Next Article in Journal
Does It Help Carbon Reduction in China? A Research Paper about the Mediating Role of Production Automation Based on the Carbon Kuznets Curve
Previous Article in Journal
Teachers’ Frequency of ICT Use in Providing Sustainable Opportunity to Learn: Mediation Analysis Using a Reading Database
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ansys-Based Evaluation of Natural Fiber and Hybrid Fiber-Reinforced Composites

1
Department of Fundamentals of Machinery Design, Silesian University of Technology, 44-100 Gliwice, Poland
2
Faculty of Aerospace Engineering, Delft University of Technology, Kluyverweg 1, 2629 HS Delft, The Netherlands
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(23), 15992; https://doi.org/10.3390/su142315992
Submission received: 30 September 2022 / Revised: 17 November 2022 / Accepted: 21 November 2022 / Published: 30 November 2022

Abstract

:
In this research, we analyzed natural composite structures that optimize the material and weight of the structure. Green composites are made of natural fibers and epoxy resin that are biodegradable, recyclable, and eco-friendly. Core material failures include wrinkling, failure in compression, and buckling. To address these issues, this work attempted to create CAD models using jute fiber, glass fiber, and epoxy resin with various ply sequences using angle orientations of 0°, 30°, and 45°, and 2–4 mm thick laminates were produced. After creating CAD models, the material strength, stiffness, deformation of samples, shear strength, strain, and other mechanical properties of the natural-fiber-reinforced composite laminates were analyzed. The samples were based on two layers of glass fiber as a core with natural fiber plies below and above this core. The natural fiber with epoxy resin, the hybrid composite with jute fiber, and the glass fiber with epoxy resin were prepared and mechanical properties of the samples were evaluated with Ansys. The results indicated that the 0° ply orientation of 3 mm thickness had a low deformation (0.237 mm) and was the best material. The tensile test was performed for natural-fiber-reinforced composite and hybrid natural reinforced composite laminates at various thicknesses and at various ply orientations using a tensile load of 2500 N. In this investigation, the best material was the one with the thickness of 3 mm with the Young modulus 35.59 GPa at 0.149 strain with 5303 Pa stress conditions. Further, the above conditions were noted with low deformation (0.237 mm) at 0° ply orientation and tensile strength was noted as 1188 GPa at 3 mm with 45° ply orientation. This hybrid composite material can be considered for unmanned aerial vehicle applications.

1. Introduction

Natural fiber composite materials are widely used in automotive, construction, textile, and aerospace industries. Natural reinforcement fibers are relatively low cost, easy to recycle, biodegradable, and eco-friendly [1]. Jute fiber and polyester resin were combined as layers with unidirectional (90°) reinforcement to form bio-composites, improving the tensile and flexural strengths. This final material was used for the interior opening handle design for a car door [2]. Sisal, roselle, and banana fibers were used in various proportions and combinations in fabricated bio-composites. These hybrid reinforced composite materials were used for the manufacture of car parts, i.e., car roofs, dashboards, and door handles [3]. The traditional composite reinforced materials have been replaced by balsa, hemp, bamboo, coir, jute fibers, etc. [4,5,6], and have become a potential market as world production reached 33.3 million tons in 2021 and is projected to reach 33.7 million tons in 2022 [7]. Nowadays, many industries are using natural-fiber-reinforced polymer composites, with the industry worth approximately USD 2.1 billion in 2010 [8] and growing to nearly USD 6.50 billion in 2021 [9]. Every year, the usage of natural fibers is more than 80,000–1,600,000 tons [10]. Some examples of natural fibers are sisal, coir, jute, leather, wool, flax, borassus, and phoenix silvestre [11,12]. Many natural fibers are available in the market; however, some of the natural fibers have limited strength. In particular, material strength is needed to meet the material-bearable criteria. Natural fibers are easily available in the Asian continent, but this is not the case in Europe. Hasan et al. (2021) revealed the natural fiber value currently used in industry and the natural fiber composite market was projected to be worth USD 74 billion in 2020. This may increase by the end of 2025 to USD 112.8 billion with an 8.8% compound annual growth rate (CAGR) [1].
Due to current global environmental issues, synthetic polymer composites are being replaced by bio-degradable composite materials. From a material-usage point of view, there is demand for the creation of a new lightweight material with optimized strength and a positive environmental impact that can be used in the current market and can replace conventional material in aerospace applications [13]. The main issues concerning the new material are material stability, the need for it to be lightweight, fuel efficiency, reduction of CO2 emissions, and protection of renewable sources. On the other hand, transport industries, especially the automobile and aerospace industries, are seeking new advanced materials with a low cost and maximum load-bearing capacity. Furthermore, the cultivation of natural fibers is very important from an economic point of view and it can impact the current environmental situation [14]. Composite-reinforced fibers are widely used in the aerospace, automobile, and marine industries which all have high-strength and lightweight requirements for structures. In aerospace applications, especially in wing construction, the sandwich type of composite structure is used extensively. These composite laminates are in the form of thin, lightweight, strong materials. Such materials can withstand ultimate loads. Composite sandwich structures are the main structures utilized in lightweight aircraft.
Composite sandwich structures consist of upper and lower layers of skins with high strength (faces), low mass, and density for the inside layers (core). There are many core materials available, such as balsa wood, polyvinyl chloride (PVC) honeycomb, polyurethane (PU) foam, corrugated core, vinyl sheet foam, polyisocyanurate sheet foam, and honeycomb core. Nowadays, advanced industries are producing honeycomb cores made of Nomex for sandwich face materials with thin aluminum alloys and composite laminates with glass, carbon, or Kevlar fibers. Many researchers have examined composites over the decades and have found many issues during the construction of sandwich structures. Natural fiber composites mainly depend on the chemical constituents and the distribution of fibers along with other intrinsic factors, i.e., fiber orientation, fiber geometry, fiber volume, and matrix arrangement. Fiber volume, thickness, and type of bonding agent are also important parameters [15]. As a result of many possible causes of failure such as material debonding, delamination, matrix cracks, face-sheet buckling, fiber breaking, fiber pullout, cell-wall crushes, and face-sheet core debonding, an analysis to investigate all possible failure modes has to be conducted. In the process of optimization of the sandwich structure, all possible failure modes have to be analyzed and parameters influencing the failure mode should be examined. In our investigation, the ply sequences in the laminates were changed alternately but in the same proportions (50:50) and the laminates were analyzed for mechanical properties.
There has been no proper research on natural fibers as synthetic fibers are used for the construction of light materials. There is an essential need to carry out research on natural fibers because they are low cost and biodegradable which are useful in unmanned aerial vehicles. Further, natural fibers are sustainable and have a good life cycle, which is necessary for UAVs. Research work was reported on a flax and glass fiber hybrid composite laminate with two alternate layers of glass fiber [16]. No research reports have been found on the use of jute fibers which are available in nature. There is no bridge in the research on these natural fibers for the construction of wings and other accessory parts of the UAVs.
Another work attempted to design the CAD models using hybrid Basalt/E-glass/S2 fiber, glass-carbon fiber, and integrating them into the Ansys FEM analysis [17]. Computer-aided design with 3D models together with the finite element method concept can avoid the experimental cost and lower time consumption. It can predict the material failure behavior, facilitates optimization of material characteristics, and can easily be accommodated in modeling from element to element [18,19,20]. This approach can reduce time consumption and costs and improve corrosion resistance, thermal stability, modulus, and strength. The novelty of the work includes the preparation of hybrid composites based on jute fiber and glass fiber with various thicknesses and at various ply orientations. No relevant works have been observed with the jute and glass fiber with various proportions and various ply sequences. In this work, primarily a 3D CAD model was used for the weight optimization, and it was integrated into Ansys for FEM analysis. Only two core layers are composed of glass fibers, and the remaining layers are composed of jute fibers. These materials were checked for deformation, stress, strain, and tensile strength. This research is mainly focused on the NFRC structures stability, material characteristics, and identification of the current failure modes as well as other issues connected with them.

