Next Article in Journal
Comparison of Coronal Discoloration Induced by White MTA and CEM Cement
Next Article in Special Issue
Mesoscale Analysis of Rubber Particle Effect on Indirect Tensile and Flexural Tensile Strength of Crumb Rubber Mortar
Previous Article in Journal
Dielectric Spectroscopy of Melt-Mixed Polypropylene and Pyrolytically Stripped Carbon Nanofiber Composites
Previous Article in Special Issue
Finite Element Analysis for Nonlinear Unbonded Circular Fiber-Reinforced Elastomeric Bearings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Natural Rubber Composites Reinforced with Green Silica from Rice Husk: Effect of Filler Loading on Mechanical Properties

by
Nicha Choophun
1,
Nattapat Chaiammart
1,4,
Kantavee Sukthavon
1,
Chatchai Veranitisagul
2,
Apirat Laobuthee
1,
Anyarat Watthanaphanit
3 and
Gasidit Panomsuwan
1,4,*
1
Department of Materials Engineering, Faculty of Engineering, Kasetsart University, Bangkok 10900, Thailand
2
Department of Materials and Metallurgical Engineering, Faculty of Engineering, Rajamangala University of Technology Thanyaburi, Khlong Hok 12110, Thailand
3
Department of Chemistry, Faculty of Science, Mahidol University, Nakhon Pathom 73170, Thailand
4
Special Research Unit for Biomass Conversion Technology for Energy and Environmental Materials, Kasetsart University, Bangkok 10900, Thailand
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2022, 6(12), 369; https://doi.org/10.3390/jcs6120369
Submission received: 21 October 2022 / Revised: 17 November 2022 / Accepted: 29 November 2022 / Published: 5 December 2022
(This article belongs to the Special Issue Progress in Rubber Blends and Composites Technology)

Abstract

:
Natural rubber (NR) composites filled with silica are typically used for tire tread applications owing to their low energy consumption and low rolling resistance. Tire tread properties vary broadly depending on the compound formulation and curing conditions. Silica loading is recognized as a critical factor influencing the mechanical properties of the composites. In this work, we aim to investigate the effect of silica loading (10–50 phr) on the mechanical properties of NR composites. Silica was prepared from rice husk waste via chemical treatment and subsequent calcination at 600 °C. Prior to the compound mixing process, silica was modified by a silane coupling agent to improve compatibility with the NR matrix. The NR compounds reinforced with silane-modified silica from rice husk were prepared using a two-roll mill machine. The scorch and cure times increased as the silica loading increased. The mechanical properties of the NR composites, including tensile strength, elongation at break, modulus, hardness, and abrasion loss, were examined as a function of silica loading. Tensile strength increased and reached the maximum value at 20 phr but decreased at high loading owing to the agglomeration of silica in the NR matrix. With increasing silica loading, hardness and modulus increased, whereas elongation at break and abrasion resistance decreased slightly. These results indicate that NR composites filled with silica are stiffer and harder at a higher silica loading due to the strong interaction between silica and the NR matrix, inhibiting the segmental mobility of rubber chains. We anticipate that the compound formulation presented in this work could potentially be adapted to tire tread applications.

Graphical Abstract

1. Introduction

The tire industry has grown continuously worldwide, with increased global demand every year [1]. Tire tread is one of the most important parts of tires and determines the performance of vehicles since it directly contacts the pavement surface for a long operation time under a heavy load [2,3]. Therefore, tire tread is easily worn out and damaged compared to other parts. Moreover, tire tread–pavement interaction creates a frictional force called skid resistance, which is an important factor that limits vehicle speed and the design of pavement geometry for driving safety [4,5,6,7]. To meet industrial requirements and ensure durability based on safety standards, further significant development of tire tread with a low rolling resistance coefficient, high traction, and good wear resistance is necessary [8,9,10].
In general, natural rubber (NR) is used as the major constituent in tread compounds owing to its low heat build-up level and good mechanical strength [11,12]. Tire tread compounds consist of several components depending on the formulation, in which reinforcement filler is an essential component that helps to improve their mechanical properties [13,14,15]. The primary reinforcing fillers in the rubber industry are carbon black (CB) and silica (SiO2) [16,17,18]. Each filler possesses its own advantages. CB provides a greater modulus and hardness, improving the rubber’s resistance to tearing, abrasion, and flexing fatigue, than silica. However, its rolling resistance and wet traction are constrained. Besides, manufacturing CB by incomplete combustion of heavy petroleum products produces toxic gases and hazardous waste, causing long-term negative effects to human health and environmental problems [19,20]. Silica, with a similar specific surface area and loading to CB, provides less reinforcing power than CB because of its weaker polymer–filler interaction. Mixing silica and rubber compounds without compatibilizers could result in poor dispersion and agglomeration since silica (polar and hydrophilic surface) is incompatible with non-polar rubber [21]. Nonetheless, silica improves the processability of rubber and yields lower rolling resistance, resulting in reduced fuel consumption and low carbon emissions. Moreover, it can be derived from renewable bioresources, such as rice straw [22], rice husk [23], corn cob [24], sugarcane bagasse [25], and bamboo leaf [26]. Therefore, silica could be used as a partial or complete replacement for CB fillers. Replacing CB fillers entirely with silica is enchanting; silica is attractive since silica is an eco-friendly renewable feedstock that can be used to produce fuel-saving tires [27,28]. Nonetheless, surface modification of silica before mixing is required to enhance the compatibility between silica and the rubber matrix. Silane coupling agents are usually used to achieve this aim [29,30]. Despite significant progress in silica-reinforced NR composites, the optimal silica loading to achieve the desired mechanical properties varies broadly depending on the silica properties, mixing procedure, and formulation [31,32,33,34,35].
In this work, silica was extracted from rice husk via a chemical treatment and subsequent calcination process. Rice husk was chosen herein as a bio-resource of silica due to the vast amount of rice husk waste in Thailand. Furthermore, among several bioresources, rice husk contains the greatest silica content [36,37]. Prior to mixing with NR to make a composite, silica derived from rice husk was modified by a silane coupling agent to enhance its compatibility with NR. Silica loading in the composites varied from 10 to 50 phr, while the other components were fixed. The effect of silica loading on the mechanical properties of the NR composites is investigated and discussed in this study.

