III.1.Characterization of the nanopowders
- III.1.1.Infrared analysis of the nanocomposites
Figure 3 shows the FTIR spectra of thenanocompositescalcinated at 1000 °C and 1200 °C for 2hours. Figure 3a presents FTIR spectra the composite with 25 % Si-75 % Ti.The bands around 1086cm-1and 1049cm-1would confirm the presence of Si-O-Si bound. The absorption band around 2850-3640cm-1would correspond to the fundamental stretching vibration of hydroxyl O-H groups caused by the presence of H₂O completely evaporated.This band would be further confirmed by the weak band at 1634 cm-1 as was rightly demonstrated by Arun Kumar,Merline-Shyla& Xavier [19]. Figure 3b exhibits the FTIR spectra of the composite with 50 % Si-50 % Ti nanocomposite. The bands appearing at 1100 and918cm-1 would be associated with asymmetric Si-O-Si stretching vibration and Si-O-Ti vibration, respectively. The band centered at 1600 cm-1would correspond to the vibration of hydroxyl groups. These findings would confirm previous results reported by Aziz, Asyikin and Sopyan [43] and Quyen et al. [20] who agreed on the fact that the band at around 2490-3650 cm-1 observed before heat treatment was due to the presence of O–Hstretching and bending vibrations. Figure3c shows the FTIR spectra of the composite with 75 % Si-25 % Ti. Three characteristic bands appeared at around 2486-3589 cm-1, 1100 cm-1, and 940 cm-1. The peak from 1100 cm-1could bean indication of the presence of asymmetric stretching vibrations of Si-O-Si bond as Du et al. [44] explained. Moreover, in line with Quyen et al. [20], the small peak observed at 940 cm-1 could be attributed to Ti-O-Si linkages affirming the bonding between Ti and Si At 1200°C.Thus, the Ti-O-Si bond would improve the thermal stability of TiO2 and suppress the phase transformation from anatase to rutile. In addition, the FTIR measurements correlated with Yu et al.[21] who argued that the presence of a peak at 650 cm-1 indicated the vibration absorption of the Ti-O-Ti bond, Furthermore, in line with Quyen et al. [20],the broad band located between 2486-3589 cm-1could be assigned to the feature stretching vibration of hydroxyl O-H groups. However, the band intensity of hydroxyl groups at around 2600-3660 cm-1 increased with the increase of TiO2 mass fraction and disappeared after calcination. This would confirms the control of the synthesis method based on the sol-gel process.
III.1.2. XRD analysis of the nanocomposites
Figure 4 shows the XRD spectra of the nanocomposites before and after annealing at 1000°C and 1200°C for.2 hours. Considering all nanocomposites, at 243°C, anatase structure was identified in accordance with the JCPDS files NO. 00-004-0477. Moreover, in line with (2015).Hui et al.[22], the small peak at around 2θ=13° would reveal the presence of amorphous SiO2. The transition from anatase to rutile structure was observed forpure TiO2 at 1000°C.Nonetheless, this phase transformation was not observed for the TiO2-SiO2 nanocomposites treated at the same temperature. It would be worth mentioning that the crystallinity of amorphous SiO2 increased with calcination temperature and that the addition of silica to TiO2nanopoudres did not have an impact on the crystalline structure of anatase TiO2. These results correlated with Pakdeland.Daoud’s [23] previous findings.Additionally, the anatase-rutile transformation with calcination at 1000°C and 1200 °C was highly delayed in the TiO2-SiO2 nanocomposites. This phenomenon was also observed even at low Si:Ti ratioss can be observed in figure 4,.Moreno et al. [24] explained this by the fact that the formation of this amorphous SiO2 phase was considered to be important in retarding the anatase-to-rutile phase transition because the aerogels synthesized with ethyl orthosilicate (TEOS) and acetic acid had a good thermal stability at high-temperature of anatase. Finally, the amorphous SiO2was concluded to be a stronger inhibitor of the phase transformation of TiO2 nanopowders at temperatures greater than 1200°C which wouldaffect the mechanical properties of the resulting nanocomposites as was rightly explained by Vega, Olaya, and Ruiz [25].
