragauskas, A.J. Recent advancements of plant-based natural fiber–reinforced composites and their applications. Compos. Part B Eng. 2020, 200, 108254. [Google Scholar] [CrossRef]
  16. Sanivada, U.K.; Mármol, G.; Brito, F.P.; Fangueiro, R. PLA Composites Reinforced with Flax and Jute Fibers—A Review of Recent Trends, Processing Parameters and Mechanical Properties. Polymers 2020, 12, 2373. [Google Scholar] [CrossRef]
  17. Vinod, A.; Sanjay, M.; Siengchin, S.; Jyotishkumar, P. Renewable and sustainable biobased materials: An assessment on biofibers, biofilms, biopolymers and biocomposites. J. Clean. Prod. 2020, 258, 120978. [Google Scholar] [CrossRef]
  18. Zhang, Z.; Cai, S.; Li, Y.; Wang, Z.; Long, Y.; Yu, T.; Shen, Y. High performances of plant fiber reinforced composites—A new insight from hierarchical microstructures. Compos. Sci. Technol. 2020, 194, 108151. [Google Scholar] [CrossRef]
  19. Gholampour, A.; Ozbakkaloglu, T. A review of natural fiber composites: Properties, modification and processing techniques, characterization, applications. J. Mater. Sci. 2020, 55, 829–892. [Google Scholar] [CrossRef]
  20. Pantaloni, D.; Bourmaud, A.; Baley, C.; Clifford, M.J.; Ramage, M.H.; Shah, D.U. A Review of Permeability and Flow Simulation for Liquid Composite Moulding of Plant Fibre Composites. Materials 2020, 13, 4811. [Google Scholar] [CrossRef]
  21. Chang, B.P.; Mohanty, A.K.; Misra, M. Studies on durability of sustainable biobased composites: A review. RSC Adv. 2020, 10, 17955–17999. [Google Scholar] [CrossRef]
  22. González-López, M.E.; Del Campo, A.S.M.; Robledo-Ortíz, J.R.; Arellano, M.; Pérez-Fonseca, A.A. Accelerated weathering of poly(lactic acid) and its biocomposites: A review. Polym. Degrad. Stab. 2020, 179, 109290. [Google Scholar] [CrossRef]
  23. Chaudhary, V.; Ahmad, F. A review on plant fiber reinforced thermoset polymers for structural and frictional composites. Polym. Test. 2020, 91, 106792. [Google Scholar] [CrossRef]
  24. Gokulkumar, S.; Thyla, P.; Prabhu, L.; Sathish, S. Measuring Methods of Acoustic Properties and Influence of Physical Parameters on Natural Fibers: A Review. J. Nat. Fibers 2020, 17, 1719–1738. [Google Scholar] [CrossRef]
  25. Milosevic, M.; Valášek, P.; Ruggiero, A. Tribology of Natural Fibers Composite Materials: An Overview. Lubricants 2020, 8, 42. [Google Scholar] [CrossRef] [Green Version]
  26. Saleem, A.; Medina, L.; Skrifvars, M.; Berglin, L. Hybrid Polymer Composites of Bio-Based Bast Fibers with Glass, Carbon and Basalt Fibers for Automotive Applications—A Review. Molecules 2020, 25, 4933. [Google Scholar] [CrossRef] [PubMed]
  27. Summerscales, J.; Dissanayake, N.P.; Virk, A.S.; Hall, W. A review of bast fibres and their composites. Part 1—Fibres as reinforcements. Compos. Part A Appl. Sci. Manuf. 2010, 41, 1329–1335. [Google Scholar] [CrossRef] [Green Version]
  28. Summerscales, J. A review of bast fibres and their composites: Part 4 ∼ organisms and enzyme processes. Compos. Part A Appl. Sci. Manuf. 2020, 140, 106149. [Google Scholar] [CrossRef]
  29. Summerscales, J.; Gwinnett, C. Forensic Identification of Bast Fibres. In Biocomposites for High-Performance Applications; Ray, D., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 125–164. [Google Scholar]
  30. Summerscales, J.; Dissanayake, N.; Virk, A.; Hall, W. A review of bast fibres and their composites. Part 2 —Composites. Compos. Part A Appl. Sci. Manuf. 2010, 41, 1336–1344. [Google Scholar] [CrossRef] [Green Version]
  31. Summerscales, J.; Virk, A.; Hall, W. A review of bast fibres and their composites: Part 3—Modelling. Compos. Part A Appl. Sci. Manuf. 2013, 44, 132–139. [Google Scholar] [CrossRef] [Green Version]
  32. Liu, M.; Thygesen, A.; Summerscales, J.; Meyer, A.S. Targeted pre-treatment of hemp bast fibres for optimal performance in biocomposite materials: A review. Ind. Crops Prod. 2017, 108, 660–683. [Google Scholar] [CrossRef] [Green Version]
  33. Summerscales, J.; Grove, S. Manufacturing Methods for Natural Fibre Composites. In Natural Fibre Composites: Materials, Processes and Properties; Hodzic, A., Shanks, R., Eds.; Woodhead Publishing: Sawston, UK, 2014. [Google Scholar]
  34. Dissanayake, N.P.; Summerscales, J. Life Cycle Assessment for Natural Fibre Composites. In Green Composites from Natural Resources; Thakur, V., Ed.; Taylor and Francis Group: Abingdon, UK, 2014; pp. 157–186. [Google Scholar]
  35. Thomason, J.L.; Jenkins, P.G.; Yang, L. Glass Fibre Strength—A Review with Relation to Composite Recycling. Fibers 2016, 4, 18. [Google Scholar] [CrossRef]
  36. Hermann, B.; Debeer, L.; De Wilde, B.; Blok, K.; Patel, M.K. To compost or not to compost: Carbon and energy footprints of biodegradable materials waste treatment. Polym. Degrad. Stab. 2011, 96, 1159–1171. [Google Scholar] [CrossRef]
  37. Bong, C.P.C.; Lim, L.Y.; Ho, W.S.; Lim, J.S.; Klemeš, J.J.; Towprayoon, S.; Ho, C.S.; Lee, C.T. A review on the global warming potential of cleaner composting and mitigation strategies. J. Clean. Prod. 2017, 146, 149–157. [Google Scholar] [CrossRef]
  38. 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. [Google Scholar] [CrossRef] [Green Version]
  39. Siddique, N.I.; Wahid, Z.A. Achievements and perspectives of anaerobic co-digestion: A review. J. Clean. Prod. 2018, 194, 359–371. [Google Scholar] [CrossRef]
  40. Wool, R.; Sun, X.S. Bio-Based Polymers and Composites; Elsevier: Amsterdam, The Netherlands, 2011. [Google Scholar]
  41. Garrison, T.F.; Murawski, A.; Quirino, R.L. Bio-Based Polymers with Potential for Biodegradability. Polymers 2016, 8, 262. [Google Scholar] [CrossRef]