2. Experimental

2.1. Materials and Chemicals

Rice husk was obtained from a rice mill in Singburi province, Thailand. Concentrated hydrochloric acid (37%, HCl) and ethanol (CH3OH, purity > 99.5%) were purchased from RCI Labscan Ltd. (Bangkok, Thailand). Potassium bromide (KBr) was purchased from PIKE technologies. Bis [3-(triethoxysilyl) propyl] tetrasulfide (C18H42O6S4Si2, Si69, purity > 90%) as a silane coupling agent was purchased from Sigma Aldrich. Natural rubber (STR20L grade) was purchased from Thai Hua Rubber Company. Deionized (DI) water was employed throughout all experiments.

2.2. Preparation of Silica from Rice Husk

Rice husks were first washed twice with tap water and dried in an oven at 105 °C for 24 h. The dried rice husks were treated with a 1 M HCl solution at 80 °C for 1 h and then washed with DI water until the washed water was neutral. The HCl-treated rice husks were dried in an oven at 105 °C for 24 h. After that, they were loaded in a ceramic tray, calcined in a CWF 11/13 Carbolite chamber furnace at 600 °C for 1 h with a heating rate of 5 °C min–1, and naturally cooled to room temperature under an air atmosphere. The white product after calcination was obtained and is denoted as RHS. The RHS was thoroughly ground with an agate and mortar and then sieved through a 100-μm mesh sieve. The preparation process of silica from rice husk is schematically depicted in Figure 1a.

2.3. Preparation of Silane-Modified Silica

A certain amount of Si69 (5.45 mL) was mixed with ethanol (800 mL) and stirred for 30 min to obtain a homogeneous solution. The RHS powders (100 g) were soaked in the ethanol mixture containing Si69 and stirred for 60 min. The ethanol was then evaporated at 80 °C for 12 h in an oven, resulting in the Si69-modified RHS (Si69-RHS) with a light yellow color. The overall preparation process of silane-modified silica is schematically shown in Figure 1b. The Si69-RHS was further used as reinforcement fillers in the NR composites.

2.4. Characterization of RHS and Si69-RHS

The morphology was investigated using a Hitachi SU3500 scanning electron microscope at an acceleration voltage of 10 kV. The energy-dispersive spectroscopy (EDS) analysis was undertaken with an Oxford Instruments Aztec One spectrometer equipped with a scanning electron microscope. The samples were coated with a thin gold film by sputtering before the investigation. Fourier transform infrared (FTIR) spectra were acquired on a Bruker Alpha-E spectrophotometer in the wavenumber range of 500–4000 cm–1. The samples were mixed with KBr and pressed into a pellet form for measurement. Thermogravimetric analysis (TGA) was conducted using a TGA 2 Mettler Toledo Mettler in the temperature range of 50–900 °C at a heating rate of 10 °C min–1 under nitrogen flow. The specific surface area was measured on a Micromeritics 3Flex surface characterization analyzer at liquid N2 temperature (−196 °C). Before the measurement, silica was heated at 100 °C for 6 h under a vacuum. The specific surface area was calculated from the adsorption isotherm in the relative pressure range of 0.05–0.30 using the Brunauer–Emmett–Teller (BET) theory.

2.5. Preparation of NR Compounds Filled with Silica

The formulations of NR compounds filled with Si69-RHS are shown in Table 1. The NR compound formulae are designated as C-x, where x represents the Si69-RHS loading. The mixing process was performed using a conventional laboratory two-roll mill (YFTR 8″). The obtained NR compounds were stored at 25 °C for 24 h in a closed container prior to curing assessment using an M-2000 moving die rheometer (MDR). The testing procedure was carried out according to the method described in the American Society for Testing and Materials (ASTM) D5289. The respective scorch time (t2) and cure time (tc90) were obtained from an MDR at 150 °C.