III. 1.3. Mechanical characterization of the nanocomposites
the structural results represent changes in the chemical composition and structure of materials, in this test we can see the influence of this nanometric change on the macro or mechanical behavior. As can be seen in figures 5a and 5b, there was a clear drop of the Young’s modulus and shear modulus at 1200°C. This can be explained by the presence of water filling the micro-pores and forming a film of water inside them. Hence, the presence of water in the pores would lead to a low contribution and a relaxation of the' Stress Ground-frost. Que et al.[26] explained this by the fact that after filling the pores by capillary condensation, E and G would decrease because the pores of different sizes would undergo the capillary condensation one after the other. Therefore, they would have a nano-indenter and would show a dependence on the heat-treatment temperature. Figure 5c shows the Poisson’s Ratio of the nanocomposites at different temperatures. The values of Poisson’s Ratio had an inverse trend to that of the modules of elasticity. This can be explained by the different behaviors of the physical and chemical properties of the different aerogels used in this study. Indeed, these aerogels would have pores and defects accompanied by the formation of some cracks. However, after heat treatment at 1200°C, there was not a very large difference knowing that the values vary within the same error range. These findings would confirm the findings reported in section III.1.1 above about the FTIR analysis.
III.2. Thermal properties of the nanocomposites
III.2.1.DSC analysis
Figure 6a shows the DSC scans of polyester composite. As can be seen in figure 6a, the values provided by the differential scanning calorimetry of the polymer composite having 15% fiberglass were clear in the temperature range between 25 and 160°C. The vitreous transmission temperature Tg of the respective compositions and states of the nanocomposites can be clearly determined through the curves in figures 6 b, 6c and 6d. The Tg started with the state of polyester/fiberglass composite at 42.8°C. With the addition of TiO2-SiO2 system of 5%wt, a relatively small increase was observed depending on the composition of this system. This increase in Tg was of a maximum value of 45.11°C for the 75%TiO2-25%SiO2nanofiller. This finding correlated very well with Nayak, Mahato, and Ray [27]. As can be seen in figure.6, theaddition of the TiO2-SiO2 nanofillers seemed to affect the value of Tg of the composite with 75%TiO2-25%SiO2at a heat treatment of 1000°C. Indeed, this value rose to 49.66°C; i.e. an increase of 15.6% in the state not loaded. Hence, it can be said that the Tg depends directly on the property of themolecular motion. This would obviously indicate that the molecular chain of nanocomposites needs a higher temperature to undergo the glass transition process. This would be explained by the fact that the movement of the molecular chains was inhibited by the addition of the functionalized nano-system .Hence, it can be induced that the TiO2 nanofiller of this composition with a rutile structure increased the internal lassion of the composite notes. In contrast, the presence of SiO2 weakened this behavior.
III.2.2. TGAanalysis
The stability and degradation mechanisms of nanocomposites are an important property for various fields of application due to changes in viscoelastic behavior. In this study, the thermal properties of nanocomposites were investigated using TGA/DTG. The TGA data is investigated to know the decomposition characteristic and thermal stability under a wide temperature range. In these methods, the changes in thermal stability are represented as percent weight loss as a function of the temperature.
The DTG data were characterized by the maximum temperature (Tmax) and the setpoint temperature (Te). The area behind the DTG plot is proportional to the mass change and the height of the peak represents the rate of mass change at that temperature [28].
Figure 8 exhibits the various TGA and DTG curves of the composites.Following Ismail, et al.[28], the area behind the DTG plot would be proportional to the mass change and the height of the peak would represent the rate of mass change at that temperature.
As can be seen in figure 8a, the unfilled composite showed a thermal stability up to about 200°C associated with a low weight loss of 5% above this temperature until a sudden drop towards 280°C. Subsequently, a loss of mass of high amplitude – approximately 90% of the initial mass – wasobserved up to a temperature close to 400°C, this was in line with da Luz et al. [29] who detected a degradation peak at 384.79° in their DTG curve. The TGA curve and the corresponding DTG thermograms of the nanofiller composites with15%wt and with the respective composition of the TiO2-SiO2 system, with a variation of the heat treatment temperatures from untreated to 1000°C and 1200°C, are illustrated in figures 8 b-e.The TGA results revealed that there was a variation of the thermal behavior of the compositesshown in the evolution of the temperature of decomposition.
