Abstract
Mechanical properties were evaluated for bio-natural fiber-reinforced epoxy hybrid composites made with varying amounts of jute, banana stem leaves (BSL), and tamarind shell powder (TSP). Each composite design had varying weight percentages of jute and BSL (5 to 25%) and a consistent mix of TSP (10%) and epoxy resin (60%). The tensile strength, flexural strength, interlaminar shear strength (ILSS), impact strength, hardness, and water absorption were examined. Fourier transform infrared spectroscopy (FTIR) and scanning electron microscopy (SEM) were used to investigate chemical bonding and morphology. The findings indicated a relationship between fiber and filler content and mechanical properties of composites, with 20% jute fiber content resulting in the highest performance. The tensile strength of the composite increased by 24.6%, rising from 32.4 MPa for the 5% jute and 25% banana stem leaves (5J25BSL) composite to 40.4 MPa for the 20% jute and 10% banana stem leaves (20J10BSL) composite. Similarly, the flexural strength saw a 27.9% improvement, increasing from 67.2 MPa in the 5J25BSL composite to 86.0 MPa in the 20J10BSL composite. The impact strength also experienced a notable increase of 39.1%, moving from 2.56 J for the 5J25BSL composite to 3.56 J for the 20J10BSL composite. These results highlight significant improvements in all three properties, as the proportion of jute in the composite increased and the proportion of banana stem leaves decreased. This research influences material selection for engineering applications and informs the development of specialized composite materials.
Download PDF
Full Article
Improving Mechanical Performance of Hybrid Polymer Composites: Incorporating Banana Stem Leaf and Jute Fibers with Tamarind Shell Powder
Karuppusamy Manickaraj,a,* Aruchamy Karthik ,b Sivasubramanian Palanisamy ,c,* Manivannan Jayamani ,d Syed Kashif Ali ,e,f Subramanian Lakshmi Sankar ,g and Saleh A Al-Farraj h
Mechanical properties were evaluated for bio-natural fiber-reinforced epoxy hybrid composites made with varying amounts of jute, banana stem leaves (BSL), and tamarind shell powder (TSP). Each composite design had varying weight percentages of jute and BSL (5 to 25%) and a consistent mix of TSP (10%) and epoxy resin (60%). The tensile strength, flexural strength, interlaminar shear strength (ILSS), impact strength, hardness, and water absorption were examined. Fourier transform infrared spectroscopy (FTIR) and scanning electron microscopy (SEM) were used to investigate chemical bonding and morphology. The findings indicated a relationship between fiber and filler content and mechanical properties of composites, with 20% jute fiber content resulting in the highest performance. The tensile strength of the composite increased by 24.6%, rising from 32.4 MPa for the 5% jute and 25% banana stem leaves (5J25BSL) composite to 40.4 MPa for the 20% jute and 10% banana stem leaves (20J10BSL) composite. Similarly, the flexural strength saw a 27.9% improvement, increasing from 67.2 MPa in the 5J25BSL composite to 86.0 MPa in the 20J10BSL composite. The impact strength also experienced a notable increase of 39.1%, moving from 2.56 J for the 5J25BSL composite to 3.56 J for the 20J10BSL composite. These results highlight significant improvements in all three properties, as the proportion of jute in the composite increased and the proportion of banana stem leaves decreased. This research influences material selection for engineering applications and informs the development of specialized composite materials.
DOI: 10.15376/biores.20.1.1998-2025
Keywords: Mechanical characteristics; Hybrid composites; Banana stem leaf and jute fibers; Tamarind shell powder; Water absorption; Scanning electron microscopy
Contact information: a: Department of Mechanical Engineering, CMS College of Engineering and Technology, Coimbatore – 641032, Tamil Nadu, India; b: Department of Mechatronics Engineering, Akshaya College of Engineering and Technology, Coimbatore – 642109, Tamil Nadu, India; c: Department of Mechanical Engineering, PTR College of Engineering and Technology, Austinpatti, Madurai – Tirumangalam road, Madurai – 625008, Tamil Nadu. India; d: Department of Mechanical Engineering, Kalasalingam Academy of Research and Education, Krishnankoil-626126, Tamil Nadu, India; e: Department of Physical Sciences, Chemistry Division, College of Science, Jazan University, P.O. Box. 114, Jazan 45142, Kingdom of Saudi Arabia; f: Nanotechnology Research Unit, College of Science, Jazan University, P.O. Box. 114, Jazan 45142, Kingdom of Saudi Arabia; g: Department of Mechanical Engineering, Sathyabama Institute of Science and Technology, Chennai, Tamilnadu, India; h: Department of Zoology, College of Science, King Saud University, Riyadh, Saudi Arabia;
* Corresponding Authors: sivaresearch948@gmail.com; raj.manicka@gmail.com
INTRODUCTION
Hybrid polymer composites represent a groundbreaking advancement in material science, offering a synergistic combination of different reinforcing components within a polymer matrix (Prashanth et al. 2017; Santulli et al. 2023; Gurusamy et al. 2024). By incorporating various reinforcements, such as fibers, particulates, or nanoparticles, hybrid polymer composites have the potential to achieve enhanced mechanical strength, improved toughness, superior thermal stability, and tailored electrical or magnetic properties. One of the key advantages of hybrid polymer composites lies in their versatility and adaptability to diverse application requirements. Different types of reinforcements can be selectively chosen and combined to achieve specific performance targets, making these composites highly customizable for various industries and engineering applications (Parbin et al. 2019). For instance, natural fibers like jute, banana stem leaf fibers, or sisal can be used alongside synthetic fibers such as glass or carbon to create composites with a unique balance of strength, stiffness, and sustainability. Fabrication techniques for hybrid polymer composites are equally diverse, ranging from conventional methods such as compression molding and resin transfer molding to advanced techniques such as additive manufacturing and in-situ polymerization (Sathishkumar et al. 2014; Kumar et al. 2023; Aruchamy et al. 2024). Each fabrication method offers its own set of advantages and challenges, allowing users to tailor the manufacturing process to suit the desired composite properties and production scale.
Mechanical properties are a key focus in the development of hybrid polymer composites. By combining different types of reinforcements, researchers aim to improve tensile strength, flexural strength, impact resistance, and fatigue performance compared to pure polymer matrices (Saba et al. 2016; Karthik et al. 2023c; Manickaraj et al. 2024a). Thermal stability is another important aspect of hybrid polymer composites, especially in high-temperature applications. The addition of certain reinforcements can improve the heat resistance of the composite, delaying thermal degradation and enhancing overall thermal conductivity (Karthik et al. 2023a). This is particularly relevant in industries such as aerospace, automotive, and electronics, where materials are exposed to elevated temperatures during operation.
These techniques provide valuable insights into the dispersion, distribution, and interfacial bonding between the polymer matrix and reinforcing components (Karuppiah et al. 2020). Understanding the microstructural features of the composite is essential for optimizing its mechanical, thermal, and electrical properties (Karthik et al. 2023b). This opens up opportunities for innovative applications in fields such as sensors, actuators, electromagnetic shielding, and sound absorption. Overall, hybrid polymer composites represent a frontier in material science, offering a diverse range of properties and applications. Continued research and development in this field hold the potential to drive innovation across various industries, providing sustainable and high-performance solutions to complex engineering challenges (Aruchamy et al. 2023).
Jute fiber, derived from the stems of the Corchorus plant, is known for its strength, durability, and eco-friendly characteristics. Jute has found applications in various industries, including textiles, packaging, construction, and automotive (Shah et al. 2021; Malik et al. 2024). The cultivation of jute is sustainable and environmentally friendly, as it requires minimal pesticide and fertilizer use and has a relatively low carbon footprint compared to synthetic fibers. Jute fibers have several properties that make them suitable for composite materials (Ramesh and Deepa 2024). They have high tensile strength, making them ideal for reinforcing polymer matrices to enhance mechanical properties such as strength and stiffness. Additionally, jute fibers exhibit good moisture absorption and biodegradability (Karimah et al. 2021). Jute fiber composites are composite materials in which jute fibers are used as reinforcement within a polymer matrix. These composites offer a compelling alternative to traditional materials due to their favorable balance of strength, weight, and cost. Common polymer matrices used in jute fiber composites include thermosetting resins like epoxy and polyester, as well as thermoplastic polymers such as polypropylene and polyethylene (Wang et al. 2019). Table 1 presents the comprehensive comparison of the properties of jute and banana stem leaf fibers with other natural fibers.
