Next Article in Journal
Surface Passivation for Promotes Bi-Excitonic Amplified Spontaneous Emission in CsPb(Br/Cl)3 Perovskite at Room Temperature
Next Article in Special Issue
Review on the Degradation of Poly(lactic acid) during Melt Processing
Previous Article in Journal
Main Characteristics of Processed Grain Starch Products and Physicochemical Features of the Starches from Maize (Zea mays L.) with Different Genotypes
Previous Article in Special Issue
Characterization and Process Optimization for Enhanced Production of Polyhydroxybutyrate (PHB)-Based Biodegradable Polymer from Bacillus flexus Isolated from Municipal Solid Waste Landfill Site
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Simulation of Wood Polymer Composites with Finite Element Analysis

by
Satya Guha Nukala
1,
Ing Kong
1,*,
Akesh Babu Kakarla
1,
Vipulkumar Ishvarbhai Patel
1 and
Hossam Abuel-Naga
2
1
Department of Engineering, School of Computing, Engineering and Mathematical Sciences, La Trobe University, Bendigo, VIC 3552, Australia
2
Department of Engineering, School of Computing, Engineering and Mathematical Sciences, La Trobe University, Melbourne, VIC 3086, Australia
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(9), 1977; https://doi.org/10.3390/polym15091977
Submission received: 28 February 2023 / Revised: 3 April 2023 / Accepted: 18 April 2023 / Published: 22 April 2023

Abstract

:
Wood is a cellulosic material that is most abundantly available in nature. Wood has been extensively used as reinforcement in polymer composite materials. Wood polymer composite (WPC) is an environmentally friendly and sustainable material exploited in building and construction within the marine, packaging, housewares, aerospace, and automotive industries. However, the precision of testing equipment for finding the properties of WPCs becomes less feasible compared to experimental analysis due to a high degree of differences in the measurement of properties such as stress, strain and deformation. Thus, evaluating the mechanical properties of WPCs using finite element analysis (FEA) can aid in overcoming the inadequacies in measuring physical properties prior to experimental analyses. Furthermore, the prediction of mechanical properties using simulation tools has evolved to analyze novel material performance under various conditions. The current study aimed to examine the mechanical properties of saw dust-reinforced recycled polypropylene (rPP) through experimentation and FEA. A model was developed using SolidWorks, and simulation was performed in ANSYS to predict the mechanical properties of the WPCs. To validate the obtained results, the simulated static tension test results were confirmed with experimental tension tests, and both assessments were well in accordance with each other. Using FEA to predict material properties could be a cost-effective technique in studying new materials under varied load conditions.

1. Introduction

Wood polymer composites (WPCs), introduced in the 1990s, have a lesser environmental impact and lower maintenance compared to other non-sustainable glass- and carbon-reinforced composite materials [1]. Following a rapid increase in demand for sustainable materials in recent years, WPCs have attracted increasing attention from the scientific community and industries to replace non-sustainable composites [2]. The benefit of WPCs is that they possess unique characteristics such as stiffness, lower water intake, high sustainability, dimensional stability and specific strength over their lifetime, and durability against environmental impacts [3,4,5] due to their combination of wood and polymeric materials [6]. In addition, polymer composites produced using wood filler have favourable mechanical properties and a higher rigidity than unfilled polymer materials [7]. The most significant application of WPCs is in the building and construction industry [1,8]. WPCs are used in a wide range of products, such as railing and fencing for construction [9], and in the interior and exterior parts of automobiles [10,11]. Furthermore, there are some major WPC manufacturers in Australia, such as Tuff Deck—Composite Decking (Dandenong South, VIC, Australia), ModWood Technologies (Campbellfield, VIC, Australia) and Advanced Plastic Recycling (Edinburgh, SA, Australia). Prior to this, many researchers have used virgin polymers for the development of WPCs, such as polyethylene terephthalate (PET) [12], polypropylene (PP) [13,14], polyethylene (PE) [13,15], polylactic acid (PLA) [16], polyvinyl chloride (PVC) [17] and polyurethane (PU) [18], which are the most predominantly used materials as the polymer matrices in WPCs. Among the above-mentioned polymers, polypropylene (PP) reinforced with wood fillers has been used for various industrial applications extensively [19,20,21]. Over the years, the amount of plastic waste generated has been increasing and has resulted in significant amounts of municipal solid waste (MSW). Therefore, there have been attempts to recycle post-consumer plastics for the production of WPCs in order to offload their ecological impacts [22,23].
The production and usage of new composite materials should satisfy the requirements in individual industrial applications. Prototyping requires extensive trials to ensure that appropriate performance is obtained throughout its lifespan. Therefore, rather than performing a series of tests for each prototype, it is preferable to use a model that can mimic similar behaviours and conditions during simulation. This facilitates the quick correction of the digital model, evaluates results and predicts trends. Finite element analysis (FEA) is the most used simulation process among the research by the science and engineering communities [24]. Using FEA simulation to predict mechanical properties during the design phase is an effective tool for optimizing parameters.
In this context, Dickson et al. [25] reported the mechanical performance of carbon, Kevlar and glass fiber-reinforced nylon composites using FEA. The results showed that the strength of Kevlar-reinforced nylon had the highest strength followed by glass fiber- and carbon fiber-reinforced composites. Caminero et al. [26] validated the impact and damage resistance of 3D printed thermoplastic reinforced with Kelvar fibers using the fused deposition modelling technique. It was reported that the impact strength increased as the percentage of fibers increased. Alharbi et al. [27] studied the uniaxial stress–strain response of 3D printed polylactic acid (PLA) using FEA and validated it with experimental data. The results showed that the stress and displacement did not vary significantly within the element range of interest. The model showed a maximum resultant displacement of 2.227 mm and strain rate of 1.615 × 10−2 [27], respectively. Bhandari et al. [28] created a 3D printed model composed of polyetherimide material to predict the elastic response of the materials using FEA. The accuracy of the FEA using Poisson’s ratio was less accurate and showed a difference of up to 19.6% between the predicted and observed values. The internal lattice structure from the analysis significantly matched the 3D printed model values.
In addition, a small number of investigations has been conducted on WPCs using FEA. For instance, Roy et al. [1] demonstrated the mechanical analysis of composites made of wood waste and recycled polypropylene. The predicted flexural and tensile properties of the biocomposites were validated by experimental data. The results showed good agreement between the simulation and empirical data. Similarly, Bhaskar et al. [29] reported the results of a three-point bending test of WPCs produced from recycled polymer and pine wood flour. The recycled polymer matrix consisted of 80% polypropylene and 20% high-density polyethylene. The results showed that the composite at 40% of pine wood flour exhibited a maximum shear stress of 10.085 MPa and maximum equivalent stress of 44.815 MPa. Moreover, maximum deflection was observed at the point of contact of load. In another study, Ezzaraa et al. [30] demonstrated the FEA of 3D printed wood–PLA composites and compared the results with experimental data. The analysis was carried out to predict the mechanical properties of the 3D printed WPCs and to propose an efficient numerical tool to understand the effects of the internal porosity of 3D printed composites. The study reported that the elastic properties of the composites mostly varied based on the wood volume fraction and Young’s modulus was based on the addition of wood particles. The comparison between the FEA and experimental results showed that the predictions were significantly equal. Moreover, the analysis also concluded that wood particle size and shape were considerably subjective to the anisotropy of the composite which could enhance the longitudinal Young’s modulus and reduce the transverse Young’s modulus. Hartmann et al. [31] reported a numerical analysis of WPC using RVE and FEM in ANSYS Workbench V 21.1 (Canonsburg, PA, United States). The study was conducted using an FEA model developed based on a softwood tracheid mainly consisting of pine wood with 12% moisture at a temperature of 23 °C. The findings of the study stated that 10 MPa must be applied for the deformation of the softwood composites. However, experimental studies reported that pine wood compressive strength ranges from 4 to 14 MPa. Hence, analysis with the RVE approach was determined successfully with appropriate accuracy.
The present study aimed to create a microscale simulation model and perform the mechanical testing of WPCs using ANSYS Workbench. The software was used to simulate the layer-based composites. Furthermore, the mechanical properties of the WPCs composed of recycled polypropylene reinforced with sawdust were studied experimentally. The predicted results from the FEA were validated by the experimental data.

