Next Article in Journal
Macro and Micro Routes to High Performance Bioplastics: Bioplastic Biodegradability and Mechanical and Barrier Properties
Next Article in Special Issue
Investigation on the Durability of E-Glass/Epoxy Composite Exposed to Seawater at Elevated Temperature
Previous Article in Journal
Green Synthesis of Thermo-Responsive Hydrogel from Oil Palm Empty Fruit Bunches Cellulose for Sustained Drug Delivery
Previous Article in Special Issue
Microbial Degradation of Rubber: Actinobacteria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Impact of Prolonged Exposure of Eleven Years to Hot Seawater on the Degradation of a Thermoset Composite

by
Amir Hussain Idrisi
1,
Abdel-Hamid I. Mourad
1,2,3,* and
Muhammad M. Sherif
4
1
Department of Mechanical Engineering, United Arab Emirates University, Al Ain 15551, United Arab Emirates
2
National Water and Energy Center, United Arab Emirates University, Al Ain 15551, United Arab Emirates
3
Mechanical Design Department, Faculty of Engineering, Helwan University, Cairo 11795, Egypt
4
Department of Civil, Construction and Environmental Engineering, School of Engineering, University of Alabama at Birmingham, Alabama, AL 35294, USA
*
Author to whom correspondence should be addressed.
Submission received: 11 May 2021 / Revised: 3 June 2021 / Accepted: 8 June 2021 / Published: 30 June 2021
(This article belongs to the Special Issue Durability and Degradation of Polymeric Materials)

Abstract

:
This paper presents a long-term experimental investigation of E-glass/epoxy composites’ durability exposed to seawater at different temperatures. The thermoset composite samples were exposed to 23 °C, 45 °C and 65 °C seawater for a prolonged exposure time of 11 years. The mechanical performance as a function of exposure time was evaluated and a strength-based technique was used to assess the durability of the composites. The experimental results revealed that the tensile strength of E-glass/epoxy composite was reduced by 8.2%, 29.7%, and 54.4% after immersion in seawater for 11 years at 23 °C, 45 °C, and 65 °C, respectively. The prolonged immersion in seawater resulted in the plasticization and swelling in the composite. This accelerated the rate of debonding between the fibers and matrix. The failure analysis was conducted to investigate the failure mode of the samples. SEM micrographs illustrated a correlation between the fiber/matrix debonding, potholing, fiber pull-out, river line marks and matrix cracking with deterioration in the tensile characteristics of the thermoset composite.

1. Introduction

Glass fiber-reinforced polymer (GFRP) composites are commonly used in automobile, aerospace and marine industries because of their high strength to weight ratio, stiffness and durability. Furthermore, glass fibers are simple to manufacture, economical, less fragile and have high chemical resistance when compared with carbon fiber-reinforced polymers (CFRPs). However, GFRPs are prone to harsh environment conditions such as UV, seawater, alkaline and corrosive conditions [1,2,3,4,5,6]. At high temperatures, polymer-based composites trap water in voids and are often fused into hydroxyl radicals of the epoxy polymers [7,8]. The immersion of polymers in water leads to deterioration in the physical properties of composites [9,10,11,12,13]. This results in the inflation of the epoxy polymer, causing bond failure due to hydrolysis and plasticization [14,15,16]. As polymers are submerged in hot water, small cracks are initiated within the polymer matrix due to water absorption and osmotic edge rupture [17,18,19]. This increases the degradation of composites with immersion time. Therefore, it is essential to ensure the durability of FRP composites in the seawater environment. Nanofillers are an effective reinforcement that could close gaps and enhance the compositional intensity [20,21,22]. Several studies have been performed in different aging conditions to investigate the performance of epoxy-based composites. It has been observed that the flexural properties of FRP samples decreased due to seawater immersion [23,24].
Pavan et al. [25] illustrated the reduction in tensile strength of glass/epoxy from 200.83 MPa to 146.42 MPa, and 185.27 MPa for atmospheric and subzero temperatures aging for 3600 h, respectively. Yan et al. [26] noticed a similar reduction in the mechanical properties of flax-fabric/epoxy composites. The tensile strength of the specimens reduced by 31.1%, 28.3% and 22.6% under the immersion of alkaline (5% NaOH) solution, seawater and water, respectively. Silva et al. [27] studied the durability of GFRP laminates embedded in epoxy resin. The composite was exposed to saltwater at 30 °C, 50 °C and 65 °C for 5000 h (approximately 7 months). For the first 1500 h, the modulus behavior indicated that the prevailing damage mechanism was swelling. Then, between 1500 h and 2500 h, the elastic modulus behavior indicated that plasticization was the predominant damage mechanism, especially for samples aged at 30 °C. Mourad et al. [13] studied the combined effect of seawater and temperature on the durability of glass/epoxy and glass/polyurethane composites that were submerged for one year. The tensile strength of glass/epoxy and glass/polyurethane reduced by 1% and 19%, respectively, for specimens that were immersed for one year at 23 °C. The reduction in tensile strength increased to 6% and 31% for glass/epoxy and glass/polyurethane, respectively, as the solution temperature increased to 65 °C. The tensile strain and modulus of glass/epoxy were not affected significantly. However, the failure strain of glass/polyurethane reduced by 25% for the submerged specimens at 65 °C. Merah et al. [28] evaluated the effect of seawater and outdoor temperature on the tensile strength of GFRP composite for the immersion of 6 months and 12 months. The observed reduction was 9%, 28%, and 3.6% for tensile strength, tensile strain and tensile modulus, respectively, for specimens that were submerged for 6 months. While specimens exposed for 12 months experienced a decrease in strength, strain, and modulus was 12%, 32.7% and 5%, respectively. Chakraverty et al. [29] studied the effect of seawater conditioning for up to 12 months on the mechanical properties, such as the interlaminar shear strength (ILSS), tensile stress and failure strain, elastic modulus and glass transition temperature (Tg) of GFRP composites. The tensile stress was reduced to 79% after immersion for 6 months. However, a slow recovery was observed, and the tensile strength rebounded to 84% after immersion for 12 months. The ultimate strain decreased by 12% and 20% after 6 months and 12 months of seawater immersion, respectively. These reduction in values of ultimate stress and strain are due to the plasticization and swelling of the composite.
Chen Y. et al. [30] utilized the Arrhenius equation-based prediction model for predicting the durability of GFRP-reinforced concrete structures. The GFRP bars were immersed in a concrete solution at temperatures of 20 °C, 40 °C, and 60 °C. The short-term data were used for the accelerated aging tests. Mourad et al. [31] investigated the long-term effect of seawater and temperature on the durability of E-glass/epoxy and E-glass/polyurethane composite for the exposure of 7.5 years. It was observed that the tensile strengths of the E-glass/epoxy composite reduced by 6.3%, 24.1% and 48.9% after immersion in seawater at 23 °C, 45 °C, and 65 °C, respectively for the duration of 7.5 years. However, this reduction was 37.6%, 48.7% and 63.6%, respectively for E-glass/polyurethane. Table 1 summarizes the research related to E-glass/epoxy composite immersion at different temperature.
It is evident from the above literature that seawater exposure of FRP composite is limited to a few months. The main objective of this research is to conduct a long-term experimental investigation of the durability of FRP composites. E-glass/epoxy composite samples were exposed at different temperatures (23 °C, 45 °C and 65 °C) for the period of 132 months (11 years) in seawater. After exposure of the composites to the harsh environment, the tensile strength, tensile strain and young’s modulus of the specimens were measured. Furthermore, scanning electron microscopy (SEM) and differential scanning calorimetry (DSC) analysis were conducted to investigate the damage mechanism of the composites.