  42. Hottle, T.A.; Bilec, M.M.; Landis, A.E. Sustainability assessments of bio-based polymers. Polym. Degrad. Stab. 2013, 98, 1898–1907. [Google Scholar] [CrossRef]
  43. Hancox, N. High performance thermoplastic resins and their composites S. Beland. Mater. Des. 1993, 14, 208. [Google Scholar] [CrossRef]
  44. Conroy, A.; Halliwell, S.; Reynolds, T. Composite recycling in the construction industry. Compos. Part A Appl. Sci. Manuf. 2006, 37, 1216–1222. [Google Scholar] [CrossRef]
  45. Stewart, R. Thermoplastic composites—Recyclable and fast to process. Reinf. Plast. 2011, 55, 22–28. [Google Scholar] [CrossRef]
  46. Sutherland, L. A review of impact testing on marine composite materials: Part I—Marine impacts on marine composites. Compos. Struct. 2018, 188, 197–208. [Google Scholar] [CrossRef]
  47. Sutherland, L. A review of impact testing on marine composite materials: Part II—Impact event and material parameters. Compos. Struct. 2018, 188, 503–511. [Google Scholar] [CrossRef]
  48. Sutherland, L. A review of impact testing on marine composite materials: Part III—Damage tolerance and durability. Compos. Struct. 2018, 188, 512–518. [Google Scholar] [CrossRef]
  49. Hildebrand, M. Local Impact Strength of Various Boat-Building Materials; Technical Research Centre of Finland: Espoo, Finland, 1997. [Google Scholar]
  50. Otheguy, M.E. Manufacture, Repair and Recycling of Thermoplastic Composite Boats. Ph.D. Thesis, Newcastle University, Newcastle, UK, 2010. [Google Scholar]
  51. Harten, K. Production by Resin Transfer Moulding. In Composite Materials in Maritime Structures—Fundamental Aspects; Shenoi, R.A., Wellicome, J.F., Eds.; Cambridge University Press: Cambridge, UK, 1993. [Google Scholar]
  52. Rudd, C.D.; Long, A.C.; Kendall, K.; Mangin, C. Liquid Moulding Technologies: Resin Transfer Moulding, Structural Reaction Injection Moulding and Related Processing Techniques; Woodhead Publishing: Sawston, UK, 1997. [Google Scholar]
  53. Potter, K. Resin Transfer Moulding; Chapman & Hall: London, UK, 1997. [Google Scholar]
  54. Kruckenberg, T.; Paton, R. Resin Transfer Moulding for Aerospace Structures; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1998. [Google Scholar]
  55. Tucker, C.; Dessenberger, R. Resin Transfer Moulding Phenomena in Polymeric Composites. In Flow and Rheology in Polymer Composites Manufacturing; Advani, S.G., Ed.; Elsevier: Amsterdam, The Netherlands, 1994; pp. 257–323. [Google Scholar]
  56. Benjamin, W.; Beckwith, S. Resin Transfer Moulding; SAMPE Monograph 3; SAMPE: Covina, CA, USA, 1999. [Google Scholar]
  57. Parnas, R. Liquid Composite Moulding; Hanser Gardner Publications: Cincinnati, OH, USA, 2000. [Google Scholar]
  58. Pearce, N.; Guild, F.; Summerscales, J. An investigation into the effects of fabric architecture on the processing and properties of fibre reinforced composites produced by resin transfer moulding. Compos. Part A Appl. Sci. Manuf. 1998, 29, 19–27. [Google Scholar] [CrossRef]
  59. Beckwith, S. Resin infusion technology Part 1—Industry highlights* Part 2—Process definitions and industry variations* Part 3—A detailed overview of RTM and VIP infusion processing. SAMPE J. 2007, 43, 61. [Google Scholar]
  60. Cripps, D.; Searle, T.; Summerscales, J. Open Mould Techniques for Thermoset Composites. In Comprehensive Composite Materials Encyclopædia; Talreja, R., Månson, J.-A., Eds.; Elsevier: Amsterdam, The Netherlands, 2000; Volume 2. [Google Scholar]
  61. Summerscales, J. Resin Infusion under Flexible Tooling (RIFT). In Encyclopedia of Composites, 2nd ed.; John Wiley & Sons: Hoboken, NJ, USA, 2012; pp. 2648–2658. [Google Scholar]
  62. Summerscales, J.; Searle, T.J. Low-pressure (vacuum infusion) techniques for moulding large composite structures. Proc. Inst. Mech. Eng. Part L J. Mater. Des. Appl. 2005, 219, 45–58. [Google Scholar] [CrossRef]
  63. Williams, C.; Summerscales, J.; Grove, S. Resin Infusion under Flexible Tooling (RIFT): A review. Compos. Part A Appl. Sci. Manuf. 1996, 27, 517–524. [Google Scholar] [CrossRef]
  64. Parton, H.; Verpoest, I. In situ polymerization of thermoplastic composites based on cyclic oligomers. Polym. Compos. 2004, 26, 60–65. [Google Scholar] [CrossRef]
  65. Hindersmann, A. Confusion about infusion: An overview of infusion processes. Compos. Part A Appl. Sci. Manuf. 2019, 126, 105583. [Google Scholar] [CrossRef]
  66. Poorzeinolabedin, M.; Parnas, L.; Dashatan, S.H. Resin infusion under flexible tooling process and structural design optimization of the complex composite part. Mater. Des. 2014, 64, 450–455. [Google Scholar] [CrossRef]
  67. Shojaei, A.; Trochu, F.; Ghaffarian, S.R.; Karimian, S.M.H.; Lessard, L. An Experimental Study of Saturated and Unsaturated Permeabilities in Resin Transfer Molding Based on Unidirectional Flow Measurements. J. Reinf. Plast. Compos. 2004, 23, 1515–1536. [Google Scholar] [CrossRef]
  68. Sharp, J.J.M.; Simmons, C.T. The compleat Darcy: New lessons learned from the first English translation of les fontaines publiques de la Ville de Dijon. Ground Water 2005, 43, 457–460. [Google Scholar] [CrossRef]
  69. Becker, D. Tooling for Resin Transfer Moulding; Wichita State University: Wichita, KS, USA, 1991. [Google Scholar]
  70. Summerscales, J. The Effect of Permeant on the Measured Permeability of a Reinforcement. In Proceedings of the 7th International Conference on Flow Processes in Composite Materials (FPCM-7), Newark, NJ, USA, 7–9 July 2004; pp. 471–476. [Google Scholar]
  71. Park, C.H.; Krawczak, P. Unsaturated and Saturated Permeabilities of Fiber Reinforcement: Critics and Suggestions. Front. Mater. 2015, 2, 38. [Google Scholar] [CrossRef] [Green Version]