The production process of jute fiber composites usually consists of several steps. First, jute fiber is processed to remove impurities and improve its quality (Chandekar et al. 2020). Jute fibers are then combined with the polymer matrix by techniques such as compression molding, injection molding or filament winding, depending on the desired shape and properties of the final composite (Raghavendra et al. 2014). The properties of jute fiber composites can be tuned by adjusting parameters such as fiber orientation, fiber fraction, and matrix material. By optimizing these parameters, researchers and engineers can achieve composites with specific mechanical, thermal, and acoustic properties suited to their intended applications. In terms of applications, jute fiber composites have been utilized in a wide range of industries (Rangasamy et al. 2021; Manickaraj et al. 2024d). In automotive manufacturing, they are used for interior components, door panels, and seat backs to reduce vehicle weight and improve fuel efficiency. In the construction industry, jute fiber composites find applications in roofing materials, wall panels, and reinforcement for concrete structures. Additionally, jute fiber composites are used in consumer goods such as furniture, luggage, and sporting equipment due to their lightweight and durable nature (Devarajan et al. 2022). The chemical structure of the Jute and BSL fiber is shown in Fig. 1.
Fig. 1. Chemical structure (A) Jute fiber, (B) BSL
The growing interest in sustainability and eco-friendly materials has led to increased research and development efforts in jute fiber composites. As advancements continue to be made in composite processing techniques, matrix materials, and fiber treatments, jute fiber composites are expected to play an increasingly significant role in addressing the challenges of sustainability and resource conservation in the manufacturing industry (Karunakaran et al. 2022; Gokul et al. 2024).
Tamarind powder composites are innovative materials that combine tamarind powder, a by-product of the tamarind fruit, with various matrix materials to create a range of functional and sustainable materials. Tamarind powder is derived from the seeds and pulp of the tamarind fruit, which is cultivated in tropical regions for its culinary and medicinal properties (Kumar et al. 2018). The incorporation of tamarind powder into composite materials offers several advantages. Tamarind powder is rich in phenolic compounds, antioxidants, and other bioactive substances, which can impart beneficial properties to the composite, such as improved mechanical strength, thermal stability, and resistance to environmental degradation (Goudar et al. 2020). Tamarind powder is biodegradable and renewable, making it an environmentally friendly alternative to synthetic fillers and reinforcements. The fabrication of tamarind powder composites involves mixing the tamarind powder with a suitable matrix material, such as a polymer resin or cementitious binder (Srinivasan et al. 2020). The mixture is processed by compression molding, extrusion, or casting to form the desired shape and structure of the composite. The properties of the composite can be tailored by adjusting parameters such as the ratio of tamarind powder to matrix material, the particle size and distribution of the tamarind powder, and the processing conditions (Babu et al. 2020).
Tamarind powder composites have been applied in various fields, including construction, automotive, packaging, and biomedical engineering. In construction, tamarind powder composites have been used as a sustainable alternative to conventional building materials, such as cement and concrete (Naik et al. 2019; Girimurugan et al. 2022; Mylsamy et al. 2024; Manickaraj et al. 2024c). The addition of tamarind powder can improve the mechanical properties and durability of the composite, while also reducing its environmental impact and carbon footprint.
Fibrillation is the process of separating fiber bundles into finer individual fibers, increasing their surface area and bonding capacity. For BSL fibers, this is achieved through mechanical, chemical, or enzymatic treatments that break down the plant material into smaller fibrils. The outer layer of the banana stem leaf is scraped or chemically treated to isolate cellulose fibers. Fibrillation enhances the fiber’s ability to bond with the matrix in composites, improving mechanical properties such as tensile strength and flexibility. Here is a comparison of the differences between jute and BSL fibers in Table 1.
This table summarizes the key differences between jute and BSL fibers, highlighting their structural, mechanical, and chemical characteristics, which are essential for understanding their behavior and suitability in composite materials.
In the automotive industry, tamarind powder composites have been investigated for use in interior components, body panels, and structural reinforcements. The lightweight and sustainable nature of tamarind powder composites make them attractive for applications where weight reduction and environmental sustainability are priorities (Jeyaprakash et al. 2021; Sumesh et al. 2023). In packaging applications, tamarind powder composites have been explored as a biodegradable and compostable alternative to conventional plastic packaging materials. The incorporation of tamarind powder can improve the strength, barrier properties, and biodegradability of the composite, making it suitable for a range of packaging applications, including food packaging, agricultural packaging, and disposable containers (Kiruthika et al. 2012; Goutham et al. 2023). In biomedical engineering, tamarind powder composites have been investigated for use in tissue engineering, drug delivery, and medical device applications. The biocompatibility and bioactivity of tamarind powder make it an attractive candidate for use in composite materials for medical applications (Karthikeyan et al. 2022). Overall, tamarind powder composites represent a promising class of materials with diverse applications and potential benefits in terms of sustainability, functionality, and performance. Continued research and development efforts are expected to further expand the range of applications and enhance the properties of tamarind powder composites, making them an increasingly attractive option for a wide range of industries (Kiruthika and Veluraja 2017; Selvi et al. 2019).
Table 1. Jute Fiber vs. BSL (Pujari et al. 2014; Neto et al. 2022; Elfaleh et al. 2023)
Banana and jute fibers may have different strengths, stiffness, or elongation properties, and their combination could result in a composite with a balanced set of mechanical properties, potentially improving both tensile strength and impact resistance. The interaction between the fibers could result in synergistic effects that are not achievable by using one fiber alone, such as better stress distribution and enhanced fiber-matrix bonding.
This study fills a research gap by exploring the underutilized combination of jute, banana stem leaves, and tamarind shell powder in epoxy composites, focusing on how varying fiber proportions influence mechanical properties. Unlike prior studies that lack detailed insights into such hybrid configurations, this work identifies the optimal jute-to-banana stem leaf ratio for enhanced performance and employs FTIR and SEM analyses to link mechanical improvements with chemical and morphological characteristics, advancing the development of high-performance natural fiber composites for engineering applications.
EXPERIMENTAL
Materials
The fibers sourced from jute and banana stem leaf plants underwent meticulous treatment processes in the Anamalai area of Coimbatore District, Tamil Nadu, India, utilizing both water retting and mechanical techniques. Water retting, a traditional method, immersed the plant stems to facilitate microbial breakdown of non-fibrous components, ensuring cleaner and more flexible fibers, especially effective for jute fibers. Meanwhile, mechanical techniques, likely including cutting and scraping, were employed to extract resilient fibers from the tougher banana stem leaves, preserving their structural integrity. Figure 2 illustrates the mechanical extraction and subsequent retting process for banana stem leaf fibers, while Fig. 3 showcases the collection of jute fibers from jute plants.
Fig. 2. (A) Banana stem leaf (BSL), (B) BSL extraction, (C) BSL fiber
Fig. 3. (A) Jute fiber plant, (B) Jute fiber
Fiber Preparation and Composite Compatibility
Following extraction, fibers likely underwent cleaning, drying, and sizing to optimize compatibility with the epoxy resin matrix. The epoxy resin and hardener used were sourced from Covai Seenu Company, Coimbatore, with the resin grade specified as [Epoxy Resin LY556 and Hardener HY951]. Tamarind seeds were processed into a fine powder, as depicted in Fig. 4, harnessing their cellulose content and antioxidant properties for composite enhancement. The seeds were first cleaned to remove impurities and then dried thoroughly to reduce moisture content. Subsequently, they were ground using a high-speed mechanical grinder to achieve a fine powder. The resulting powder was screened using a mesh sieve to ensure uniform particle size, making it suitable for consistent integration into the composite matrix. The selection of treatment methods was based on fiber characteristics and composite requirements, ensuring the fibers were uniform, clean, and amenable to composite fabrication. These steps culminated in the production of high-quality composites with tailored mechanical properties (Nguyen and Nguyen 2021; Palanisamy et al. 2023).
Fig. 4. (A)Tamarind seed, (B) tamarind powder
Fig. 5. Composite specimens for all tests
Composite Fabrication
To make the composite, jute and BSL fibers were cut into 300 mm lengths. The composite layers were made up of four that alternated between jute and banana stem leaf fibers. The method began with the application of a mixture of TSP and epoxy resin on an aluminium foil cover. The first layer was made of jute fiber, then a mixture of epoxy resin, TSP, and hardener, and finally banana stem leaf fiber imbedded in resin. This layering sequence was repeated, allowing each layer to dry before applying the next. Subsequently, the composites were subjected to conditions of 130 °C and 30 bar pressure for 60 min before being demolded (Kar et al. 2023; Manickaraj et al. 2024b). The final dimensions of the laminate measured 300 x 300 x 5 mm. Figure 5 shows specimens cut for as per ASTM Standards to various test.