2. Experimental Procedures

In this research, the durability of unidirectional E-glass/epoxy composite reinforced with 52 vol% of glass fibers was investigated. The E-glass/epoxy was manufactured through a continuous lamination process and obtained from Gordon Composites, Inc. Samples with the thickness of 3 mm for the tensile test were fabricated as per ASTM Standard D-3039 [40]. The specimens had GFRP end-tabs bonded onto both ends to give uniform load distribution and to promote failure in the loading direction. The line diagram of the sample and its dimensions are given in Figure 1 and Table 2, respectively.
The shoulders of the samples were sealed by insulating tape to avoid any damage during the conditioning period. The samples were immersed in Gulf seawater at room temperature (23 °C), 45 °C, and 65 °C with an exposure time of 11 years. The tensile test results were recorded with an average of three test results of each group of specimens exposed to an environmental condition. The representative images of the conditioning chambers are shown in Figure 2.

3. Results and Discussion

3.1. Water Absorption

The weight of specimens was measured to calculate the absorbed moisture before immersion and after the removal of the specimens from the water tank. Figure 3 illustrates an increase in the weight of the specimen was observed at 23 °C, 45 °C, and 65 °C for E-glass/epoxy specimens. The water absorption or increase in weight was 4.1%, 5.1% and 6.7% for samples immersed for 11 years at 23 °C, 45 °C, and 65 °C, respectively. It is worth noting that the water absorption increased steeply in the first 18 months of immersion. The rate of absorption decreased from 18 months to 36 months of immersion for E-glass/epoxy composites. After 36 months of immersion, the composite nearly saturated and the increase in water absorption was negligible for the remaining period of immersion. The increase in weight is due to the water absorption that occurs through diffusive and/or capillary processes. Diffusion is considered a key process under which moisture penetrates polymeric composites and is known to be a matrix-dominated phenomenon in which water is mainly diffused into the matrix [41,42]. The capillary process involves moisture being drawn into voids and microcracks at the fiber/matrix interfaces, in which the cracks provide a transport system for moisture to penetrate [43,44].
Moreover, it was observed that an increase in absorption rate occurs at as the exposure temperature is increased. This could be due to the degradation (i.e., cracks are initiated and crack growth rate increases) of the polymeric composite with the increase in temperature. Similar findings were reported in the literature, as an increase in temperatures has influence on accelerating the moisture absorption and degradation process in polymers [45].

3.2. Tensile Testing

The tensile samples were manufactured according to the ASTM Standard D-3039 and tested at room temperature (300 K) using an MTS universal testing machine (MTS system corporation, Eden Prairie, MN, USA). The tests were conducted at a crosshead speed of 2 mm/min. The tensile strength of the samples after 11 years of exposure to seawater at different temperatures are summarized in Table 3 and Figure 4. The average of three test results for the same case is presented. The tensile strengths of the E-glass/epoxy composite after immersion in seawater at 23 °C, 45 °C, and 65 °C for the duration of 11 years was reduced from 794 MPa to 729 MPa, 558 MPa and 362 MPa, respectively. The reduction percentages are 8.2%, 29.7%, and 54.4%, respectively. These results show that exposure to seawater at elevated temperatures caused a higher reduction in the tensile strength than immersion in seawater at 23 °C [3,31,46,47,48,49,50,51]. The graph demonstrates that the tensile strength reduced by 46.6% after exposure for 5 years. However, a reduction in tensile strength of 14.6% was observed after exposure of 6 years and the immersion of specimens at 65 °C. The deteriorations in tensile strength could be ascribed to the combined effects of stress corrosion and diffusion of water molecules, resulting in the degradation of resin and weakening of the fiber/matrix interface. Furthermore, the chemical reaction between leaked metal ions (Na, K and Si) from E-glass fibers and Cl ions from seawater caused severe degradation to the tensile properties for samples submerged in seawater [52].
Figure 5 and Table 3 show the evolution of failure strain of the E-glass/epoxy composite after immersion in seawater at 23 °C, 45 °C, and 65 °C for the duration of 11 years. Figure 5 illustrates that after 2 years of exposure to seawater at 23 °C, the ultimate strain increased gradually from 2.14% to 2.50%. However, post 11 years of exposure, the specimens experienced a decrease in ultimate strain to 2.11%. Samples immersed at 45 °C had a similar behavior of an initial surge followed by a decline in failure strain, though at a faster rate. The ultimate strain for specimens submerged in seawater at 45 °C was 2.38% after 6 months of exposure which then decreased to 1.49% after 11 years of exposure. These observations illustrate that seawater uptake contributes to the plasticization of the epoxy matrix. This results in an increase in tensile strain and a decrease in tensile modulus. However, as the exposure duration increases, the matrix becomes more rigid and brittle, causing a decrease in failure strain and a slight increase in tensile modulus. This is more pronounced at 65 °C, as the tensile strain reduced to 0.86% after 11 years of immersion. Strait et al. [53] observed an increase in the impact energy of glass/epoxy due to initial plasticization at the early phase of seawater immersion. Additionally, Merah et al. [28] reported brittle fracture and a reduction in ultimate strain after prolonged exposure to seawater. In general, the prolonged seawater immersion causes plasticization and swelling in the composite material and affects the tolerance capacity of the composite material [29,54,55].
Figure 6 indicates that seawater immersion had a slight effect on the elastic modulus of the specimens exposed to various temperatures. The elastic modulus increased from 37.1 GPa to 37.9 GPa, 37.8 GPa, and 38.6 GPa for specimens immersed for 11 years at 23 °C, 45 °C, and 65 °C, respectively. The elastic modulus of the FRP specimens depended on the modulus of fibers. The immersion in the seawater environment had an insignificant effect on the durability of the fibers [56]. The data demonstrate that seawater immersion at the elevated temperature contributes to the epoxy matrix plasticization, which affects the failure strain and tensile modulus. Over a duration of 11 years of use, the matrix becomes stiff and fragile, allowing failure strain to decrease and tensile modulus to increase significantly.
The effect of temperature in seawater conditioning was examined by the analogy of stress–strain behavior of the sample immersed for 5 years and 11 years at 23 °C, 45 °C, and 65 °C, as shown in Figure 6. The strength of specimens conditioned in seawater at 23 °C decreased by 3% after 5 years of exposure and 8.2% after 11 years of exposure, as shown in Figure 7a,b, respectively. The results show that the failure strain reduced by 6.5% and 1.4% after immersion at 23 °C for 5 years and 11 years, respectively. However, the modulus increased by 0.6% and 2.15%. At 45 °C, the tensile strength of the E-glass/epoxy composite declined by 18.6% and 29.7% for 5 years and 11 years of immersion, respectively, whereas the ultimate strain was decreased by 18.2% and 30.4%, respectively. The modulus of specimens exposed to seawater was almost similar to control specimens for 5 years of immersion but increased by 1.9% after 11 years of exposure. For immersion at 65 °C, the slopes of the graphs show an increase in tensile modulus by 2% from 37.1 MPa to 37.8 MPa of samples exposed to 5 years and 4% from 37.1 MPa to 38.6 MPa for samples exposed to 11 years. The failure strain of the specimen significantly decreased by 48.6% after 5 years of exposure and 59.8% after 11 years of conditioning at 65 °C, as shown in Figure 7. The reduction in tensile strength and failure strain could be due to the reaction between water and epoxy that causes a breakdown of the polymer’s molecular weight, leading to the fragile nature of the matrix. Water can also function as an anti-plasticizer, preventing polymer segments from moving and making the matrix more brittle [29]. Merah et al. [28] observed a decrease in brittle fracture and failure strain due to prolonged immersion in seawater.