2. Materials and Methods

2.1. Material Selection

Natural fibers have been used in the automotive and aerospace industries since the early 90s. The possibility and usage of NFRC in the automotive region are excellent and optimistic. Modern vehicle and aerospace manufacturers have utilized various NFRC components. Hence, this research work addresses the best natural fiber material finding in terms of mechanical properties, lightweight and better strength, and compatibility for Unmanned Arial Vehicle (UAV). In this work, some of the natural fibers such as jute, glass fibers, and bonding agent resin are taken for analysis. The materials and methods are described in two categories, the first being material selection and the latter material description. The material selection procedure is classified and depicted in Figure 1.

2.1.1. Natural Reinforced Fiber Composites—Jute

Jute fiber is a thick, soft, and good fiber that can stretch out into grainy, strong aligned threads. These fibers belong to the Tiliaceace family. Jute fabric is used in many applications, for example, in the textile, automotive, and aerospace industries. Jute plant cultivation is suitable for subtropical climates with the best places for the cultivation of natural fibers being in India, Nepal, Bangladesh, China, Thailand, and Brazil. The primary ingredients of the jute fibers are cellulose and lignin as raw materials. Jute fibers have sporadic cross-sections with several diameters varying in the structure and size of the plant. Diameter ranges from 0.01 mm to 0.04 mm, approximately [9].
Natural jute fiber is useful for composite reinforcement which contains mechanical properties such as tensile strength (393–773 MPa), elastic modulus (10–30 GPa), and density of 1.44 g/cm3 [21]. However, some mechanical properties of composites reinforced with natural jute fibers even now require evaluation. In particular, the determination of impact behavior is a relevant characteristic that may be associated with impact conditions such as car crashes [22]. The Charpy test has been used by many researchers to investigate the impact behavior of various composite material structures. It is possible to estimate the energy absorbing capability and strength of the laminates throughout the loading and subsequent failure of the specimen [23,24]. Pereira et al. (2017) stated in his research work that energy absorption, the volume fraction of the jute fiber, increased when the strength was evaluated by Charpy impact test. The jute fiber material (longitudinal) was taken and unidirectionally aligned with reinforcement of the epoxy resin matrix. After manufacturing, the natural fiber composites were examined in the laboratory. It was noted that the energy-absorption value was 214 J/m on 30% volume of jute fiber reinforcement of the epoxy resin matrix [25]. Singh et al. (2018) summarized general values on jute fiber, i.e., tensile strength, Young’s modulus, percent elongation, density, and diameter [26]. Selected material properties are listed in Table 1.

2.1.2. E-Glass Fiber

Glass fiber is a lightweight material that is also extremely strong and robust. When compared to metals, their bulk strength, stiffness, and weight properties are also very favorable. In the epoxy resin matrix, an E-glass fiber with random orientation was used as a reinforcing material. The mechanical properties of E-glass fiber are tensile strength 3445 MPa, compressive strength 1080 MPa, elastic module 73 GPa, density 2.58 g/cm3, and Poisson’s ratio 0.22, respectively [27].

2.1.3. Adhesive Epoxy Resin

Epoxy resins are admirable bonding materials that are also called thermosetting resins. Resins are basically in a liquid or solid form. Epoxy is one of the best adhesives for the composite material manufacturing process. Nowadays, many resins are available on the market, and generally, they are also divided into many groups, such as epoxy resins, vinyl ester resins, and polyester resins [28]. In this research, pure epoxy resins are used for manufacturing natural fiber laminates. The resin is mixed with hardener at different compositions to give better mechanical properties. They are used in many applications such as painting, coatings, fabric manufacturing, wood shining, and construction. The adhesive properties are listed in Table 2, other natural fiber properties are presented in Table 1, and their chemical composition properties are also shown in Table 3.

2.2. Failure Criteria in Laminates

Composite materials are considered lightweight and high strength, and they have better mechanical properties when compared to other materials. These properties vary in transverse and longitudinal directions; nevertheless, some extreme condition composites can fail in terms of stability or can damage the structure. Composite materials are especially damaged under two conditions such as material damage starting an insignificant crack (initiation) followed by the propagation of damage, and such problems occur before damage to the composite material. The composite laminate failure criteria are expressed as follows in mathematical equations [29,32,33],
Natural-fiber-reinforced composite tension/shear (Material direction_1)
( E 1 ( ε 1 ) X 1 T ) 2 + ( G 31 ( ε 31 ) S F S ) 2 = r 7 2
Natural-fiber-reinforced composite compression (direction_1)
( E 1 ( ε 1 ) X 1 c ) 2 = r 8 2 ε 1 = ε 1 ( ε 3 ) E 3 E 1
Natural-fiber-reinforced composite shear/tension (direction_2)
( E 2 ( ε 2 ) X 2 T ) 2 + ( G 23 ( ε 23 ) S F S ) 2 = r 9 2
Natural-fiber-reinforced composite compression (direction_2)
( E 2 ( ε 2 ) X 2 c ) 2 = r 10 2 ε 2 = ε 2 ( ε 3 ) E 3 E 1
Natural-fiber-reinforced composite crush/damage
( E 3 ( ε 3 ) S F C ) 2 = r 11 2
Natural-fiber-reinforced composite matrix shear (direction_12)
( G 12 ( ε 12 ) S 12 ) 2 = r 12 2
Natural-fiber-reinforced composite matrix delamination (direction_12)
S 2 ( E 3 ( ε 3 ) S 3 T ) 2 + ( G 23 ( ε 23 ) S 23 ) 2 + ( G 31 ( ε 31 ) S 31 ) 2 = r 13 2
where X1T and X1c are longitudinal tensile and compressive material strength at direction_1, X2T and X2c are transverse material tensile and compressive strength at direction_2, SFS, SFC, and S3T are natural fiber shear/tension (direction_2), damage strength, and material tensile strength, S12, S23, and S31 are natural finer composite matrix mode shear strength at 12, 23, and 31 directions. In the final equation (Equation (7)), S is the material delamination scaling factor; r7- 13 material damage thresholds [30,31,34]. G12, G23, and G31 are the natural finer composite matrix shear modulus at 12, 23, and 31 directions. E1, E2, and E3 are the natural fiber composite matrix Young’s modulus at various directions.
The composite matrix is classified as seven columns and six rows. Every column represents the material damage mode shape threshold number (r7–13), and the row represents each loading path (damage variable quantity). The ‘q’ is the material matrix considered as follows in Equation (8) [35,36].
q = [ 1 0 1 0 1 0 0 1 0 1 1 0 0 0 0 0 1 0 1 1 1 1 1 1 0 1 0 1 1 0 1 0 1 0 1 0 ]  
The new stiffness constant is considered as ɛi. The compliance matrix S is calculated based on Equation (9).
S = [ 1 E 1 ( 1 ε 1 ) υ 12 E 1 υ 31 E 3 0 0 0 υ 12 E 1 1 E 2 ( 1 ε 2 ) υ 32 E 3 0 0 0 υ 13 E 1 υ 23 E 2 1 E 2 ( 1 ε 3 ) 0 0 0 0 0 0 1 G 12 ( 1 ε 4 ) 0 0 0 0 0 0 1 G 23 ( 1 ε 5 ) 0 0 0 0 0 0 1 G 31 ( 1 ε 6 ) ]  
The above-mentioned matrix is used for calculating the strain in orthotropic composite materials. Matrix S and stiffness matrix c are obtained using the association (S−1). Based on this Equation (10), the stresses are calculated.
δ = C ε
When the meshing elements fail, local stiffness lessens the material which is due to internal cracks, which in turn leads to material deformation in the form of elements. Such causes are called element distortion and divergence while solving the problem [37,38,39].