  72. Steenkamer, D.A.; McKnight, S.H.; Wilkins, D.J.; Karbhari, V.M. Experimental characterization of permeability and fibre wetting for liquid moulding. J. Mater. Sci. 1995, 30, 3207–3215. [Google Scholar] [CrossRef]
  73. Kim, Y.R.; McCarthy, S.; Fanucci, J.; Nolet, S.; Koppernaes, C. Resin flow through fiber reinforcements during composite processing. SAMPE Q. 1991, 22, 0036–0821. [Google Scholar]
  74. Diallo, M.; Gauvin, R.; Trochu, F. Key factors affecting the permeability measurement in continuous fiber reinforcements. In Proceedings of the ICCM-11, Gold Coast, Australia, 14–18 July 1997; pp. 441–445. [Google Scholar]
  75. Lai, Y.-H.; Khomami, B.; Kardos, J.L. Accurate permeability characterization of preforms used in polymer matrix composite fabrication processes. Polym. Compos. 1997, 18, 368–377. [Google Scholar] [CrossRef]
  76. Francucci, G.; Rodríguez, E.S.; Vázquez, A. Study of saturated and unsaturated permeability in natural fiber fabrics. Compos. Part A Appl. Sci. Manuf. 2010, 41, 16–21. [Google Scholar] [CrossRef]
  77. Masoodi, R.; Pillai, K.M.; Grahl, N.; Tan, H. Numerical simulation of LCM mold-filling during the manufacture of natural fiber composites. J. Reinf. Plast. Compos. 2012, 31, 363–378. [Google Scholar] [CrossRef]
  78. Nguyen, V.H.; Lagardère, M.; Park, C.H.; Panier, S. Permeability of natural fiber reinforcement for liquid composite molding processes. J. Mater. Sci. 2014, 49, 6449–6458. [Google Scholar] [CrossRef]
  79. Zhu, Y.; Cai, J.; Qin, Y.; Liu, D.; Yan, C.; Zhang, H. Effects of liquid absorption and swelling on the permeability of natural fiber fabrics in liquid composite moulding. Polym. Compos. 2017, 38, 996–1004. [Google Scholar] [CrossRef]
  80. Nguyen, V.H.; Deléglise, M.; Park, C.H. Modeling of resin flow in natural fiber reinforcement for liquid composite molding processes. Compos. Sci. Technol. 2015, 113, 38–45. [Google Scholar] [CrossRef]
  81. Chaishome, J.; Rattanapaskorn, S. The influence of alkaline treatment on thermal stability of flax fibres. In Proceedings of the 2nd International Conference on Mining, Material and Metallurgical Engineering, Bangkok, Thailand, 17–18 March 2017. [Google Scholar]
  82. Gassan, J.; Bledzki, A.K. Thermal degradation of flax and jute fibers. J. Appl. Polym. Sci. 2001, 82, 1417–1422. [Google Scholar] [CrossRef]
  83. Chaishome, J.; Brown, K.; Brooks, R.; Clifford, M.J. Thermal Degradation of Flax Fibres as Potential Reinforcement in Thermoplastic Composites. Adv. Mater. Res. 2014, 894, 32–36. [Google Scholar] [CrossRef]
  84. Shahzad, A. A Study in Physical and Mechanical Properties of Hemp Fibres. Adv. Mater. Sci. Eng. 2013, 2013, 1–9. [Google Scholar] [CrossRef] [Green Version]
  85. Ouajai, S.; Shanks, R. Composition, structure and thermal degradation of hemp cellulose after chemical treatments. Polym. Degrad. Stab. 2005, 89, 327–335. [Google Scholar] [CrossRef]
  86. CA Composites. Available online: https://cacomposites.com/ (accessed on 12 February 2020).
  87. Castro Composites. Available online: https://www.castrocompositesshop.com/ (accessed on 12 February 2020).
  88. Easy Composites. Available online: https://www.easycomposites.co.uk/#!/ (accessed on 12 February 2020).
  89. Tygavac Advanced Materials. Available online: https://catalogue.tygavac.co.uk/index.php?lang=EN (accessed on 12 February 2020).
  90. Fibre Glast. Available online: https://www.fibreglast.com/category/Vacuum_Bagging (accessed on 12 February 2020).
  91. Northern Composites. Available online: https://northerncomposites.com/ (accessed on 12 February 2020).
  92. 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. [Google Scholar] [CrossRef]
  93. Prabhakaran, R.; Lystrup, A.; Andersen, T. Attribute Based Selection of Thermoplastic Resin for Vacuum Infusion Process: A Decision Making Methodology. In Dynamic Methods and Process Advancements in Mechanical, Manufacturing, and Materials Engineering; Davim, J.P., Ed.; IGI Global: Hershey, PA, USA, 2013. [Google Scholar]
  94. Hedrick, R.M.; Richard, J.W.R. Reinforced Polyamides and Process of Preparation Thereof. U.S. Patent 3,419,517, 31 December 1968. [Google Scholar]
  95. Pillay, S.; Vaidya, U.K.; Janowski, G.M. Liquid Molding of Carbon Fabric-reinforced Nylon Matrix Composite Laminates. J. Thermoplast. Compos. Mater. 2005, 18, 509–527. [Google Scholar] [CrossRef]
  96. Gong, Y.; Liu, A.; Yang, G. Polyamide single polymer composites prepared via in situ anionic polymerization of ε-caprolactam. Compos. Part A Appl. Sci. Manuf. 2010, 41, 1006–1011. [Google Scholar] [CrossRef]
  97. Van Rijswijk, K.; Koppes, K.; Bersee, H.; Beukers, A. Processing Window for Vacuum Infusion of Fiber-Reinforced Anionic Polyamide-6. In Proceedings of the FPCM-7 Conference, Newark, DE, USA, 7–9 July 2004. [Google Scholar]
  98. Van Rijswijk, K.; Bersee, H.; Beukers, A.; Picken, S.; Van Geenen, A. Optimisation of anionic polyamide-6 for vacuum infusion of thermoplastic composites: Influence of polymerisation temperature on matrix properties. Polym. Test. 2006, 25, 392–404. [Google Scholar] [CrossRef]
  99. Van Rijswijk, K.; Bersee, H.; Jager, W.; Picken, S. Optimisation of anionic polyamide-6 for vacuum infusion of thermoplastic composites: Choice of activator and initiator. Compos. Part A Appl. Sci. Manuf. 2006, 37, 949–956. [Google Scholar] [CrossRef]
  100. Van Rijswijk, K.; Joncas, S.; Bersee, H.E.N.; Bergsma, O.K.; Beukers, A. Sustainable Vacuum-Infused Thermoplastic Composites for MW-Size Wind Turbine Blades—Preliminary Design and Manufacturing Issues. J. Sol. Energy Eng. 2005, 127, 570–580. [Google Scholar] [CrossRef]