Methodology
Figure 6 shows a general scheme for the experimental work. As shown, the jute and banana materials were placed in the category of fibers, whereas the tamarind seen powder was regarded as a filler. After mixing the ingredients, they were pressed in the molding process, forming the composites, which were subject to testing.
Fig. 6. Simple diagram of methodology process
Composite Designation
Table 2 outlines the composition of composite samples, detailing the fiber content (Jute and BSL), filler content (TSP), and epoxy resin content for each sample. Additionally, it includes a designation for each composite based on its composition. The final column presents the blend designations.
Table 2. Mix Compositions and Designations
Mechanical Testing
A test bench machine (Model: MTS 810, Zwick/Roell Z100) was used to conduct mechanical tests. Tensile performance was evaluated in accordance with ASTM D3039 (2014) (Pinnell et al. 2005) standards using samples measuring 250 mm x 25 mm x 5mm at a speed of 2 mm/min. Flexural tests were also performed on the same machine following ASTM D7264 (2015) (Kumar et al. 2019) requirements. The size of the specimen was 125 mm x 13 mm x 5 mm. Impact energy values were determined using the test bench machine (Model: Instron CEAST 9050) in accordance with ASTM D256 (2023) (Adhikari and Gowda 2017) standards, with samples sized 60 x 12 x 5 mm. Five samples were used for each test, and the outcomes were recorded (Balaji et al. 2020). Hardness, an important mechanical attribute that indicates a material’s resistance to scratching or indentation, was measured using a Shore D durometer (Model: Mitutoyo HH-338). The mean hardness values of the composites were determined by making two indentations in each sample. Specimens were prepared in compliance with ASTM D2240 (2021) (Bahtiyar et al. 2023).
Interlaminar Shear Strength
Fiber-reinforced polymer composites are prone to a key type of failure known as shear failure. The Interlaminar Shear Strength (ILSS) method assesses a sample’s resistance to shear loading by measuring layer strength and providing an in-plane shear modulus. ILSS samples are heated for five h, from 80 to 120 °C, and then cooled gradually before being tested and uses an ASTM D2344 2022 (Kotik and Ipina 2021). An equation was utilised to calculate the ILSS, and testing was performed at a cross-head velocity of 5 mm/min (Manickaraj et al. 2023). Standard specimen size for this test was 25 mm x 25 mm with a thickness of 5 mm. ILSS was determined through the application of Eq. 1,
(1)
where d is the sample diameter, F is a ratio of the observed force to the force needed for debonding, and l represents the length of the fiber implanted in the matrix.
Water Absorption Test
To find the water absorption percentage of the hybrid composites containing jute/banana stem leaf (BSL) fibers, the samples were left at room temperature for 5 days while submerged in water. The ASTM D570 (2022) criteria () were followed in the preparation of the test specimens (Hassan et al. 2019; Manickaraj et al. 2022; Ramasubbu et al. 2024). The standard size for the water absorption test is typically 50 mm x 50 mm (2 in x 2 in) with a thickness of 5 mm. Once the specified soaking time had passed, the samples were removed from the water and gently cleaned with a soft cloth before being allowed to dry. After that, the samples were weighed to see how well the composites absorbed moisture. The equation (2) used to calculate water absorption in composite materials according to ASTM D570 is,
W = (W2−W1) / W1×100 (2)
where W is the water absorption percentage (%), W1 is the initial weight of the specimen before immersion (g), and W2 is the weight of the specimen after immersion for a specified period (grams).
FTIR Characterization
In this experiment, FTIR analysis was used to analyse the chemical makeup and components of both jute and banana stem leaf (BSL) fiber (Almeshaal et al. 2022; Johny et al. 2023). To create an infrared spectrum, the disc sample preparation method was employed. The FTIR analysis was performed using a Fourier Transform Infrared Spectrometer (Model: Thermo Scientific Nicolet]. The spectra were recorded in the range of 4000 cm⁻¹ to 400 cm⁻¹ with a resolution of 4 cm⁻¹. The samples were analyzed using the standard transmission mode.
SEM
SEM was used to evaluate the morphological representations of the generated composites (Manickaraj et al. 2023). The Zeiss EVO 18 scanning electron microscope was used to examine the fracture surfaces of the composites. The specimens were examined under an acceleration voltage of 20 kV and at a working distance of 10 mm. High-resolution images were captured to observe the surface features and fiber-matrix interactions of the samples
RESULTS AND DISCUSSION
Tensile Test
The tensile strength results for composite designations (5J25BSL, 10J20BSL, 15J15BSL, 20J10BSL, 25J5BSL) demonstrated a clear and consistent increase with higher composite designation, indicating more content of particles. Beginning with a base value of 17.5 MPa for neat epoxy, the tensile strengths progress systematically from 32.4 MPa for 5J25BSL to 40.4 MPa for 20J10BSL. This progression indicates a positive correlation between composite designation and tensile strength, suggesting that as the designation advances, the material’s ability to withstand tensile forces improves significantly (Salman 2020; Nurazzi et al. 2021). The observed trend reflects enhanced structural integrity and resistance to deformation, which are critical in applications requiring robust mechanical performance. Composites corresponding to 20J10BSL often involve tailored material compositions or refined manufacturing processes aimed at reinforcing mechanical properties. Notably, designations 20J10BSL and 25J5BSL exhibited superior tensile strength compared to lower designations, highlighting their potential for high-performance applications (Ban et al. 2021). The systematic increase in tensile strength with 20J10BSL can be attributed to optimized material compositions, refined manufacturing processes, improved interfacial bonding, and reduced structural weaknesses. These factors collectively contributed to the enhanced mechanical performance of the composites, making them suitable for demanding applications.
Figure 7 illustrates the variation in tensile strength across different composite designations, reinforcing the significance of these results in the field of materials science and engineering.
Fig. 7. Tensile strength
Fig. 8. Tensile strain plot
Strain in the data range from 0.01% to 0.48% indicated the response of the materials respond to increasing forces of deformation. As strain increases, tensile stress generally rises, reaching a peak for each specimen before potentially declining. This pattern indicates how the materials handle increasing deformation, with variations in peak strength and behavior under higher strain levels (Suwas et al. 2020; Zhang et al. 2021; Chithra et al. 2024). Figure 8 shows the various in the tensile strength and strain curve of the various composite specimens.
Flexural Strength
The flexural strength results for composite designations (5J25BSL, 10J20BSL, 15J15BSL, 20J10BSL, 25J5BSL) demonstrated a consistent increase with higher composite designation, starting from a baseline value of 34.0 MPa for neat epoxy. The data showed a progressive increase from 67.2 MPa for 5J25BSL to 86.0 MPa for 20J10BSL, highlighting a positive correlation between designation and flexural strength. This trend underscores the material’s enhanced ability to withstand bending forces, indicating superior structural robustness and resistance to deformation (Anike et al. 2022; De Souza et al. 2020). Specifically, the results highlight that 20J10BSL exhibited the highest flexural strength among the tested designations, positioning it favorably for applications demanding exceptional resistance to bending stresses. These findings provide valuable insights into how composite designation influences flexural strength, empowering material engineers and designers with critical information for selecting materials tailored to specific application requirements. Figure 9 illustrates the variation in flexural strength across different composite designations, reinforcing the significance of these results in advancing materials science and engineering.
Fig. 9. Flexural strength
The data reveals how flexural strength (in MPa) of five specimens changed with increasing strain (0.01% to 0.45%). As strain rose, flexural stress generally increased for all specimens, reflecting better resistance to bending stress. Each specimen showed a unique pattern, with some reaching higher peak values of flexural strength before eventually declining, while others exhibiting more gradual increases (Zidan et al. 2020). This variation highlights differences in material properties and their ability to withstand bending under increasing deformation. Figure 10 shows the flexural strain plot with respect to flexural strength.