2. Materials and Methods

2.1. Materials

Recycled polypropylene (rPP) was collected from the Bendigo Recycling Centre, Eaglehawk, VIC, Australia. The collected rPP originated from used milk and yoghurt bottles. Sawdust (SD) was obtained from Raw Boards Pty. Ltd., Bendigo, VIC, Australia. The SD comprised different types of hardwood species (red gum, ironbark and yellow gum wood) collected from the various construction and demolition (C&D) activities with different particle sizes. The chemicals sodium stearate (C18H35NaO2) and sodium hydroxide (NaOH) were purchased from Bunnings, Bendigo, VIC, Australia. Hydrochloric acid (HCl) was procured from Sigma-Aldrich Pty. Ltd., Melbourne, VIC, Australia.

2.2. Methodology

Initially, rPP was shredded using a plastic crusher (DongGuan ZhongLi Instrument Technology Co., Ltd., Dongguan, China), and it was further cleaned with NaOH solution (5%) for 60 min. Afterwards, the rPP was washed using sodium stearate twice, followed by rinsing with water for the removal of excess dirt and other debris. The SD waste was collected from the various C&D activities in various sizes. Furthermore, the SD was sieved according to the ASTM E11 sieve method to obtain 0.05 mm particles [32]. Later, the SD was washed with a 20% NaOH solution and followed by 10 M HCl to remove excess dirt, alkaline and residues on the surface of the SD. Furthermore, the SD was rinsed three times with deionized water and oven-dried for 24 h at 70 °C as per the methodology reported by Medupin et al. [33]. Once the cleaning and segregation processes of rPP and SD were completed, they were pre-mixed in a Ziploc bag according to the concentrations shown in Table 1. The pre-mixture was later fed into a co-rotating batch mixer (ZL-3011 Rubber Lab Banbury Kneader Mixer; DongGuan ZhongLi Instrument Technology Co., Ltd., Dongguan, China). The mixer parameters, such as the hopper temperature (190 °C) and spindle speed (8 rpm), were maintained as constant and the hopper was rotated clockwise and anti-clockwise to obtain a good consistency of the WPCs. Finally, the obtained composite was crushed into tiny pieces with a plastic crusher (ZL-9031 Digital Crushing Strength Tester; DongGuan ZhongLi Instrument Technology Co., Ltd., Dongguan, China) and hot-pressed (CY-PCH-600D Laboratory Hydraulic Press; Zhengzhou CY Scientific Instrument Co., Ltd., Zhengzhou, China) at 190 °C with a pressure of 20 kPa for 15 min and cooled to room temperature [34,35]. Dog bone specimens were produced with an ASTM D638 type IV standard [36,37]. The experimental work used in determining each specimen’s mechanical properties was performed by Nukala et al. [34]. The averages of the data from the experimental work were evaluated for three replicates of each dataset using statistical analysis. The analysis was carried out with GraphPad Prism 9.0 (GraphPad Software, Inc., San Diego, CA, USA) through the ANOVA method.