This temperature was almost stable at the maximum value for the composite with 75%TiO2-25%SiO2 of heat treatment at 1200°C with a weight loss ratio of 5% .These findings differed withda Luz et al. [29] who revealed that the incorporation of the bagasse fiber nanofillers into a polymer matrix at different compositions and heat treatment temperatures, did not significantly modify the thermal behavior of the composite
2.2. Dynamic mechanical analysis (DMA)
Figure 9 shows the plots of the storage modulus (E'), the loss modulus (E'') and the ratio of the loss and storage modulus (tanδ) of the fiberglass polyester composite with a value of E'= 2.849 GPa, E'' =257 MPa at a room temperature of 27°C and a maximum value of tanδ' of 0.85. Compared to the unfilled composite, the incorporation of 5% nanofillershowed an increase in the storage modulus.Maharramov et al. [30] considered this increase as a contribution to the improvement of the interfacial bond between the nanofillers and the polyester matrix. It can be observed that the highest storage modulus was obtained with 5 %wt of 25%TiO2-75%SiO2of untreated composite. The storage modulus of nanocomposites decreased when the temperature increased until the material reached its glass transition temperature of T=44.53°C. This could contribute to the improvement of the thermal stability and the resistance of the treated samples. The thermal capacity would increase by the evolution of the the increase of the Tg value. However, a small reduction would be observed at 1000°C and then a sharp decrease at 1200°C.
Similarly, it was observed that the increase of Tg decreased the storage modulus and caused the transition from glassy state to the rubbery state. Therefore, the molecules would start to lose their packed structure and a rise of vibration al movements would occur leading to a steep loss of storage modulus. Indeed, Bendaoued et al. [31].recommended that a better reinforcing effect in the composites would require a limitation of the amount of vibration occurring within the molecules of a materia to maintain its solid state. Finally, the DMA test coupled with the mechanical and thermal resistance revealed that the most durable compositions would be made of 25% Ti-75% Si and 75% Ti-25Si treated at 1200°C.
Figure 10 exhibits the DMA analysis of the composite with different amounts of fillers. As can be seen in figure 10a, the loss modulus (E”) indicated the viscous response of the samples measured and provided information on their tendency to disperse the applied energy. The structural mobility of composite increased with the addition of nanofillers. This phenomenon was explained by Lertwimolnun, and Vergnes’ [32]as a sign of interactions of a high surface area and a higher ratio of nanofiller/polyester.These results were in agreement with Choi, Kim and Laine [33] findings. Figure 10b illustrates the ratio of the loss and storage modulus (tanδ). The loss modulus increased in function of the increase of TiO2 particles because of their thermal resistance and their stability. This observation confirmed previous results by Gonzalez-Calderon et al.[34] who revealed that this effect becomes clearer with the heat treatment at 1200°C. Moreover, in the present study, the nanocomposite (COMP+5 Wt % 75Ti-25Si) treated at 1200°C were the most stable and efficient. From a thermo-mechanical point of view, the increase in the temperature peaks of the loss modulus could be attributed to the improvement in thermal stability and to the segmental immobilization of the polyester chainsreinforced withglassfibersand the nanofillers.Lertwimolnun, and Vergnes’ [32] and Alamri,andLow [35]reported that these mechanical properties were due to the higher tendency of nanocomposites to absorb energy due to the addition of nanofillers. The peak tan (δ) of the nanocomposites decreased despite the improvement of the temperature and the addition of nanofillers. This finding would confirms the high rigidity and the interfacial adhesion with the nanofillers and the polymer matrix. When the temperature was raised, the storage modulus of the nanocomposites increased more than the loss modulus, indicating that the structure was thermally stable. With the addition of nanofillers, there was no longer any change in the Tg of the samples. Finally, the viscoelastic behavior and dynamic mechanical properties of nanocomposites were strongly related to the physical and chemical properties of the aerogels used in this study.