2.6. Preparation of the NR Composites and Testing

The NR compounds were compressed using a compression-molding machine (Yong-Fong Machinery, Ltd., Model YFY HM-100, Samut Sakorn, Thailand) at 150 °C according to their respective cure times to obtain vulcanized NR composites (Figure 2). The fractured surfaces of the vulcanized NR compounds were investigated by a scanning electron microscope at an acceleration voltage of 10 kV. The dispersion of Si69-RHS in the NR matrix was examined by EDS mapping analysis. The NR composites were cut into a specific shape according to the ASTM for each mechanical test as follows: The stress–strain curves of the composites were measured on a Hounsfield H50KS testing machine according to ASTM D412 at a crosshead speed of 500 mm min–1; hardness was measured using a Shore type A durometer according to ASTM D2240; and abrasion resistance was tested using a QC-615A AKRON abrasion test machine according to ASTM D2228. All mechanical values were obtained from the average of 5 samples.

3. Results and Discussion

3.1. Characterization of RHS and Si69-RHS

Figure 3a,c displays the SEM images of RHS and Si69-RHS, respectively. These images reveal irregular-shaped particles broken into a micron scale of less than 100 μm. No significant morphological change is observed between RHS and Si69-RHS, indicating no effect of surface modification on the RHS morphology. The EDS analysis was conducted to confirm the presence of Si69 on RHS by measuring the elemental composition in the sample. From the EDS spectra (Figure 3b,d), only silicon (Si K) and oxygen (O K) signals were detected for the RHS, while an additional small signal of sulfur (S K) was found for Si69-RHS. This result confirms the presence of Si69 on RHS.
The FTIR spectra of RHS and Si69-RHS shown in Figure 4 reveal similar absorption features. A broad peak covering approximately 3200–3600 cm–1 and the peak at 1650 cm–1 are attributed to O–H stretching and O–H bending, respectively, indicating the presence of water molecules. The absorption at 1096 and 810 cm–1 can be assigned to the Si–O–Si stretching bond, which is the characteristic peak of silica [38]. However, the O–H stretching and bending peaks of Si69-RHS became weaker, while a very small peak of the C–H stretching bond from Si69 appeared at 2995 cm–1 (shown in the inset) [39,40,41]. The suppression of the O–H peak and the emergence of the C–H peak of Si69-RHS indicate its greater hydrophobicity and reduced water absorption [40].
To further confirm the presence of Si69 on RHS, TGA thermograms were recorded and are depicted in Figure 5. For the first stage (50–150 °C), the RHS exhibited a significant weight loss of approximately 7%, while a weight loss of less than 2% was observed for Si69-RHS. The lower weight loss of Si69-RHS compared to RHS was due to less water absorption on Si69-RHS caused by its hydrophobic surface, as discussed in the FTIR analysis [40]. With an increase in temperature, the weight of RHS slightly decreased from 92.8% at 150 °C to 91.4% at 900 °C. This small weight loss of RHS was due to the almost complete decomposition of organic compounds during the calcination process. In contrast, a noticeable weight loss of approximately 4.2% was observed for Si69-RHS with a temperature range of 150–700 °C, which was attributed to the decomposition of Si69 [42,43]. According to the EDS, FTIR, and TGA analyses, the results are in good agreement to confirm that RHS was successfully modified by Si69.