2.3. Optimization Procedure

The preliminary examination was carried out for material density and cross-sectional area for optimization of material weight by 3D CAD models. This research mainly focused on the material strength at various geometrical configurations and was carried out in Ansys. Optimization of the natural-fiber-reinforced composites (NFRC) was taken into account. In the optimization, the 3D CAD geometrical models were integrated into Ansys for ply orientation. These geometrical CAD models are represented to understand the technical configuration of composite materials. The second procedure is hybrid natural reinforced composites (HNRC), and it is considered to increase the material strength and other mechanical properties. The natural reinforced fiber composite is designed, and the optimization procedure is shown in Figure 2.

2.4. Research Methodology

The research was conducted in terms of material optimization and lightweight structure strength analysis, and the procedure is classified into 6 cases. Each case optimizes configuration and verifies the mechanical properties and behavior. Case_1 represents the total thickness of the NFRC at 2 mm and ply orientation of 0°, 30°, and 45°. In case_2 and case_3, it is 3 and 4 mm, but in all the cases, the material strength is toward the longitudinal fiber direction. Similarly, in case_4, case_5, and case_6, instead of NFRC they were tested with the HNRC. The difference in the thickness due to the additional natural fiber layers are added for case_1, case_2, and case_3 without E-glass fiber. In contrast, in case_4, case_5, and case_6, the middle core layers are fabricated with E-glass fiber, but the thickness is increased by the additional natural fiber layers. The design classification case analysis is classified as follows in Figure 3.
  • Case:1 NFRC optimization with 2 mm thickness.
  • Case:2 NFRC optimization with 3 mm thickness.
  • Case:3 NFRC optimization with 4 mm thickness.
  • Case:4 HNRC optimization with 2 mm thickness.
  • Case:5 HNRC optimization with 3 mm thickness.
  • Case:6 HNRC optimization with 4 mm thickness.

3. Geometrical Modeling

Natural-fiber-reinforced composite laminate consists of different types of ply orientation. The ply sequence is the key point to increase the material strength. This was also considered for weight optimization. Case one is considered as 0°, 30°, and 45° ply orientation (balanced laminate) 8 [0°/0°/0°/0°/0°/0°/0°/0°], the second ply—sequence of 30° (angle-symmetric ply is laminate) 8 [0°/30°/0°/−30°/−30°/0°/30°/0°], and the third ply—sequence as 8 [0°/45°/0°/−45°/−45°/0°/45°/0°]. The total thickness of the laminates considered for analysis is 2–4 mm. Hence, epoxy resin is used for bonding between each ply. The natural fiber-woven type sheet is approximately 300 mm in length and 50 mm in width. The natural fibers are standard of 0.25 to 0.33 mm in diameter [10]. These laminates are very thin layers and lightweight structures, and the combination of the two different materials, such as natural fibers and glass fibers, is called hybrid composite. The master material surface represents natural fibers, and the slave material surface is called glass fibers, and it is presented in Figure 4. The application perspective for natural reinforced fiber composites is in aerospace applications (unmanned aerial vehicles with entire wing structures with fuselage).

Geometrical Model_1 and _2

The natural reinforced fibers are constructed along with epoxy resin at various thicknesses and different ply orientations. The created geometrical models are a combination of natural fiber and synthetic fiber (E-glass). The geometrical CAD model of 2D and 3D diagrams is presented in Figure 5.

4. Analysis

Evaluating the material strength and mechanical properties is a very challenging task. However, in this research, different types of laminates with different ply orientation configurations were designed as per the engineering data. The analysis of natural reinforced fiber composite laminates was performed using Ansys. The analysis is conducted by two methods: the first method considers the analysis of natural fiber with adhesive, and the latter one consists of hybrid natural fiber along with the epoxy resin. The load and boundary conditions are also subject to 2500 N tension. The static load was applied in the longitudinal direction. The fiber orientation is considered as far as the increase in the material integrity as well the stability of the material. The results attained by the analysis of several synthetic fibers could be replaced by natural reinforced fiber composites. This material is also suitable for automotive and aerospace applications, especially lightweight structures for unmanned aerial vehicles. Stress–strain values, deformation shear stress, and von Mises stress are verified in this research. All the results are plotted and notified by their names. The load and boundary conditions are presented in Figure 6.

4.1. Engineering Material Data

Before conducting the analysis, precautionary measurements were taken towards material selection and engineering data. Most of the data was considered from recent research articles published in highly ranked journals.

4.2. Boundary Conditions and Meshing

One of the big challenges in assessing the material strength is the fact that natural fibers have high water absorption rate, which is also their disadvantage. However, fibers can be dried at room temperature in order to avoid such issues. Many studies have conducted different tests; nevertheless, there is still a research gap. The load is applied in the longitudinal direction, the fiber orientation is also considered in the unidirectional fiber orientation. The specimen is designed 300 mm long, 50 mm wide and with thicknesses of 2–4 mm. A boundary condition is applied at one end (Dof-0), and at another end a static load of 2500 N is applied in terms of tension. The meshing of the NRFC model is discretized model into small elements. Selecting the mesh type also plays a very important role since it is basically made up of elements which contain nodes. The sweep mesh is used for this research. The meshing element size is 2 mm, the number of nodes is 3926, and there are elements 3750 of shell 181 type.

5. Results

5.1. Case_1, 2, and 3 (Natural Reinforced Fiber Composite Laminate)

The analysis was carried out in two stages as described earlier. The first stage of analysis is toward natural reinforced fiber composite laminates with different plies- orientation. The maximum material stress (von Mises stress), total deformation, shear stress and strain are plotted in Figure 7, Figure 8, Figure 9 and Figure 10. The analysis has been carried out for the natural reinforced fiber composites at different configurations as shown in Figure 6. The maximum deformation value (0.950 mm) was noted at 30° at 2 mm thickness. The minimum deformation was (0.105 mm) at 45° at 2 mm.

5.1.1. Maximum Deformation Results of 0°, 30°, and 45° with 2, 3, and 4 mm, Respectively (NFRC)

The NFRC maximum deformation value is 0.950 mm, and it is noted at 30° with a 2 mm thickness of the laminate. The minimum deformation value is 0.105 mm, and it is noted at 45° with 2 mm thickness. All configurations of laminate results are presented in Table 4, and analysis results are shown in Figure 7.

5.1.2. Maximum von Mises Stress Results of 0°, 30°, and 45° with 2, 3, and 4 mm, Respectively (NFRC)

The NFRC maximum von Mises stress value is 2384 Pa, and it is noted at 45° with a 2 mm thickness of the laminate. The minimum von Mises stress value is 1021.11 Pa, and it is noted at 0° with 4 mm thickness. All configurations of laminate results are presented in Table 5, and analysis results are shown in Figure 8.