III. 3.Mechanical properties of the composite
III.3.1. Microhardness of the nanocomposite
Figure 12 shows the effect of adding nanopowders on the impact strength of the nanocomposites The addition of nanoparticles seemed to increase the microhardness of the 25 TiO2 %-75 SiO2 % nanocompositemicrohardness to around 26hv at 1200°C. This would represent a 60% increase. However, it seemed that the other compositions did not gain a significant increase in microhardness. Furthermore, the the microhardness of the 75%TiO2-25%SiO2nanocomposite (in green) decresed to 21 hv at 1200 °C. This would imply that the more SiO2 was present in the nanocomposite, the higher its hardness was after undergoing a thermal treatment.
This improvement could be attributed to a low agglomeration of the nanofillers grains presented in the polymer matrix which would lead to a strong polymer-filler interaction, thus yielding an increase in microhardness. Rao, Singh and Dwivedi [36].reported that nanocomposites exhibited an improved microhardness with lower loading rate.
III.3.2. Tensile test of the nanocomposite
Figure 13 presents the stress-strain diagram of the elaborated nanocomposites. These tensile tests were conducted to assess the evolution of the tensility of the nanocomposite in function of the addition of nanofillers at different contents. The curve of the 85% wt polyester 15% wt fiber glass composite showed an elastic behavior. With the addition of nanofillers, a very clear plastic zone was observed with an increase in the ductility of the produced material.
Additionally, the tensile test allowed to determinethe tensile modulus, ultimate tensile strength (UTS) and breaking stress (SB). Figure 14 illustrates the effect of the nanoparticles reinforcements on the tensile strength of the prepared samples. Figure 14a shows the value of Yong's modulus for the deferent compositions. The tensile modulus of the composite made of 85% polyester + 15% fiberglasswas 6.6 GPa.With the addition of 5% weight of System nanoparticles made of 25Ti-75Si, there was an increase of 15.15%, while this increase was of only 4.5%, for the composite + 25 Si -75Ti. Therefore, it seemed that the presence of a higher percentage of Si was responsible for the enhancement of the mechanical properties of the composite This finding correlated with the reported DMA analysis and the microhardness results.
Figure 14 shows the effects of heat treatment on the mechanical properties of the composite. Hence, figure 14a shows the composition+5 % wt (25%TiO2-75SiO2) hadthe highest Young's modulus value after heat treatment or internal structural modification of r. However, the strain to break test and the UTS shown in figures 13 b and 13 c, respectively revealed that the composition having 75 %TiO2 exhibited the highest elasticity values. Therefore, SiO2enhanced the rigidity and TiO2 enhanced the plasticity of the nanocomposite.
III.4. Morphological analysis through SEM observations
Figure 15 shows the SEM micrographs of the composites. The SEM images of the fracture surface of the tested samples showed the deformation mechanisms that occurred after performing the tensile tests on the prepared nanocomposites. As can be seen in figure 15a, the unfilled sample showed a failure section across the matrix surface and the glass fibers orientation with noticeable porosity which could explain the fragility and low mechanical resistance of the composite. Behera et al. [37] rightly explainedthis behavior bythe existence of potholes, fractures and fiber detachment among the failure mechanisms of composites. In addition, in this study, the close inspection of the interface between the glass fabric and the polyester matrix revealed a weak fiber-matrix adhesion. Figures 15b, 15c and 15d show the presence of slightly agglomerated nanofillers proving the importance of the dispersion methods to distribute the nanoparticles TiO2andSiO2homogenously in the polyester matrix. This observation correlates well with Saberianet al.’s[38] finding. Indeed, these scholars reported that the well-dispersed TiO2-SiO2 system nanoparticles were attributed to better dispersion of the nanoparticles that would improve the interface area leading to a good creation of matrix/fiber interface bonds. Figure 15d shows a visible agglomeration of the nanoparticles. In total agreement with Sahnesarayi, Sarpoolaky, and Rastegari [39], it can be concluded that under these conditions, the dispersion became more difficult to reach a very high concentrations of nanoparticles. These agglomerates caused defects in the nanocomposites and acted as stress concentration sites, generating early cracks and fractures. With the heat treatment,the problem improved. The comparison between the mechanical performances obtained for the different nanocomposites would indicate that the use of 5% wt of nanofillers with the heat treatment, a more resistant fibre-polyester matrix interface was created. As a result, the glass fibers could transfer their reinforcing potential to the nanocomposites and fiber pull-out would be reduced