Fig. 10. Flexural strain variations plot
Interlaminar Shear Strength (ILSS)
The interlaminar shear strength (ILSS) results for composite designations (5J25BSL, 10J20BSL, 15J15BSL, 20J10BSL, 25J5BSL) demonstrated a consistent increase in ILSS with higher designation. Starting at 3.52 MPa for 5J25BSL and reaching 4.52 MPa for 20J10BSL, the data suggests a positive correlation between composite designation and interlaminar shear strength (Ashok and Kani 2022). In comparison, the hybrid composites studied here demonstrated ILSS values ranging from 3.52 to 4.52 MPa, with the highest ILSS observed for the 20J10BSL designation. These results indicate competitive interlaminar shear strength compared to previous studies, showcasing the effectiveness of the fabrication techniques and material compositions employed in this research (Zhang et al. 2002; Graupner and Müssig 2022). This upward trend indicates an improvement in the material’s ability to resist internal shear forces between its layers, reflecting enhanced interlaminar bonding and structural integrity. The observed progression is in line with expectations, implying that 20J10BSL may involve modifications in resin matrix properties or fabrication techniques aimed at strengthening the interlaminar regions (Kim et al. 1992; Malik et al. 2024). The superior ILSS of the 20J10BSL composite suggests promising potential for applications requiring robust resistance to shear forces. Overall, this comparison underscores the significance of the ILSS results obtained in this study, validating the effectiveness of the hybrid composite formulations in enhancing interlaminar bonding and structural integrity. The results highlight 20J10BSL as having the highest ILSS, making it a promising candidate for applications where resistance to shear forces is crucial (Rao et al. 2021; Kumar et al. 2023). These findings provide valuable insights for material engineers and manufacturers, aiding in the selection of composite materials based on interlaminar shear strength requirements. The results of interlaminar shear strength (ILSS) are displayed in Fig. 11.
Fig. 11. Interlaminar shear strength (ILSS)
Impact Strength
The impact strength results demonstrated a consistent enhancement with increasing composite designation, starting from a baseline value of 1.05 J for neat epoxy. There was a clear progression from 2.56 J for 5J25BSL to 3.56 J for 20J10BSL, indicating improved toughness and resistance to fracture (Ahmed et al. 2021; Sharma et al. 2021). This trend suggests that 20J10BSL led to better absorption and endurance of impact forces, which is crucial in applications where impact resistance is a critical requirement. These improvements are expected with 20J10BSL due to adjustments in material composition or processing methods aimed at strengthening impact properties. For example, the 20J10BSL composite, exhibiting the highest impact strength, likely benefited from enhanced fiber-matrix interactions and optimized fiber distribution within the matrix. A comprehensive discussion on impact strength should delve into the mechanisms of energy absorption. This involves explaining how factors such as fiber orientation, distribution, and interfacial bonding influence the dissipation of impact energy (El-Abbasy 2023; Taghizadeh et al. 2024; Vescovini et al. 2024). Well-aligned fibers and uniform distribution effectively distribute stress, thereby hindering crack propagation and enhancing overall toughness. Moreover, strong bonding between fibers and the matrix facilitates efficient stress transfer, enabling the composite to withstand impact forces effectively. Additionally, factors such as fiber length and aspect ratio contribute, with longer fibers bridging cracks and reinforcing the matrix structure. By elucidating these mechanisms, the discussion provides insights into the factors driving the observed increase in impact strength with higher composite designation. This information can guide engineers in optimizing composite formulations for applications requiring superior resistance to impact loading. Figure 12 illustrates the impact energy results across different composite designations, reinforcing the significance of these findings in materials science and engineering.
Fig. 12. Impact strength
Hardness
The hardness results for composite designations (5J25BSL, 10J20BSL, 15J15BSL, 20J10BSL, 25J5BSL) revealed a consistent increase in hardness with higher designation, with a baseline value of 6.93 SD for neat epoxy. Starting at 73.2 for 5J25BSL and reaching 83.75 for 20J10BSL, the data showed a clear positive correlation between composite designation and hardness. This progression signifies that as the designation advances, the materials become progressively harder, indicating improved resistance to indentation and wear (Domagała et al. 2021; Rajak et al. 2021; Mohd Bakhori et al. 2022). The observed increase in hardness suggests enhanced durability, making these composites well-suited for applications requiring robust performance over extended periods. This enhancement aligns with expectations, as higher composite designations typically involve adjustments in filler content or reinforcement strategies aimed at bolstering material hardness. Notably, the results underscore 20J10BSL as having the highest hardness among the tested composites, highlighting its potential for applications demanding superior resistance to abrasion and wear. For example, in automotive manufacturing, such composites could be utilized for components such as gears, bearings, and brake parts that endure substantial frictional rubbing. Similarly, in construction, they could be applied to flooring, decking, and structural elements to withstand heavy foot traffic and environmental conditions. In aerospace contexts, where materials face extreme operational demands, the enhanced hardness can ensure prolonged reliability and performance of critical components. Figure 13 illustrates the variation in hardness across different composite designations, emphasizing the significance of these findings in optimizing material selection for various industrial applications.
Fig. 13. Hardness
Water Absorption
The water absorption outcomes across composite designations (5J25BSL, 10J20BSL, 15J15BSL, 20J10BSL, 25J5BSL) demonstrate discrepancies in the absorbed water percentage. By analyzing the initial and final weights and calculating the absorbed water percentage, valuable insights into the materials’ moisture vulnerability are obtained (Al-Hajaj et al. 2018). The data reveals a consistent rise in water absorption with increasing composite designation, from 48.0% for 5J25BSL to 58.9% for 25J5BSL. This suggests varying porosity and hydrophilic characteristics influenced by composition or fabrication methods. The higher water absorption in 20J10BSL indicates a higher proportion of hydrophilic components; this is in accord with jute fibers’ known tendency of moisture absorption due to their cellulose structure’s hydroxyl groups (Sathiyamoorthy and Senthilkumar 2020). Conversely, lower designations may feature a higher polymer matrix ratio, with typically lower water absorption. Factors affecting absorption include fiber-matrix interface quality and treatment methods, where better bonding reduces moisture ingress pathways. Fiber treatments, such as surface modifications, can alter hydrophilicity, impacting overall absorption. Long-term durability may be compromised by water absorption, leading to dimensional instability, mechanical property reduction, and interface degradation. Understanding these influences is vital for designing moisture-resistant composites for sustained performance in practical applications. (Abd El-baky and Attia 2019; Xu et al. 2024). The findings of water absorption are displayed in Fig. 14.
Fig. 14. Water absorption
Fourier Transform Infrared Spectroscopy
The results of Fourier transform infrared spectroscopy are shown in Fig. 15. The prominent peak residing at 1650 cm-1 represents the amide-I band—a resonance stemming from the stretching vibration of the C=O bond within the amide group (Schweitzer‐Stenner 2021). In the context of composite materials, such as bio-natural fiber-reinforced epoxy hybrids, this peak often arises from the presence of proteins or nitrogen-containing compounds inherent in the natural fibers. Between 1600 and 1000 cm-1, multiple peaks correspond to specific functional groups and molecular vibrations intrinsic to the analyzed materials (Radha and Wang 2023). The signal at 2794 cm-1 signifies C-H asymmetric bending vibrations, while the discernible peak at 832 cm-1 (Indran et al. 2018) is attributed to out-of-plane stretching motions. The peak at 2251 cm-1 represents wax or oil constituents (Manimaran et al. 2021). The stacked FTIR spectra allow for a clear comparison of peak shifts, which indicate the interactions between the epoxy matrix and the reinforcing fibers such as jute fibers and BSL. Notable changes in the peaks may suggest the formation of hydrogen bonds, ester linkages, or other chemical interactions that contribute to the strengthening of the composite. By analyzing these spectral changes, the chemical bonding mechanisms between the fibers and the matrix can be better understood, providing insight into the structural integrity and performance of the hybrid composites.
Fig. 15. FTIR analysis. (A) 5J25BSL, (B) 10J20BSL, (C) 15J15BSL, (D) 20J10BSL, (E) 25J5BSL
Scanning Electron Microscopy
Figure 16 shows the SEM image of 20J10BSL. Figure 16A shows the presence of numerous fiber pullout and voids, indicating poor bonding between the fibers and the epoxy matrix. This weak interfacial adhesion likely leads to reduced load transfer between the fibers and the matrix, resulting in lower tensile strength and an increased potential for mechanical failure (Palanisamy et al. 2022; Kurien et al. 2023). The fiber pullout observed suggests that the fibers had not been effectively anchored within the matrix, weakening the composite’s overall integrity. In contrast, Fig. 16B illustrates fiber dispersion and the presence of microcracks. The good adhesion between the fibers and the matrix is evident in the reduced occurrence of fiber pullout and the resistance to crack propagation. This strong interfacial bonding enhances the composite’s tensile strength by facilitating efficient load transfer between the fibers and the epoxy matrix (Amir et al. 2017; Ashraf et al. 2023). However, the microcracks and agglomerations observed in the image suggest potential issues with the matrix curing process, which could affect the composite’s mechanical performance. Despite these signs, the adhesion in these composites appeared better than in unfilled sisal epoxy composites. The observed agglomerations suggest potential issues with matrix curing, which could impact mechanical performance.