5.1.3. Maximum von Mises Strain Results of 0°, 30°, and 45° with 2, 3, and 4 mm, Respectively (NFRC)

The NFRC maximum von Mises strain value is 0.631, and it is noted at 45° with a 2 mm thickness of the laminate. The minimum von Mises strain value is 0.276, and it is noted at 0° with 4 mm thickness. All configurations of laminate results are presented in Table 6, and analysis results are shown in Figure 9.

5.1.4. Maximum von Mises Shear Stress Results of 0°, 30°, and 45° with 2, 3, and 4 mm, Respectively (NFRC)

The NFRC maximum shear stress value is 798.20 Pa, and it is noted at 45° with a 2 mm thickness of the laminate. The minimum shear stress value is 79.67 Pa, and it is noted at 0° with 4 mm thickness. All configurations of laminate results are presented in Table 7, and analysis results are shown in Figure 10.

5.1.5. Stress–Strain Results of 0°, 30°, and 45° with 2, 3, and 4 mm, Respectively

Natural reinforced fiber composite laminate and hybrid composite, shown in Figure 11.

5.1.6. Stress–Strain Graph Comparison of 0°, 30°, and 45° with 2, 3, and 4 mm, Respectively

Natural reinforced fiber composite laminate, shown in Figure 12.

5.2. Case_4, 5, and 6 (Hybrid Natural Reinforced Fiber Composite Laminate)

The natural fiber composite laminate results were presented in the first part, nonetheless, in order to continue the second part of the analysis hybrid composites were considered. It consists of natural fiber and E-Glass with epoxy resin.

5.2.1. Maximum Deformation Results of 0°, 30°, and 45° with 2, 3, and 4 mm, Respectively (HNRC)

The HNRC maximum deformation value is 0.626 mm, and it is noted at 45° with a 2 mm thickness of the laminate. The minimum deformation value is 0.209 mm, and it is noted at 0° with 4 mm thickness. All configurations of laminate results are presented in Table 8, and analysis results are shown in Figure 13.

5.2.2. Maximum von Mises Stress Results of 0°, 30°, and 45° with 2–4 mm, Respectively (HNRC)

The HNRC maximum von Mises stress value is 6125.26 Pa, and it is noted at 0° with a 2 mm thickness of the laminate. The minimum von Mises stress value is 3257.8 Pa, and it is noted at 45° with 4 mm thickness. All configurations of laminate results are presented in Table 9, and analysis results are shown in Figure 14.

5.2.3. Maximum Strain Results of 0°, 30°, and 45° with 2–4 mm, Respectively (HNRC)

The HNRC maximum von Mises strain value is 0.377, and it is noted at 45° with a 2 mm thickness of the laminate. The minimum von Mises strain value is 0.131, and it is noted at 0° with 4 mm thickness. All configurations of laminate results are presented in Table 10, and analysis results are shown in Figure 15.

5.2.4. Maximum Shear Stress Results at 0°, 30°, and 45° with 2, 3, and 4 mm, Respectively (HNRC)

The HNRC maximum shear stress value is 1188.39 Pa, and it is noted at 45° with a 3 mm thickness of the laminate. The minimum shear stress value is 198.755 Pa, and it is noted at 0° with 4 mm thickness. All configurations of laminate results are presented in Table 11, and analysis results are shown in Figure 16.

5.2.5. Stress–Strain Graph Comparison at 0°, 30°, and 45° with 2, 3, and 4 mm, Respectively

Hybrid natural reinforced fiber composite laminate) shown in Figure 17.

6. Discussion

The tensile test was performed for NFRC laminates at various thicknesses and at various ply orientations using the tensile load of 2500 N. The deformation results of 0° ply orientation were noted as 0.897, 0.594, and 0.446 mm for 2–4 mm thickness, respectively. The deformation results of 30° ply orientation were noted as 0.95, 0.646, and 0.713 mm for 2–4 mm thickness, respectively. The deformation results at 45° ply orientation were noted as 0.105, 0.713, and 0.591 mm for 2–4 mm thickness, respectively. These higher values indicate that the deformation is high. Similarly, the same conditions were applied for HFRC laminates, the results of 0° ply orientation were noted as 0.274, 0.237, and 0.209 mm for 2–4 mm thickness, respectively. The results of 30° ply orientation were noted as 0.543, 0.414, and 0.327 mm for 2–4 mm thickness, respectively. The deformation results of 45° ply orientation were noted as 0.626, 0.488, and 0.382 mm for 2–4 mm thickness, respectively. The results indicated that the 0° ply orientation of 4 mm thickness is the best material with low deformation (0.209 mm). This natural fiber (Jute fiber) can be used as skin for wings in unmanned aerial vehicles.
The above conditions were utilized to test the stress and strain on the same material and the results indicated that Young’s modulus 35.717 GPa was noted for 0° ply orientation at 4 mm thickness. Similar results were observed in the report of Serra-Parareda et al. (2021) on Henequen fiber reinforced PP composites [40]. The jute fiber was used for the fabrication of bio-composite and found that the results reported by several researchers (Mitra (2014); Abilash and Siva Pragash, (2013) support this investigation [41,42]. Ammurullah et al. (2022) addressed the importance of the computational simulation-based study on three different ceramic materials and analyzed (ceramic-on-ceramic couplings) them using Tresca stresses [43]. Saravannan et al. (2021) was investigated with different natural fibers along with aluminum material as a combination of hybrid material. He concluded that natural fiber (flax) has superior properties compared to other materials [9]. In this investigation, the best material is the one with the thickness of 3 mm with the Young modulus 35.59 GPa at 0.149 strain with 5303 Pa stress conditions. These hybrid composites can be used in wing construction and sub parts of spars, ribs, and also for skin in unmanned aerial vehicles. Further, the above conditions were noted with low deformation (0.0237) of 0° ply orientation.

6.1. Comparison of Stress–Strain Plot at Natural Reinforced Fiber Composite and Hybrid Composite Shown in Figure 18

The stress–strain plots of NRFC and HNRC composites neatly showed that the HRFC material had very high yield strength. Among them, 3 mm thickness of 0° ply orientation the deformation is very low, and at the same time, it can be observed that the stress–strain plot gives an excellent Young modulus value of 35.59 GPa. The tensile strength of 3 mm thickness was noted as 1188 GPa. The comparison between NRFC and HNRC results is presented in Figure 19.

6.2. Comparison with Results of All Cases (Natural Reinforced Fiber Composite Laminate and Hybrid Composite)

The maximum deformation between NRFC and HNRC is 0.95 mm, and it is noted at (HNRC) 0° with a 4 mm thickness of the laminate. The minimum deformation value is 0.105 mm, and it is noted at (NRFC) 45° with 2 mm thickness. All laminate results are presented in Table 12.
The maximum von Mises stress between NRFC and HNRC is 6124.26 Pa, and it is noted at (HNRC) 0° with a 2 mm thickness of the laminate. The minimum von Mises stress value is 1021.11 Pa, and it is noted at (NRFC) 0° with 4 mm thickness. All laminate results are presented in Table 13.
The maximum strain between NRFC and HNRC is 0.631, and it is noted at (NRFC) 45° with a 2 mm thickness of the laminate. The minimum strain value is 0.131, and it is noted at (HNRC) 0° with 4 mm thickness. All laminate results are presented in Table 14.
The maximum shear stress between NRFC and HNRC is 1188.39 Pa, and it is noted at (HNRC) 45° with a 3 mm thickness of the laminate. The minimum shear stress value is 79.67 Pa, and it is noted at (NRFC) 0° with 4 mm thickness. All laminate results are presented in Table 15.
Based on the Ansys results, the composite laminate is fabricated with different thicknesses and is ready for laboratory testing. This material can be used for UAV structures. The fabricated laminates are shown in Figure 20.