  101. Bio Based Press. Available online: https://www.biobasedpress.eu/2014/10/bionylon/ (accessed on 1 December 2020).
  102. Jenkins, S. A Bio-Based Caprolactam Joint-Development Project is Now Underway. Chemical Engineering, 1 March 2018. [Google Scholar]
  103. Bomgardner, M. Genomatica, Aquafil Partner for Biobased Nylon. Chemical and Engineering News, 23 January 2018. [Google Scholar]
  104. Lee, Y.; Lin, K.-Y.A.; Kwon, E.E.; Lee, J. Renewable routes to monomeric precursors of nylon 66 and nylon 6 from food waste. J. Clean. Prod. 2019, 227, 624–633. [Google Scholar] [CrossRef]
  105. Luisier, A.; Bourban, P.; Månson, J. In Situ Polymerization of Polyamide 12 for Thermoplastic Composites. In Proceedings of the 12th International Conference on Composite Materials, Paris, France, 5–9 July 1999; pp. 33–45. [Google Scholar]
  106. Zingraff, L.; Michaud, V.; Bourban, P.-E.; Månson, J.-A. Resin transfer moulding of anionically polymerised polyamide 12. Compos. Part A Appl. Sci. Manuf. 2005, 36, 1675–1686. [Google Scholar] [CrossRef]
  107. Chauvel, A.; Lefebvre, G. Petrochemical Processes; OPHRYS Editions: Paris, France, 2001; Volume 80. [Google Scholar]
  108. Rusu, G.; Rusu, E. Caprolactam-laurolactam (nylon 6/12) copolymers: Synthesis and characterization. High Perform. Polym. 2004, 16, 569–584. [Google Scholar] [CrossRef]
  109. Wu, W.; Klunker, F.; Xie, L.; Jiang, B.; Ziegmann, G. Simultaneous binding and ex situ toughening concept for textile reinforced pCBT composites: Influence of preforming binders on interlaminar fracture properties. Compos. Part A Appl. Sci. Manuf. 2013, 53, 190–203. [Google Scholar] [CrossRef]
  110. Agirregomezkorta, A.; Sánchez-Soto, M.; Aretxaga, G.; Sarrionandia, M.; Aurrekoetxea, J. Effects of vacuum infusion processing parameters on the impact behavior of carbon fiber reinforced cyclic butylene terephthalate composites. J. Compos. Mater. 2013, 48, 333–344. [Google Scholar] [CrossRef]
  111. Baets, J.; Godara, A.; Devaux, J.; Verpoest, I. Toughening of isothermally polymerized cyclic butylene terephthalate for use in composites. Polym. Degrad. Stab. 2010, 95, 346–352. [Google Scholar] [CrossRef]
  112. Agirregomezkorta, A.; Martínez, A.; Sánchez-Soto, M.; Aretxaga, G.; Sarrionandia, M.; Aurrekoetxea, J. Impact behaviour of carbon fibre reinforced epoxy and non-isothermal cyclic butylene terephthalate composites manufactured by vacuum infusion. Compos. Part B Eng. 2012, 43, 2249–2256. [Google Scholar] [CrossRef]
  113. Prabhakaran, R.D.; Andersen, T.L.; Lystrup, A. Glass/CBT Laminate Processing and Quality Aspects. In Proceedings of the 10th International Conference on Flow Processes in Composite Materials (FPCM10), Monte Verita, Switzerland, 12–15 July 2010. [Google Scholar]
  114. Brunelle, D.J.; Bradt, J.E.; Serth-Guzzo, J.; Takekoshi, T.; Evans, T.L.; Pearce, E.J.; Wilson, P.R. Semicrystalline Polymers via Ring-Opening Polymerization: Preparation and Polymerization of Alkylene Phthalate Cyclic Oligomers. Macromolecules 1998, 31, 4782–4790. [Google Scholar] [CrossRef]
  115. Brunelle, D.J. Cyclic oligomer chemistry. J. Polym. Sci. Part A Polym. Chem. 2007, 46, 1151–1164. [Google Scholar] [CrossRef]
  116. East, G.; Girshab, A. Cyclic oligomers in poly(1,4-butylene terephthalate). Polymer 1982, 23, 323–324. [Google Scholar] [CrossRef]
  117. Ishak, Z.M.; Gatos, K.; Karger-Kocsis, J. On the in-situ polymerization of cyclic butylene terephthalate oligomers: DSC and rheological studies. Polym. Eng. Sci. 2006, 46, 743–750. [Google Scholar] [CrossRef]
  118. Steeg, M.; Mitschang, P.; Chakraborty, P.; Hartmann, T. Modeling the viscosity and conversion of in-situ polymerizing PBT using empirical data. In Proceedings of the 17th ICCM, Edinburgh, UK, 27–31 July 2009. [Google Scholar]
  119. Pang, K.; Kotek, R.; Tonelli, A. Review of conventional and novel polymerization processes for polyesters. Prog. Polym. Sci. 2006, 31, 1009–1037. [Google Scholar] [CrossRef]
  120. Abt, T.; Sánchez-Soto, M. A Review of the Recent Advances in Cyclic Butylene Terephthalate Technology and its Composites. Crit. Rev. Solid State Mater. Sci. 2016, 42, 173–217. [Google Scholar] [CrossRef] [Green Version]
  121. Baets, J.; Dutoit, M.; Devaux, J.; Verpoest, I. Toughening of glass fiber reinforced composites with a cyclic butylene terephthalate matrix by addition of polycaprolactone. Compos. Part A Appl. Sci. Manuf. 2008, 39, 13–18. [Google Scholar] [CrossRef]
  122. Brunelle, D.J.; Shannon, T.G. Preparation and polymerization of bisphenol A cyclic oligomeric carbonates. Macromolecules 1991, 24, 3035–3044. [Google Scholar] [CrossRef]
  123. Salem, A.; Stewart, K.; Gifford, S.; Berenbaum, A. Fabrication of thermoplastic matrix structural composites by resin transfer molding of cyclic bisphenol-A polycarbonate oligomers. SAMPE J. 1991, 27, 17–22. [Google Scholar]
  124. Gonçalves, C.; Gonçalves, I.C.; Magalhães, F.D.; Pinto, A.M. Poly (lactic acid) Composites Containing Carbon-Based Nanomaterials: A Review. Polymers 2017, 9, 269. [Google Scholar] [CrossRef] [Green Version]
  125. Avinc, O.; Khoddami, A. Overview of Poly (lactic acid) (PLA) Fibre. Fibre Chem. 2009, 41, 391–401. [Google Scholar] [CrossRef]
  126. Kim, H.; Kobayashi, S.; Abdurrahim, M.A.; Zhang, M.J.; Khusainova, A.; Hillmyer, M.A.; Abdala, A.A.; Macosko, C.W. Graphene/polyethylene nanocomposites: Effect of polyethylene functionalization and blending methods. Polymer 2011, 52, 1837–1846. [Google Scholar] [CrossRef]
  127. Zhuang, W.; Liu, J.; Zhang, J.; Hu, B.X.; Shen, J. Preparation, characterization, and properties of TiO2/PLA nanocomposites by in situ polymerization. Polym. Compos. 2009, 30, 1074–1080. [Google Scholar] [CrossRef]
  128. Yang, J.-H.; Lin, S.-H.; Lee, Y.-D. Preparation and characterization of poly(l-lactide)–graphene composites using the in situ ring-opening polymerization of PLLA with graphene as the initiator. J. Mater. Chem. 2012, 22, 10805. [Google Scholar] [CrossRef]
  129. Louisy, E.; Samyn, F.; Bourbigot, S.; Fontaine, G.; Bonnet, F. Preparation of Glass Fabric/Poly(l-lactide) Composites by Thermoplastic Resin Transfer Molding. Polymers 2019, 11, 339. [Google Scholar] [CrossRef] [Green Version]
  130. ARKEMA. Elium® Resin for Composites. Available online: https://www.arkema.com/global/en/products/product-finder/product-range/incubator/elium_resins/ (accessed on 1 December 2020).