Fig. 16. (A) Fiber pullout, voids, good bonding; (B) Fiber dispersion
Overall, SEM analysis aids in correlating micro structural features, such as fiber-matrix bonding and crack propagation, with the mechanical properties observed in tensile testing, providing valuable insights into fracture mechanisms and overall composite performance.
CONCLUSIONS
- Tensile strength represents the maximum amount of stress that a material can withstand before breaking under tension. In this dataset, the composite designation with 20% jute and 10% banana stalk leaf (BSL) fibers offered the highest tensile strength (40.4 MPa). This implies that the composition containing 20% BSL fibers and 10% jute fibers had the best resilience to stretching pressures of the combinations tested. As the percentage of BSL fibers declined or jute fibers increases, the tensile strength decreases slightly, as demonstrated by the lower values in the other categories.
- Flexural strength assesses a material’s capacity to withstand distortion during bending. Similar to tensile strength, the 20% jute and 10% BSL composite showed the highest flexural strength at 86.0 MPa, demonstrating more resilience to bending forces than other compositions. Again, the trend indicates that a mix of 20% BSL fibers and 10% jute fibers improved flexural characteristics.
- Impact strength refers to a material’s ability to absorb energy during fracture under abrupt loading. Once again, the 20% jute and 10% BSL composite had the maximum impact strength at 3.56 J, indicating that it was better able to sustain rapid loads or impacts than the other classifications.
- Interlamellar shear strength (ILSS) assesses the strength of the bond between layers of a composite material. Consistently, the 20% jute and 10% BSL composite showed the highest ILSS value at 4.52 MPa, indicating stronger interlaminar bonding than the other compositions.
- Hardness measures a material’s resistance to indentation and scratching. The 20% jute and 10% BSL composite once again had the highest hardness value (83.8 SD), suggesting higher resilience to surface deformation when compared to other designations.
- Water absorption results varied across composites and showed no correlation with fiber %. However, the 20% jute and 10% BSL composite had a moderate water absorption of 52.4%, indicating adequate moisture resistance.
- The present study used Fourier transform infrared (FTIR) and scanning electron microscopy (SEM) investigations to explore chemical bonding and morphological properties, respectively, offering further information about the composites’ composition and structure.
- Overall, the 20% jute and 10% BSL composite surpassed other compositions in terms of tensile strength, flexural strength, impact strength, ILSS, and hardness. In addition, owing to dispersion, fiber content, homogeneity, interfacial bonding, and synergistic effects, 25% jute and 5% BSL performed worse than 20% jute and 10% BSL. This implies that the combination of 20% jute and 10% BSL fibers gave the most favorable mechanical qualities among the examined compositions, making it a suitable contender for applications that require high-performance materials.
ACKNOWLEDGMENTS
This project was supported by Researchers Supporting Project Number (RSP-2025R7) at King Saud University, Riyadh, Saudi Arabia.
Data Availability Statement
Data are available on request from the authors.
Declaration of Conflicting Interests
The authors declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.
REFERENCES CITED
Abd El-baky, M. A., and Attia, M. A. (2019). “Water absorption effect on the in-plane shear properties of jute-glass-carbon-reinforced composites using Iosipescu test,” Journal of Composite Materials 53(21), 3033-3045. DOI: 10.1177/0021998318809525
Adhikari, R. K., and Gowda, B. K. (2017). “Exploration of mechanical properties of banana/jute hybrid polyester composite,” Materials Today: Proceedings 4(8), 7171-7176. DOI: 10.1016/j.matpr.2017.07.043
Ahmed, M. M., Dhakal, H. N., Zhang, Z. Y., Barouni, A., and Zahari, R. (2021). “Enhancement of impact toughness and damage behaviour of natural fibre reinforced composites and their hybrids through novel improvement techniques: A critical review,” Composite Structures 259, article 113496. DOI: 10.1016/j.compstruct.2020.113496
Al-Hajaj, Z., Zdero, R., and Bougherara, H. (2018). “Mechanical, morphological, and water absorption properties of a new hybrid composite material made from 4 harness satin woven carbon fibres and flax fibres in an epoxy matrix,” Composites Part A: Applied Science and Manufacturing 115, 46-56. DOI: 10.1016/j.compositesa.2018.09.015
Almeshaal, M., Palanisamy, S., Murugesan, T. M., Palaniappan, M., and Santulli, C. (2022). “Physico-chemical characterization of Grewia Monticola Sond (GMS) fibers for prospective application in biocomposites,” Journal of Natural Fibers 19(17), 15276-15290. DOI: 10.1080/15440478.2022.2123076
Amir, N., Abidin, K. A. Z., and Shiri, F. B. M. (2017). “Effects of fibre configuration on mechanical properties of banana fibre/PP/MAPP Natural fibre reinforced polymer composite,” Procedia Engineering 184, 573-580. DOI: 10.1016/j.proeng.2017.04.140
Anike, E. E., Saidani, M., Olubanwo, A. O., and Anya, U. C. (2022). “Flexural performance of reinforced concrete beams with recycled aggregates and steel fibres,” Structures 39, 1264-1278. DOI: 10.1016/j.istruc.2022.03.089
Aruchamy, K., Palaniappan, S. K., Lakshminarasimhan, R., Mylsamy, B., Dharmalingam, S. K., Ross, N. S., and Pavayee Subramani, S. (2023). “An experimental study on drilling behavior of silane-treated cotton/bamboo woven hybrid fiber reinforced epoxy polymer composites,” Polymers 15(14), article 3075. DOI: 10.3390/polym15143075
Aruchamy, K., Karuppusamy, M., Krishnakumar, S., Palanisamy, S., Jayamani, M., Sureshkumar, K., Ali, S. K., and Al-Farraj, S. A. (2024). “Enhancement of mechanical properties of hybrid polymer composites using palmyra palm and coconut sheath fibers: The role of tamarind shell powder,” BioResources 20(1), 698-724. DOI:
10.15376/biores.20.1.698-724
Ashok, K. G., and Kani, K. (2022). “Experimental studies on interlaminar shear strength and dynamic mechanical analysis of luffa fiber epoxy composites with nano PbO addition,” Journal of Industrial Textiles 51(3_suppl), 3829S-3854S. DOI: 10.1177/15280837211052317
Ashraf, W., Ishak, M. R., Khan, T., Sebaey, T. A., Singh, B., and Alshahrani, H. (2023). “Study of impact performance of composite sandwich structure fabricated partially with natural fiber facesheet,” Polymer Composites 45(4), 3422-3435. DOI: 10.1002/pc.27999
ASTM D256 (2023). “Standard test methods for determining the Izod pendulum impact resistance of plastics,” ASTM International, West Conshohocken, PA, USA.
ASTM D570 (2022). “Standard test method for water absorption of plastics,” ASTM International, West Conshohocken, PA, USA.
ASTM D2240 (2021). “Standard test method for rubber property—Durometer hardness,” ASTM International, West Conshohocken, PA, USA.
ASTM D3039 (2014). “Standard test method for tensile properties of polymer matrix composite materials,” ASTM International, West Conshohocken, PA, USA.
ASTM D7264 (2015). “Standard test method for flexural properties of polymer matrix composite materials,” ASTM International, West Conshohocken, PA, USA.