7. Limitations

Indeed, this investigation demonstrated that HFRC laminates have significantly better mechanical properties than natural jute fibers. Despite this, there are a few limitations that must be addressed in future research. The simulation model for HFRC laminates made of jute fibers and glass fibers disregarded the effect of moisture absorption as well as the microbial degradation of fibers over time [44]. However, by chemically treating fibers and modifying their surfaces, it is possible to reduce moisture absorption and microbial degradation [45]. These issues can be addressed in future research investigations.

8. Conclusions and Future Work

In this work, the research is conducted on materials of different strengths of various types of ply sequences concerning thickness. The maximum stress, strain, deformation, and shear stress were obtained at different ply configurations using Ansys. Natural reinforced fiber composite (Jute) is considered a laminate 2 mm with different ply sequences, and the total number of plies is 8. After assessing the simulation results, it was found that mechanical properties such as maximum stress values are noted in the longitudinal direction with fiber various orientations. HFRC laminates, the results of 0° ply orientation were noted as 0.274, 0.237, and 0.209 mm for 2–4 mm thickness, respectively. The other orientations and ply sequences did not show positive results. However, at 45° ply orientation at 3 mm thickness, the tensile strength, and deformation were noted as 1188 GPa and 0.237, respectively. Hybrid natural-fiber-reinforced composites have better mechanical properties than natural fibers, and the simulation results were analyzed at different configurations. The jute and glass fiber with epoxy resin composite is the best material at 3 mm thickness with 45° ply orientation. These hybrid composites can be used in wing construction and sub parts of spars, ribs, and also for skin in unmanned aerial vehicles.
The results from this study indicated that the material yield strength dramatically increased when the material is a combination of natural fibers with synthetic fibers. There is less deformation even in natural fibers; however, the material characteristic is identified as yield strength, and the material has linier elastic behavior. Material strength depends upon the fiber orientation. Nevertheless, with these results, the specimens have to be tested at the laboratory. Finally, it has been concluded that natural hybrid composites are suitable for the fabrication of prototype unmanned aerial vehicle structures based on simulation results. The results of simulation and tests were proposed for future work to fabricate the test specimens as per the ASTM standards. After analyzing the laboratory tests, this material can be used for manufacturing in UAV applications.

Author Contributions

The basic conceptualization, R.K.; methodology, W.S.; software, M.S. and R.K.; writing, R.K. and M.S.; review and editing, W.S. and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledges the financial support from the university, under the excellence of project grants provided by the University (BKM_10/060/BKM22/2023). Department of Fundamentals of Machinery Design, Faculty of Mechanical Engineering, Silesian University of Technology, Gliwice, Poland. The research reported in this paper was also co-financed by the European Union from the European Social Fund in the framework of the project “Silesian University of Technology as a Center of Modern Education based on research and innovation” POWR.03.05.00-00-Z098/17.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to express our thanks to the following researchers for their contribution to the research and carrying out some of the preparatory work as part of Project-Based Learning—the supervisors: Maciej Dyzia, Roman Niestrój, and the students: Adam Jabłoński, Kamil Kalarus, Agnieszka Gębka, Maciej Znojkiewicz, Mateusz Duleba, Miłosz Kościelniak.

Conflicts of Interest

The authors declare that there are no conflict of interest.