  131. ARKEMA. Liquid Thermoplastic Resin for Tougher Composites. Available online: https://www.arkema.com/export/shared/.content/media/downloads/products-documentations/incubator/brochure-elium-2017.pdf (accessed on 24 October 2019).
  132. Raponi, O.D.A.; De Souza, B.R.; Barbosa, L.C.M.; Ancelotti, A.C., Jr. Thermal, rheological, and dielectric analyses of the polymerization reaction of a liquid thermoplastic resin for infusion manufacturing of composite materials. Polym. Test. 2018, 71, 32–37. [Google Scholar] [CrossRef]
  133. Bhudolia, S.K.; Perrotey, P.; Joshi, S.C. Mode I fracture toughness and fractographic investigation of carbon fibre composites with liquid Methylmethacrylate thermoplastic matrix. Compos. Part B Eng. 2018, 134, 246–253. [Google Scholar] [CrossRef]
  134. Obande, W.; Mamalis, D.; Ray, D.; Yang, L.; Brádaigh, C.M. Mechanical and thermomechanical characterisation of vacuum-infused thermoplastic- and thermoset-based composites. Mater. Des. 2019, 175, 107828. [Google Scholar] [CrossRef]
  135. Bhudolia, S.K.; Perrotey, P.; Joshi, S.C. Optimizing Polymer Infusion Process for Thin Ply Textile Composites with Novel Matrix System. Materials 2017, 10, 293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Monti, A.; El Mahi, A.; Jendli, Z.; Guillaumat, L. Mechanical behaviour and damage mechanisms analysis of a flax-fibre reinforced composite by acoustic emission. Compos. Part A Appl. Sci. Manuf. 2016, 90, 100–110. [Google Scholar] [CrossRef] [Green Version]
  137. Haggui, M.; Jendli, Z.; Akrout, A.; El Mahi, A.; Haddar, M. Damage identification in flax fibre composite with thermoplastic matrix under quasi-static loading. In Proceedings of the International Conference on Advanced Materials, Mechanics and Manufacturing A3M, Hammamet, Tunisia, 19–21 December 2016. [Google Scholar]
  138. Tison, F. Thermoplastic composites production by room-temperature vacuum infusion. JEC Compos. 2016, 102, 57–58. [Google Scholar]
  139. GVR. Synthetic and Bio-Based Methyl Methacrylate (MMA) Market Size, Application Analysis, Regional Outlook, Competitive Strategies and Forecasts, 2014 to 2020; Grand View Research Report GVR1280; GVR: San Francisco, CA, USA, 2014. [Google Scholar]
  140. Youk, J.H.; Kambour, R.P.; Macknight, W.J. Polymerization of Ethylene Terephthalate Cyclic Oligomers with Antimony Trioxide. Macromolecules 2000, 33, 3594–3599. [Google Scholar] [CrossRef] [Green Version]
  141. Youk, J.H.; Boulares, A.; Kambour, R.P.; Macknight, W.J. Polymerization of Ethylene Terephthalate Cyclic Oligomers with a Cyclic Dibutyltin Initiator. Macromolecules 2000, 33, 3600–3605. [Google Scholar] [CrossRef] [Green Version]
  142. Nagahata, R.; Sugiyama, J.; Goyal, M.; Goto, M.; Honda, K.; Asai, M.; Ueda, M.; Takeuchi, K. Thermal polymerization of uniform macrocyclic ethylene terephthalate dimer. Polymer 2001, 42, 1275–1279. [Google Scholar] [CrossRef]
  143. Nagahata, R.; Sugiyama, J.I.; Goyal, M.; Asai, M.; Ueda, M.; Takeuchi, K. Solid-phase thermal polymerization of macrocyclic ethylene terephthalate dimer using various transesterification catalysts. J. Polym. Sci. Pol. Chem. 2000, 38, 3360–3368. [Google Scholar] [CrossRef]
  144. Pini, N.; Zaniboni, C.; Busato, S.; Ermanni, P. Perspectives for Reactive Molding of PPA as Matrix for High-performance Composite Materials. J. Thermoplast. Compos. Mater. 2006, 19, 207–216. [Google Scholar] [CrossRef]
  145. Goodfellow. PLLA. Available online: http://www.goodfellow.com/E/Poly-L-lactic-acid-Biopolymer.html (accessed on 24 October 2019).
  146. Farah, S.; Anderson, D.G.; Langer, R. Physical and mechanical properties of PLA, and their functions in widespread applications—A comprehensive review. Adv. Drug Deliv. Rev. 2016, 107, 367–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Tham, W.L.; Poh, B.T.; Ishak, Z.A.M.; Chow, W.S. Characterisation of water absorption of biodegradable poly(lactic acid)/halloysite nanotube nanocomposites at different temperatures. J. Eng. Sci. 2016, 12, 13. [Google Scholar]
  148. Zhang, L.; Lv, S.; Sun, C.; Wan, L.; Tan, H.; Zhang, Y. Effect of MAH-g-PLA on the Properties of Wood Fiber/Polylactic Acid Composites. Polymers 2017, 9, 591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Polymer Properties Database. Polyurethane. Available online: https://polymerdatabase.com/polymers/hexamethylenediisocyanatediethyleneglycol.html (accessed on 24 October 2019).