2.3. Tensile Test

Hot-pressed dog bone specimens with a thickness of 3 mm were used for tensile testing. A Zhongli ZL-8001A tensile tester (DongGuan ZhongLi Instrument Technology Co., Ltd., Dongguan, China) at a crosshead speed of 3 mm/min with a load of 500 kN was used [38,39]. The samples were stored at a room temperature under dry conditions before testing [40,41,42].

2.4. Microstructural Analysis

The microstructural analysis of the fractured cross-sectional surfaces of the composites was conducted using a benchtop SEM (Hitachi Benchtop SEM 3030; Tokyo, Japan) [1]. The samples were sputter coated with platinum at 10 kV for 30 s before the examination [40]. The micrographs were obtained under variable pressures ranging from 10 to 15 kV.

2.5. Simulation

The 3D model was developed using SolidWorks software V 21.0 (Dassault Systèmes SE, Vélizy-Villacoublay, France) as per the ASTM D638 standard for tensile tests. The specimen boundary conditions were applied with one end of the model fixed (A) and the other with a load applied (B), as shown in Figure 1a. A load of 500 kN was used to predict the maximum stress, strain and deformation. The model was imported into the ANSYS Composite PrepPost (ACP) of ANSYS Workbench (Canonsburg, PA, USA) [31,43]. The ACP plugin was used to analyze the layered composites, as shown in Figure 1b. The model was stacked up in five layers with a total thickness of 3 mm. The materials used in the simulation were SD and rPP. Additionally, a customized material library was created in ANSYS Workbench using experimental data obtained from the literature [44,45,46,47]. The properties of the rPP and SD are listed in Table 2.
Based on a study by Roy et al. [1], it was assumed that the material would behave as isotropic in nature; it was further required to define its elastic coefficients in ANSYS. Therefore, the acquired engineering stress and strain values were converted into true stress and strain values according to Equations (1) and (2) [51] and the plastic strain values were calculated according to Equation (3) [1].
σ t r u e = σ e n g i n e e r i n g × 1 + e n g i n e e r i n g  
t r u e   =   ln 1 + e n g i n e e r i n g
p l a s t i c   =   t r u e σ t r u e E  
where is the strain, σ is the stress and E is the Young’s modulus of the material. The engineering stress ( σ engineering) and engineering strain ( engineering) values are the experimental tensile test values [52]. The average Young’s modulus of the composite material was calculated according to Hooke’s law, i.e., stress is directly proportional to strain.
The calculated values from Equations (1) and (3) were used to define the isotropic plastic hardening characteristics in the simulation software. The parameters for the deformation rate used in the FEA were the same as per the experimental conditions. To achieve accurate results, mesh generation and element size were adjusted accordingly. The simulation results were in accordance with the experimental results since all the constraints and boundary conditions were similar. Therefore, it could be concluded that material definition was accurate.

3. Results and Discussion

3.1. Morphology

The micrographs of the rPP-SD composites are shown in Figure 2. It can be seen that the SD was randomly distributed in the rPP, with no clear gap, no significant alignment in a particular direction and good interfacial bonding. The following bonding is an indication of the isotropic behaviour of the composite and stress transfer from the weaker matrix to the stronger SD fiber as reported by Adhikary et al. [53]. Furthermore, Renner et al. [54] reported that the strength and interfacial interactions confirmed the composite failure mode and micromechanical deformation.

3.2. Mesh Generation

Mesh sensitivity is a reference for how much a solution can be changed in terms of the entities such as mesh density, element type, number of elements and nodes used for the individual problem under study [55]. The most commonly used sensitivity is type I sensitivity in FEA, and to obtain an accurate solution, a finer mesh is preferred. Alternatively, type II sensitivity is another type of mesh that can be used as a refined mesh [56,57]. As per the discussion, an evaluation of mesh sensitivity was also performed in the present study to assess the effect of mesh size on von Mises stress [58] and the resultant displacement. Smooth refinement-based mesh was chosen over the standard mesh, as shown in Figure 3a. As a result, more nodes allowed the composite specimens to obtain highly accurate results and perform large deformations [59]. Furthermore, to improve the analysis of the mesh quality, mesh plots were created. It can also be noted that the maximum stress point was located on the rim of the specimen and failed nodal points are shown with red arrows in Figure 3b,c. It is worth mentioning that there were no failed nodal points at the centerline of the specimen where necking occurred which shows that the type I meshing obtained was very fine and smooth [60]. The subsequent results proved that the mesh quality was acceptable, as the aspect ratio (Figure 3b) was less than five and the Jacobian ratio (Figure 3c) was less than two [61,62]. The Jacobian ratio measures the shape of the given element, which is compared to the ideal element, and the ideal shape of an element depends on the element type [63]. Furthermore, the ideal Jacobian ratio should be between 1 and 10 for most elements, which correlates with the present Jacobian result [64,65]. Moreover, the results indicate that the element edge lengths were close to each other and had more positive values. In addition, the mesh generated a maximum number of nodes, as displayed in Figure 3b,c [27,66].