References

  1. Faridul Hasan, K.M.; György Horváth, P.; Bak, M.; Alpár, T. A state-of-the-art review on coir fiber-reinforced bio-composites. R. Soc. Chem. 2021, 11, 10548–10571. [Google Scholar]
  2. Ali, L.M.; Ansari, M.N.M.; Pua, G.; Jawaid, M.; Islam, M.S. A review on natural fiber reinforced polymer composite and its applications. Int. J. Polym. Sci. 2015, 2015, 243947. [Google Scholar]
  3. Chandramohan, D.; Bharanichandar, J. Natural Fiber Reinforced Polymer Composites for Automobile Accessories. Am. J. Environ. Sci. 2013, 9, 494–504. [Google Scholar] [CrossRef]
  4. Brischetto, S. Analysis of natural fibre composites for aerospace structures. Aircr. Eng. Aerosp. Technol. 2018, 9, 1372–1384. [Google Scholar] [CrossRef]
  5. Pickering, K. (Ed.) Properties and Performance of Natural-Fibre Composites; Woodhead Publishing: Cambridge, UK, 2008. [Google Scholar]
  6. Velasco-Parra, J.A.; Ramón-Valencia, B.A.; Mora-Espinosa, W.J. Mechanical characterization of jute fiber-based bio composite to manufacture automotive components. J. Appl. Res. Technol. 2021, 19, 472–491. [Google Scholar] [CrossRef]
  7. Textile Technology. Available online: https://www.textiletechnology.net/fibers/news/dnfi-stable-world-natural-fiber-production-in-2022-32691 (accessed on 16 September 2022).
  8. Pickering, K.L.; Efendy, M.G.A.; Le, T.M. A review of recent developments in natural fibre composites and their mechanical performance. Compos. Part A Appl. Sci. Manuf. 2016, 83, 98–112. [Google Scholar] [CrossRef] [Green Version]
  9. Saravanan, K.G.; Prabu, R.; Sivapragasam, A.; Daniel, N. Comparative Analysis of Natural Fibre Reinforced Composite Material Using ANSYS. Adv. Mater. Sci. Eng. 2021, 2021, 9391237. [Google Scholar] [CrossRef]
  10. Ramli, N.; Mazalan, N.; Ando, Y. Natural fiber for green technology in automotive industry a brief review. In IOP Conference Series: Materials Science and Engineering, Proceeding of the Wood and Biofiber International Conference (WOBIC 2017), Selangor, Malaysia, 21–23 November 2017; IOP Science: Bristol, UK, 2017. [Google Scholar]
  11. Jayamani, E.; Hamdan, S.; Rezaur Rahman, K.G.; Bin Bakria, M.K. Comparative study of dielectric properties of hybrid natural fiber composites. In Proceedings of the 12th Global Congress on Manufacturing and Management GCCM, Vellore, India, 8–10 December 2014. [Google Scholar]
  12. Erden, S.; Ho, K. Fiber Reinforced Composite in Fiber Technology for Fiber Reinforced Composites; Woodhead Publishing Series in Composites Science and Engineering; Woodhead Publishing: Cambridge, UK, 2017; pp. 51–79. [Google Scholar]
  13. Bale, J.; Boimau, K.; Nenobesi, M. Natural Composite Reinforced by Lontar (Borassus flabellifer) Fiber: An Experimental Study on Open-Hole Tensile Strength. Int. J. Biomater. 2017, 2017, 7685047. [Google Scholar] [CrossRef] [Green Version]
  14. Khan, M.A.; Hinrichsen, M.A.; Drzal, L.T. Influence of novel coupling agents on mechanical properties of jute reinforced polypropylene composite. J. Mater. Sci. Lett. 2001, 20, 1711–1713. [Google Scholar] [CrossRef]
  15. Saravana Bavan, D.; Mohan Kumar, G.C. Potential use of natural fiber composite materials in India. J. Reinf. Plast. Compos. 2010, 29, 3600–3613. [Google Scholar] [CrossRef]
  16. Liu, Z.; Wang, H.; Yang, L.; Du, J. Research on mechanical properties and durability of flax/glass fiber bio-hybrid FRP composites laminates. Compos. Struct. 2022, 290, 115566. [Google Scholar] [CrossRef]
  17. Shaik, A.A.; Pradhan, A.A.; Kotsthane, A.M.; Patil, S.; Karuppan, S. Comparative analysis of basalt/E-Glass/S2-fiberglass-carbon fiber reinforced epoxy laminates suing finite element method. Mater. Today-Proc. 2022, 63, 630–638. [Google Scholar] [CrossRef]
  18. Laurin, F.; Carrere, N.; Maire, J.-F. Laminated composite structures subjected to compressive loading: A material and structural buckling analysis. Compos. Struct. 2007, 80, 172–182. [Google Scholar] [CrossRef]
  19. Reichert, S.; Schwinn, T.; La Magna, R.; Waimer, F.; Knippers, J.; Menges, A. Fibrous structures: An integrative approach to design computation, simulation and fabrication for lightweight, glass and carbon fibre composite structures in architecture based on biomimetic design principles. Comput. Aided-Des. 2014, 52, 27–39. [Google Scholar] [CrossRef]
  20. Ammarullah, M.I.; Santoso, G.; Sugiharto, S.; Supriyono, T.; Kurdi, O.; Tauviqirrahman, M.; Winarni, T.I.; Jamari, J. Tresca stress study of CoCrMo-on-CoCrMo bearings based on body mass index using 2D computational model. J. Tribol. 2022, 33, 31–38. [Google Scholar]
  21. Celino, A.; Freour, S.; Jacquemin, F.; Casari, P. The hygroscopic behavior of plant fibers: A review. Front. Chem. 2014, 1, 43. [Google Scholar] [CrossRef] [Green Version]
  22. Holbery, J.; Houston, D. Natural-fiber-reinforced polymer composites applications in automotive. JOM 2006, 58, 80–86. [Google Scholar] [CrossRef]
  23. Adams, D.F.; Perry, J.L. Instrumented Charpy impact tests of several unidirectional composite materials. Fibre Sci. Technol. 1975, 8, 275–302. [Google Scholar] [CrossRef]
  24. Adams, D.F.; Perry, J.L. Low level Charpy impact of graphite epoxy hybrid composites. J. Eng. Mater. Technol. ASME 1977, 99, 257–263. [Google Scholar] [CrossRef]
  25. Pereira, A.C.; Monteiro, S.N.; de Assis, F.S.; Margem, F.M.; da Luz, F.S.; de Oliveira Braga, F. Charpy impact tenacity of epoxy matrix composites reinforced with aligned jute fibers. J. Mater. Res. Technol. 2017, 6, 312–316. [Google Scholar] [CrossRef]
  26. Singh, H.; Inder, J.; Singh, P.; Singh, S.; Dhawan, V.; Tiwari, S.K. A brief review of jute fiber and its composites. Mater. Today Proc. 2018, 5, 28427–28437. [Google Scholar] [CrossRef]
  27. EL-Wazerya, M.S.E.; EL-Elamya, M.I.E.; Zoalfakarb, S.H. Mechanical Properties of Glass Fiber Reinforced Polyester Composites. Int. J. Appl. Sci. Eng. 2017, 14, 121–131. [Google Scholar]
  28. Kabir, M.; Wang, H.; Lau, K.; Cardona, F. Chemical treatments on plant-based natural fibre reinforced polymer composites: An overview. Compos. Part B Eng. 2012, 43, 2883–2892. [Google Scholar] [CrossRef]
  29. Goumgh, A.; Assarara, M.; Zouaria, W.; Azouaouib, K.; El Mahic, A.; Ayada, R. Study of the fatigue behavior of hybrid flax-glass/epoxy composites. Compos. Struct. 2022, 294, 115790. [Google Scholar] [CrossRef]
  30. Mohantya, A.K.; Misraa, B.; Hinrichsen, G. Bio-fibres biodegradable polymers and bio-composites. Macromol. Mater. Eng. 2000, 276–277, 1–24. [Google Scholar] [CrossRef]
  31. Stamboulisa, A.; Bailliea, C.A.; Peijsb, T. Effects of environmental conditions on mechanical and physical properties of flax fibers. Compos. Part A Appl. Sci. Manuf. 2001, 32, 1105–1115. [Google Scholar] [CrossRef]
  32. Alawar, A.; Hamed, A.M.; Al-Kaabi, K. Characterization of treated date palm tree fiber as composite reinforcement. Compos. Part B Eng. 2007, 40, 601–606. [Google Scholar] [CrossRef]
  33. Alves, C.; Silva, A.; Reis, L.; Ferrão, P.; Freitas, M. Sustainable design of automotive components through jute fiber composites: An integrated approach. New Trends Dev. Automot. Ind. 2011, 14, 223–254. [Google Scholar]