  150. Goodfellow. PEK. Available online: http://www.goodfellow.com/catalogue/GFCat4.php?ewd_token=hoDZrxdrFfmNKKrxxD1M5IWFzhfmZo&n=1QxXsTWTaq3A26aehcP1wUGGfOrqUF&ewd_urlNo=GFCatSeab1&CatLook=PEK (accessed on 24 October 2019).
  151. PerkinElmer. DSC Transitions of Common Thermoplastics. Available online: https://thermalsupport.com/wp-content/uploads/2018/05/PETech-44.pdf (accessed on 29 June 2020).
  152. RTP Imagineering Plastics. RTP2201A Polyetherketone (PEK) Glass Fiber Product Data Sheet & General Processing Conditions. Available online: http://web.rtpcompany.com/info/data/2200A/RTP2201A.htm (accessed on 29 June 2020).
  153. Davies, P. Environmental degradation of composites for marine structures: New materials and new applications. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2016, 374, 20150272. [Google Scholar] [CrossRef] [PubMed]
  154. Davies, P.; Arhant, M. Fatigue Behaviour of Acrylic Matrix Composites: Influence of Seawater. Appl. Compos. Mater. 2018, 26, 507–518. [Google Scholar] [CrossRef] [Green Version]
  155. Ishak, Z.A.M.; Lim, N.C. Effect of moisture absorption on the tensile properties of short glass fiber reinforced poly(butylene terephthalate). Polym. Eng. Sci. 1994, 34, 1645–1655. [Google Scholar] [CrossRef]
  156. Tanaka, K.; Fukushima, Y.; Kashihara, K.; Katayama, T. Effect of water absorption on the mechanical properties of continuous carbon fibre reinforced polycarbonate composites. High Perform. Struct. Mater. V 2010, 112, 153–162. [Google Scholar] [CrossRef]
  157. Prabhakaran, R.D.; Andersen, T.L.; Lystrup, A. Influence of moisture absorption on properties of fiber reinforced polyamide 6 composites. In Proceedings of the 2nd Joint US-Canada Conference on Composites, Montreal, QC, Canada, 26–28 September 2011; pp. 500–510. [Google Scholar]
  158. Colin, X.; Verdu, J. Humid Ageing of Organic Matrix Composites. In Durability of Composites in a Marine Environment; Davies, P., Rajapakse, Y., Eds.; Springer: Dordrecht, The Netherlands, 2014; pp. 47–114. [Google Scholar]
  159. Xie, Y.; Hill, C.A.; Xiao, Z.; Militz, H.; Mai, C. Silane coupling agents used for natural fiber/polymer composites: A review. Compos. Part A Appl. Sci. Manuf. 2010, 41, 806–819. [Google Scholar] [CrossRef]
  160. Mohit, H.; Selvan, V.A.M. A comprehensive review on surface modification, structure interface and bonding mechanism of plant cellulose fiber reinforced polymer based composites. Compos. Interfaces 2018, 25, 629–667. [Google Scholar] [CrossRef]
  161. Graupner, N.; Rößler, J.; Ziegmann, G.; Müssig, J. Fibre/matrix adhesion of cellulose fibres in PLA, PP and MAPP: A critical review of pull-out test, microbond test and single fibre fragmentation test results. Compos. Part A Appl. Sci. Manuf. 2014, 63, 133–148. [Google Scholar] [CrossRef]
  162. Anbupalani, M.S.; Venkatachalam, C.D.; Rathanasamy, R. Influence of coupling agent on altering the reinforcing efficiency of natural fibre-incorporated polymers—A review. J. Reinf. Plast. Compos. 2020, 39, 520–544. [Google Scholar] [CrossRef]
  163. Wright, W. The effect of diffusion of water into epoxy resins and their carbon-fibre reinforced composites. Composites 1981, 12, 201–205. [Google Scholar] [CrossRef]
  164. Smith, L.; Schmitz, V. The effect of water on the glass transition temperature of poly(methyl methacrylate). Polymer 1988, 29, 1871–1878. [Google Scholar] [CrossRef]
  165. Passerini, N.; Craig, D.Q. An investigation into the effects of residual water on the glass transition temperature of polylactide microspheres using modulated temperature DSC. J. Control. Release 2001, 73, 111–115. [Google Scholar] [CrossRef]
  166. Summerscales, J. Durability of Composites in the Marine Environment. In Durability of Composites in a Marine Environment; Davies, P., Rajapakse, Y., Eds.; Springer: Dordrecht, The Netherlands, 2014; pp. 1–13. [Google Scholar]
  167. Parodi, E.; Peters, G.; Govaert, L.E. Structure–Properties Relations for Polyamide 6, Part 1: Influence of the Thermal History during Compression Moulding on Deformation and Failure Kinetics. Polymers 2018, 10, 710. [Google Scholar] [CrossRef] [Green Version]