Babu, S. R., Karthikeyan, S., Senthilkumar, P., and Koodalingam, B. (2020). “Mechanical behavior on tamarind & dates seeds powder, prawn shells powder with Arundo donax L. Leaf reinforced epoxy composite,” Materials Today: Proceedings 33, 3031-3036. DOI: 10.1016/j.matpr.2020.03.192
Bahtiyar, G., Ekrem, M., Ünal, B. and Ak, S. (2023). “Mechanical properties and damage behaviours of polyurethane composites reinforced with BNNP and MWCNT hybrid nanoparticles,” Journal of Elastomers & Plastics 55(4), 613-625. DOI: 10.1177/00952443231165427
Balaji, A., Purushothaman, R., Udhayasankar, R., Vijayaraj, S., and Karthikeyan, B. (2020). “Study on mechanical, thermal and morphological properties of banana fiber-reinforced epoxy composites,” Journal of Bio-and Tribo-Corrosion 6, 1-10. DOI: 10.1007/s40735-020-00357-8
Ban, H., Zhou, G., Yu, H., Shi, Y., and Liu, K. (2021). “Mechanical properties and modelling of superior high-performance steel at elevated temperatures,” Journal of Constructional Steel Research 176, article 106407. DOI: 10.1016/j.jcsr.2020.106407
Chandekar, H., Chaudhari, V., and Waigaonkar, S. (2020). “A review of jute fiber reinforced polymer composites,” Materials Today: Proceedings 26 (2), 2079-2082. DOI: 10.1016/j.matpr.2020.02.449
Chithra, N.V., Karuppasamy, R., Manickaraj, K. and Ramakrishnan, T., (2024). “Effect of reinforcement addition on mechanical behavior of Al MMC – A critical review,” Journal of Environmental Nanotechnology 13(2), 65-79. DOI: 10.13074/jent.2024.06.242632
De Souza, R. H., Kaizer, M. R., Borges, C. E. P., Fernandes, A. B. F., Correr, G. M., Diógenes, A. N., Zhang, Y., and Gonzaga, C. C. (2020). “Flexural strength and crystalline stability of a monolithic translucent zirconia subjected to grinding, polishing and thermal challenges,” Ceramics International 46(16), 26168-26175. DOI: 10.1016/j.ceramint.2020.07.114
Devarajan, B., LakshmiNarasimhan, R., Venkateswaran, B., Mavinkere Rangappa, S., and Siengchin, S. (2022). “Additive manufacturing of jute fiber reinforced polymer composites: A concise review of material forms and methods,” Polymer Composites 43(10), 6735-6748. DOI: 10.1002/pc.26789
Domagała, I., Przystupa, K., Firlej, M., Pieniak, D., Gil, L., Borucka, A., Naworol, I., Biedziak, B., and Levkiv, M. (2021). “Analysis of the statistical comparability of the hardness and wear of polymeric materials for orthodontic applications,” Materials 14(11), article 2925. DOI: 10.3390/ma14112925
El-Abbasy, A. A. (2023). “Tensile, flexural, impact strength, and fracture properties of ultra-high-performance fiber-reinforced concrete – A comprehensive review,” Construction and Building Materials 408, article 133621. DOI: 10.1016/j.conbuildmat.2023.133621
Elfaleh, I., Abbassi, F., Habibi, M., Ahmad, F., Guedri, M., Nasri, M., and Garnier, C. (2023). “A comprehensive review of natural fibers and their composites: An eco-friendly alternative to conventional materials,” Results in Engineering 19, article 101271. DOI: 10.1016/j.rineng.2023.101271
Girimurugan, R., Shilaja, C., Mayakannan, S., Rajesh, S., and Aravinth, B. (2022). “Experimental investigations on flexural and compressive properties of epoxy resin matrix sugarcane fiber and tamarind seed powder reinforced bio-composites,” Materials Today: Proceedings 66 (3), 822-828. DOI: 10.1016/j.matpr.2022.04.386
Gokul, S, Ramakrishnan, T, Manickaraj, K, Devadharshan, P, Mathew, M. K. and Prabhu, T. V. (2024). “Analyzing challenges and prospects for sustainable development with green energy: A comprehensive review,” AIP Conference Proceedings 3221(1), 020043. DOI: 10.1063/5.0235884
Goudar, S., Jain, R. K., and Das, D. (2020). “Physico‐mechanical properties of tamarind pod shell‐based composite,” Polymer Composites 41(2), 505-521. DOI: 10.1002/pc.25383
Goutham, E. R. S., Hussain, S. S., Muthukumar, C., Krishnasamy, S., Kumar, T. S. M., Santulli, C., Palanisamy, S., Parameswaranpillai, J., and Jesuarockiam, N. (2023). “Drilling parameters and post-drilling residual tensile properties of natural-fiber-reinforced composites: A review,” Journal of Composites Science 7(4), article 136. DOI: 10.3390/jcs7040136
Graupner, N., and Müssig, J. (2022). “Interfacial and interlaminar shear strength of unidirectional viscose fibre-reinforced epoxy composites—An overview of the comparability of results obtained by different test methods,” Frontiers in Materials 9, article 709845. DOI: 10.3389/fmats.2022.709845
Gurusamy, M., Soundararajan, S., Karuppusamy, M., and Ramasamy, K. (2024). “Exploring the mechanical impact of fine powder integration from ironwood sawdust and COCO dust particles in epoxy composites,” Matéria (Rio de Janeiro) 29(3), article e20240216. DOI: 10.1590/1517-7076-RMAT-2024-0216
Hassan, M. M., Le Guen, M. J., Tucker, N., and Parker, K. (2019). “Thermo-mechanical, morphological and water absorption properties of thermoplastic starch/cellulose composite foams reinforced with PLA,” Cellulose 26, 4463-4478. DOI: 10.1007/s10570-019-02393-1
Indran, S., Edwin Raj, R. D., Daniel, B. S. S., and Binoj, J. S. (2018). “Comprehensive characterization of natural Cissus quadrangularis stem fiber composites as an alternate for conventional FRP composites,” Journal of Bionic Engineering 15, 914-923. DOI: 10.1007/s42235-018-0078-9
Jeyaprakash, P., Moshi, A. A. M., Rathinavel, S., and Babu, A. G. (2021). “Mechanical property analysis on powderized tamarind seed-palm natural fiber hybrid composites,” Materials Today: Proceedings 43, 1919-1923. DOI: 10.1016/j.matpr.2020.10.930
Johny, V., Kuriakose Mani, A., Palanisamy, S., Rajan, V. K., Palaniappan, M., and Santulli, C. (2023). “Extraction and physico-chemical characterization of pineapple crown leaf fibers (PCLF),” Fibers 11(1), article 5. DOI: 10.3390/fib11010005
Kar, A., Saikia, D., Palanisamy, S., Santulli, C., Fragassa, C., and Thomas, S. (2023). “Effect of alkali treatment under ambient and heated conditions on the physicochemical, structural, morphological, and thermal properties of Calamus tenuis cane fibers,” Fibers 11(11), article 92. DOI: 10.3390/fib11110092
Karimah, A., Ridho, M. R., Munawar, S. S., Ismadi, Amin, Y., Damayanti, R., Lubis, M. A. R., Wulandari, A. P., Nurindah, and Iswanto, A. H. (2021). “A comprehensive review on natural fibers: Technological and socio-economical aspects,” Polymers 13(24), article 4280. DOI: 10.3390/polym13244280
Karthik, A., Bhuvaneswaran, M., and Sampath, P. S. (2023a). “Study the mechanical characteristics of NaOH & SLS treated cotton-kenaf fabric reinforced epoxy composites laminates,” in: International Symposium on Lightweight and Sustainable Polymeric Materials, 32, 65-77. DOI: 10.1007/978-981-99-5567-1_6
Karthik, A., Jeyakumar, R., Sampath, P. S., Soundararajan, R., and Manikandan, G. K. (2023b). “Study and fabrication of fan blade using coconut leaf sheath fibre/epoxy-reinforced composite materials,” Journal of The Institution of Engineers (India): Series D, 105, 1-8. DOI: 10.1007/s40033-023-00478-7
Karthik, A., Sampath, P. S., Thirumurugan, V., and Prakash, C. (2023c). “Effect of sample cutting angle on mechanical properties of jute/cotton fabric epoxy composite laminates,” Biomass Conversion and Biorefinery 14, 1-11. DOI: 10.1007/s13399-023-04166-0
Karthikeyan, R., Shilaja, C., Sivalingam, A., and Gopinath, P. (2022). “Experimental investigations on mechanical and water absorption properties of epoxy resin-banana fiber-tamarind seed particles hybrid biocomposites,” Materials Today: Proceedings 68, 2220-2225. DOI: 10.1016/j.matpr.2022.08.436
Karunakaran, K., Pugazhenthi, R., Anbuchezhiyan, G., Saravanan, R., and Reddy, M. V. (2022). “Experimental investigations on synthesis and characterization of tamarind seed powder reinforced bio-composites,” Materials Today: Proceedings 64, 760-764. DOI: 10.1016/j.matpr.2022.05.210
Karuppiah, G., Kuttalam, K. C., and Palaniappan, M. (2020). “Multiobjective optimization of fabrication parameters of jute fiber / polyester composites with egg shell powder and nanoclay filler,” Molecules 25(23), article 5579. DOI: 10.3390/molecules25235579
Kim, J., Baillie, C., Poh, J., and Mai, Y.-W. (1992). “Fracture toughness of CFRP with modified epoxy resin matrices,” Composites Science and Technology 43(3), 283-297. DOI: 10.1016/0266-3538(92)90099-O
Kiruthika, A. V, Priyadarzini, T. R. K., and Veluraja, K. (2012). “Preparation, properties and application of tamarind seed gum reinforced banana fibre composite materials,” Fibers and Polymers 13, 51-56. DOI: 10.1007/s12221-012-0051-x
Kiruthika, A. V, and Veluraja, K. (2017). “Physical properties of plant fibers (sisal, coir) and its composite material with tamarind seed gum as low-cost housing material,” Journal of Natural Fibers 14(6), 801-813. DOI: 10.1080/15440478.2017.1279104
Kotik, H. G., and Ipina, J. E. P. (2021). “Suggested modifications of the ASTM D2344-16 short-beam shear test method to be applied to fiber metal laminates,” Journal of Testing and Evaluation 49(2), 1213-1221. DOI: 10.1520/JTE20170399
Kumar, A., Sharma, K., and Dixit, A. R. (2023). “Tensile, flexural and interlaminar shear strength of carbon fiber reinforced epoxy composites modified by graphene,” Polymer Bulletin 80(7), 7469-7490. DOI: 10.1007/s00289-022-04413-w
Kumar, G. N., Kumar, C. S. and Rao, G. S. (2019. “An experimental investigation on mechanical properties of hybrid polymer nanocomposites” Materials Today: Proceedings 19, 691-699. DOI: 10.1016/j.matpr.2019.07.755
Kumar, P., Sivasubramanian, P., Pradeepkumar, C., Yadav, N., and Santulli, C. (2023). “Advancements in integrated additive manufacturing for composite materials: Techniques, challenges, case studies, and applications,” in: Hybrid Metal Additive Manufacturing, CRC Press, Boca Raton, FL, pp. 87-102.