  34. Fakirov, S.; Bhattacharyya, D. Homopolymers, blends and composites. In Engineering Biopolymers; Hanser Publishers: Munich, Germany, 2007; ISBN 978-1-56990-405-3. [Google Scholar]
  35. Nguyen, L.H.; Leassig, T.; Ryan, S. A methodology for hydrocode analysis of ultra-high molecular weight polyethylene composite under ballistic impact. Compos. Part A Appl. Sci. Manuf. 2016, 84, 224–235. [Google Scholar] [CrossRef]
  36. Yen, C.F.; Caiazzo, A.A. Innovative Processing of Multifunctional Composite Armor for Ground Vehicles; Technical Report ARL-CR-484; U.S. Army Research Laboratory ARL: Aberdeen Proving Ground, MD, USA, 2001. [Google Scholar]
  37. Gama, B.A. A Progressive Composite Damage Model for Unidirectional and Woven Fabric; Newark Materials Sciences Corporation (MSC): Canton, MI, USA; University of Delaware Center for Composite Materials (UD-CCM): Newark, DE, USA, 2015; No. 215. [Google Scholar]
  38. Yen, C.F. A ballistic material model for continuous-fiber reinforced composites. Int. J. Impact Eng. 2012, 46, 11–22. [Google Scholar] [CrossRef] [Green Version]
  39. Hassan, M.; Mahnaz, Z. Strain-rate-dependent progressive damage modelling of UHMWPE composite laminate subjected to impact loading. Int. J. Damage Mech. 2021, 31, 1–31. [Google Scholar]
  40. Serra-Parareda, F.; Vilaseca, F.; Aguado, R.; Espinach, F.X.; Tarrés, Q.; Delgado-Aguilar, M. Effective Young’s Modulus Estimation of Natural Fibers through Micromechanical Models: The Case of Henequen Fibers Reinforced-PP Composites. Polymers 2021, 1322, 3947. [Google Scholar] [CrossRef]
  41. Mitra, B.C. Environment friendly composite material: Bio composite and Green Composite. Déf. Sci. J. 2014, 64, 244–261. [Google Scholar]
  42. Abilash, N.; Sivapragash, M. Environmental benefits of eco-friendly natural fiber reinforced polymeric composite materials. Int. J. Appl. Innov. Eng. Manag. 2013, 2, 53–59. [Google Scholar]
  43. Ammarullah, M.I.; Santoso, G.; Sugiharto, S.; Supriyono, T.; Wibowo, D.B.; Kurdi, O.; Tauviqirrahman, M.; Jamari, J. Minimizing Risk of Failure from Ceramic-on-Ceramic Total Hip Prosthesis by Selecting Ceramic Materials Based on Tresca Stress. Sustainability 2022, 14, 13413. [Google Scholar] [CrossRef]
  44. Hodzic, A.; Shanks, R. Natural Fibre Composites: Materials, Processes and Properties; Woodhead Publishing: Oxford, UK, 2014; ISBN 0857099221. [Google Scholar]
  45. Ku, H.; Wang, H.; Pattarachaiyakoop, N.; Trada, M. A review on the tensile properties of natural fiber reinforced polymer composites. Compos. Part B Eng. 2011, 42, 856–873. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of experimental work methodology for the evaluation of load-bearing capacity of materials.
Figure 1. Schematic diagram of experimental work methodology for the evaluation of load-bearing capacity of materials.
Sustainability 14 15992 g001
Figure 2. Natural reinforced fibers composite design and optimization layout.
Figure 2. Natural reinforced fibers composite design and optimization layout.
Sustainability 14 15992 g002
Figure 3. Configuration of ply orientation for natural fiber jute laminates, (a,d,g) ply orientation of 0° with 2–4 mm, (b,e,h) ply orientation of 30° with 2–4 mm, (c,f,i) ply orientation of 45° with 2–4 mm.
Figure 3. Configuration of ply orientation for natural fiber jute laminates, (a,d,g) ply orientation of 0° with 2–4 mm, (b,e,h) ply orientation of 30° with 2–4 mm, (c,f,i) ply orientation of 45° with 2–4 mm.
Sustainability 14 15992 g003
Figure 4. Configuration of hybrid composite laminates, (a,d,g) ply orientation of 0° with 2–4 mm, (b,e,h) ply orientation of 30° with 2–4 mm, (c,f,i) ply orientation of 45° with 2–4 mm.
Figure 4. Configuration of hybrid composite laminates, (a,d,g) ply orientation of 0° with 2–4 mm, (b,e,h) ply orientation of 30° with 2–4 mm, (c,f,i) ply orientation of 45° with 2–4 mm.
Sustainability 14 15992 g004
Figure 5. (a) Natural fiber: Jute 3D-CAD model; (b) hybrid natural fiber 3D-CAD model; and (c) geometrical 2D-CAD model with 2, 3 and 4 mm.
Figure 5. (a) Natural fiber: Jute 3D-CAD model; (b) hybrid natural fiber 3D-CAD model; and (c) geometrical 2D-CAD model with 2, 3 and 4 mm.
Sustainability 14 15992 g005
Figure 6. (a) Geometrical CAD model and (b) FEM model with boundary conditions.
Figure 6. (a) Geometrical CAD model and (b) FEM model with boundary conditions.
Sustainability 14 15992 g006
Figure 7. Deformation at 0°, 30°, and 45° with different configurations (Jute—2–4 mm).
Figure 7. Deformation at 0°, 30°, and 45° with different configurations (Jute—2–4 mm).
Sustainability 14 15992 g007
Figure 8. Maximum stress at 0°, 30°, and 45° with different configurations (Jute—2–4 mm).
Figure 8. Maximum stress at 0°, 30°, and 45° with different configurations (Jute—2–4 mm).
Sustainability 14 15992 g008
Figure 9. Maximum strain at 0°, 30°, and 45° with different configurations (Jute—2–4 mm).
Figure 9. Maximum strain at 0°, 30°, and 45° with different configurations (Jute—2–4 mm).
Sustainability 14 15992 g009
Figure 10. Shear stress of 0°, 30°, and 45° with different configurations (Jute—2–4 mm).
Figure 10. Shear stress of 0°, 30°, and 45° with different configurations (Jute—2–4 mm).
Sustainability 14 15992 g010
Figure 11. Stress–strain graph of 0°, 30°, and 45° with different configurations (Jute and hybrid—2–4 mm).
Figure 11. Stress–strain graph of 0°, 30°, and 45° with different configurations (Jute and hybrid—2–4 mm).
Sustainability 14 15992 g011
Figure 12. Stress–strain graph comparison of 0°, 30°, and 45° with deferent configurations (Jute 2–4 mm).
Figure 12. Stress–strain graph comparison of 0°, 30°, and 45° with deferent configurations (Jute 2–4 mm).
Sustainability 14 15992 g012
Figure 13. Maximum deformation of 0°, 30°, and 45° with different configurations (Hybrid—2–4 mm).
Figure 13. Maximum deformation of 0°, 30°, and 45° with different configurations (Hybrid—2–4 mm).
Sustainability 14 15992 g013
Figure 14. Maximum stress of 0°, 30°, and 45° with different configurations (Hybrid—2–4 mm).
Figure 14. Maximum stress of 0°, 30°, and 45° with different configurations (Hybrid—2–4 mm).
Sustainability 14 15992 g014
Figure 15. Maximum Strain at 0°, 30°, and 45° with different configurations (Hybrid—2–4 mm).
Figure 15. Maximum Strain at 0°, 30°, and 45° with different configurations (Hybrid—2–4 mm).
Sustainability 14 15992 g015
Figure 16. Maximum shear stress at 0°, 30°, and 45° with different configurations (Hybrid—2–4 mm).
Figure 16. Maximum shear stress at 0°, 30°, and 45° with different configurations (Hybrid—2–4 mm).
Sustainability 14 15992 g016
Figure 17. Stress–strain graph comparison of 0°, 30°, and 45° with different configurations (Jute—2–4 mm).
Figure 17. Stress–strain graph comparison of 0°, 30°, and 45° with different configurations (Jute—2–4 mm).
Sustainability 14 15992 g017
Figure 18. Stress–strain graph comparison between (a) hybrid composite and (b) natural reinforced fiber composite (Jute and Hybrid—2–4 mm).