3.2. Characterization of Si69-RHS/NR Composites

The scorch and cure times of the NR compounds filled with various Si69-RHS loadings are shown in Figure 6. Both scorch and cure times tended to be longer with an increase in the Si69-RHS loading. This increasing trend was in agreement with several previous reports in the literature [44,45,46]. The high specific surface area (~280 m2 g–1) of Si69-RHS could result in an increase in interaction with other components in compounds causing a disturbance in the vulcanization process and, thus, delaying scorch and cure times [47].
Figure 7 presents the SEM images taken from the fractured surface of the NR composites. The NR composites without Si69-RHS had a relatively smooth surface (Figure 7a). The fractured surface had no obvious formation of voids and became rougher as the Si69-RHS loading increased (Figure 7b–f). However, it is difficult to identify and evaluate the dispersion of Si69-RHS in the NR matrix from the SEM images. To solve this issue, EDS mapping analysis was conducted by measuring the dispersion of the Si element (blue) along with C (yellow) and O (green) elements, as shown in Figure 8. The Si element revealed uniform dispersion without agglomeration in the NR matrix at the Si69-RHS loadings of 10 and 20 phr. Good dispersion of the Si69-RHS in the NR matrix with good compatibility could help improve the mechanical properties of the composites. However, agglomeration began to be observed when the Si69-RHS loading reached 30 phr and became larger at higher loadings (i.e., 40 and 50 phr).
The mechanical properties of the NR composites filled with Si69-RHS were examined. Tensile strength, modulus, and elongation at break can be determined from the stress–strain curves, as shown in Figure 9a. From Figure 9b, tensile strength increased with increasing Si69-RHS loading until reaching a maximum value of 20.9 ± 1.8 MPa at 20 phr, which was an approximately 26% improvement from that of NR (16.6 ± 1.5 MPa). The increase in tensile strength was attributed to good dispersion of Si69-RHS in the NR matrix and good interfacial interaction between the matrix and fillers, which can facilitate a more efficient energy-transfer process [44,48]. However, a further increase in Si69-RHS loading greater than 20 phr had an adverse effect on the tensile strength of the NR composites. The decrease in tensile strength is caused by the agglomeration of fillers in the NR matrix (>20 phr), as shown in the EDS mapping analysis (Figure 8). The change in tensile strength in this study is similar to the findings of Prasertsri and Rattanasom [44] and da Costa et al. [49]. With an increase in Si69-RHS loading from 0 to 50 phr, the elongation at break gradually decreased from 734% to 599%, as shown in Figure 9c. In contrast, the opposite trend was observed for the modulus at 100% of elongation (100% modulus). The 100% modulus increased from 1.04 ± 0.11 MPa at 0 phr to 1.38 ± 0.08 MPa at 40 phr but slightly dropped to 1.30 ± 0.06 MPa at 50 phr (Figure 9d).
As shown in Figure 10a, the hardness (shore A) of the composites showed a similar trend with the 100% modulus, which continuously increased from 33.2 ± 0.8 to 44.9 ± 0.1 with increased Si69-RHS loading. According to the reports by Jacques et al. [50] and da Costa et al. [49,51], the hardness and modulus of vulcanized NR composites normally increase with the use of particulate fillers. From our results, it is clear that the NR composites became stiffer and harder when reinforcing with the Si69-RHS fillers since the interaction between NR and Si69-RHS could restrict the segmental mobility and movement of rubber chains, which decreased the elasticity of the rubber composites [45,52].
Abrasion resistance is another important property for practical applications of rubber tires since it reflects a tire’s ability to resist the progressive removal of material at the surface from mechanical rubbing and scraping [53]. It was found that a high Si69-RHS loading reduced abrasion resistance, as reflected by the higher value of abrasion loss (Figure 10b). The abrasion loss of the NR composites filled with Si69-RHS loading of 50 phr was 0.24 ± 0.02%, which was about five-times greater than without Si69-RHS (0.06 ± 0.01%). Although the abrasion loss of the NR composites filled with Si69-RHS increased significantly, it showed satisfactory results for practical use. The mechanical properties of the NR composites discussed above are summarized in Table 2.

4. Conclusions

NR composites were successfully prepared at different Si69-RHS loadings from 0 to 50 phr. High Si69-RHS loadings could disturb the NR vulcanization process, thus, prolonging both the scorch and cure times. The Si69-RHS fillers were uniformly dispersed in the NR matrix with good compatibility at low loading (i.e., 10 and 20 phr); however, agglomeration began to be observed at 30 phr. For the mechanical properties of the NR composites, tensile strength reached the maximum value at a Si69-RHS loading of 20 phr and decreased at high loadings (>20 phr) owing to the agglomeration of fillers in the NR composites. With increasing Si69-RHS loading, hardness and 100% modulus tended to increase, whereas the opposite trend was observed for elongation at break and abrasion resistance. These results indicate that the addition of Si69-RHS could inhibit the mobility of the rubber chain, making the composites stiffer and harder. The results presented in this work could be helpful in further optimizing and developing a suitable formula for the NR composites for tire tread applications.

Author Contributions

Conceptualization, G.P. and A.L.; methodology, G.P., A.L. and K.S.; validation, G.P. and A.L.; formal analysis, N.C. (Nicha Choophun), N.C. (Nattapat Chaiammart) and K.S.; investigation, N.C. (Nicha Choophun), N.C. (Nattapat Chaiammart) and K.S.; resources, G.P., A.L. and C.V.; data curation, N.C. (Nicha Choophun), N.C. (Nattapat Chaiammart) and K.S.; writing—original draft preparation, N.C. (Nicha Choophun); writing—review and editing, G.P., A.W. and A.L.; visualization, N.C. (Nicha Choophun); supervision, G.P.; project administration, G.P.; funding acquisition, G.P. All authors have read and agreed to the published version of the manuscript.