3.3. Failure Analysis

The analysis of the fractured surface was carried out using the JEOL-JSM 7610F SEM (JOEL Ltd., Tokyo, Japan) and micrographs are presented in Figure 8 and Figure 9. The strong bonding between the fibers and the matrix was observed in E-glass/epoxy samples immersed for 5 years at 23 °C and 65 °C, as illustrated in Figure 8a,b. Slight deterioration in the matrix phase was noticed after the immersion of 5 years at room temperature, as shown in Figure 8a. At 65 °C, in Figure 8b, the bonding at the fiber–matrix interface further reduced and matrix flow and river line marks were observed; however, the fibers remained unaffected. This implies matrix breakdown occurred due to the hydrolysis reaction and accelerated at high-temperature immersion. Furthermore, it resulted in a non-uniform load distribution between fibers [13]. Few fibers were pulled out for specimens exposed for 5 years in seawater at 23 °C. After 11 years of exposure, significant fiber pullout and potholing was observed for composites immersed at temperatures of 23 °C and 65 °C, as shown in Figure 9a,b, which indicate the deterioration of the fiber/matrix interface. These images indicate a reduction in the bonding strength between the matrix and the fiber. Furthermore, as the immersion time increased, it resulted in a degradation in the tensile strength of the composite. It has been reported by Chen et al. [57] that the degradation of E-glass fibers involves mainly the etching of free hydroxyl ions (OH-) and leaching of water molecules. The free hydroxyl ions (OH-) break the Si-O-Si bond of the glass fibers, especially for etching in alkaline solution [57,58]. The damage at the fiber/matrix interface involves a complex mechanism as the fiber/matrix interface is a heterogeneous area [59,60,61,62,63]. Fracture at the fiber–matrix interface is usually caused by the debonding between fiber and resin which occurs mainly in two stages. The first stage includes the breakage of chemical bonding due to chemical corrosion between fiber and resin. The second stage occurs due to the poor interlocking between fibers and resin as the resin swells through water absorption [64].
Further degradation at the interface of the fiber–matrix is due to the large and prevalent gaps between fibers and matrices. The reaction between water and epoxy believed to cause a breakdown of the polymer’s molecular weight, leading to the fragile nature of the matrix. Water can also function as an anti-plasticizer, preventing polymer segments from moving and making the matrix more brittle [29]. Damage to glass fibers is not evident, but its smooth cross-sectional surface of the pulled-out fibers reveals brittle failure of the fiber. This could be due to the loss in ductility of the fiber, lack of support from the brittle matrix, or unhindered crack propagation into the fiber [13].

3.4. Differential Scanning Calorimetry (DSC) Test

The differential scanning calorimetry (DSC) for the E-glass/epoxy composite was conducted on samples immersed at different temperatures and exposure times. This test was carried out using a TA-Instruments DSC Q200 device (TA-Instruments, New Castle, DE, USA). The test was performed in an inert environment using nitrogen gas. The experiment was run between 25 °C and 250 °C with a heating rate of 10.0 °C/min. Three samples for each condition were considered to determine the glass transition temperature. Figure 10 represents the DSC curves for the control sample and specimens immersed at 23 °C, 45 °C, and 65 °C for 5 years and 11 years. The glass transition temperatures (Tg) of the samples immersed at 23 °C and 45 °C for 5 years and 11 years were found to be similar to the Tg of the control specimens (118.2 °C). The glass transition temperature of specimens immersed for 11 years at 23 °C and 45 °C were 117.72 °C and 117.11 °C, respectively. It indicates that the seawater immersion at 23 °C and 45 °C had a negligible effect on the Tg of the composite. However, a slight reduction in Tg was observed for specimens immersed in seawater at 65 °C. The Tg of samples reduced to 114.21 °C and 114.27 °C after immersion of 5 years and 11 years, respectively, at 65 °C. The shift in Tg could be due to the combined effect of high temperature and seawater aging. The behavior could be associated with the degree of cross-linking or the degree of polymerization after curing. The degradation in an epoxy composite is owing to the existence of hydrophilic groups which form weak hydrogen bonds by reacting with water molecules during immersion at 65 °C [65]. The results demonstrate that the moisture uptake acts as a plasticizer and decreases Tg. The reduction in Tg represents thermal degradation and loss in mechanical properties, which limits the service temperature of the polymer. The mechanisms of the long chain of the polymer could be isolated at high-temperature immersion and react with each other to alter the properties of the polymer [66].