Kumar, T. S. M., Rajini, N., Jawaid, M., Rajulu, A. V., and Jappes, J. W. (2018). “Preparation and properties of cellulose/tamarind nut powder green composites,” J. Nat. Fibers 15(1), 11-20. DOI: 10.1080/15440478.2017.1302386
Kurien, R. A., Selvaraj, D. P., Sekar, M., Koshy, C. P., Paul, C., Palanisamy, S., Santulli, C., and Kumar, P. (2023). “A comprehensive review on the mechanical, physical, and thermal properties of abaca fibre for their introduction into structural polymer composites,” Cellulose 30, 1-22. DOI: 10.1007/s10570-023-05441-z
Malik, K., Ahmad, F., Dawood, M. S. I. S., Islam, M. S., Ali, S., Raza, A., and Shahed, C. A. (2024). “Mechanical property enhancement of graphene-kenaf-epoxy multiphase composites for automotive applications,” Composites Part A: Applied Science and Manufacturing 177, article 107916. DOI: 10.1016/j.compositesa.2023.107916
Manickaraj, K., Ramamoorthi, R., Karuppasamy, R., Sakthivel, K. R., and Vijayaprakash, B. (2024a). “A review of natural biofiber‐reinforced polymer matrix composites,” Evolutionary Manufacturing, Design and Operational Practices for Resource and Environmental Sustainability, 135-141. DOI: 10.1002/9781394198221.ch11
Manickaraj, K., Ramamoorthi, R., Ramakrishnan, T., and Karuppasamy, R. (2024b). “Enhancing solid waste sustainability with iroko wooden sawdust and african oil bean shell particle-strengthened epoxy composites,” Global Nest Journal 26(1) 1-5. DOI: 10.30955/gnj.005467
Manickaraj, K., Karuppasamy, R., Vijayaprakash, B., and Sakthivel, K. R. (2024c). “Effect of fiber length on the mechanical properties of unsaturated polyester composites enhanced by chemically modified borassus stalk leaf fiber,” in: International Conference on Recent Advancements in Materials Science and Technology 2, 81-88. DOI: 10.1007/978-3-031-69966-5_8
Manickaraj, K., Ramamoorthi, R., Sathish, S., and Johnson Santhosh, A. (2023). “A comparative study on the mechanical properties of African teff and snake grass fiber-reinforced hybrid composites: Effect of bio castor seed shell/glass/SiC fillers,” International Polymer Processing 38(5), 551-563. DOI: 10.1515/ipp-2023-4343
Manickaraj, K., Nithyanandhan, T., Sathish, K., Karuppasamy, R., and Sachuthananthan, B. (2024d). “An experimental investigation of volume fraction of natural java jute and sponge gourd fiber reinforced polymer matrix composite,”10th International Conference on Advanced Computing and Communication Systems (ICACCS), 1, 2373-2378. DOI: 10.1109/ICACCS60874.2024.10717221
Manickaraj, K., Ramamoorthi, R., Sathish, S., and Makeshkumar, M. (2022). “Effect of hybridization of novel African teff and snake grass fibers reinforced epoxy composites with bio castor seed shell filler: Experimental investigation,” Polymers & Polymer Composites, 30.
Manimaran, P., Sanjay, M. R., Senthamaraikannan, P., Saravanakumar, S. S., Siengchin, S., Pitchayyapillai, G., and Khan, A. (2021). “Physico-chemical properties of fiber extracted from the flower of Celosia argentea plant,” Journal of Natural Fibers 18(3), 464-473. DOI: 10.1080/15440478.2019.1629149
Mohd Bakhori, S. N., Hassan, M. Z., Mohd Bakhori, N., Jamaludin, K. R., Ramlie, F., Md Daud, M. Y., and Abdul Aziz, S. (2022). “Physical, mechanical and perforation resistance of natural-synthetic fiber interply laminate hybrid composites,” Polymers 14(7), article 1322. DOI: 10.3390/polym14071322
Mylsamy, B., Shanmugam, S. K. M., Aruchamy, K., Palanisamy, S., Nagarajan, R., and Ayrilmis, N. (2024). “A review on natural fiber composites: Polymer matrices, fiber surface treatments, fabrication methods, properties, and applications,” Polymer Engineering & Science 64(6), 2345-2373. DOI: 10.1002/pen.26713
Naik, S., Halemani, B., and Raju, G. U. (2019). “Investigation of the mechanical properties of tamarind seed particles reinforced epoxy composites,” in: AIP Conference Proceedings 2057 (1). DOI: 10.1063/1.5085594
Neto, J., Queiroz, H., Aguiar, R., Lima, R., Cavalcanti, D., and Banea, M. D. (2022). “A review of recent advances in hybrid natural fiber reinforced polymer composites,” Journal of Renewable Materials 10(3), article 561. DOI: 10.32604/jrm.2022.017434
Nguyen, T. A., and Nguyen, T. H. (2021). “Banana fiber-reinforced epoxy composites: mechanical properties and fire retardancy,” International Journal of Chemical Engineering 2021, 1-9. DOI: 10.1155/2021/1973644
Nurazzi, N. M., Asyraf, M. R. M., Fatimah Athiyah, S., Shazleen, S. S., Rafiqah, S. A., Harussani, M. M., Kamarudin, S. H., Razman, M. R., Rahmah, M., and Zainudin, E. S. (2021). “A review on mechanical performance of hybrid natural fiber polymer composites for structural applications,” Polymers 13(13), article 2170. DOI: 10.3390/polym13132170
Palanisamy, S., Kalimuthu, M., Santulli, C., Palaniappan, M., Nagarajan, R., and Fragassa, C. (2023). “Tailoring epoxy composites with Acacia caesia bark fibers: Evaluating the effects of fiber amount and length on material characteristics,” Fibers 11(7), article 63. DOI: 10.3390/fib11070063
Palanisamy, S., Mayandi, K., Dharmalingam, S., Rajini, N., Santulli, C., Mohammad, F., and Al-Lohedan, H. A. (2022). “Tensile properties and fracture morphology of Acacia caesia bark fibers treated with different alkali concentrations,” Journal of Natural Fibers 19(15), 11258-11269. DOI: 10.1080/15440478.2021.2022562
Parbin, S., Waghmare, N. K., Singh, S. K., and Khan, S. (2019). “Mechanical properties of natural fiber reinforced epoxy composites: A review,” Procedia Computer Science 152, 375-379. DOI: 10.1016/j.procs.2019.05.003
Pinnell, M., Fields, R. and Zabora, R. (2005). “Results of an interlaboratory study of the ASTM standard test method for tensile properties of polymer matrix composites D 3039,” Journal of Testing and Evaluation 33(1), 27-31. DOI: 10.1520/JTE12521
Prashanth, S., Subbaya, K. M., Nithin, K., and Sachhidananda, S. (2017). “Fiber reinforced composites – A review,” J. Mater. Sci. Eng. 6(03), 2-6. DOI: 10.4172/2169-0022.1000341
Pujari, S., Ramakrishna, A., and Kumar, M. S. (2014). “Comparison of jute and banana fiber composites: A review,” International Journal of Current Engineering and Technology 2(2), 121-126. DOI: 10.14741/ijcet/spl.2.2014.22
Radha, A., and Wang, S.-F. (2023). “Rationalization of the electrochemical behavior of Praseodymium molybdates in the sensitive electrochemical reduction of nitrofuran antibiotics-furaltadone,” Microchemical Journal 195, article 109509. DOI: 10.1016/j.microc.2023.109509