Funding

This project was funded by the National Research Council of Thailand (NRCT). The APC was funded by the Kasetsart University Research and Development Institute (KURDI, grant no. FF(KU) 25.64).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to gratefully thank the Faculty of Engineering, Kasetsart University for the research assistant scholarship (no. 63/12/MATE/M.Eng) and the Department of Materials and Metallurgical Engineering, Faculty of Engineering, Rajamangala University of Technology Thanyaburi for the rubber processing equipment and machineries. The authors would also like to thank Jularpar Suthiprapar, a scientist at the Materials Innovation Center (MIC), Kasetsart University, for her help in the SEM and EDS measurements of the NR composites.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bijina, V.; Jandas, P.J.; Joseph, S.; Gopu, J.; Abhitha, K.; John, H. Recent trends in industrial and academic developments of green tyre technology. Polym. Bull. 2022. [Google Scholar] [CrossRef]
  2. Noh, I.; Hong, S.; Kim, J. Tensile test analysis of tire tread rubber’s tendency to resist cracks. Appl. Sci. 2022, 12, 7931. [Google Scholar] [CrossRef]
  3. Hopmann, C.; Limper, A.; Kammer, S.; Schön, M.; Drach, M.; Facklam, M. The role of rubber rheology in tire tread extrusion: A review. High Temp. High Press. 2020, 49, 299–342. [Google Scholar] [CrossRef]
  4. Kumar, A.; Gupta, A. Review of factors controlling skid resistance at tire-pavement interface. Adv. Civ. Eng. 2021, 2021, 2733054. [Google Scholar] [CrossRef]
  5. Peng, J.; Chu, L.; Fwa, T.F. Determination of safe vehicle speeds on wet horizontal pavement curves. Road Mater. Pavement Des. 2021, 22, 2641–2653. [Google Scholar] [CrossRef]
  6. Praticò, F.G.; Giunta, M. Quantifying the effect of present, past and oncoming alignment on the operating speeds of a two-lane rural road. Balt. J. Road Bridge Eng. 2012, 7, 181–190. [Google Scholar] [CrossRef]
  7. Praticò, F.G.; Giunta, M. Speed Distribution on Low-Volume Roads in Italy: From Inferences to Rehabilitation Design Criteria. Transp. Res. Rec. J. Transp. Res. Board 2011, 2203, 79–84. [Google Scholar] [CrossRef]
  8. Shah, V.; Bhaliya, J.; Patel, G.M.; Deshmukh, K. Advances in polymeric nanocomposites for automotive applications: A review. Polym. Adv. Technol. 2022, 33, 3023–3048. [Google Scholar] [CrossRef]
  9. Weng, P.; Tang, Z.; Guo, B. Solving “magic triangle” of tread rubber composites with phosphonium-modified petroleum resin. Polymer 2020, 190, 12244. [Google Scholar] [CrossRef]
  10. Qin, X.; Wang, J.; Han, B.; Wang, B.; Mao, L.; Zhang, L. Novel design of eco-friendly super elastomer materials with optimized hard segments micro-structure: Toward next-generation high-performance tires. Front. Chem. 2018, 6, 240. [Google Scholar] [CrossRef] [PubMed]
  11. Martin, P.J.; Brown, P.; Chapman, A.V.; Cook, S. Silica-reinforced epoxidized natural rubber tire treads—performance and durability. Rubber Chem. Technol. 2015, 88, 390–411. [Google Scholar] [CrossRef]
  12. Bera, A.; Ganguly, D.; Ghorai, S.K.; Rath, J.P.; Ramakrishnan, S.; Kuriakose, J.; Amarnath, S.K.P.; Chattopadhyay, S. Treatment of natural rubber with bio-based components: A green endeavor to diminish the silica agglomeration for tyre tread application. Chem. Eng. J. Adv. 2022, 11, 100349. [Google Scholar] [CrossRef]
  13. Hamed, G.R. Reinforcement of rubber. Rubber Chem. Technol. 2000, 73, 524–533. [Google Scholar] [CrossRef]
  14. Kohl, D.J.; Beaucage, G. Rational design of reinforced rubber. Curr. Opin. Solid State Mater. Sci. 2002, 6, 183–194. [Google Scholar] [CrossRef]
  15. Sarkar, P.; Bhowmick, A.K. Sustainable rubbers and rubber additives. J. Appl. Polym. Sci. 2018, 135, 45701. [Google Scholar] [CrossRef] [Green Version]
  16. Zhang, Y.; Ge, S.; Tang, B.; Koga, T.; Rafailovich, M.H.; Sokolov, J.C.; Peiffer, D.G.; Li, Z.; Dias, A.J.; McElrath, K.O.; et al. Effect of carbon black and silica fillers in elastomer blends. Macromolecules 2001, 34, 7056–7065. [Google Scholar] [CrossRef]
  17. Rattanasom, N.; Saowapark, T.; Deeprasertkul, C. Reinforcement of natural rubber with silica/carbon black hybrid filler. Polym. Test. 2007, 26, 369–377. [Google Scholar] [CrossRef]
  18. Akhmar, Z.; Salim, S.A.; Hassan, A.; Ismail, H. A review on hybrid fillers in rubber composites. Polym. Plast. Technol. Eng. 2018, 57, 523–539. [Google Scholar]