3.5. Fourier Transform Infrared Spectroscopy (FTIR) Analysis

The Fourier transform infrared spectroscopy (FTIR) was performed on a Perkin Elmer Spectrum 100 FTIR spectrometer (PerkinElmer Life and Analytical Sciences, Shelton, CT, USA) at room temperature in the transmission mode. FTIR spectra were logged in between 600 cm−1 and 4000 cm−1 at a resolution of 2 cm−1 with 10 scans. Before testing the samples, background spectra were taken in the empty chamber to eliminate the influence of moisture and CO2 in air. Figure 11 presents a comparison of typical spectra for the control specimens of E-glass/epoxy and specimens immersed at 23 °C, 45 °C, and 65 °C for 11 years. The seawater exposure led to the presence of equivalent FTIR bands on the residue spectrum due to leaching of functional groups from the resin of composite.
The control E-glass/epoxy specimens had two peaks at 2296.3 cm−1 and 2353.2 cm−1. It indicates the presence of unreacted R-N=C=O groups [67,68]. The two peaks became weak after immersion of 11 years at 23 °C and 45 °C, but disappeared for specimens exposed to 65 °C. The vanishing of the two peaks at 2296.3 cm−1 and 2353.2 cm−1 after immersion for 11 years at 65 °C is due to the reaction between the water molecules and unreacted R-N=C=O groups, which can be presented as
R-N=C=O + H2O→R − NH2 + CO2
Therefore, the degradation at the fiber/matrix interface during immersion at 65 °C can be described as the release of CO2, shown in Equation (1). At this immersion temperature, the sustained release of CO2 is responsible for the increase in mass loss. This results in a reduction in the tensile strength of the composite [69,70,71,72].

4. Conclusions

In this research, E-glass/epoxy was conditioned at different temperatures (23 °C, 45 °C, and 65 °C) and exposure times (up to 11 years) to investigate the durability of the composite. The results show that seawater immersion slightly affects the durability of the composite at 23 °C; however, it reduced to more than half at 65 °C. The tensile strength of the E-glass/epoxy composites reduced by 8.2%, 29.7%, and 54.4% after immersion in seawater for 11 years at 23 °C, 45 °C, and 65 °C, respectively. The tensile strength reduced by 46.6% after exposure to seawater for 5 years, but the reduction was only 14.6% in the remaining 6 years of immersion at 65 °C. Similarly, the failure strain of the composites reduced with the increase in immersion temperature and exposure time. The degradation behavior was justified by the analysis of SEM micrographs, DSC curves and FTIR spectra. The deterioration was observed mainly owing to breakage of chemical bonding between fiber and resin due to chemical corrosion, and poor interlocking between fibers and resin due to resin swelling through water absorption.

Author Contributions

Conceptualization, A.H.I. and A.-H.I.M.; methodology, A.H.I. and A.-H.I.M.; formal analysis, A.H.I.; investigation, A.H.I. and A.-H.I.M.; resources, A.-H.I.M.; data curation, A.H.I. and M.M.S.; writing—original draft preparation, A.H.I.; writing—review and editing, A.-H.I.M. and M.M.S.; supervision, A.-H.I.M.; funding acquisition, A.-H.I.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the UPAR research program at the United Arab Emirates University, UAE [grant number 31N225].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data and models used during the study appear in the submitted article.

Acknowledgments

The authors would also like to acknowledge the United Arab Emirates University for providing the research facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jamali, J.; Mourad, A.-H.I.; Fan, Y.; Wood, J.T. Through-thickness fracture behavior of unidirectional glass fibers/epoxy composites under various in-plane loading using the CTS test. Eng. Fract. Mech. 2016, 156, 83–95. [Google Scholar] [CrossRef]
  2. Mourad, A.-H.I.; Al Mansoori, M.S.; Al Marzooqi, L.A.; Genena, F.A.; Cherupurakal, N. Optimization of curing conditions and nanofiller incorporation for production of high performance laminated Kevlar/epoxy nanocomposites. In Proceedings of the ASME 2018 Pressure Vessels and Piping Conference, Prague, Czech Republic, 15–20 July 2018. [Google Scholar]
  3. Idrisi, A.H.; Ismail Mourad, A.-H.I.; Abdel-Magid, B.; Mozumder, M.; Afifi, Y. Impact of the Harsh Environment on E-Glass Epoxy Composite. In Proceedings of the ASME 2019 Pressure Vessels & Piping Conference, San Antonio, TX, USA, 14–19 July 2019; Volume 6A. [Google Scholar] [CrossRef]
  4. Sathishkumar, T.P.; Satheeshkumar, S.; Naveen, J. Glass fiber-reinforced polymer composites—A review. J. Reinf. Plast. Compos. 2014, 33, 1258–1275. [Google Scholar] [CrossRef]
  5. Fouad, H.; Mourad, A.-H.I.; ALshammari, B.A.; Hassan, M.K.; Abdallah, M.Y.; Hashem, M. Fracture toughness, vibration modal analysis and viscoelastic behavior of Kevlar, glass, and carbon fiber/epoxy composites for dental-post applications. J. Mech. Behav. Biomed. Mater. 2020, 101, 103456. [Google Scholar] [CrossRef] [PubMed]
  6. Mourad, A.-H.I.; Abdel-Magid, B. Durability of Thermoset Composites in Seawater Environment. In Proceedings of the CAMX 2015, Dallas, TX, USA, 26–29 October 2015. [Google Scholar]
  7. Maggana, C.; Pissis, P. Water sorption and diffusion studies in an epoxy resin system. J. Polym. Sci. Part B Polym. Phys. 1999, 37, 1165–1182. [Google Scholar] [CrossRef]
  8. Gonon, P.; Sylvestre, A.; Teysseyre, J.; Prior, C. Combined effects of humidity and thermal stress on the dielectric properties of epoxy-silica composites. Mater. Sci. Eng. B 2001, 83, 158–164. [Google Scholar] [CrossRef]
  9. Yilmaz, T.; Sinmazcelik, T. Effects of hydrothermal aging on glass–fiber/polyetherimide (PEI) composites. J. Mater. Sci. 2010, 45, 399–404. [Google Scholar] [CrossRef]
  10. Gautier, L.; Mortaigne, B.; Bellenger, V. Interface damage study of hydrothermally aged glass-fibre-reinforced polyester composites. Compos. Sci. Technol. 1999, 59, 2329–2337. [Google Scholar] [CrossRef]
  11. Al-Kuwaiti, M.H.; Mourad, A.-H.I. Thermomechanical characteristics of compacted and noncompacted plain weave woven laminated composites. In Proceedings of the ASME 2017 Pressure Vessels and Piping Conference, Waikoloa, HI, USA, 16–20 July 2017; Volume 3A. [Google Scholar] [CrossRef]
  12. Al-Kuwaiti, M.H.H.; Mourad, A.-H.I. Effect of different environmental conditions on the mechanical behavior of plain weave woven laminated composites. Procedia Eng. 2015, 130, 638–643. [Google Scholar] [CrossRef] [Green Version]
  13. Mourad, A.-H.I.; Abdel-Magid, B.M.; El-Maaddawy, T.; Grami, M.E. Effect of seawater and warm environment on glass/epoxy and glass/polyurethane composites. Appl. Compos. Mater. 2010, 17, 557–573. [Google Scholar] [CrossRef]
  14. Zheng, Q.; Morgan, R.J. Synergistic thermal-moisture damage mechanisms of epoxies and their carbon fiber composites. J. Compos. Mater. 1993, 27, 1465–1478. [Google Scholar] [CrossRef]
  15. De’Nève, B.; Shanahan, M.E.R. Water absorption by an epoxy resin and its effect on the mechanical properties and infra-red spectra. Polymer 1993, 34, 5099–5105. [Google Scholar] [CrossRef]
  16. Xiao, G.Z.; Delamar, M.; Shanahan, M.E.R. Irreversible interactions between water and DGEBA/DDA epoxy resin during hygrothermal aging. J. Appl. Polym. Sci. 1997, 65, 449–458. [Google Scholar] [CrossRef]
  17. Hodzic, A.; Kim, J.K.; Lowe, A.E.; Stachurski, Z.H. The effects of water aging on the interphase region and interlaminar fracture toughness in polymer–glass composites. Compos. Sci. Technol. 2004, 64, 2185–2195. [Google Scholar] [CrossRef]
  18. Ellyin, F.; Maser, R. Environmental effects on the mechanical properties of glass-fiber epoxy composite tubular specimens. Compos. Sci. Technol. 2004, 64, 1863–1874. [Google Scholar] [CrossRef]
  19. Mourad, A.-H.I. Thermo-mechanical characteristics of thermally aged polyethylene/polypropylene blends. Mater. Des. 2010, 31, 918–929. [Google Scholar] [CrossRef]
  20. Mourad, A.-H.I.; Idrisi, A.H.; Zaaroura, N.; Sherif, M.M.; Fouad, H. Damage assessment of nanofiller-reinforced woven kevlar KM2plus/Epoxy resin laminated composites. Polym. Test. 2020, 86, 106501. [Google Scholar] [CrossRef]
  21. Fan, X.J.; Lee, S.W.R.; Han, Q. Experimental investigations and model study of moisture behaviors in polymeric materials. Microelectron. Reliab. 2009, 49, 861–871. [Google Scholar] [CrossRef]
  22. Mourad, A.-H.I.; Zaaroura, N. Impact of nanofillers incorporation on laminated nanocomposites performance. J. Mater. Eng. Perform. 2018, 27, 4453–4461. [Google Scholar] [CrossRef]
  23. Karbhari, V.M.; Chu, W. Degradation kinetics of pultruded E-glass/vinylester in alkaline media. ACI Mater. J. 2005, 102, 34. [Google Scholar]
  24. Gellert, E.P.; Turley, D.M. Seawater immersion ageing of glass-fibre reinforced polymer laminates for marine applications. Compos. Part A Appl. Sci. Manuf. 1999, 30, 1259–1265. [Google Scholar] [CrossRef]
  25. Pavan, A.; Dayananda, P.; Vijaya, K.M.; Hegde, S.; Hosagade, P.N. Influence of seawater absorption on vibrational and tensile characteristics of quasi-isotropic glass/epoxy composites. J. Mater. Res. Technol. 2019, 8, 1427–1433. [Google Scholar] [CrossRef]
  26. Yan, L.; Chouw, N. Effect of water, seawater and alkaline solution ageing on mechanical properties of flax fabric/epoxy composites used for civil engineering applications. Constr. Build. Mater. 2015, 99, 118–127. [Google Scholar] [CrossRef]
  27. Silva, M.A.G.; da Fonseca, B.S.; Biscaia, H. On estimates of durability of FRP based on accelerated tests. Compos. Struct. 2014, 116, 377–387. [Google Scholar] [CrossRef]
  28. Merah, N.; Nizamuddin, S.; Khan, Z.; Al-Sulaiman, F.; Mehdi, M. Effects of harsh weather and seawater on glass fiber reinforced epoxy composite. J. Reinf. Plast. Compos. 2010, 29, 3104–3110. [Google Scholar] [CrossRef]
  29. Chakraverty, A.P.; Mohanty, U.K.; Mishra, S.C.; Satapathy, A. Sea water ageing of GFRP composites and the dissolved salts. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2015; Volume 75, p. 12029. [Google Scholar]
  30. Chen, Y.; Davalos, J.F.; Ray, I. Durability prediction for GFRP reinforcing bars using short-term data of accelerated aging tests. J. Compos. Constr. 2006, 10, 279–286. [Google Scholar] [CrossRef]
  31. Mourad, A.-H.I.; Idrisi, A.H.; Wrage, M.C.; Abdel-Magid, B.M. Long-term durability of thermoset composites in seawater environment. Compos. Part B Eng. 2019, 168, 243–253. [Google Scholar] [CrossRef]
  32. Hu, Y.; Li, X.; Lang, A.W.; Zhang, Y.; Nutt, S.R. Water immersion aging of polydicyclopentadiene resin and glass fiber composites. Polym. Degrad. Stab. 2016, 124, 35–42. [Google Scholar] [CrossRef]
  33. Guermazi, N.; Tarjem, A.B.; Ksouri, I.; Ayedi, H.F. On the durability of FRP composites for aircraft structures in hygrothermal conditioning. Compos. Part B Eng. 2016, 85, 294–304. [Google Scholar] [CrossRef]
  34. Bobbaa, S.; Lemana, Z.; Zainudina, E.S.; Sapuana, S.M. The influence of hydrothermal aging on E-glass and S-glass fiber/epoxyreinforced composite pipes. J. Mater. Environ. Sci. 2019, 10, 790–804. [Google Scholar]
  35. Feng, P.; Wang, J.; Wang, Y.; Loughery, D.; Niu, D. Effects of corrosive environments on properties of pultruded GFRP plates. Compos. Part B Eng. 2014, 67, 427–433. [Google Scholar] [CrossRef]
  36. Deniz, M.E.; Karakuzu, R. Seawater effect on impact behavior of glass–epoxy composite pipes. Compos. Part B Eng. 2012, 43, 1130–1138. [Google Scholar] [CrossRef]
  37. Wei, B.; Cao, H.; Song, S. Degradation of basalt fibre and glass fibre/epoxy resin composites in seawater. Corros. Sci. 2011, 53, 426–431. [Google Scholar] [CrossRef]
  38. Antunes, M.B.; Almeida, J.H.S., Jr.; Amico, S.C. Curing and seawater aging effects on mechanical and physical properties of glass/epoxy filament wound cylinders. Compos. Commun. 2020, 22, 100517. [Google Scholar] [CrossRef]
  39. Ghabezi, P.; Harrison, N. Mechanical behavior and long-term life prediction of carbon/epoxy and glass/epoxy composite laminates under artificial seawater environment. Mater. Lett. 2020, 261, 127091. [Google Scholar] [CrossRef]
  40. D3039 ASTM. Standard Test Method for Tensile Properties of Polymer Matrix Composite Materials; ASTM Int: West Conshohocken, PA, USA, 2008. [Google Scholar]
  41. Ray, B.C. Temperature effect during humid ageing on interfaces of glass and carbon fibers reinforced epoxy composites. J. Colloid Interface Sci. 2006, 298, 111–117. [Google Scholar] [CrossRef]
  42. Akbar, S.; Zhang, T. Moisture diffusion in carbon/epoxy composite and the effect of cyclic hygrothermal fluctuations: Characterization by Dynamic Mechanical Analysis (DMA) and Interlaminar Shear Strength (ILSS). J. Adhes. 2008, 84, 585–600. [Google Scholar] [CrossRef]
  43. Patel, S.R.; Case, S.W. Durability of hygrothermally aged graphite/epoxy woven composite under combined hygrothermal conditions. Int. J. Fatigue 2002, 24, 1295–1301. [Google Scholar] [CrossRef]
  44. Li, Y.; Li, R.; Huang, L.; Wang, K.; Huang, X. Effect of hygrothermal aging on the damage characteristics of carbon woven fabric/epoxy laminates subjected to simulated lightning strike. Mater. Des. 2016, 99, 477–489. [Google Scholar] [CrossRef]
  45. Almudaihesh, F.; Holford, K.; Pullin, R.; Eaton, M. The influence of water absorption on unidirectional and 2D woven CFRP composites and their mechanical performance. Compos. Part B Eng. 2020, 182, 107626. [Google Scholar] [CrossRef]
  46. Mourad, A.-H.I.; Alnaqbi, A.H. Load Frame Testing Device Adapted to Test Different Environmental Conditions. U.S. Patent 10,132,731, 20 November 2018. [Google Scholar]
  47. Mourad, A.-H.I.; Alnaqbi, A.H. Load Testing System and Device for Sustain Load Measurement over Time in Different Environmental Conditions. U.S. Patent 10,101,251, 16 October 2018. [Google Scholar]
  48. Sabry, I.; Ghafaar, M.A.; Mourad, A.-H.I.; Idrisi, A.H. Stir casted SiC-Gr/Al6061 hybrid composite tribological and mechanical properties. SN Appl. Sci. 2020, 2, 943. [Google Scholar] [CrossRef] [Green Version]
  49. Mourad, A.-H.I.; Greish, Y.E.; Ayad, O.G. Processing and mechanical performance of Hydroxyapatite-UHMWPE composites. In Proceedings of the IEEE 2019 Advances in Science and Engineering Technology International Conferences, Dubai, United Arab Emirates, 26 March–10 April 2019; pp. 1–5. [Google Scholar]
  50. Iftikhar, S.H.; Mourad, A.-H.I.; Sheikh-Ahmad, J. An overview of friction stir welding of high-density polyethylene. In Proceedings of the IEEE 2020 Advances in Science and Engineering Technology International Conferences, Dubai, United Arab Emirates, 4 February–9 April 2020; pp. 1–6. [Google Scholar]
  51. Idrisi, A.H.; Mourad, A.-H.I. Impact of the harsh environment on GFRE and GFRPol composite. In Proceedings of the Ocean 2019 MTS/IEEE SEATTLE, Seattle, WA, USA, 27–31 October 2019; pp. 1–5. [Google Scholar] [CrossRef]
  52. Wang, J.; GangaRao, H.; Liang, R.; Zhou, D.; Liu, W.; Fang, Y. Durability of glass fiber-reinforced polymer composites under the combined effects of moisture and sustained loads. J. Reinf. Plast. Compos. 2015, 34, 1739–1754. [Google Scholar] [CrossRef]
  53. Strait, L.H.; Karasek, M.L.; Amateau, M.F. Effects of seawater immersion on the impact resistance of glass fiber reinforced epoxy composites. J. Compos. Mater. 1992, 26, 2118–2133. [Google Scholar] [CrossRef]
  54. Mejia, E.B.; Mourad, A.-H.I.; Faqer, A.S.B.; Halwish, D.F.; Al Hefeiti, H.O.; Al Kashadi, S.M.; Cherupurakal, N.; Mozumder, M.S. Impact on HDPE mechanical properties and morphology due to processing. In Proceedings of the IEEE 2019 Advances in Science and Engineering Technology International Conferences, Dubai, United Arab Emirates, 26 March–10 April 2019; pp. 1–5. [Google Scholar]
  55. Mourad, A.-H.I.; Cherupurakal, N.; Hafeez, F.; Barsoum, I.; A Genena, F.; S Al Mansoori, M.; A Al Marzooqi, L. Impact Strengthening of Laminated Kevlar/Epoxy Composites by Nanoparticle Reinforcement. Polymers 2020, 12, 2814. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, Z.; Zhao, X.-L.; Xian, G.; Wu, G.; Raman, R.K.S.; Al-Saadi, S.; Haque, A. Long-term durability of basalt-and glass-fibre reinforced polymer (BFRP/GFRP) bars in seawater and sea sand concrete environment. Constr. Build. Mater. 2017, 139, 467–489. [Google Scholar] [CrossRef]
  57. Chen, Y.; Davalos, J.F.; Ray, I.; Kim, H.-Y. Accelerated aging tests for evaluations of durability performance of FRP reinforcing bars for concrete structures. Compos. Struct. 2007, 78, 101–111. [Google Scholar] [CrossRef]
  58. Charles, R.J. The strength of silicate glasses and some crystalline oxides. In Proceedings of the International Conference on Atomic Mechanisms of Fracture, Swampscott, MA, USA, 12–16 April 1959; MIT Press: Cambridge, MA, USA, 1959; pp. 225–250. [Google Scholar]
  59. Robert, M.; Benmokrane, B. Combined effects of saline solution and moist concrete on long-term durability of GFRP reinforcing bars. Constr. Build. Mater. 2013, 38, 274–284. [Google Scholar] [CrossRef]
  60. Mourad, A.-H.I.; Fouad, H.; Elleithy, R. Impact of some environmental conditions on the tensile, creep-recovery, relaxation, melting and crystallinity behaviour of UHMWPE-GUR 410-medical grade. Mater. Des. 2009, 30, 4112–4119. [Google Scholar] [CrossRef]
  61. Mourad, A.-H.I.; Elsayed, H.F.; Barton, D.C. Semicrystalline polymers deformation and fracture behaviour under quasistatic strain rates and triaxial states of stress. Strength Fract. Complex. 2004, 2, 149–162. [Google Scholar]
  62. Mourad, A.-H.I.; Dehbi, A. On use of trilayer low density polyethylene greenhouse cover as substitute for monolayer cover. Plast. Rubber Compos. 2014, 43, 111–121. [Google Scholar] [CrossRef]
  63. Mourad, A.-H.I.; Akkad, R.O.; Soliman, A.A.; Madkour, T.M. Characterisation of thermally treated and untreated polyethylene–polypropylene blends using D.S.C.; TGA and IR techniques. Plast. Rubber Compos. 2009, 38, 265–278. [Google Scholar] [CrossRef]
  64. Wang, Z.; Huang, X.; Xian, G.; Li, H. Effects of surface treatment of carbon fiber: Tensile property, surface characteristics, and bonding to epoxy. Polym. Compos. 2016, 37, 2921–2932. [Google Scholar] [CrossRef]
  65. Zafar, A.; Bertocco, F.; Schjødt-Thomsen, J.; Rauhe, J.C. Investigation of the long term effects of moisture on carbon fibre and epoxy matrix composites. Compos. Sci. Technol. 2012, 72, 656–666. [Google Scholar] [CrossRef]
  66. Kawagoe, M.; Takeshima, M.; Nomiya, M.; Qiu, J.; Morita, M.; Mizuno, W.; Kitano, H. Microspectroscopic evaluations of the interfacial degradation by absorbed water in a model composite of an aramid fibre and unsaturated polyester. Polymer 1999, 40, 1373–1380. [Google Scholar] [CrossRef]
  67. Babaghayou, M.I.; Mourad, A.-H.I.; Lorenzo, V.; Chabira, S.F.; Sebaa, M. Anisotropy evolution of low density polyethylene greenhouse covering films during their service life. Polym. Test. 2018, 66, 146–154. [Google Scholar] [CrossRef]
  68. Guo, F.; Al-Saadi, S.; Raman, R.K.S.; Zhao, X.L. Durability of fiber reinforced polymer (FRP) in simulated seawater sea sand concrete (SWSSC) environment. Corros. Sci. 2018, 141, 1–13. [Google Scholar] [CrossRef]
  69. Mozumder, M.S.; Mourad, A.-H.I.; Mairpady, A.; Pervez, H.; Haque, M.E. Effect of TiO 2 Nanofiller Concentration on the Mechanical, Thermal and Biological Properties of HDPE/TiO 2 Nanocomposites. J. Mater. Eng. Perform. 2018, 27, 2166–2181. [Google Scholar] [CrossRef]
  70. Babaghayou, M.I.; Mourad, A.-H.I.; Ochoa, A.; Beltrán, F.; Cherupurakal, N. Study on the thermal stability of stabilized and unstabilized low-density polyethylene films. Polym. Bull. 2020, 1–17. [Google Scholar] [CrossRef]
  71. Zhou, J.; Lucas, J.P. Hygrothermal effects of epoxy resin. Part I: The nature of water in epoxy. Polymer 1999, 40, 5505–5512. [Google Scholar] [CrossRef]
  72. Zhou, J.; Lucas, J.P. Hygrothermal effects of epoxy resin. Part II: Variations of glass transition temperature. Polymer 1999, 40, 5513–5522. [Google Scholar] [CrossRef]
Figure 1. Line diagram of tensile test specimen.
Figure 1. Line diagram of tensile test specimen.
Polymers 13 02154 g001
Figure 2. Representative images of the conditioning chambers (a) 23 °C, (b) 45 and 65 °C.
Figure 2. Representative images of the conditioning chambers (a) 23 °C, (b) 45 and 65 °C.
Polymers 13 02154 g002
Figure 3. Variation in weight of E-glass/epoxy composite with conditioning duration and temperature.
Figure 3. Variation in weight of E-glass/epoxy composite with conditioning duration and temperature.
Polymers 13 02154 g003
Figure 4. Tensile strength versus exposure time at different temperatures for E-glass/epoxy composite.
Figure 4. Tensile strength versus exposure time at different temperatures for E-glass/epoxy composite.
Polymers 13 02154 g004
Figure 5. Tensile strain versus exposure time at different temperatures for E-glass/epoxy composite.
Figure 5. Tensile strain versus exposure time at different temperatures for E-glass/epoxy composite.
Polymers 13 02154 g005
Figure 6. Effect of seawater immersion on the modulus of elasticity.
Figure 6. Effect of seawater immersion on the modulus of elasticity.
Polymers 13 02154 g006
Figure 7. Stress–strain behavior of specimen immersed in seawater for (a) 5 years, (b) 11 years.
Figure 7. Stress–strain behavior of specimen immersed in seawater for (a) 5 years, (b) 11 years.
Polymers 13 02154 g007
Figure 8. Fracture surfaces of the E-glass/epoxy composite conditioned for 5 years at (a) 23 °C (b) 65 °C.
Figure 8. Fracture surfaces of the E-glass/epoxy composite conditioned for 5 years at (a) 23 °C (b) 65 °C.