  168. Greene, J. PLA and PHA Biodegradation in the Marine Environment; CalRecycle: Sacramento, CA, USA, 2012.
  169. Pelegrini, K.; Donazzolo, I.; Brambilla, V.; Grisa, A.M.C.; 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, 133. [Google Scholar] [CrossRef]
  170. Soroudi, A.; Jakubowicz, I. Recycling of bioplastics, their blends and biocomposites: A review. Eur. Polym. J. 2013, 49, 2839–2858. [Google Scholar] [CrossRef]
  171. Álvarez-Chávez, C.R.; Edwards, S.E.; Moure-Eraso, R.; Geiser, K. Sustainability of bio-based plastics: General comparative analysis and recommendations for improvement. J. Clean. Prod. 2012, 23, 47–56. [Google Scholar] [CrossRef]
  172. Yates, M.R.; Barlow, C.Y. Life cycle assessments of biodegradable, commercial biopolymers—A critical review. Resour. Conserv. Recycl. 2013, 78, 54–66. [Google Scholar] [CrossRef]
  173. Murphy, R.; Bartle, I. Biodegradable Polymers and Sustainability: Insights from Life Cycle Assessment. In Proceedings of the National Non-Food Crops Centre Seminar, London, UK, 25 May 2004. [Google Scholar]
  174. Krzan, A.; Hemjinda, S.; Miertus, S.; Corti, A.; Chiellini, E. Standardization and certification in the area of environmentally degradable plastics. Polym. Degrad. Stab. 2006, 91, 2819–2833. [Google Scholar] [CrossRef]
  175. ASTM. Standard Test Method for Determining Aerobic Biodegradation of Plastic Materials Under Controlled Composting Conditions, Incorporating Thermophilic Temperatures; D5338-15; ASTM: West Conshohocken, PA, USA, 2015. [Google Scholar]
  176. ASTM. Standard Specification for Labeling of Plastics Designed to be Aerobically Composted in Municipal or Industrial Facilities; D6400-19; ASTM: West Conshohocken, PA, USA, 2019. [Google Scholar]
  177. ASTM. Standard Specification for Labeling of End Items that Incorporate Plastics and Polymers as Coatings or Additives with Paper and Other Substrates Designed to be Aerobically Composted in Municipal or Industrial Facilities; D6868-19; ASTM: West Conshohocken, PA, USA, 2019. [Google Scholar]
  178. BSI. Packaging. Requirements for Packaging Recoverable Through Composting and Biodegradation. Test Scheme and Evaluation Criteria for the Final Acceptance of Packaging; BS EN 13432; BSI: London, UK, 2000. [Google Scholar]
  179. ISO. Determination of the Ultimate Aerobic Biodegradability of Plastic Materials in an Aqueous Medium—Method by Analysis of Evolved Carbon Dioxide; ISO 14852; ISO: Geneva, Switzerland, 2018. [Google Scholar]
  180. ASTM. Standard Test Method for Determining Aerobic Biodegradation of Plastic Materials in the Marine Environment by a Defined Microbial Consortium or Natural Sea Water Inoculum; D6691-17; ASTM: West Conshohocken, PA, USA, 2017. [Google Scholar]
  181. ASTM. Standard Test Method for Weight Attrition of Plastic Materials in the Marine Environment by Open System Aquarium Incubations; D7473-12; ASTM: West Conshohocken, PA, USA, 2012. [Google Scholar]
  182. ASTM. Standard Test Method for Determining Aerobic Biodegradation of Plastics Buried in Sandy Marine Sediment under Controlled Laboratory Conditions; D7991-15; ASTM: West Conshohocken, PA, USA, 2015. [Google Scholar]
  183. BSI. Plastics. Determination of the Degree of Disintegration of Plastic Materials in Marine Habitats under Real Field Conditions; BS ISO 22766; BSI: London, UK, 2000. [Google Scholar]
  184. Zumstein, M.T.; Narayan, R.; Kohler, H.-P.E.; McNeill, K.; Sander, M. Dos and Do Nots When Assessing the Biodegradation of Plastics. Environ. Sci. Technol. 2019, 53, 9967–9969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Naumann, S.; Buchmeiser, M.R. Latent and Delayed Action Polymerization Systems. Macromol. Rapid Commun. 2014, 35, 682–701. [Google Scholar] [CrossRef]
  186. Endo, T.; Sanda, F. Design of latent catalysts and their application to polymer synthesis. Macromol. Symp. 1996, 107, 237–242. [Google Scholar] [CrossRef]
  187. Groote, R.R.; Jakobs, R.T.M.; Sijbesma, R.P. Mechanocatalysis: Forcing latent catalysts into action. Polym. Chem. 2013, 4, 4846–4859. [Google Scholar] [CrossRef]
  188. Alibaba. Available online: https://www.alibaba.com/ (accessed on 7 July 2020).
  189. Goodship, V. Management, Recycling and Reuse of Waste Composites; Elsevier: Amsterdam, The Netherlands, 2009. [Google Scholar]
  190. ARKEMA. Recycling Composite Parts Made of Elium® Resin is Possible! Available online: https://www.arkema.com/en/products/product-finder/range-viewer/Elium-resins-for-composites/ (accessed on 24 October 2019).
  191. Cousins, D.S.; Suzuki, Y.; Murray, R.E.; Samaniuk, J.R.; Stebner, A.P. Recycling glass fiber thermoplastic composites from wind turbine blades. J. Clean. Prod. 2019, 209, 1252–1263. [Google Scholar] [CrossRef]
  192. Silcock, A.; Hinkley, P. Report on the Spread of Fire at Summerland in Douglas on the Isle of Man, 2 August 1973; BRE Current Paper CP74/74; Building Research Establishment: Watford, UK, 1974. [Google Scholar]
  193. Le Duigou, A.; Pillin, I.; Bourmaud, A.; Davies, P.; Baley, C. Effect of recycling on mechanical behaviour of biocompostable flax/poly(l-lactide) composites. Compos. Part A Appl. Sci. Manuf. 2008, 39, 1471–1478. [Google Scholar] [CrossRef] [Green Version]
  194. Patel, M. Cumulative energy demand (CED) and cumulative CO2 emissions for products of the organic chemical industry. Energy 2003, 28, 721–740. [Google Scholar] [CrossRef]
  195. Morão, A.; De Bie, F. Life Cycle Impact Assessment of Polylactic Acid (PLA) Produced from Sugarcane in Thailand. J. Polym. Environ. 2019, 27, 2523–2539. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic of the monomer infusion under flexible tooling (RIFT/MIFT).
Figure 1. Schematic of the monomer infusion under flexible tooling (RIFT/MIFT).
Polymers 12 02928 g001
Table 1. Commercial consumables for RIFT, and their maximum working temperatures [86,87,88,89,90,91].
Table 1. Commercial consumables for RIFT, and their maximum working temperatures [86,87,88,89,90,91].