  19. Tsai, P.-J.; Shieh, H.-Y.; Lee, W.-J.; Lai, S.-O. Health-risk assessment for workers exposed to polycyclic aromatic hydrocarbons (PAHs) in a carbon black manufacturing industry. Sci. Total Environ. 2001, 278, 137–150. [Google Scholar] [CrossRef]
  20. Reijnders, L. Cleaner nanotechnology and hazard reduction of manufactured nanoparticles. J. Clean. Prod. 2006, 14, 124–133. [Google Scholar] [CrossRef]
  21. Hussain, M.N.M.; Ali, Z.; Maafa, I.M.M.; Akhter, P.; Majeed, K.; Ahmed, A.; Shehzad, N. Influence of silica materials on synthesis of elastomer nanocomposites: A review. J. Elastomer Plast. 2019, 52, 747–771. [Google Scholar]
  22. Lu, P.; Hsieh, Y.-L. Highly pure amorphous silica nano-disk from rice straw. Powder Technol. 2012, 225, 149–155. [Google Scholar] [CrossRef]
  23. Steven, S.; Restiawaty, E.; Bindar, Y. Routes for energy and bio-silica production from rice husk: A comprehensive review and emerging prospect. Renew. Sustain. Energy Rev. 2021, 149, 111329. [Google Scholar] [CrossRef]
  24. Shim, J.; Velmurugan, P.; Oh, B.-T. Extraction and physical characterization of amorphous silica made from corn cob ash at variable pH conditions via sol-gel processing. J. Ind. Eng. Chem. 2015, 30, 249–253. [Google Scholar] [CrossRef] [Green Version]
  25. Seroka, N.S.; Taziwa, R.T.; Khotseng, L. Extraction and synthesis of silicon nanoparticles (SiNPs) from sugarcane bagasse ash: A mini-review. Appl. Sci. 2022, 12, 2310. [Google Scholar] [CrossRef]
  26. Joshi, U.R.; Roy, R.; Satsangi, P.G. Investigation of bamboo leaves as an alternative source of silica: Extraction, characterization and its application as an adsorbent for methylene blue sequestration. ChemistrySelect 2022, 7, e202200011. [Google Scholar] [CrossRef]
  27. Lolage, M.; Parida, P.; Chaskar, M.; Gupta, A.; Rautaray, D. Green silica: Industrially scalable & sustainable approach towards achieving improved “nanofiller–elastomer” interaction and reinforcement in tire tread compounds. Sustain. Mater. Technol. 2020, 26, e00232. [Google Scholar]
  28. Shoul, B.; Marfavi, Y.; Sadeghi, B.; Kowsari, E.; Sadeghi, P.; Ramakrishna, S. Investigating the potential of sustainable use of green silica in the green tire industry: A review. Environ. Sci. Pollut. Res. 2022, 29, 51298–51317. [Google Scholar] [CrossRef]
  29. Silane Coupling Agents for Silica-Filled Tire-Tread Compounds: The Link between Chemistry and Performance. In Silanes and Other Coupling Agents; ByKash, L.; Mittal, K.L. (Eds.) CRC Press: Boca Raton, FL, USA, 2004; Volume 3. [Google Scholar]
  30. Sarkawi, S.S.; Dierkes, W.K.; Noordermeer, J. Morphology of silica-reinforced natural rubber: The effect of silane coupling agent. Rubber Chem. Technol. 2015, 88, 359–372. [Google Scholar] [CrossRef] [Green Version]
  31. Yatsuyanagi, F.; Suzuki, N.; Ito, M.; Kaidou, H. Effects of surface chemistry of silica particles on the mechanical properties of silica filled styrene–butadiene rubber systems. Polym. J. 2002, 34, 332–339. [Google Scholar] [CrossRef]
  32. Kaewsakul, W.; Sahakaro, K.; Dierkes, W.K.; Noordermeer, J.W.M. Optimization of mixing conditions for silica-reinforced natural rubber tire tread compounds. Rubber Chem. Technol. 2012, 85, 277–294. [Google Scholar] [CrossRef] [Green Version]
  33. Pourhossaini, M.-R.; Razzaghi-Kashani, M. Effect of silica particle size on chain dynamics and frictional properties of styrene butadiene rubber nano and micro composites. Polymer 2014, 55, 2279–2284. [Google Scholar] [CrossRef]
  34. Padmanathan, H.R.; Federico, C.E.; Addiego, F.; Rommei, R.; Kotecky, O.; Westermann, S.; Fleming, Y. Influence of silica specific surface area on the viscoelastic and fatigue behaviors of silica-filled SBR composites. Polymers 2021, 13, 3094. [Google Scholar] [CrossRef]
  35. Han, D.; Zhang, S.; Wang, K.; Pan, Y.; Zhu, D.; Wang, C.; Pan, Y. A comparison of the effects of traditional and wet mixing processes of rubber on metal friction and wear. J. Appl. Polym. Sci. 2021, 138, 50761. [Google Scholar] [CrossRef]
  36. Werther, J.; Saenger, M.; HArtge, E.-U.; Ogada, T.; Siagi, Z. Combustion of agricultural residues. Prog. Energy Combust. Sci. 2000, 26, 1–27. [Google Scholar] [CrossRef]
  37. Prabha, S.; Durgalakshmi, D.; Rajendran, S.; Lichtfouse, E. Plant-derived silica nanoparticles and composites for biosensors, bioimaging, drug delivery and supercapacitors: A review. Environ. Chem. Lett. 2021, 19, 1667–1691. [Google Scholar] [CrossRef] [PubMed]