Figure 18. Stress–strain graph comparison between (a) hybrid composite and (b) natural reinforced fiber composite (Jute and Hybrid—2–4 mm).
Sustainability 14 15992 g018
Figure 19. Summarized results (Jute and Hybrid—2–4 mm).
Figure 19. Summarized results (Jute and Hybrid—2–4 mm).
Sustainability 14 15992 g019
Figure 20. Fabricated laminates: (a) natural jute fibers specimens; (b) hybrid natural composite specimens.
Figure 20. Fabricated laminates: (a) natural jute fibers specimens; (b) hybrid natural composite specimens.
Sustainability 14 15992 g020
Table 1. Mechanical properties of natural fibers.
Table 1. Mechanical properties of natural fibers.
Natural
Fibers
Types
Dia. (μm)Tensile Strength (MPa)Young’s
Modulus (GPa)
Elongation at Break (%)References
Jute fiber25–200200, 393, 773–111013–25, 26, 26.5 0.7, 1.16–1.5[9,27,28,29,30]
Flax fiber10–40110, 343, 600–200012, 30, 60, 85–1200.7, 1, 1.16, 1.5, 2.1–4[9,27,28,29,30]
Sisal fiber50–200468–6409.4–22.03–7[29,30]
Cotton fiber287–8005.5–12.67–8[29,30]
Coir fiber100–450131–1754–615–40[29,30]
Raw date palm fiber (DPF)100–100058–2032–7.55–10[29,30]
Table 2. Properties of epoxy resins.
Table 2. Properties of epoxy resins.
DescriptionValueReferences
Young’s modulus (MPa)35,000[9]
Poisson’s ration0.35[9]
Density (kg/m3)1280[9]
Table 3. Chemical properties of natural fiber.
Table 3. Chemical properties of natural fiber.
Natural Fibers
Types
CelluloseLigninHemicellulosePectin/WaxMoisture ContentReferences
Jute61, 67–71.5, 12–1313.6–20.40.2/0.512.6[1,27,29,31]
Flax712.218.6–20.62.3/1.710.0[1,27,29,31]
Sisal67–68, 67–788.0–11.010.0–14.2010.0/2.0/20.11.0[1,27,29,31]
Coir43.448.34.0 10.2[1,27]
Table 4. Total deformation results at various configurations of 2–4 mm.
Table 4. Total deformation results at various configurations of 2–4 mm.
Description0° Deformation (mm)30° Deformation (mm)45° Deformation
(mm)
Jute fiber (2 mm)0.8970.9500.105
Jute fiber (3 mm)0.5940.6460.713
Jute fiber (4 mm)0.4460.4750.591
Table 5. The total maximum stress (von Mises) results at various configuration 2–4 mm.
Table 5. The total maximum stress (von Mises) results at various configuration 2–4 mm.
Description0° Maximum Stress (Von-Mises) (Pa)30° Maximum Stress (Von-Mises) (Pa)45° Maximum Stress (Von-Mises) (Pa)
Jute fiber (2 mm)2043.952157.172384.01
Jute fiber (3 mm)1362.151502.571650.00
Jute fiber (4 mm)1021.111087.961364.00
Table 6. The total maximum strain results at various configurations 2–4 mm.
Table 6. The total maximum strain results at various configurations 2–4 mm.
Description0° Max. Von. Strain30° Max. Von. Strain45° Max. Von. Strain
Jute fiber (2 mm)0.5510.5710.631
Jute fiber (3 mm)0.3670.4770.522
Jute fiber (4 mm)0.2760.3430.431
Table 7. The total shear stress results at various configuration 2–4 mm.
Table 7. The total shear stress results at various configuration 2–4 mm.
Description0° Shear Stress (XY) (Pa)30° Shear Stress (XY) (Pa)45° Shear Stress (XY) (Pa)
Jute fiber (2 mm)155.00591.00798.20
Jute fiber (3 mm)102.92505.00467.00
Jute fiber (4 mm)79.67322.00498.00
Table 8. The total deformation results at various configurations 2–4 mm.
Table 8. The total deformation results at various configurations 2–4 mm.
Description0° Deformation (mm)30° Deformation (mm)45° Deformation (mm)
Hybrid (J/G/J)Hybrid (J/G/J)Hybrid (J/G/J)
Thickness (2 mm)0.2740.5430.626
Thickness (3 mm)0.2370.4140.488
Thickness (4 mm)0.2090.3270.382
Table 9. The total Max. Stress results at various configuration 2–4 mm.
Table 9. The total Max. Stress results at various configuration 2–4 mm.
Description0° Maximum Stress (von Mises) (Pa)30° Maximum Stress (von Mises) (Pa)45° Maximum Stress (von Mises) (Pa)
Hybrid (J/G/J)Hybrid (J/G/J)Hybrid (J/G/J)
Thickness (2 mm)6124.265247.634860.69
Thickness (3 mm)5303.474447.023635.47
Thickness (4 mm)4679.53946.653257.8
Table 10. The total maximum von Mises strain results at various configurations of 2–4 mm.
Table 10. The total maximum von Mises strain results at various configurations of 2–4 mm.
Description0° Max. Von. Strain30° Max. Von. Strain45° Max. Von. Strain
Hybrid (J/G/J)Hybrid (J/G/J)Hybrid (J/G/J)
Thickness (2 mm)0.1740.3220.377
Thickness (3 mm)0.1490.2560.295
Thickness (4 mm)0.1310.2020.231
Table 11. The shear stress results at various configuration 2–4 mm.
Table 11. The shear stress results at various configuration 2–4 mm.
Description0° Shear Stress (XY) (Pa)30° Shear Stress (XY) (Pa)45° Shear Stress (XY) (Pa)
Hybrid (J/G/J)Hybrid (J/G/J)Hybrid (J/G/J)
Thickness (2 mm)332.824356.55437.1
Thickness (3 mm)249.661141.731188.39
Thickness (4 mm)198.755215.69277.711
Table 12. The total deformation (in mm) at various configurations 2–4 mm.
Table 12. The total deformation (in mm) at various configurations 2–4 mm.
Description0° Deformation30° Deformation45° Deformation
JuteHybrid (J/G/J)JuteHybrid (J/G/J)JuteHybrid (J/G/J)
Thickness (2 mm)0.8970.2740.950.5430.1050.626
Thickness (3 mm)0.5940.2370.6460.4140.7130.488
Thickness (4 mm)0.4460.2090.4750.3270.5910.382
Table 13. The total stresses (von Mises in Pa) at various configuration 2–4 mm.
Table 13. The total stresses (von Mises in Pa) at various configuration 2–4 mm.
Description0° Maximum Stress 30° Maximum Stress45° Maximum Stress
JuteHybrid (J/G/J)JuteHybrid (J/G/J)JuteHybrid (J/G/J)
Thickness (2 mm)2043.956124.262157.175247.632384.014860.69
Thickness (3 mm)1362.155303.471502.574447.0216503635.47
Thickness (4 mm)1021.114679.51087.963946.6513643257.8
Table 14. The total strain at various configuration 2, 3, and 4 mm.
Table 14. The total strain at various configuration 2, 3, and 4 mm.
Description0° Max. Von. Strain30° Max. Von. Strain45° Max. Von. Strain
JuteHybrid (J/G/J)JuteHybrid (J/G/J)JuteHybrid (J/G/J)
Thickness (2 mm)0.5510.1740.5710.3220.6310.377
Thickness (3 mm)0.3670.1490.4770.2560.5220.295
Thickness (4 mm)0.2760.1310.3430.2020.4310.231
Table 15. The total shear stresses (Pa) at various configuration 2–4 mm.
Table 15. The total shear stresses (Pa) at various configuration 2–4 mm.
Description0° Shear Stress (XY)30° Shear Stress (XY)45° Shear Stress (XY)
JuteHybrid (J/G/J)JuteHybrid (J/G/J)JuteHybrid (J/G/J)
Thickness (2 mm)155332.824591356.55798.2437.1
Thickness (3 mm)102.92249.665051141.734671188.39
Thickness (4 mm)79.67198.755322215.69498277.711
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kumpati, R.; Skarka, W.; Skarka, M. Ansys-Based Evaluation of Natural Fiber and Hybrid Fiber-Reinforced Composites. Sustainability 2022, 14, 15992. https://doi.org/10.3390/su142315992

AMA Style

Kumpati R, Skarka W, Skarka M. Ansys-Based Evaluation of Natural Fiber and Hybrid Fiber-Reinforced Composites. Sustainability. 2022; 14(23):15992. https://doi.org/10.3390/su142315992

Chicago/Turabian Style

Kumpati, Ramesh, Wojciech Skarka, and Michał Skarka. 2022. "Ansys-Based Evaluation of Natural Fiber and Hybrid Fiber-Reinforced Composites" Sustainability 14, no. 23: 15992. https://doi.org/10.3390/su142315992

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