ConsumablesMaterialsMaximum Temperature (°C)
TapePolyamide (PA) film tape rubber adhesive204
Polyester tape with rubber/acrylic/silicone adhesive177–220
Polytetrafluoroethylene (PTFE) tape with rubber/acrylic/silicone adhesive204, 260
Fluoropolymer tape with rubber/silicone adhesive204, 260
Polytetrafluoroethylene (PTFE) coated glass fabric tape with silicone adhesive260
Polyimide (PI) tape with silicone adhesive399
Bagging filmPolyethylene (PE)48, 82
Polypropylene (PP)135
Polyvinyl chloride (PVC)121
Polyolefin121, 140
Polyamide (PA) and polyolefin multilayer155
Polyamide (PA) + polyethylene (PE) + polyamide (PA) multilayer170
Polyurethane (PU)135, 176
Polyamide (PA) and polypropylene (PP) multilayer180
Polymethyl pentene (PMP) and polyamide (PA) multilayer190
Thermoplastic elastomer (TPE)121–195
Ethylene tetrafluoroethylene (ETFE) and polyamide (PA) multilayer230
Polyamide (PA)120–246
Polytetrafluoroethylene (PTFE)315
Polyimide (PI)399, 426
Peel plyPolyester121–249
Polyamide (PA)160–260
Fiberglass coated with polytetrafluoroethylene (PTFE)288, 316
Fiberglass coated with silicone427
Flow mediumLow-density polyethylene (LDPE)65
Low-density polyethylene (LDPE) and high-density polyethylene (HDPE)65, 100
Polyethylene (PE)120
High-density polyethylene (HDPE)93, 125
Polypropylene (PP)100, 150
Polyester170, 200
Polyamide (PA)177–216
Release filmPolypropylene (PP)100
Polyolefin and high-density polyethylene (HDPE)120
Polyolefin121–157
Polyethylene (PE)125
Polymethyl pentene (PMP)193, 200
Fluorinated ethylene propylene (FEP)210, 260
Ethylene tetrafluoroethylene (ETFE)220, 260
Fluoropolymer260, 315
Polyimide (PI)405
Sealant tapeSynthetic Rubber90–232
Silicone400, 427
Hose/PipePolyvinyl chloride (PVC)50
Polyurethane (PU)60
Low-density polyethylene (LDPE)90
High-density polyethylene (HDPE)90
Polyethylene (PE)90, 121
Silicone230, 260
SpiralPolyethylene (PE)80, 121
Polypropylene (PP)60, 120
Table 2. Key parameters of thermoplastic monomers for in situ polymerization infusion.
Table 2. Key parameters of thermoplastic monomers for in situ polymerization infusion.
PolymerMonomer Viscosity (mPa·s)Processing Temperature (°C)Bio-Based Monomer Available?Ref.
Polyamide-6 (PA6)~5130–200Yes[94,95,96,97]
Polyamide-12 (PA12)23180–240No[92,106]
Polybutylene terephthalate (PBT)20–150180–260No[64,92,109,110,111,112,113,115,117]
Polycarbonate (PC)1000250–300No[123]
Polylactide (PLA)-150–185Yes[127,128,129]
Polymethyl methacrylate (PMMA, Elium®)10020–100Yes *[131]
Thermoplastic polyurethane (TPU)800300No[93]
Polyether ketone (PEK)80340–390No[92]
Polyethylene terephthalate (PET)30250–325No[140,141,142,143]
Polyphthalamide (PPA)1000200–290No[93,144]
* Should be commercially available in the near future.
Table 3. Overview of properties of monomer-infusible thermoplastic polymers.
Table 3. Overview of properties of monomer-infusible thermoplastic polymers.
PolymerDensity (g/cm3)Tensile Strength (MPa)Tensile Modulus (GPa)Strain to Failure (%)Tg (°C)Tm (°C)Moisture Absorption (%)Ref.
PA61.13852.0–3.81940–60219–2306–11[93]
PA121.0450–601.430040–501801.1–1.8[93,144]
PBT1.31851.8–2.73025–602250.09[93]
PC1.20602.2>1001503000.16[93]
PLA1.25703.62.455–65170–200<2[145,146,147,148]
PMMA
(Elium®)
1.20663.172.6107-0.5[131,135,138]
TPU1.20400.2–2.3>500−8–171400.1[93,149]
PEK1.301153.720228-0.07[92,150]
PET1.3869313722550.5[93]
PPA1.18902.5–3.56121–138310–3300.36[93,144]
Table 4. The relative fit to requirements of the potential monomers for thermoplastic matrix marine composites (green is good, amber is marginal, red unlikely to be usable) [151,152]. Bio-based colors are based on Table 2; numbers are Google “results” searching for “bio-based xxx” on 14 June 2020 to indicate the level of interest.
Table 4. The relative fit to requirements of the potential monomers for thermoplastic matrix marine composites (green is good, amber is marginal, red unlikely to be usable) [151,152]. Bio-based colors are based on Table 2; numbers are Google “results” searching for “bio-based xxx” on 14 June 2020 to indicate the level of interest.
PolymerEssential CriteriaDesirable CriteriaPass/Fail
Monomer Process Viscosity (mPa·s)Process Temperature (°C)Tg (°C)Moisture Absorption (%)Bio-BasedRecyclable
PA6~5130–20040–606–11✓ 196kTm = 219–230 °C
PA1223180–24040–50<2✓ 136kTm = 180 °C
PBT20–150180–26025–600.09✓ 1110kTm = 225 °C
PC250–300250–3001500.16✓ 286 Mamorphous (Tp *~235 °C)
PLA-150–18555–65<2✓ 22 MTm = 170–200 °C
PMMA (Elium®)10020–1001070.5✓ 220kdepolymerize
TPU800300−8–170.1✓ 1360kTm = 140 °C
PEK340–390340–3902280.07✓ 1140kTm = 385–413 °C
PET250–325250–325730.5✓ 217 MTm = 255 °C
PPA1000200–290121–1380.36✓ 1230kTm = 310–330 °C
* Tp represents the processing temperature.
Table 5. Cumulative process energy requirements (CPR), cumulative energy demand (CED) and the related cumulative CO2 emissions (CCO2) for polymeric products [194,195].
Table 5. Cumulative process energy requirements (CPR), cumulative energy demand (CED) and the related cumulative CO2 emissions (CCO2) for polymeric products [194,195].
PolymerCPR (GJ/ton)CED (GJ/ton)CCO2 (kg CO2/ton)
Thermoplastics
PA690.7122.74680
PC49.380.33110
PET33.459.42070
PLA~89.2501 (2334 *)
PMA **55.682.63740
PU48.575.53050
Thermosets
Epoxy73.6107.14680
Polyester37.564.52390
* Total fossil global warming potential excluding CO2 uptake. ** PMA environmental burdens below were for “polymethyl acrylate” (may not be PMMA).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Qin, Y.; Summerscales, J.; Graham-Jones, J.; Meng, M.; Pemberton, R. Monomer Selection for In Situ Polymerization Infusion Manufacture of Natural-Fiber Reinforced Thermoplastic-Matrix Marine Composites. Polymers 2020, 12, 2928. https://doi.org/10.3390/polym12122928

AMA Style

Qin Y, Summerscales J, Graham-Jones J, Meng M, Pemberton R. Monomer Selection for In Situ Polymerization Infusion Manufacture of Natural-Fiber Reinforced Thermoplastic-Matrix Marine Composites. Polymers. 2020; 12(12):2928. https://doi.org/10.3390/polym12122928

Chicago/Turabian Style

Qin, Yang, John Summerscales, Jasper Graham-Jones, Maozhou Meng, and Richard Pemberton. 2020. "Monomer Selection for In Situ Polymerization Infusion Manufacture of Natural-Fiber Reinforced Thermoplastic-Matrix Marine Composites" Polymers 12, no. 12: 2928. https://doi.org/10.3390/polym12122928

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop