Next Article in Journal
Cotton Textile with Antimicrobial Activity and Enhanced Durability Produced by L-Cysteine-Capped Silver Nanoparticles
Previous Article in Journal
Multicriteria Decision Making of a Life Cycle Engineered Rack and Pinion System
Previous Article in Special Issue
Fabrication and Thermal Performance of 3D Copper-Mesh-Sintered Foam/Paraffin Phase Change Materials for Solar Thermal Energy Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quasi-Static Flexural Behavior of Epoxy-Matrix-Reinforced Crump Rubber Composites

1
Department of Mechanical Engineering, School of Mechanical, Chemical and Materials Engineering, Adama Science and Technology University, Adama 1888, Ethiopia
2
Space and Aeronautic Research Institute, King Abdulaziz City for Science and Technology (KACST), Riyadh 11442, Saudi Arabia
3
Department of Mechanical Engineering, College of Engineering, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
4
Mechanical Engineering Department, College of Engineering, King Khalid University, Abha 61421, Saudi Arabia
5
Research Center for Advanced Materials Science (RCAMS), King Khalid University, P.O. Box 9004, Guraiger, Abha 61413, Saudi Arabia
6
Department of Mechanical and Industrial Engineering, College of Engineering, Majmaah University, Al-Majmaah, Riyadh 11952, Saudi Arabia
7
Department of Mechanical Engineering, Nitte Meenakshi Institute of Technology Bengaluru, Bangalore 560064, India
*
Author to whom correspondence should be addressed.
Submission received: 1 April 2022 / Revised: 22 April 2022 / Accepted: 26 April 2022 / Published: 11 May 2022
(This article belongs to the Special Issue Particulate-Filled Advanced Polymer Composites)

Abstract

:
Waste tires have emerged as a severe environmental threat worldwide as they create a number of disposal and landfill burden issues. In the present study, environmental pollutant crump rubber derived from waste discarded tires was reinforced with epoxy resin and prepared by means of an open-mold casting method to assess its mechanical properties. The impact of crump rubber content (0, 10, 20 and 30 vol.%) on the mechanical behavior of the composites was assessed using three-point bending tests at a constant strain rate of 0.1 and 0.01 mm/s. The stress–strain profiles of the 0.01 mm/s specimens revealed higher strains to failure compared with the 0.1 mm/s tested specimens and all the specimens showed brittle failure. Irrespective of the strain rates, tests revealed a marginal increase in the strength values of the composites and a significant increase in the modulus of all the composites compared with neat epoxy specimens. The results suggest that crump rubber can be effectively used in utilitarian composites requiring good flexural modulus and strength properties. Crump rubber epoxy composites with 30 vol.% of crump rubber showed higher modulus and strength compared with neat epoxy and other composites owing to the toughening phase induced by the crump rubber particles. The failure and fracture features of the specimens were analyzed using scanning electron microscopy.

1. Introduction

Polymeric matrixes involving reinforcement with particulate or functional components are called particulate polymer matrix composites. The mechanical and thermal response of particulate polymer matrix composites are often reported to be enhanced owing to the presence of particulates [1,2]. Polymer matrix composites have gained significant importance in the last few decades mainly attributed to their ability to be tuned for desired properties [3]. Polymer matrix composites are widely employed in the manufacturing of products ranging from small devices to heavy parts and are used in many applications in the aerospace sector (e.g., structural work in commercial aircraft, satellite systems), the marine sector (e.g., body components of boats, kayak, and canoes), in sports equipment (e.g., footwear, sports goods), the bio-medical sector (e.g., medical imbeds, fabrication of devices for MRI scanners and X-rays), electrical sectors (e.g., printed circuit boards, electrical panels and electronic devices), automotive sectors, and so on [3,4,5]. Thus, there is always a need to develop new material systems specially aimed at meeting the envisaged demands of applications. Material development with unwanted constituents has attracted significant attention from the research community around the world to address and reduce disposal and environmental pollution issues arising from the difficulties highlighted. The major advantage of utilizing waste materials with widely available matrix materials is to enable the proper utilization of waste materials and to reduce polymer consumption [6,7]. The ready accessibility of different types of waste materials, combined with the low cost of components, increase the feasibility of utilizing these materials in commonly used composites.
A substantial number of studies are available on the usage of waste reinforcements in widely used matrix resins, such as fly ash cenosphere [8,9,10,11,12,13], rice husk [14,15,16], blast furnace slag [17,18], waste tire rubber [19,20,21,22,23,24], walnut shell powder [25,26], etc. Rubber particles derived from discarded automobile tires have attracted substantial interest from researchers to mitigate the problems associated with waste tire rubbers, including landfill burden, disposal issues and severe environmental pollution. Thus, more focus is being directed to looking for viable options to address these problems [27]. The non-biodegradability and difficult disintegration of waste tires, mainly attributed to their characteristic cross- linking formation, has led to more attention directed to tackling the problems in an effective way [28,29,30,31]. Recycled rubber particles derived from discarded automobile tires are termed crump rubber [32]. Carbon, sulfur, calcium and oxygen constitute a major percentage of crump rubber, with small amounts of silica, aluminum, titanium, iron and zinc also present [33,34,35,36]. Epoxy resins are an extensively used thermosetting polymer credited to their distinctive features when manufactured, including small pressure requirements to prepare products, small cure shrinkage and residual stresses [37,38]. Thereby, composites fabricated with epoxy and crump rubber for envisaged applications can benefit from effective waste utilization and reduction in cost. A number of studies have reported the use of waste tire particles as an effective means of strengthening in epoxy-based composites.
Studies related to quasi-static flexural behavior are very rare, and no investigations have been reported on the quasi-static flexural response of crump rubber epoxy composites. Strain-rate sensitivity is a very important issue in the transport segment with considerable attention devoted to the crashworthiness of systems [39]. The mechanical properties of epoxy-based composites are viewed as strain-rate sensitive. Numerous studies on epoxy-base composites indicate that the strength of composites increases with strain rate [39,40]. Though strain-rate sensitivity is mainly attributed to the matrix, the percentage volume of reinforcements also significantly affects the failure mechanism. Furthermore, quasi-static tests on composites assist in understanding the response of materials in more detail as the materials’ response is significantly decreased during the tests [41,42,43]. In the present study, quasi-static flexural tests were performed on four varieties of crump rubber epoxy composites consisting of varying crump rubber volume fractions (0, 10, 20 and 30 vol.%). The volume fractions of crump rubber were selected based on previous studies [44,45]. Quasi-static tests were performed at two strain rates of 0.1 and 0.01 mm/s conforming to ASTM standards and significantly lower than the normal strain rate of 1.42 mm/min. The effect of crump rubber on the flexural properties and specific properties of composites was evaluated. Finally, the properties of the composites were related to their structures using scanning electron micrographs.

2. Materials and Methods

2.1. Materials

LAPOX L-12 epoxy resin (diglycidyl ether of bisphenol A) and K6 hardener (triethylene tetra amine) manufactured by Atul Industries (Ahmedabad, India) were used as matrix material and curing agent, respectively. The matrix resin and K6 hardener had densities of 1192 and 954 kg/m3, respectively. Crump rubber particles obtained from Arihant Chemicals Ltd. (Kolkata, India) were used as reinforcement with a density of 1451 kg/m3. The properties of the crump rubber particles and the epoxy resin are presented in Table 1. The chemical structures of diglycidyl ether of bisphenol A and triethylene tetra amine are presented in Figure 1. K-6 hardener is a diamine that reacts with and opens the epoxide rings to form a 3D network polymer [46].

2.2. Fabrication

An open-mold casting method was used to process particulate polymer composites with variable crump rubber (0–30 vol.%) volume percentage. The desired quantity of matrix resin and crump rubber were mixed in a beaker at a steady rate to achieve a consistent slurry. Trapped air bubbles produced during mechanical mixing of constituents were released by degassing the prepared slurry for several minutes. K6 hardener by 10 wt.% was poured into the slurry and mixed further to start the polymerization reaction and was finally poured into a mold layered with silicone. The curing of samples was performed for 24 h. Samples were cut into the desired shape of specimens according to ASTM standards using a water jet cutting machine. A schematic representation of the fabrication of specimens is presented in Figure 2. The EC-VP convention was used to represent the specimen configuration, wherein E, C and VV indicate epoxy, crump rubber and volume fraction of reinforcement, respectively.

2.3. Experimental Program

Flexural specimens with dimensions of 127 × 12.7 × 3.2 mm were prepared according to ASTM D-790 standards [48]. A Zwick Universal testing machine with 20 kN load cell was utilized to perform quasi-static flexural tests at two strain rates of 0.1 and 0.01 mm/s, with a constant span length of 52 mm. The average values of three specimens per composition are presented in the graphs. A scanning electron microscope JEOL (JSM 6380 LA) was used to investigate the microstructure of the tested compositions. The surfaces of the specimens were sputter-coated to a thickness of up to 9 nm with gold-palladium using an auto fine coater to prevent charge build-up.

3. Results and Discussions

3.1. Material Processing

The densities of the rump rubber epoxy composites were calculated theoretically and were reported experimentally in research carried out by Shahapurkar K. [23]. The densities of the composites for various compositions is presented in Figure 3a. The densities of the composites increased with increase in crump rubber content and were higher than for neat epoxy. Details of crump rubber and epoxy resin were reported by Shahapurkar K. [23]. The crump rubber mean particle size was reported as 182.24 μm, available in the work reported by Shahapurkar K. [22]. A micrograph of crump rubber is shown in Figure 3b. The images of testing are presented in Figure 4a,b.

3.2. Stress–Strain Curves

Stress–strain curves of composites subjected to 0.1 and 0.01 mm/s strain rate are shown in Figure 5a,b, respectively. Irrespective of the strain rates, the stress–strain curves in both cases were identical in nature and revealed linear stress strain curves until the elastic region and showed brittle failure thereafter. However, with decrease in the strain rate, all the composites, including the neat epoxy specimen, revealed higher strain to failure indicating good absorption of energy, mainly attributed to the plasticizing effect of the crump rubber [49,50,51]. A marked rise in stress can also be seen with increasing crump rubber content in the flexural profiles. Similar stress strain profiles can be observed in the work carried out by Tagliavia et al. and Garcia et al. [52,53]. Decreasing the strain rate provided more time for the system to adapt to the load and thereby the failure of the composites was prolonged. Such occurrences necessitate the study of composites under varying strain rates to understand the behavior of the systems in more detail.

3.3. Flexural Modulus

The flexural modulus of all the composites, irrespective of the strain rate, increased linearly with increase in the crump rubber content and was lower than for the neat epoxy specimens (Figure 6). The percentage increase in crump rubber further enhanced the capacity of composites to withstand more load and thereby resulted in a higher modulus. Reinforcing a distributed hardening phase in the form of crump rubber in matrix considerably improved the stiffness of the composites [54]. The higher entanglement of polymer chains of crump rubber particles with the epoxy matrix resulted in tightly creased bonding of constituents, contributing to the enhancement of the modulus of the composites [55,56]. As the behavior of the crump rubber particles is inherently elastic in nature, they assist the epoxy matrix to effectively attenuate the deformation arising by stretching to large strains [57]. As a result, the ability of composites to bridge the flaws arising due to deformation increased with increase in the crump rubber content. Additionally, the toughness of the composites reinforced with crump rubber was also dependent on the chemical composition and cure schedule [22,46,58]. The EC-30 composites revealed the highest modulus compared with the other compositions, regardless of the strain rate. The specific modulus of all the composites increased with increase in the crump rubber content and was lower than for the neat epoxy specimen. The higher specific modulus of composites, despite possessing higher density, demonstrated the weight-saving potential of composites. For the 0.1 and 0.01 mm/s strain rates, the modulus of the EC-30 composites showed increments of 73% and 56%, while the specific modulus increased by 64% and 48%, respectively.

3.4. Flexural Strength

A marginal increase in the composite strength was observed with increase in the crump rubber volume composition in contrast to the neat specimens, irrespective of the strain rates (Figure 7). Epoxy matrixes are highly cross-linked and, therefore, are susceptible to brittle failures—reinforcing the matrix with crump rubber toughened the composites marginally. In the present study, all the composites showed a slight increment in strength compared to the neat specimens. For the 0.1 and 0.01 mm/s strain rates, the strength of the EC-30 composites showed increments of 9% and 7%, while the specific strength increased by 4% and 2%, respectively. The particle bridging mechanisms of the crump rubber particles to stretch and bridge the crack surfaces were mainly credited with the increase in the strength of the composites [20,59,60]. The neat epoxy specimens were brittle in nature and, therefore, revealed low strength. On the other hand, the low filler content of the crump rubber particles tended to make cracks advance in the matrix rich phase and showed lower strength, while higher crump rubber volume had a higher tendency to restrict splitting of crack planes and to strain harden swiftly due to inherent elastic behavior. Therefore, higher fracture energy was needed for cracks to propagate, resulting in overall strength enhancement of the composites [57].

3.5. Fracture Features

Post-flexural-test micrographs of specimens subjected to different strain rates are shown in Figure 8. Neat epoxy specimens are presented in Figure 8a,b. Striation marks evident on the surface revealed brittle failure of the neat specimens. Reinforcing with a low amount of crump rubber also revealed striation marks (Figure 8c,d); however, bridging of the cracks and a reduced level of striation was evident from the micrographs. The brittle mode of failure of specimens was governed by the formation of flaws on the tensile side of specimens resulting in a small quantity of debris [61]. Micrographs with 30 vol.% at 0.1 and 0.01 mm/s are shown in Figure 8e,f, respectively. Increasing the crump rubber content significantly reduced the striation marks on the surface, revealing a ductile response of the composites, compared with a brittle response at lower filler content and for the neat epoxy specimens. The EC-30 composites also produced higher levels of debris than E0 and EC-10. The debris implies that specimens failed on the compression side. The EC-30 composites underwent more deformation due to the higher crump rubber content and, therefore, revealed higher modulus and strength values in comparison with the other compositions. Furthermore, the crump rubber particles could be observed to be closely bonded with the epoxy matrix and uniformly distributed across the surface. These factors assisted in increasing the stiffness of the composites.

4. Conclusions

In the present investigation, quasi-static flexural tests were performed on four different types of crump rubber epoxy composites consisting of varying crump rubber volume fractions (0, 10, 20 and 30 vol.%). Quasi- static tests were performed at two strain rates of 0.1 and 0.01 mm/s.
The main conclusions drawn from the study are summarized below:
  • The density of all the composites and the void percentages increased with increase in the crump rubber volume percentage and were higher than for neat epoxy.
  • The stress–strain profiles of the specimens subjected to 0.01 mm/s showed higher strains to failure compared with 0.1 mm/s tested specimens, while all the specimens, irrespective of the strain rates, showed brittle failure.
  • The modulus of the composites increased with increasing crump rubber content and was noted to be higher than for the neat epoxy specimens. For the 0.1 and 0.01 mm/s strain rates, the modulus of the EC-30 composites showed increments of 73% and 56%, while the specific modulus increased by 64% and 48%, respectively, compared with the other compositions.
  • A marginal increase in strength was observed for all the composites in contrast with the neat specimens. For the 0.1 and 0.01 mm/s strain rates, the strength of the EC-30 composites showed increments of 9% and 7%, while the specific strength increased by 4% and 2%, respectively, in comparison with the other composites.
  • Irrespective of the strain rates, the tests revealed a marginal increase in strength and a significant increase in the modulus of the composites in contrast to the neat specimens.
  • Scanning electron micrographs of the fractured surfaces revealed a uniform distribution of constituents and a higher deformation ability of crump rubber particles, which assisted in enhancing the composites’ performance. The higher crump rubber percentage in the composites reduced the deformation of specimens significantly and helped in effective stress transfer between the constituents.
  • The results presented in the study imply that epoxy composites consisting of crump rubber can be effectively utilized in specific envisaged applications demanding a higher flexural modulus and strength properties.
The present study successfully demonstrated the use of crump rubber particles in the most widely used epoxy matrix. The dual benefits of addressing environmental issues and reducing polymer consumption were realized by the study.

Author Contributions

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

Funding

Taif University Researchers Supporting Project number (TURSP-2020/349), Taif University, Taif, Kingdom of Saudi Arabia. and the Deanship of Scientific Research, King Khalid University (KKU), Abha-61421, Asir, Kingdom of Saudi Arabia under the grant number RGP.1/74/42.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the information is available within the article.

Acknowledgments

The research was supported by the Taif University Researchers Supporting Project number (TURSP-2020/349), Taif University, Taif, Kingdom of Saudi Arabia. The authors also gratefully acknowledge the Deanship of Scientific Research, King Khalid University (KKU), Abha-61421, Asir, Kingdom of Saudi Arabia for funding this research work under the grant number RGP.1/74/42. The authors also thank Department of Mechanical Engineering, Adama Science and Technology University, Ethiopia for their support and facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Karimi, D.; Milani, A.S. SRVE modeling of particulate polymer matrix composites with irregularly shaped inclusions: Application to a green stone composite. Compos. Struct. 2019, 228, 111331. [Google Scholar] [CrossRef]
  2. Damadzadeh, B.; Jabari, H.; Skrifvars, M.; Airola, K.; Moritz, N.; Vallittu, P.K. Effect of ceramic filler content on the mechanical and thermal behaviour of poly-l-lactic acid and poly-l-lactic-co-glycolic acid composites for medical applications. J. Mater. Sci. Mater. Med. 2010, 21, 2523–2531. [Google Scholar] [CrossRef] [PubMed]
  3. Devaraju, S.; Hariharan, A.; Balaji, K.; Alagar, M. Thermal and Morphological Analyses of Polymer Matrix Composites. In Encyclopedia of Materials: Composites; Brabazon, D., Ed.; Elsevier: Oxford, UK, 2021; pp. 1038–1068. [Google Scholar] [CrossRef]
  4. Oladele, I.O.; Omotosho, T.F.; Adediran, A.A. Polymer-Based Composites: An Indispensable Material for Present and Future Applications. Int. J. Polym. Sci. 2020, 2020, 8834518. [Google Scholar] [CrossRef]
  5. Gupta, N.; Zeltmann, S.E.; Shunmugasamy, V.C.; Pinisetty, D. Applications of polymer matrix syntactic foams. Jom 2014, 66, 245–254. [Google Scholar] [CrossRef]
  6. Abdel-Shafy, H.I.; Mansour, M.S.M. Solid waste issue: Sources, composition, disposal, recycling, and valorization. Egypt. J. Pet. 2018, 27, 1275–1290. [Google Scholar] [CrossRef]
  7. Awoyera, P.O.; Adesina, A. Plastic wastes to construction products: Status, limitations and future perspective. Case Stud. Constr. Mater. 2020, 12, e00330. [Google Scholar] [CrossRef]
  8. Shahapurkar, K.; Garcia, C.D.; Doddamani, M.; Mohan Kumar, G.C.; Prabhakar, P. Compressive behavior of cenosphere/epoxy syntactic foams in arctic conditions. Compos. Part B Eng. 2018, 135, 253–262. [Google Scholar] [CrossRef]
  9. Shahapurkar, K.; Chavan, V.B.; Doddamani, M.; Kumar, G.C.M. Influence of surface modification on wear behavior of fly ash cenosphere/epoxy syntactic foam. Wear 2018, 414–415, 327–340. [Google Scholar] [CrossRef]
  10. Shahapurkar, K.; Doddamani, M.; Mohan Kumar, G.C.; Gupta, N. Effect of cenosphere filler surface treatment on the erosion behavior of epoxy matrix syntactic foams. Polym. Compos. 2019, 40, 2109–2118. [Google Scholar] [CrossRef]
  11. Hegde, S.; Padmaraj, N.H.; Siddesh, V.; Sunaya, T.S.; Kini, K.A.; Sanil, V.K. Experimental investigation of mechanical sustainability and acoustic performance of fly ash cenosphere/epoxy polymer composites. J. King Saud. Univ.-Eng. Sci. 2021. [Google Scholar] [CrossRef]
  12. Dalbehera, S.; Acharya, S.K. Effect of cenosphere addition on the mechanical properties of jute-glass fiber hybrid epoxy composites. J. Ind. Text. 2015, 46, 177–188. [Google Scholar] [CrossRef]
  13. Nguyen, T.A.; Pham, T.M.H. Study on the properties of epoxy composites using fly ash as an additive in the presence of nanoclay: Mechanical properties, flame retardants, and dielectric properties. J. Chem. 2020, 2020, 8854515. [Google Scholar] [CrossRef]
  14. Satapathy, A.; Alok Kumar, J.; Mantry, S.; Singh, S.K.; Patnaik, A. Processing and Characterization of Jute-Epoxy Composites Reinforced with SiC Derived from Rice Husk. J. Reinf. Plast. Compos. 2009, 29, 2869–2878. [Google Scholar] [CrossRef]
  15. Kumar, S.; Mer, K.K.S.; Gangil, B.; Patel, V.K. Synergy of rice-husk filler on physico-mechanical and tribological properties of hybrid Bauhinia-vahlii/sisal fiber reinforced epoxy composites. J. Mater. Res. Technol. 2019, 8, 2070–2082. [Google Scholar] [CrossRef]
  16. Jain, N.; Somvanshi, K.S.; Gope, P.C.; Singh, V.K. Mechanical characterization and machining performance evaluation of rice husk/epoxy an agricultural waste based composite material. J. Mech. Behav. Mater. 2019, 28, 29–38. [Google Scholar] [CrossRef]
  17. Patnaik, P.K.; Swain, P.T.R.; Biswas, S. Investigation of mechanical and abrasive wear behavior of blast furnace slag-filled needle-punched nonwoven viscose fabric epoxy hybrid composites. Polym. Compos. 2019, 40, 2335–2345. [Google Scholar] [CrossRef]
  18. Patnaik, P.K.; Biswas, S. Effect of blast furnace slag content on mechanical and slurry abrasion behavior of needle-punched nonwoven fabric reinforced epoxy composites. Adv. Polym. Technol. 2018, 37, 1764–1773. [Google Scholar] [CrossRef]
  19. Shahapurkar, K.; Soudagar, M.E.M.; Shahapurkar, P.; Mathapathi, M.; Khan, T.M.Y.; Mujtaba, M.A.; Ali, M.D.I.; Thanaiah, K.; Siddiqui, M.I.H.; Ali, M.A. Effect of crump rubber on the solid particle erosion response of epoxy composites. J. Appl. Polym. Sci. 2022, 139, 51470. [Google Scholar] [CrossRef]
  20. Shahapurkar, K.; Chenrayan, V.; Soudagar, M.E.M.; Badruddin, I.A.; Shahapurkar, P.; Elfasakhany, A.; Mujtaba, M.A.; Siddiqui, M.I.; Ali, M.A.; Mahlia, T.M. Leverage of Environmental Pollutant Crump Rubber on the Dry Sliding Wear Response of Epoxy Composites. Polymers 2021, 13, 2894. [Google Scholar] [CrossRef]
  21. Shahapurkar, K.; Chenrayan, V.; Tesfamarium, B.B.; Soudagar, M.E.M.; Hossain, N.; Rajhi, A.A.; Alamri, S.; Alarifi, I.M.; Shahapurkar, P.; Mujtaba, M.A.; et al. Parametric Analysis of Epoxy/Crumb Rubber Composite by Using Taguchi—GRA Hybrid Technique. Polymers 2021, 13, 3441. [Google Scholar] [CrossRef]
  22. Shahapurkar, K. Tensile behavior of environmental pollutant crumb rubber filled epoxy composites. Mater. Res. Express 2021, 8, 095505. [Google Scholar] [CrossRef]
  23. Shahapurkar, K. Compressive behavior of crump rubber reinforced epoxy composites. Polym. Compos. 2021, 42, 329–341. [Google Scholar] [CrossRef]
  24. Alblalaihid, K.; Shahapurkar, K.; Chennarayan, V.; Beyecha, H.; Tirth, V.; Algahtani, A.; Soudagar, M.E.M.; Talugeri, V. Experimental investigation on the flexural response of epoxy composites filled with environmental pollutant crump rubber. Mater. Res. Express 2022, 9, 035503. [Google Scholar] [CrossRef]
  25. Salasinska, K.; Barczewski, M.; Górny, R.; Kloziński, A. Evaluation of highly filled epoxy composites modified with walnut shell waste filler. Polym. Bull. 2018, 75, 2511–2528. [Google Scholar] [CrossRef] [Green Version]
  26. Singh, V.K. Mechanical behavior of walnut (Juglans L.) shell particles reinforced bio-composite. Sci. Eng. Compos. Mater. 2015, 22, 383–390. [Google Scholar] [CrossRef]
  27. Bulei, C.; Todor, M.P.; Heput, T.; Kiss, I. Directions for material recovery of used tires and their use in the production of new products intended for the industry of civil construction and pavements. IOP Conf. Ser. Mater. Sci. Eng. 2018, 294, 012064. [Google Scholar] [CrossRef]
  28. Wu, B.; Zhou, M.H. Recycling of waste tyre rubber into oil absorbent. Waste Manag. 2009, 29, 355–359. [Google Scholar] [CrossRef]
  29. Yilmaz, A.; Degirmenci, N. Possibility of using waste tire rubber and fly ash with Portland cement as construction materials. Waste Manag. 2009, 29, 1541–1546. [Google Scholar] [CrossRef]
  30. Rowhani, A.; Rainey, T.J. Scrap Tyre Management Pathways and Their Use as a Fuel—A Review. Energies 2016, 9, 888. [Google Scholar] [CrossRef]
  31. Sofi, A. Effect of waste tyre rubber on mechanical and durability properties of concrete—A review. Ain Shams Eng. J. 2018, 9, 2691–2700. [Google Scholar] [CrossRef]
  32. Khed, V.C.; Mohammed, B.S.; Nuruddin, M.F. Effects of Different Crumb Rubber Sizes on the Flowability and Compressive Strength of Hybrid Fibre Reinforced ECC; IOP Publishing: Bristol, UK, 2018. [Google Scholar]
  33. Mashaan, N.S.; Ali, A.H.; Karim, M.R.; Abdelaziz, M. A Review on Using Crumb Rubber in Reinforcement of Asphalt Pavement. Sci. World J. 2014, 2014, 214612. [Google Scholar] [CrossRef] [PubMed]
  34. Jie, X.U.; Yao, Z.; Yang, G.; Han, Q. Research on crumb rubber concrete: From a multi-scale review. Constr. Build. Mater. 2020, 232, 117282. [Google Scholar]
  35. Lapkovskis, V.; Mironovs, V.; Kasperovich, A.; Myadelets, V.; Goljandin, D. Crumb rubber as a secondary raw material from waste rubber: A short review of end-of-life mechanical processing methods. Recycling 2020, 5, 32. [Google Scholar] [CrossRef]
  36. Irmawaty, R.; Noor, N.M.; Muhaimin, A.A. Feasibility of crumb rubber as fine aggregate in concrete. IOP Conf. Ser. Earth Environ. Sci. 2020, 419, 012054. [Google Scholar] [CrossRef]
  37. Sukanto, H.; Raharjo, W.W.; Ariawan, D.; Triyono, J.; Kaavesina, M. Epoxy resins thermosetting for mechanical engineering. Open Eng. 2021, 11, 797–814. [Google Scholar] [CrossRef]
  38. Drake, R.S.; Egan, D.R.; Murphy, W.T. Elastomer-modified epoxy resins in coatings applications. ACS Symp. Ser. 1983, 221, 1–20. [Google Scholar]
  39. Bharath Kumar, B.R.; Singh, A.K.; Doddamani, M.; Luong, D.D.; Gupta, N. Quasi-Static and High Strain Rate Compressive Response of Injection-Molded Cenosphere/HDPE Syntactic Foam. JOM 2016, 68, 1861–1871. [Google Scholar] [CrossRef]
  40. Song, B.; Chen, W.W.; Lu, W.Y. Mechanical characterization at intermediate strain rates for rate effects on an epoxy syntactic foam. Int. J. Mech. Sci. 2007, 49, 1336–1343. [Google Scholar] [CrossRef]
  41. Sutherland, L.S.; Soares, C.G. The use of quasi-static testing to obtain the low-velocity impact damage resistance of marine GRP laminates. Compos. Part B Eng. 2012, 43, 1459–1467. [Google Scholar] [CrossRef]
  42. Jayavardhan, M.L.; Doddamani, M. Quasi-static compressive response of compression molded glass microballoon/HDPE syntactic foam. Compos. Part B Eng. 2018, 149, 165–177. [Google Scholar] [CrossRef]
  43. Doddamani, M. Effect of surface treatment on quasi-static compression and dynamic mechanical analysis of syntactic foams. Compos. Part B Eng. 2019, 165, 365–378. [Google Scholar] [CrossRef]
  44. Shivamurthy, B.; Doreswamy, D.; Nishanth, J.; HC, S.P. Physical and tribo-mechanical properties of waste rubber tyre/epoxy composites. Mater. Res. Express 2019, 6, 095321. [Google Scholar] [CrossRef]
  45. Adesina, A.Y.; Zainelabdeen, I.H.; Dalhat, M.A.; Mohammed, A.S.; Sorour, A.A.; Al-Badour, F.A. Influence of micronized waste tire rubber on the mechanical and tribological properties of epoxy composite coatings. Tribol. Int. 2020, 146, 106244. [Google Scholar] [CrossRef]
  46. Verma, A.; Negi, P.; Singh, V.K. Experimental investigation of chicken feather fiber and crumb rubber reformed epoxy resin hybrid composite: Mechanical and microstructural characterization. J. Mech. Behav. Mater. 2018, 27, 3–4. [Google Scholar] [CrossRef]
  47. Licari, J.J.; Swanson, D.W. Chapter 3-Chemistry, Formulation, and Properties of Adhesives. In Adhesives Technology for Electronic Applications, 2nd ed.; Licari, J.J., Swanson, D.W., Eds.; William Andrew Publishing: Oxford, UK, 2011; pp. 75–141. [Google Scholar] [CrossRef]
  48. ASTM D790-17; Standard Test Methods for Flexural Properties of Unreinforced and Reinforced Plastics and Electrical Insulating Materials. ASTM International: West Conshohocken, PA, USA, 2010.
  49. Tripathi, G.; Srivastava, D. Studies on the physico-mechanical and thermal characteristics of blends of DGEBA epoxy, 3,4 epoxy cyclohexylmethyl, 3′,4′-epoxycylohexane carboxylate and carboxyl terminated butadiene co-acrylonitrile (CTBN). Mater. Sci. Eng. A 2008, 496, 483–493. [Google Scholar] [CrossRef]
  50. Dadfar, M.R.; Ghadami, F. Effect of rubber modification on fracture toughness properties of glass reinforced hot cured epoxy composites. Mater. Des. 2013, 47, 16–20. [Google Scholar] [CrossRef]
  51. Calabrese, L.; Valenza, A. Effect of CTBN rubber inclusions on the curing kinetic of DGEBA–DGEBF epoxy resin. Eur. Polym. J. 2003, 39, 1355–1363. [Google Scholar] [CrossRef]
  52. Tagliavia, G.; Porfiri, M.; Gupta, N. Analysis of flexural properties of hollow-particle filled composites. Compos. Part B Eng. 2010, 41, 86–93. [Google Scholar] [CrossRef]
  53. Garcia, C.D.; Shahapurkar, K.; Doddamani, M.; Kumar, G.C.M.; Prabhakar, P. Effect of arctic environment on flexural behavior of fly ash cenosphere reinforced epoxy syntactic foams. Compos. Part B Eng. 2018, 151, 265–273. [Google Scholar] [CrossRef]
  54. Kargarzadeh, H.; Ahmad, I.; Abdullah, I. Liquid rubbers as toughening agents. In Micro-Nanostructured Epoxy/Rubber Blends; John Wiley & Sons: Hoboken, NJ, USA, 2014; pp. 31–52. [Google Scholar]
  55. Mark, J.E. Rubber elasticity. J. Chem. Educ. 1981, 58, 898. [Google Scholar] [CrossRef] [Green Version]
  56. Wang, H.; Liu, X.; Zhang, H.; Apostolidis, P.; Erkens, S.; Skarpas, A. Micromechanical modelling of complex shear modulus of crumb rubber modified bitumen. Mater. Des. 2020, 188, 108467. [Google Scholar] [CrossRef]
  57. Kunz-Douglass, S.; Beaumont, P.W.R.; Ashby, M.F. A model for the toughness of epoxy-rubber particulate composites. J. Mater. Sci. 1980, 15, 1109–1123. [Google Scholar] [CrossRef]
  58. Meeks, A.C. Fracture and mechanical properties of epoxy resins and rubber-modified epoxy resins. Polymer 1974, 15, 675–681. [Google Scholar] [CrossRef]
  59. D, K.; Shahapurkar, K.; Venkatesh, C.; R, M.; Tirth, V.; Manivannan, C.; Alarifi, I.M.; Soudagar, M.E.M.; El-Shafay, A.S. Influence of Graphene Nano Fillers and Carbon Nano Tubes on the Mechanical and Thermal Properties of Hollow Glass Microsphere Epoxy Composites. Processes 2022, 10, 40. [Google Scholar] [CrossRef]
  60. Yee, A.F.; Pearson, R.A. Toughening mechanisms in elastomer-modified epoxies. J. Mater. Sci. 1986, 21, 2462–2474. [Google Scholar] [CrossRef] [Green Version]
  61. Gupta, N.; Woldesenbet, E. Microscopic Studies of Syntactic Foams Tested Under Three-Point Bending Conditions. In Proceedings of the ASME 2002 Engineering Technology Conference on Energy, Houston, TX, USA, 4–5 February 2002; pp. 147–152. [Google Scholar]
Figure 1. Chemical structure of (a) diglycidyl ether of bisphenol A, and (b) triethylene tetra amine [19,47].
Figure 1. Chemical structure of (a) diglycidyl ether of bisphenol A, and (b) triethylene tetra amine [19,47].
Processes 10 00956 g001
Figure 2. Schematic representation of the fabrication of composites.
Figure 2. Schematic representation of the fabrication of composites.
Processes 10 00956 g002
Figure 3. (a) Density of the composites and (b) Micrograph of crump rubber particles.
Figure 3. (a) Density of the composites and (b) Micrograph of crump rubber particles.
Processes 10 00956 g003
Figure 4. Images of test conduction on the specimens (a) before loading, and (b) during loading.
Figure 4. Images of test conduction on the specimens (a) before loading, and (b) during loading.
Processes 10 00956 g004
Figure 5. Characteristic stress–strain curves of all the composites at strain rates of (a) 0.1 mm/s and (b) 0.01 mm/s.
Figure 5. Characteristic stress–strain curves of all the composites at strain rates of (a) 0.1 mm/s and (b) 0.01 mm/s.
Processes 10 00956 g005
Figure 6. Experimentally measured flexural modulus at (a) 0.1 mm/s and (b) 0.01 mm/s, and specific flexural modulus at (c) 0.1 mm/s and (d) 0.01 mm/s.
Figure 6. Experimentally measured flexural modulus at (a) 0.1 mm/s and (b) 0.01 mm/s, and specific flexural modulus at (c) 0.1 mm/s and (d) 0.01 mm/s.
Processes 10 00956 g006
Figure 7. Experimentally measured flexural strength at (a) 0.1 mm/s and (b) 0.01 mm/s, and specific flexural strength at (c) 0.1 mm/s and (d) 0.01 mm/s.
Figure 7. Experimentally measured flexural strength at (a) 0.1 mm/s and (b) 0.01 mm/s, and specific flexural strength at (c) 0.1 mm/s and (d) 0.01 mm/s.
Processes 10 00956 g007
Figure 8. Scanning electron microscopy of E0. (a) 0.1 and (b) 0.01 mm/s, EC-10 (c) 0.1 and (d) 0.01 mm/s, EC-30 (e) 0.1 and (f) 0.01 mm/s.
Figure 8. Scanning electron microscopy of E0. (a) 0.1 and (b) 0.01 mm/s, EC-10 (c) 0.1 and (d) 0.01 mm/s, EC-30 (e) 0.1 and (f) 0.01 mm/s.
Processes 10 00956 g008
Table 1. Properties of crump rubber and epoxy matrix.
Table 1. Properties of crump rubber and epoxy matrix.
Crump RubberEpoxy
Physical propertiesUnitDescriptionUnitValue
Density1451 kg/m3ColorGrey scale0.8
Young’s modulus2600–2900 MPaEpoxy valueEq./kg5.35
Tensile strength40–70 MPaViscosity at 25 °CMPa11,850
Elongation at break25–50%Volatile content at 105 °C/h%0.4
Melting point200 °CHydrolysable chlorinewt.%0.08
Colorblack/blueMarten’s value°C150
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shahapurkar, K.; Alblalaihid, K.; Chenrayan, V.; Alghtani, A.H.; Tirth, V.; Algahtani, A.; Alarifi, I.M.; Kiran, M.C. Quasi-Static Flexural Behavior of Epoxy-Matrix-Reinforced Crump Rubber Composites. Processes 2022, 10, 956. https://0-doi-org.brum.beds.ac.uk/10.3390/pr10050956

AMA Style

Shahapurkar K, Alblalaihid K, Chenrayan V, Alghtani AH, Tirth V, Algahtani A, Alarifi IM, Kiran MC. Quasi-Static Flexural Behavior of Epoxy-Matrix-Reinforced Crump Rubber Composites. Processes. 2022; 10(5):956. https://0-doi-org.brum.beds.ac.uk/10.3390/pr10050956

Chicago/Turabian Style

Shahapurkar, Kiran, Khalid Alblalaihid, Venkatesh Chenrayan, Abdulaziz H. Alghtani, Vineet Tirth, Ali Algahtani, Ibrahim M. Alarifi, and M. C. Kiran. 2022. "Quasi-Static Flexural Behavior of Epoxy-Matrix-Reinforced Crump Rubber Composites" Processes 10, no. 5: 956. https://0-doi-org.brum.beds.ac.uk/10.3390/pr10050956

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