3.3. Mechanical Properties

3.3.1. Stress and Strain Distributions

The specimen’s equivalent stress and strain distribution while being stretched is displayed in Figure 4a,b. Concentration areas of maximum and minimum stress are indicated with red and blue arrows in Figure 4a. The analysis showed that the specimens failed at a maximum stress region of 21.75 MPa and a maximum strain region of 0.0075 mm at a strain rate of 29.5%.
The stress–strain curve from the experimental data and FEA are shown in Figure 5. The maximum experimental stress in rPP-SD3 and rPP-SD4 was obtained at 21.65 MPa and 23.15 MPa, respectively. The maximum strain rate was approximately 29.5% in both cases. The stress–strain curve for the experiment and FEA showed a linear elastic region at the beginning of tension, followed by a yield point indicating buckling and collapse. Furthermore, the composites showed rupture and decreased stress after the yield point. The high concentration region corresponded to composite failure at a maximum stress region 21.75 MPa for FEA and 22.85 MPa for the experimental analysis. According to the studies conducted by Bhaskar et al. [29] and Tiwari et al. [67], tensile strength and durability highly depend on a specimen’s parameters such as its dimensions, thickness and pressure loading. Another study conducted by Das Lala et al. [68] demonstrated the utilization of biodegradable polymer composites such as rubber seed shell and epoxy resin. The obtained results of the composites were validated using both experimental results and FEA. The results stated that with an increase in the filler content, the tensile strength of the composite increased. Furthermore, it was stated that a similar trend was observed in the results from the finite element technique. Navaneethakrishnan et al. [69] studied the structural analysis of a natural fiber-reinforced polymer matrix composite using vinyl ester resin, sisal and luffa fiber. The results indicated that a 20% sisal and 10% luffa natural fiber composite showed higher tensile strength. The experimental results were validated against ANSYS results and the tensile strength of the natural fiber-reinforced composite was 0.05 MPa higher than the experimental result. Thus, in relation to the above-mentioned discussion, it was evident that rPP-SD composites displayed a good agreement between the FEA and experimental results and were validated against each other as shown in Figure 5. Similar observations were made on the stress- and strain-related analysis using ANSYS software on different WPCs. Values were evaluated and validated experimentally and using ANSYS [70,71].

3.3.2. Deformation

The total deformation of the specimen occurred at a higher concentration area, shown with a red arrow in Figure 6a. The variation in deformation for the experimental and simulation values was 0.54 mm for rPP-SD3 and 0.68 mm for rPP-SD4. This variation in deformation can be considered as the tolerance of the machine [18]. Fang et al. [72] studied the mechanical performance of glass fiber-reinforced polymer (GFRP)–bamboo wood sandwich beams. The beams were investigated experimentally and validated using ANSYS Workbench. The experimental results showed that with the increase in the thickness of the bamboo and GRPF layers, there was a significant increase in the tensile strength and deformation capacity of the beams. Furthermore, the experimental deformation results were considerably similar to the FEA. The deformation and structural analysis of a sisal fiber-reinforced polymer composite-based wind turbine were investigated by Appadurai et al. [73]. In their study, the wind turbine profile was modelled in CATIA V5 (Dassault Systèmes SE, Vélizy-Villacoublay, France) and evaluated for its mechanical properties in ANSYS. The results stated that the sisal fiber-reinforced polymer composite-based wind turbine had superior mechanical properties compared to structural steel and other natural fiber composite wind turbines. Rostampour-Haftkhani et al. [74] measured and predicted the deformation performance of WPC profiles. The deformation of the WPC profiles was analyzed experimentally and with ANSYS software in order to decrease the cost and time of measurement. The results stated that FEA predicted deformation values with a mean absolute percentage of less than 3% and it was suggested as an efficient method for predicting the deformation properties of WPC profiles. In conclusion, the simulation result correlates with the experimental result as shown in Figure 6b,c, as it has a similar fracture concentration area and exhibits the same deformation rate.

4. Conclusions

This study aimed to investigate the mechanical properties of sawdust-reinforced recycled polypropylene composites. The morphology of the composites showed a homogeneous dispersion of SD in rPP, which aided in ensuring good mechanical properties. Von Mises stress, equivalent elastic strain and deformation were investigated. The results showed that the stress concentration region was due to the force applied according to simulation parameters. Similarly, the strain rate depended on the thickness and the force applied, while the deformation rate depended on the elongation of the specimen with respect to the stacked layers and the applied force. The location of the fracture point was predicted using FEA analysis. The analysis was validated using experimental data. Both the FEA and experimental results showed a similar trend in the stress and strain plot. Furthermore, the orientation of the fibers and damage locations need to be further studied to obtain a better understanding and develop an accurate analytical model.

Author Contributions

Conceptualization, S.G.N. and I.K.; methodology, S.G.N.; software, S.G.N.; validation, S.G.N., A.B.K. and I.K.; formal analysis, S.G.N. and I.K.; investigation, S.G.N.; resources, I.K.; data curation, S.G.N.; writing—original draft preparation, S.G.N.; writing—review and editing, I.K., A.B.K., V.I.P. and H.A.-N.; visualization, S.G.N.; supervision, I.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roy, H.; Pahlevani, F.; Cholake, S.; Echeverria, C.; Banerjee, A.; Sahajwalla, V. Simulation of marine bio-composite using empirical data combined with finite element technique. J. Compos. Sci. 2018, 2, 48. [Google Scholar] [CrossRef]
  2. Echeverria, C.; Pahlevani, F.; Gaikwad, V.; Sahajwalla, V. The effect of microstructure, filler load and surface adhesion of marine bio-fillers, in the performance of Hybrid Wood-Polypropylene Particulate Bio-composite. J. Clean. Prod. 2017, 154, 284–294. [Google Scholar] [CrossRef]
  3. Ashori, A.; Matini Behzad, H.; Tarmian, A. Effects of chemical preservative treatments on durability of wood flour/HDPE composites. Compos. Part B Eng. 2013, 47, 308–313. [Google Scholar] [CrossRef]
  4. Chan, C.M.; Vandi, L.J.; Pratt, S.; Halley, P.; Richardson, D.; Werker, A.; Laycock, B. Composites of Wood and Biodegradable Thermoplastics: A Review. Polym. Rev. 2018, 58, 444–494. [Google Scholar] [CrossRef]
  5. Kalali, E.N.; Zhang, L.; Shabestari, M.E.; Croyal, J.; Wang, D.-Y.Y.; Naderikalali, E.; Zhang, L.; Entezar, M.; Croyal, J.; Wang, D.-Y.Y. Flame-retardant wood polymer composites (WPCs) as potential fire safe bio-based materials for building products: Preparation, flammability and mechanical properties. Fire Saf. J. 2017, 107, 210–216. [Google Scholar] [CrossRef]
  6. Kakarla, A.B.; Nukala, S.G.; Kong, I. Biodegradable materials. In Materials for Lightweight Constructions; CRC Press: Boca Raton, FL, USA, 2022; pp. 161–190. [Google Scholar]
  7. Oksman, K. Mechanical Properties of Natural Fibre Mat Reinforced Thermoplastic. Appl. Compos. Mater. 2000, 7, 403–414. [Google Scholar] [CrossRef]
  8. Väntsi, O.; Kärki, T. Environmental assessment of recycled mineral wool and polypropylene utilised in wood polymer composites. Resour. Conserv. Recycl. 2015, 104, 38–48. [Google Scholar] [CrossRef]
  9. Jubinville, D.; Esmizadeh, E.; Tzoganakis, C.; Mekonnen, T. Thermo-mechanical recycling of polypropylene for the facile and scalable fabrication of highly loaded wood plastic composites. Compos. Part B Eng. 2021, 219, 108873. [Google Scholar] [CrossRef]
  10. Moreno, D.D.P.; de Camargo, R.V.; dos Santos Luiz, D.; Branco, L.T.P.; Grillo, C.C.; Saron, C. Composites of Recycled Polypropylene from Cotton Swab Waste with Pyrolyzed Rice Husk. J. Polym. Environ. 2021, 29, 350–362. [Google Scholar] [CrossRef]
  11. Burgada, F.; Fages, E.; Quiles-Carrillo, L.; Lascano, D.; Ivorra-Martinez, J.; Arrieta, M.P.; Fenollar, O. Upgrading Recycled Polypropylene from Textile Wastes in Wood Plastic Composites with Short Hemp Fiber. Polymers 2021, 13, 1248. [Google Scholar] [CrossRef]
  12. Rahman, K.S.; Islam, M.N.; Rahman, M.M.; Hannan, M.O.; Dungani, R.; Khalil, H.P.S.A. Flat-pressed wood plastic composites from sawdust and recycled polyethylene terephthalate (PET): Physical and mechanical properties. Springerplus 2013, 2, 629. [Google Scholar] [CrossRef]
  13. Mo, X.; Zhang, X.; Fang, L.; Zhang, Y. Research Progress of Wood-Based Panels Made of Thermoplastics as Wood Adhesives. Polymers 2021, 14, 98. [Google Scholar] [CrossRef]
  14. Gill, Y.Q.; Abid, U.; Irfan, M.S.; Saeed, F.; Shakoor, A.; Firdaus, A. Fabrication, Characterisation, and Machining of Polypropylene/Wood Flour Composites. Arab. J. Sci. Eng. 2022, 47, 5973–5983. [Google Scholar] [CrossRef]
  15. Diouf, P.M.; Thiandoume, C.; Abdulrahman, S.T.; Ndour, O.; Jibin, K.P.; Maria, H.J.; Thomas, S.; Tidjani, A. Mechanical and rheological properties of recycled high-density polyethylene and ronier palm leaf fiber based biocomposites. J. Appl. Polym. Sci. 2022, 139, 51713. [Google Scholar] [CrossRef]
  16. Ilyas, R.A.; Zuhri, M.Y.M.; Aisyah, H.A.; Asyraf, M.R.M.; Hassan, S.A.; Zainudin, E.S.; Sapuan, S.M.; Sharma, S.; Bangar, S.P.; Jumaidin, R.; et al. Natural Fiber-Reinforced Polylactic Acid, Polylactic Acid Blends and Their Composites for Advanced Applications. Polymers 2022, 14, 202. [Google Scholar] [CrossRef]
  17. Mirowski, J.; Oliwa, R.; Oleksy, M.; Tomaszewska, J.; Ryszkowska, J.; Budzik, G. Poly(vinyl chloride) Composites with Raspberry Pomace Filler. Polymers 2021, 13, 1079. [Google Scholar] [CrossRef]
  18. Mirski, R.; Dukarska, D.; Walkiewicz, J.; Derkowski, A. Waste Wood Particles from Primary Wood Processing as a Filler of Insulation PUR Foams. Materials 2021, 14, 4781. [Google Scholar] [CrossRef]
  19. Sanvezzo, P.B.; Branciforti, M.C. Recycling of industrial waste based on jute fiber-polypropylene: Manufacture of sustainable fiber-reinforced polymer composites and their characterisation before and after accelerated aging. Ind. Crops Prod. 2021, 168, 113568. [Google Scholar] [CrossRef]
  20. Da Ferreira, E.S.B.; Luna, C.B.B.; Araújo, E.M.; Siqueira, D.D.; Wellen, R.M.R. Polypropylene/wood powder composites: Evaluation of PP viscosity in thermal, mechanical, thermomechanical, and morphological characters. J. Thermoplast. Compos. Mater. 2022, 35, 71–92. [Google Scholar] [CrossRef]
  21. Venkatesh, G.S.; Deb, A.; Karmarkar, A.; Gurumoorthy, B. Eco-Friendly Wood Polymer Composites for Sustainable Design Applications. In CIRP Design 2012; Springer: London, UK, 2013; pp. 399–408. [Google Scholar]
  22. Petchwattana, N.; Covavisaruch, S.; Sanetuntikul, J. Recycling of wood–plastic composites prepared from poly(vinyl chloride) and wood flour. Constr. Build. Mater. 2012, 28, 557–560. [Google Scholar] [CrossRef]
  23. Ayrilmis, N.; Jarusombuti, S.; Fueangvivat, V.; Bauchongkol, P. Effect of thermal-treatment of wood fibres on properties of flat-pressed wood plastic composites. Polym. Degrad. Stab. 2011, 96, 818–822. [Google Scholar] [CrossRef]
  24. Ameen, M. Boundary Element Analysis: Theory and Programming; CRC Press: Boca Raton, FL, USA, 2001; ISBN 0849310016. [Google Scholar]
  25. Dickson, A.N.; Barry, J.N.; McDonnell, K.A.; Dowling, D.P. Fabrication of continuous carbon, glass and Kevlar fibre reinforced polymer composites using additive manufacturing. Addit. Manuf. 2017, 16, 146–152. [Google Scholar] [CrossRef]
  26. Caminero, M.A.; Chacón, J.M.; García-Moreno, I.; Rodríguez, G.P. Impact damage resistance of 3D printed continuous fibre reinforced thermoplastic composites using fused deposition modelling. Compos. Part B Eng. 2018, 148, 93–103. [Google Scholar] [CrossRef]
  27. Alharbi, M.; Kong, I.; Patel, V.I. Simulation of uniaxial stress–strain response of 3D-printed polylactic acid by nonlinear finite element analysis. Appl. Adhes. Sci. 2020, 8, 5. [Google Scholar] [CrossRef]
  28. Bhandari, S.; Lopez-Anido, R. Finite element analysis of thermoplastic polymer extrusion 3D printed material for mechanical property prediction. Addit. Manuf. 2018, 22, 187–196. [Google Scholar] [CrossRef]
  29. Mothilal, T.; Ragothaman, G.; Manuel, D.J.; Socrates, S.; Mathavan, S.; Bhaskar, K.; Jayabalakrishnan, D.; Vinoth Kumar, M.; Sendilvelan, S.; Prabhahar, M. Analysis on mechanical properties of wood plastic composite. Mater. Today Proc. 2020, 45, 5886–5891. [Google Scholar] [CrossRef]
  30. Ezzaraa, I.; Ayrilmis, N.; Abouelmajd, M.; Kuzman, M.K.; Bahlaoui, A.; Arroub, I.; Bengourram, J.; Lagache, M.; Belhouideg, S. Numerical Modeling Based on Finite Element Analysis of 3D-Printed Wood-Polylactic Acid Composites: A Comparison with Experimental Data. Forests 2023, 14, 95. [Google Scholar] [CrossRef]
  31. Hartmann, R.; Puch, F. Numerical Simulation of the Deformation Behavior of Softwood Tracheids for the Calculation of the Mechanical Properties of Wood–Polymer Composites. Polymers 2022, 14, 2574. [Google Scholar] [CrossRef]
  32. Nukala, S.G.; Kong, I.; Kakarla, A.B.; Kong, W.; Kong, W. Development of Wood Polymer Composites from Recycled Wood and Plastic Waste: Thermal and Mechanical Properties. J. Compos. Sci. 2022, 6, 194. [Google Scholar] [CrossRef]
  33. Medupin, R. Mechanical Properties of Wood Waste Reinforced Polymer Matrix Composites. Am. Chem. Sci. J. 2013, 3, 507–513. [Google Scholar] [CrossRef]
  34. Nukala, S.G.; Kong, I.; Kakarla, A.B.; Tshai, K.Y.; Kong, W. Preparation and Characterisation of Wood Polymer Composites Using Sustainable Raw Materials. Polymers 2022, 14, 3183. [Google Scholar] [CrossRef]
  35. Najafi, S.K.; Hamidinia, E.; Tajvidi, M. Mechanical properties of composites from sawdust and recycled plastics. J. Appl. Polym. Sci. 2006, 100, 3641–3645. [Google Scholar] [CrossRef]
  36. American Society for Testing and Materials. Standard Test Method for Tensile Properties of Plastics 1; American Society for Testing and Materials: West Conshohocken, PA, USA, 2006. [Google Scholar]
  37. Reddy, G.; Krishna, V.; Shanker, K. Tensile and Water Absorption Properties of FRP Composite Laminates without Voids and with Voids. Procedia Eng. 2017, 173, 1684–1691. [Google Scholar] [CrossRef]
  38. Sood, A.; Ramarao, S.; Carounanidy, U. Influence of different crosshead speeds on diametral tensile strength of a methacrylate based resin composite: An in-vitro study. J. Conserv. Dent. 2015, 18, 214–217. [Google Scholar] [CrossRef]
  39. Kim, D.B.; Lee, G.T.; Lee, I.H.; Cho, H.Y. Finite Element Analysis for Fracture Criterion of PolyJet Materials. J. Korean Soc. Manuf. Process Eng. 2015, 14, 134–139. [Google Scholar] [CrossRef]
  40. Nukala, S.G.; Kong, I.; Patel, V.I.; Kakarla, A.B.; Kong, W.; Buddrick, O. Development of Biodegradable Composites Using Polycaprolactone and Bamboo Powder. Polymers 2022, 14, 4169. [Google Scholar] [CrossRef]
  41. Ratanawilai, T.; Taneerat, K. Alternative polymeric matrices for wood-plastic composites: Effects on mechanical properties and resistance to natural weathering. Constr. Build. Mater. 2018, 172, 349–357. [Google Scholar] [CrossRef]
  42. Dhal, J.P.; Mishra, S.C. Processing and Properties of Natural Fiber-Reinforced Polymer Composite. J. Mater. 2013, 2013, 297213. [Google Scholar] [CrossRef]
  43. Waqas, H.M.; Shi, D.; Imran, M.; Khan, S.Z.; Fathallah, E.; Helal, M. Design Optimization Of Composite Wood Sandwiched Submersible Pressure Hulls. J. Appl. Sci. Eng. 2022, 26, 1295–1304. [Google Scholar] [CrossRef]
  44. Fracz, W.; Janowski, G. Strength Analysis of Molded Pieces Produced from Wood-Polymer Composites (Wpc) Including Their Complex Structures. Compos. Theory Pract. 2016, 16, 260–265. [Google Scholar]
  45. Nithiyakumar, M.; Gopalakrishnan, D. Development and analysis of jute and coir reinforced composites. Textile 2007, 2348. [Google Scholar]
  46. Abbas, S.J.; Ali, M.M.; Al-Mosawi, A.I. Using of Ansys Program to Calculate the Mechanical Properties of Advanced Fibers Reinforced Composite. Iraqi J. Mech. Mater. Eng. 2012, 12, 673–679. [Google Scholar]
  47. Mohapatra, R.; Mishra, A.; Choudhury, B. Determination of thermal conductivity of pine wood dust filled epoxy composites. Therm. Sci. 2017, 21, 199–210. [Google Scholar] [CrossRef]
  48. Stasiak, M.; Molenda, M.; Bańda, M.; Gondek, E. Mechanical properties of sawdust and woodchips. Fuel 2015, 159, 900–908. [Google Scholar] [CrossRef]
  49. Soury, E.; Behravesh, A.H.; Rouhani Esfahani, E.; Zolfaghari, A. Design, optimisation and manufacturing of wood-plastic composite pallet. Mater. Des. 2009, 30, 4183–4191. [Google Scholar] [CrossRef]
  50. Horabik, J.; Bańda, M.; Józefaciuk, G.; Adamczuk, A.; Polakowski, C.; Stasiak, M.; Parafiniuk, P.; Wiącek, J.; Kobyłka, R.; Molenda, M. Breakage strength of wood sawdust pellets: Measurements and modelling. Materials 2021, 14, 3273. [Google Scholar] [CrossRef]
  51. Kakarla, A.B.; Kong, I.; Nukala, S.G.; Kong, W. Mechanical Behaviour Evaluation of Porous Scaffold for Tissue-Engineering Applications Using Finite Element Analysis. J. Compos. Sci. 2022, 6, 46. [Google Scholar] [CrossRef]
  52. Faridmehr, I.; Hanim Osman, M.; Bin Adnan, A.; Farokhi Nejad, A.; Hodjati, R.; Amin Azimi, M. Correlation between Engineering Stress-Strain and True Stress-Strain Curve. Am. J. Civ. Eng. Archit. 2014, 2, 53–59. [Google Scholar] [CrossRef]
  53. Adhikary, K.B.; Pang, S.; Staiger, M.P. Dimensional stability and mechanical behaviour of wood–plastic composites based on recycled and virgin high-density polyethylene (HDPE). Compos. Part B Eng. 2008, 39, 807–815. [Google Scholar] [CrossRef]
  54. Renner, K.; Móczó, J.; Pukánszky, B. Deformation and failure of PP composites reinforced with lignocellulosic fibers: Effect of inherent strength of the particles. Compos. Sci. Technol. 2009, 69, 1653–1659. [Google Scholar] [CrossRef]
  55. Barbero, E.J.; Shahbazi, M. Determination of material properties for ANSYS progressive damage analysis of laminated composites. Compos. Struct. 2017, 176, 768–779. [Google Scholar] [CrossRef]
  56. Martinez, X.; Oller, S. Numerical Simulation of Matrix Reinforced Composite Materials Subjected to Compression Loads. Arch. Comput. Methods Eng. 2009, 16, 357–397. [Google Scholar] [CrossRef]
  57. Oliver, J. A consistent characteristic length for smeared cracking models. Int. J. Numer. Methods Eng. 1989, 28, 461–474. [Google Scholar] [CrossRef]
  58. Frey, P.J. Generation and adaptation of computational surface meshes from discrete anatomical data. Int. J. Numer. Methods Eng. 2004, 60, 1049–1074. [Google Scholar] [CrossRef]
  59. Martinez, X.; Oller, S.; Barbu, L.G.; Barbat, A.H.; de Jesus, A.M.P. Analysis of Ultra Low Cycle Fatigue problems with the Barcelona plastic damage model and a new isotropic hardening law. Int. J. Fatigue 2015, 73, 132–142. [Google Scholar] [CrossRef]
  60. Burkhart, T.A.; Andrews, D.M.; Dunning, C.E. Finite element modeling mesh quality, energy balance and validation methods: A review with recommendations associated with the modeling of bone tissue. J. Biomech. 2013, 46, 1477–1488. [Google Scholar] [CrossRef]
  61. Zambaldi, E.; Magalhães, R.R.; Dias, M.C.; Mendes, L.M.; Tonoli, G.H.D. Numerical simulation of poly(lactic acid) polymeric composites reinforced with nanofibrillated cellulose for industrial applications. Polym. Eng. Sci. 2022, 62, 4043–4054. [Google Scholar] [CrossRef]
  62. Charupeng, N.; Kunthong, P. A novel finite element algorithm for predicting the elastic properties of wood fibers. Int. J. Comput. Mater. Sci. Eng. 2022, 11, 2150027. [Google Scholar] [CrossRef]
  63. Guo, Z.; Shi, X.; Chen, Y.; Chen, H.; Peng, X.; Harrison, P. Mechanical modeling of incompressible particle-reinforced neo-Hookean composites based on numerical homogenisation. Mech. Mater. 2014, 70, 17. [Google Scholar] [CrossRef]
  64. Bucki, M.; Lobos, C.; Payan, Y.; Hitschfeld, N. Jacobian-based repair method for finite element meshes after registration. Eng. Comput. 2011, 27, 285–297. [Google Scholar] [CrossRef]
  65. Bucki, M.; Lobos, C.; Payan, Y. A fast and robust patient specific Finite Element mesh registration technique: Application to 60 clinical cases. Med. Image Anal. 2010, 14, 303–317. [Google Scholar] [CrossRef] [PubMed]
  66. Madeo, A.; Casciaro, R.; Zagari, G.; Zinno, R.; Zucco, G. A mixed isostatic 16 dof quadrilateral membrane element with drilling rotations, based on Airy stresses. Finite Elem. Anal. Des. 2014, 89, 52–66. [Google Scholar] [CrossRef]
  67. Tiwari, S.K.; Umamaheswara Rao, A.; Reddy, N.; Sharma, H.; Pandey, J.K. Synthesis, characterisation and finite element analysis of polypropylene composite reinforced by jute and carbon fiber. Mater. Today Proc. 2021, 46, 10884–10891. [Google Scholar] [CrossRef]
  68. Das Lala, S.; Sadikbasha, S.; Deoghare, A.B. Prediction of elastic modulus of polymer composites using Hashin–Shtrikman bound, mean field homogenisation and finite element technique. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2020, 234, 1653–1659. [Google Scholar] [CrossRef]
  69. Navaneethakrishnan, G.; Karthikeyan, T.; Saravanan, S.; Selvam, V.; Parkunam, N.; Sathishkumar, G.; Jayakrishnan, S. Structural analysis of natural fiber reinforced polymer matrix composite. Mater. Today Proc. 2020, 21, 7–9. [Google Scholar] [CrossRef]
  70. Shih, Y.-F.; Huang, C.-C.; Chen, P.-W. Biodegradable green composites reinforced by the fiber recycling from disposable chopsticks. Mater. Sci. Eng. A 2010, 527, 1516–1521. [Google Scholar] [CrossRef]
  71. Satyanarayana, K.G.; Sukumaran, K.; Mukherjee, P.S.; Pavithran, C.; Pillai, S.G.K. Natural fibre-polymer composites. Cem. Concr. Compos. 1990, 12, 117–136. [Google Scholar] [CrossRef]
  72. Fang, H.; Sun, H.; Liu, W.; Wang, L.; Bai, Y.; Hui, D. Mechanical performance of innovative GFRP-bamboo-wood sandwich beams: Experimental and modelling investigation. Compos. Part B Eng. 2015, 79, 182–196. [Google Scholar] [CrossRef]
  73. Appadurai, M.; Raj, E.F.I.; LurthuPushparaj, T. Sisal fiber-reinforced polymer composite-based small horizontal axis wind turbine suited for urban applications—A numerical study. Emergent Mater. 2022, 5, 565–578. [Google Scholar] [CrossRef]
  74. Haftkhani, A.R.; Chavooshi, A.; Arabi, M. Simulation and prediction of bending performance of wood plastic composite profiles using finite element method by ANSYS Workbench software. Iran. J. Wood Pap. Ind. 2022, 13, 325–344. [Google Scholar]
Figure 1. (a) Boundary conditions of the specimen and (b) stacking sequence of the specimen.
Figure 1. (a) Boundary conditions of the specimen and (b) stacking sequence of the specimen.
Polymers 15 01977 g001
Figure 2. SEM micrographs of WPC samples at cross-sectional failure depicting the random distribution of SD for (a) rPP-SD3 and (b) rPP-SD4.
Figure 2. SEM micrographs of WPC samples at cross-sectional failure depicting the random distribution of SD for (a) rPP-SD3 and (b) rPP-SD4.
Polymers 15 01977 g002
Figure 3. (a) Mesh generated in ANSYS, (b) mesh aspect ratio of WPCs and (c) Jacobian ratio of WPCs.
Figure 3. (a) Mesh generated in ANSYS, (b) mesh aspect ratio of WPCs and (c) Jacobian ratio of WPCs.
Polymers 15 01977 g003
Figure 4. (a) Equivalent stress distribution and (b) equivalent strain distribution.
Figure 4. (a) Equivalent stress distribution and (b) equivalent strain distribution.
Polymers 15 01977 g004
Figure 5. Stress–strain curve from the FEA and experimental data.
Figure 5. Stress–strain curve from the FEA and experimental data.
Polymers 15 01977 g005
Figure 6. Total deformation in the (a) FEA, (b) rPP-SD3 and (c) rPP-SD4.
Figure 6. Total deformation in the (a) FEA, (b) rPP-SD3 and (c) rPP-SD4.
Polymers 15 01977 g006
Table 1. Composition of rPP-SD composites.
Table 1. Composition of rPP-SD composites.
CompositesrPP (wt.%)SD (wt.%)
rPP-SD37030
rPP-SD46040
Table 2. Properties of recycled polypropylene and sawdust.
Table 2. Properties of recycled polypropylene and sawdust.
PropertiesSDrPP
Young’s modulus (MPa)0.30 [48]6 [49]
Poisson’s ratio0.33 [48]0.38 [49]
Ultimate tensile strength (MPa)0.5 [50]21 [49]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nukala, S.G.; Kong, I.; Kakarla, A.B.; Patel, V.I.; Abuel-Naga, H. Simulation of Wood Polymer Composites with Finite Element Analysis. Polymers 2023, 15, 1977. https://doi.org/10.3390/polym15091977

AMA Style

Nukala SG, Kong I, Kakarla AB, Patel VI, Abuel-Naga H. Simulation of Wood Polymer Composites with Finite Element Analysis. Polymers. 2023; 15(9):1977. https://doi.org/10.3390/polym15091977

Chicago/Turabian Style

Nukala, Satya Guha, Ing Kong, Akesh Babu Kakarla, Vipulkumar Ishvarbhai Patel, and Hossam Abuel-Naga. 2023. "Simulation of Wood Polymer Composites with Finite Element Analysis" Polymers 15, no. 9: 1977. https://doi.org/10.3390/polym15091977

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