Polymers 13 02154 g008
Figure 9. Fracture surfaces of the E-glass/epoxy composite conditioned for 11 years at (a) 23 °C (b) 65 °C.
Figure 9. Fracture surfaces of the E-glass/epoxy composite conditioned for 11 years at (a) 23 °C (b) 65 °C.
Polymers 13 02154 g009
Figure 10. DSC curves for E-glass/epoxy control sample and conditioned samples.
Figure 10. DSC curves for E-glass/epoxy control sample and conditioned samples.
Polymers 13 02154 g010
Figure 11. FTIR spectrum of E-glass/epoxy control sample and conditioned samples.
Figure 11. FTIR spectrum of E-glass/epoxy control sample and conditioned samples.
Polymers 13 02154 g011
Table 1. Glass/epoxy composite immersed at different environment conditions.
Table 1. Glass/epoxy composite immersed at different environment conditions.
AuthorsCompositeConditioningDurationTests Performed
Silva et al. [27]E-glass/epoxySaltwater at 30 °C, 45 °C and 55 °C750–5000 hrs (Approx 7 months)Water absorption, tensile properties
Chakraverty et al. [29]E-glass/epoxySeawater at room temperature2, 4, 6, 8, 10, and12 months (1 year)Water absorption, ILSS properties, DSC, failure analysis
Hu et al. [32]Glass/polydicycl-opentadiene and glass/epoxySaltwater and deionized water at 60 °C1, 3, 6, and 12 months (1 year)Water absorption, DMA, tensile strength, ILSS,
Guermazi et al. [33]Glass/epoxy, carbon/epoxy and glass/carbon/epoxyTapwater at 24 ± 3, 70 and 90 °C3 monthsWater absorption, DMA, tensile and flexural test, abrasive wear test
Bobba et al. [34]E-glass and S-glass fiber–epoxyTap water at 90 °C600, 1200, and 1800 hrs (2.5 months)Impact test
Feng et al. [35]Glass/epoxyH2SO4, NaOH and NaCl at 60 and 90 °C7, 15, 30, and 90 days (3 months)Water absorption, flexural properties, hardness, SEM analysis
Merah et al. [28]Glass fiber-reinforced epoxy (GFRE)Seawater at outdoor temperature6 and 12 monthsTensile properties, failure analysis
Mourad et al. [13]Glass/epoxy and glass/polyurethaneSeawater at 23 °C and 65 °C3, 6, 9, and 12 monthsWater absorption, tensile properties, DSC, failure analysis
Mourad et al. [31]E-glass/epoxy and E-glass/polyurethaneSeawater at 23 °C and 65 °CUp to 7.5 yearsTensile properties, failure analysis
EminDeniz et al. [36]Glass/epoxy compositeSeawater at 20 °C3, 6, 9, and 12 monthsImpact test
Pavan et al. [25]E-glass/epoxy laminatesArtificial seawater in sub-zero and ambient temperatures3600 h(5 months)Water absorption, tensile properties, and failure analysis
Wei et al. [37]Basalt fiber-reinforced plastic (BFRP) and glass fiber-reinforced plastic (GFRP)Artificial seawater at 25 °C10, 20, 30, 60, and 90 daysWater absorption, tensile and flexural properties, and microstructural analysis
Antunes et al. [38]Glass/epoxy filament wound cylindersSeawater at 80 °C7–28 daysHoop stress, failure analysis
Ghabezi et al. [39]Carbon/epoxy and glass/epoxyArtificial seawater at room temperature and 60 °C45 daysTensile and 3-point bending
Table 2. Tensile specimen geometry.
Table 2. Tensile specimen geometry.
ParametersDimensions
Specimen length, L250 mm
Specimen width, b15 mm
Specimen thickness, h3 mm
Gauge length, S150 mm
Tab length, Ltab50 mm
Tab thickness, htab4 mm
Tab bevel angle90°
Table 3. Tensile properties of E-glass/epoxy after conditioning in seawater.
Table 3. Tensile properties of E-glass/epoxy after conditioning in seawater.
Conditioning Duration (Months)23 °C45 °C65 °C
Tensile Strength (MPa)Failure Strain (%)Tensile Modulus (GPa)Tensile Strength (MPa)Failure Strain (%)Tensile Modulus (GPa)Tensile Strength (MPa)Failure Strain (%)Tensile Modulus (GPa)
0 (Control Sample)794 ± 462.14 ± 0.0237.1 ± 2.5794 ± 462.14 ± 0.0237.1 ± 2.5794 ± 462.14 ± 0.0237.1 ± 2.5
6791 ± 542.4 ± 0.033.33 ± 2.5786 ± 402.38 ± 0.133.3 ± 1.6761 ± 202.3 ± 0.2033.1 ± 0.6
12789 ± 412.1 ± 0.2338.49 ± 5.4754 ± 492.17 ± 1.16536.9 ± 4.3721 ± 442.2 ± 0.134.7 ± 3.1
15787 ± 382.4 ± 0.2034.73 ± 0.8712 ± 431.97 ± 0.1535.1 ± 0.5630 ± 471.5 ± 0.135.3 ± 0.3
18781 ± 652.3 ± 0.133.82 ± 1.7697 ± 381.92 ± 0.01432.2 ± 1.4565 ± 481.5 ± 0.1836.9 ± 1.1
24778 ± 872.5 ± 0.3832.16 ± 1.7687 ± 352.02 ± 0.27532.8 ± 1.3502 ± 511.5 ± 0.1733.0 ± 0.9
36775 ± 92 ± 0.4937.83 ± 3.2659 ± 311.85 ± 0.2637.9 ± 1.8484 ± 221.3 ± 0.0337.1 ± 0.4
60770 ± 532 ± 0.2537.31 ± 1.5646 ± 431.75 ± 0.16537.2 ± 1.6424 ± 531.1 ± 0.0837.8 ± 1.8
72758 ± 221.98 ± 0.0337.68 ± 2.2625 ± 301.81 ± 0.0539.1 ± 1.8413 ± 271.1 ± 0.0739.7 ± 1.4
84750 ± 341.9 ± 0.2738.81 ± 0.8612 ± 421.67 ± 0.2635.4 ± 1.1409 ± 851.2 ± 0.2533.7 ± 1.4
90744 ± 61.85 ± 0.0834.96 ± 1.3602 ± 211.54 ± 0.0635.8 ± 1.1406 ± 271.1 ± 0.0436.9 ± 0.9
96741 ± 282.06 ± 0.1538.88 ± 2.8596 ± 341.55 ± 0.07537.5 ± 1.8401 ± 321.13 ± 0.0339.2 ± 1.1
102740 ± 652.03 ± 0.0938.7 ± 1.7588 ± 391.61 ± 0.1638.4 ± 1.5398 ± 331.09 ± 0.1839.8 ± 1.7
108739 ± 422.03 ± 0.1738.15 ± 2.6583 ± 411.58 ± 0.16537.4 ± 1.7391 ± 210.93 ± 0.0736.4 ± 0.9
114738 ± 392.05 ± 0.1238.21 ± 1.2579 ± 371.52 ± 0.1337.4 ± 1.1381 ± 380.98 ± 0.1539.0 ± 1.2
120733 ± 412.1 ± 0.1637.99 ± 1.4572 ± 251.5 ± 0.15638.1 ± 1.4373 ± 310.95 ± 0.0439.5 ± 1.6
126731 ± 452.08 ± 0.1438.4 ± 1.5564 ± 281.53 ± 0.1638.5 ± 1.6369 ± 220.9 ± 0.139.4 ± 1.4
132729 ± 482.11 ± 0.1137.9 ± 2.3558 ± 311.49 ± 0.0937.8 ± 1.9362 ± 250.86 ± 0.0838.6 ± 1.8
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Idrisi, A.H.; Mourad, A.-H.I.; Sherif, M.M. Impact of Prolonged Exposure of Eleven Years to Hot Seawater on the Degradation of a Thermoset Composite. Polymers 2021, 13, 2154. https://0-doi-org.brum.beds.ac.uk/10.3390/polym13132154

AMA Style

Idrisi AH, Mourad A-HI, Sherif MM. Impact of Prolonged Exposure of Eleven Years to Hot Seawater on the Degradation of a Thermoset Composite. Polymers. 2021; 13(13):2154. https://0-doi-org.brum.beds.ac.uk/10.3390/polym13132154

Chicago/Turabian Style

Idrisi, Amir Hussain, Abdel-Hamid I. Mourad, and Muhammad M. Sherif. 2021. "Impact of Prolonged Exposure of Eleven Years to Hot Seawater on the Degradation of a Thermoset Composite" Polymers 13, no. 13: 2154. https://0-doi-org.brum.beds.ac.uk/10.3390/polym13132154

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