  38. Shahrokh Abadi, M.H.; Delbari, A.; Fakoor, Z.; Baedi, J. Effects of annealing temperature on infrared spectra of SiO2 extracted from rice husk. J. Ceram. Sci. Technol. 2015, 6, 41–46. [Google Scholar]
  39. Yao, H.; Weng, G.; Liu, Y.; Fu, K.; Chang, A.; Chen, Z.-R. Effect of silane coupling agent on the fatigue crack propagationof silica-filled natural rubber. J. Appl. Polym. Sci. 2015, 132, 41980. [Google Scholar] [CrossRef]
  40. Chu, P.; Zhang, H.; Zhao, J.; Gao, F.; Guo, Y.; Dang, B.; Zhang, Z. On the volume resistivity of silica nanoparticle filled epoxy with different surface modifications. Compos. Part A 2017, 99, 139–148. [Google Scholar] [CrossRef]
  41. Kumkrong, N.; Dittanet, P.; Saeoui, P.; Loykulnant, S.; Prapainainar, P. Properties of silica/natural rubber composite film and foam: Effects of silica content and sulfur vulcanization system. J. Polym. Res. 2022, 29, 302. [Google Scholar] [CrossRef]
  42. Yu, K.; Yang, L.; Wang, J.; Zhu, Z.; Wang, T.-J. Modification of nanosilica particles with hydrophobic modifier bis [3-(triethoxysilyl)propyl]tetrasulfide by using micro-injection in aqueous solutions. Colloids Surf. A Physicochem. Eng. Asp. 2020, 599, 124852. [Google Scholar] [CrossRef]
  43. Wang, D.; Tang, Z.; Huang, R.; Li, H.; Zhang, C.; Guo, B. Inverse vulcanization of vinyltriethoxysilane: A novel interfacial coupling agent for silica-filled rubber composites. Macromolecules 2022, 55, 8485–8494. [Google Scholar] [CrossRef]
  44. Prasertsri, S.; Rattanasom, N. Mechanical and damping properties of silica/natural rubber composites prepared from latex system. Polym. Test. 2011, 30, 515–526. [Google Scholar] [CrossRef]
  45. Arayapranee, W.; Na-Ranong, N.; Rempel, G.L. Application of rice husk ash as fillers in the natural rubber industry. J. Appl. Polym. Sci. 2005, 98, 34–41. [Google Scholar] [CrossRef]
  46. Buakhlee, W.; Suwanna, P.; Kaewwattana, W. Cure characteristic and mechanical properties of silica masterbatch prepared from fresh natural rubber latex mixing. Plast. Rubber Compos. 2020, 49, 127–133. [Google Scholar] [CrossRef]
  47. Wagner, M.P. Reinforcing silicas and silicates. Rubber Chem. Technol. 1976, 49, 703–774. [Google Scholar] [CrossRef]
  48. Ramasamy, S.; Ismail, H.; Munusamy, Y. Tensile and morphological properties of rice husk powder filled natural rubber latex foam. Polym. Plast. Technol. Eng. 2012, 51, 1524–1529. [Google Scholar] [CrossRef]
  49. da Costa, H.M.; Visconte, L.L.Y.; Nunes, R.C.R.; Furtado, C.R.G. The Effect of coupling agent and chemical treatment on rice husk ash- filled natural rubber composites. J. Appl. Polym. Sci. 2000, 76, 1019–1027. [Google Scholar] [CrossRef]
  50. Jacques, J.E. Rubber Technology and Manufacture, 2nd ed.; Blow, C.M., Hepburn, C., Eds.; Butterworth Scientific: London, UK, 1982. [Google Scholar]
  51. da Costa, H.M.; Visconte, L.L.Y.; Nunes, R.C.R.; Furtado, C.R.G. Rice husk ash filled natural rubber compounds the use of rheometric data to qualitatively estimate optimum filler loading. Int. J. Polym. Mater. 2004, 53, 475–497. [Google Scholar] [CrossRef]
  52. Gurovich, D.; Macosko, C.W.; Tirrell, M. The Influence of filler-filler and filler-polymer interactions on the physical properties of silica-filled liquid polyisoprene. Rubber Chem. Technol. 2004, 77, 1–12. [Google Scholar] [CrossRef]
  53. Arayapranee, W. Chapter 8: Rubber Abrasion Resistance. In Abrasion Resistance of Materials; Adamiak, M., Ed.; InTech: London, UK, 2012; pp. 147–166. [Google Scholar]
Figure 1. Schematic illustration showing the preparation processes of (a) silica from rice husk and (b) silane-modified silica.
Figure 1. Schematic illustration showing the preparation processes of (a) silica from rice husk and (b) silane-modified silica.
Jcs 06 00369 g001
Figure 2. Schematic illustration showing the preparation of silica-filled NR composites. a N-Cyclohexyl-2-benzothiazolesulfenamide.
Figure 2. Schematic illustration showing the preparation of silica-filled NR composites. a N-Cyclohexyl-2-benzothiazolesulfenamide.
Jcs 06 00369 g002
Figure 3. SEM images and the corresponding EDS spectra of RHS (a,b) and Si69-RHS (c,d). Note: The peak at 2.12 keV corresponds to the Au K signal from the sputtered gold films.
Figure 3. SEM images and the corresponding EDS spectra of RHS (a,b) and Si69-RHS (c,d). Note: The peak at 2.12 keV corresponds to the Au K signal from the sputtered gold films.
Jcs 06 00369 g003
Figure 4. Comparative FTIR spectra of RHS and Si69-RHS.
Figure 4. Comparative FTIR spectra of RHS and Si69-RHS.
Jcs 06 00369 g004
Figure 5. Comparative TGA curves of RHS and Si69-RHS under N2 flow.
Figure 5. Comparative TGA curves of RHS and Si69-RHS under N2 flow.
Jcs 06 00369 g005
Figure 6. Scorch and cure times of the NR compounds filled with various Si69-RHS loadings.
Figure 6. Scorch and cure times of the NR compounds filled with various Si69-RHS loadings.
Jcs 06 00369 g006
Figure 7. SEM images of the NR composites filled with various Si69-RHS loadings: (a) 0 phr, (b) 10 phr, (c) 20 phr, (d) 30 phr, (e) 40 phr, and (f) 50 phr.
Figure 7. SEM images of the NR composites filled with various Si69-RHS loadings: (a) 0 phr, (b) 10 phr, (c) 20 phr, (d) 30 phr, (e) 40 phr, and (f) 50 phr.
Jcs 06 00369 g007
Figure 8. SEM-EDS mapping analysis of the NR composites filled with various Si69-RHS loadings: (a) 10 phr, (b) 20 phr, (c) 30 phr, (d) 40 phr, and (e) 50 phr. The yellow, blue, and green represent the carbon (C K signal), silicon (Si K signal), and oxygen (O K signal) elements, respectively.
Figure 8. SEM-EDS mapping analysis of the NR composites filled with various Si69-RHS loadings: (a) 10 phr, (b) 20 phr, (c) 30 phr, (d) 40 phr, and (e) 50 phr. The yellow, blue, and green represent the carbon (C K signal), silicon (Si K signal), and oxygen (O K signal) elements, respectively.
Jcs 06 00369 g008
Figure 9. Mechanical properties of the NR composites filled with various Si69-RHS loadings (10–50 phr): (a) stress–strain curves, (b) tensile strength, (c) elongation at break, and (d) 100% modulus.
Figure 9. Mechanical properties of the NR composites filled with various Si69-RHS loadings (10–50 phr): (a) stress–strain curves, (b) tensile strength, (c) elongation at break, and (d) 100% modulus.
Jcs 06 00369 g009
Figure 10. (a) Hardness (Shore A) and (b) abrasion loss of the NR composites filled with various Si69-RHS loadings.
Figure 10. (a) Hardness (Shore A) and (b) abrasion loss of the NR composites filled with various Si69-RHS loadings.
Jcs 06 00369 g010
Table 1. Formulations of the NR compounds filled with Si69-RHS.
Table 1. Formulations of the NR compounds filled with Si69-RHS.
MaterialsPart Per Hundred (phr)
C-0C-10C-20C-30C-40C-50
Natural rubber100100100100100100
Zinc oxide555555
Stearic acid222222
CBS a111111
Si69-RHS01020304050
Sulfur1.51.51.51.51.51.5
Aromatic oil101010101010
a N-Cyclohexyl-2-benzothiazolesulfenamide.
Table 2. Mechanical properties of the NR composites filled with various Si69-RHS loadings.
Table 2. Mechanical properties of the NR composites filled with various Si69-RHS loadings.
Si69-RHS
Loading (phr)
Tensile Strength
(MPa)
Young Modulus
(MPa)
Elongation at Break (%)Hardness
(Shore A)
Abrasion Loss
(%)
016.6 ± 1.51.04 ± 0.11734 ± 3533.2 ± 0.80.06 ± 0.01
1019.5 ± 1.51.17 ± 0.12711 ± 1738.3 ± 0.70.06 ± 0.01
2020.9 ± 1.81.13 ± 0.09731 ± 4240.6 ± 0.70.13 ± 0.01
3019.7 ± 0.81.23 ± 0.08686 ± 541.6 ± 0.50.15 ± 0.02
4018.5 ± 0.81.38 ± 0.06635 ± 843.6 ± 0.50.17 ± 0.02
5015.5 ± 0.51.30 ± 0.08589 ± 1544.9 ± 0.10.24 ± 0.02
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Choophun, N.; Chaiammart, N.; Sukthavon, K.; Veranitisagul, C.; Laobuthee, A.; Watthanaphanit, A.; Panomsuwan, G. Natural Rubber Composites Reinforced with Green Silica from Rice Husk: Effect of Filler Loading on Mechanical Properties. J. Compos. Sci. 2022, 6, 369. https://doi.org/10.3390/jcs6120369

AMA Style

Choophun N, Chaiammart N, Sukthavon K, Veranitisagul C, Laobuthee A, Watthanaphanit A, Panomsuwan G. Natural Rubber Composites Reinforced with Green Silica from Rice Husk: Effect of Filler Loading on Mechanical Properties. Journal of Composites Science. 2022; 6(12):369. https://doi.org/10.3390/jcs6120369

Chicago/Turabian Style

Choophun, Nicha, Nattapat Chaiammart, Kantavee Sukthavon, Chatchai Veranitisagul, Apirat Laobuthee, Anyarat Watthanaphanit, and Gasidit Panomsuwan. 2022. "Natural Rubber Composites Reinforced with Green Silica from Rice Husk: Effect of Filler Loading on Mechanical Properties" Journal of Composites Science 6, no. 12: 369. https://doi.org/10.3390/jcs6120369

Article Metrics

Back to TopTop