Raghavendra, G., Ojha, S., Acharya, S. K., and Pal, S. K. (2014). “Jute fiber reinforced epoxy composites and comparison with the glass and neat epoxy composites,” Journal of Composite Materials 48(20), 2537-2547. DOI: 10.1177/0021998313499955
Rajak, D. K., Wagh, P. H., and Linul, E. (2021). “Manufacturing technologies of carbon/glass fiber-reinforced polymer composites and their properties: A review,” Polymers 13(21), article 3721. DOI: 10.3390/polym13213721
Ramasubbu, R., Kayambu, A., Palanisamy, S., and Ayrilmis, N. (2024). “Mechanical properties of epoxy composites reinforced with Areca catechu fibers containing silicon carbide,” BioResources 19(2), 2353-2370. DOI: 10.15376/biores.19.2.2353-2370
Ramesh, M., and Deepa, C. (2024). “Processing and properties of jute (Corchorus olitorius L.) fibres and their sustainable composite materials: A review,” Journal of Materials Chemistry A 12, 1923-1997. DOI: 10.1039/D3TA05481K
Rangasamy, G., Mani, S., Kolandavelu, S. K. S., Alsoufi, M. S., Ibrahim, A. M. M., Muthusamy, S., Panchal, H., Sadasivuni, K. K., and Elsheikh, A. H. (2021). “An extensive analysis of mechanical, thermal and physical properties of jute fiber composites with different fiber orientations,” Case Studies in Thermal Engineering 28, article 101612. DOI: 10.1016/j.csite.2021.101612
Rao, Y. S., Mohan, N. S., Shetty, N., and Shivamurthy, B. (2021). “Effects of solid lubricant fillers on the flexural and shear strength response of carbon fabric-epoxy composites,” Polymer Testing 96, article 107085. DOI: 10.1016/j.polymertesting.2021.107085
Saba, N., Jawaid, M., Alothman, O. Y., Paridah, M. T., and Hassan, A. (2016). “Recent advances in epoxy resin, natural fiber-reinforced epoxy composites and their applications,” Journal of Reinforced Plastics and Composites 35(6), 447-470. DOI: 10.1177/0731684415618459
Salman, S. D. (2020). “Effects of jute fibre content on the mechanical and dynamic mechanical properties of the composites in structural applications,” Defence Technology 16(6), 1098-1105. DOI: 10.1016/j.dt.2019.11.013
Santulli, C., Palanisamy, S., and Dharmalingam, S. (2023). “Natural fibres-based bio-epoxy composites: Mechanical and thermal properties,” in: Epoxy-Based Biocomposites (1st Ed.), pp. 163-176. DOI: 10.1201/9781003271017
Sathishkumar, T. P., Naveen, J. and, and Satheeshkumar, S. (2014). “Hybrid fiber reinforced polymer composites – A review,” Journal of Reinforced Plastics and Composites 33(5), 454-471. DOI: 10.1177/0731684413516393
Sathiyamoorthy, M., and Senthilkumar, S. (2020). “Mechanical, thermal, and water absorption behaviour of jute/carbon reinforced hybrid composites,” Sādhanā 45(1), article 278. DOI: 10.1007/s12046-020-01514-y
Schweitzer‐Stenner, R. (2021). “The combined use of amide I bands in polarized Raman, IR, and vibrational dichroism spectra for the structure analysis of peptide fibrils and disordered peptides and proteins,” Journal of Raman Spectroscopy 52(12), 2479-2499. DOI: 10.1002/jrs.6137
Selvi, B. J., Natarajan, M., and Palampalle, B. P. (2019). Characterization of Areca and Tamarind Fiber Reinforced Hybrid Polymer Composites for Structural Applications. 2019-28-0168. DOI: 10.4271/2019-28-0168
Shah, S. S., Shaikh, M. N., Khan, M. Y., Alfasane, M. A., Rahman, M. M., and Aziz, M. A. (2021). “Present status and future prospects of jute in nanotechnology: A review,” The Chemical Record 21(7), 1631-1665. DOI: 10.1002/tcr.202100135
Sharma, P., Mali, H. S., and Dixit, A. (2021). “Mechanical behavior and fracture toughness characterization of high strength fiber reinforced polymer textile composites,” Iranian Polymer Journal 30(2), 193-233. DOI: 10.1007/s13726-020-00884-8
Srinivasan, T., Kumar, S. B., Suresh, G., Ravi, R., Srinath, S. R. L., Paul, A. I., and Vishweshwaran, M. (2020). “Experimental investigation and fabrication of palmyra palm natural fiber with tamarind seed powder reinforced composite,” in: IOP Conference Series: Materials Science and Engineering 988, article 12022. DOI: 10.1088/1757-899X/988/1/012022
Sumesh, K. R., Ajithram, A., Palanisamy, S., and Kavimani, V. (2023). “Mechanical properties of ramie/flax hybrid natural fiber composites under different conditions,” Biomass Conversion and Biorefinery, 1-12. DOI: 10.1007/s13399-023-04628-5
Suwas, S., Bisht, A., and Jagadeesh, G. (2020). “Microstructural changes in materials under shock and high strain rate processes: Recent updates,” in: Mechanics of Materials in Modern Manufacturing Methods and Processing Techniques 361-392. DOI: 10.1016/B978-0-12-818232-1.00012-6
Taghizadeh, S., Niknejad, A., Maccioni, L., and Concli, F. (2024). “Investigating the mechanical behavior and energy absorption characteristics of empty and foam-filled glass/epoxy composite sections under lateral indentation,” Materials 17(15), article 3847. DOI: 10.3390/ma17153847
Vescovini, A., Cruz, J. A., Ma, D., Colombo, C., Salerno, A., Bianchi, O., Amico, S. C., and Manes, A. (2024). “Experimental investigation on low-velocity impact behavior of glass, Kevlar, and hybrid composites with an elastomeric polyurethane matrix,” Composites Part C: Open Access 13, article 100426. DOI: 10.1016/j.jcomc.2023.100426
Wang, H., Memon, H., AM Hassan, E., Miah, M. S., and Ali, M. A. (2019). “Effect of jute fiber modification on mechanical properties of jute fiber composite,” Materials 12(8), artilce 1226. DOI: 10.3390/ma12081226
Xu, Q., Xiao, S., Wang, Y.-Q., Peng, C., and Gao, H. (2024). “Wear-induced variation of surface roughness in grinding 2.5 D Cf/SiC composites,” International Journal of Mechanical Sciences 264, article 108811. DOI: 10.1016/j.ijmecsci.2023.108811
Zhang, C., Hoa, S. V, and Ganesan, R. (2002). “Experimental characterization of interlaminar shear strengths of graphite/epoxy laminated composites,” Journal of Composite Materials 36(13), 1615-1652. DOI: 10.1177/0021998302036013582
Zhang, K., Xiao, J., and Zhang, Q. (2021). “Experimental study on stress-strain curves of seawater sea-sand concrete under uniaxial compression with different strain rates,” Advances in Structural Engineering 24(6), 1124-1137. DOI: 10.1177/1369433220958765
Zidan, S., Silikas, N., Haider, J., Alhotan, A., Jahantigh, J., and Yates, J. (2020). “Evaluation of equivalent flexural strength for complete removable dentures made of zirconia-impregnated PMMA nanocomposites,” Materials 13(11), article 2580. DOI: 10.3390/ma13112580
Article submitted: November 12, 2024; Peer review completed: December 14, 2024; Revised version received: December 27, 2024; Accepted: December 28, 2024; Published: January 13, 2025.
DOI: 10.15376/biores.20.1.1998-2025