Figure 4 - uploaded by Yiming Jiang
Content may be subject to copyright.
The FTIR spectra of the garnet before and after modification. 

The FTIR spectra of the garnet before and after modification. 

Source publication
Article
Full-text available
The present study investigated the tribological performance of epoxy (EP) matrix composites enhanced with natural garnet. The garnet was surface-modified with sodium stearate for optimal performance. Composites comprising different contents and particle sizes of modified garnet (MG) were prepared with a mixture of EP and MG. The sodium stearate-bon...

Contexts in source publication

Context 1
... with modification of nanoparticles [24,25], MG microparticles had a higher water contact angle. Figure 4 shows the FTIR spectra of unMG and MG. The garnets before and after modification exhibited the same stretching vibration bands at 1463.3, 913.5, 849.0, 690.3, 526.8, and 461.1 cm −1 As shown in Figure 3, contact angles of unMG and MG in water were approximately 20 • and 135 • , respectively. ...
Context 2
... with modification of nanoparticles [24,25], MG microparticles had a higher water contact angle. Figure 4 shows the FTIR spectra of unMG and MG. The garnets before and after modification exhibited the same stretching vibration bands at 1463.3, 913.5, 849.0, 690.3, 526.8, and 461.1 cm −1 [26][27][28][29]. ...
Context 3
... radicals reacted with Ca 2+ and Fe 3+ of garnet and formed a hydrophobic structure on the garnet surface. Figure 4. The FTIR spectra of the garnet before and after modification. ...
Context 4
... showed a strong hydrophilic surface that led to poor compatibility between itself and the matrix. Therefore, the hydrophilic surface of garnet powder must be modified to become oleophilic before preparation of composite materials. MG surface became nonpolar, which contributed to extreme hydrophobicity and lipophilicity. As a consequence, the contact angle ( Figure 3) exhibited remarkable variations due to surface modification with sodium stearate. As shown in Figure 3, contact angles of unMG and MG in water were approximately 20° and 135°, respectively. At the same time, contact angles of unMG and MG in benzene containing 2 wt % epoxy were 104° and 23°, respectively. This result indicates that unMG exhibited strong hydrophilicity and oleophobicity, whereas MG was extremely hydrophobic and lipophilic. The water contact angle of MG is larger than the 82.03° angle of EP [23]. The significant transformation of the contact angle confirms the successful grafting of sodium stearate on the surface of the garnet. Sodium stearate, with its excellent surface activation effect, remarkably enhances the hydrophobic properties of garnet. Compared with modification of nanoparticles [24,25], MG microparticles had a higher water contact angle. Figure 4 shows the FTIR spectra of unMG and MG. The garnets before and after modification exhibited the same stretching vibration bands at 1463.3, 913.5, 849.0, 690.3, 526.8, and 461.1 cm −1 As shown in Figure 3, contact angles of unMG and MG in water were approximately 20 • and 135 • , respectively. At the same time, contact angles of unMG and MG in benzene containing 2 wt % epoxy were 104 • and 23 • , respectively. This result indicates that unMG exhibited strong hydrophilicity and oleophobicity, whereas MG was extremely hydrophobic and lipophilic. The water contact angle of MG is larger than the 82.03 • angle of EP [23]. The significant transformation of the contact angle confirms the successful grafting of sodium stearate on the surface of the garnet. Sodium stearate, with its excellent surface activation effect, remarkably enhances the hydrophobic properties of garnet. Compared with modification of nanoparticles [24,25], MG microparticles had a higher water contact angle. Figure 4 shows the FTIR spectra of unMG and MG. The garnets before and after modification exhibited the same stretching vibration bands at 1463.3, 913.5, 849.0, 690.3, 526.8, and 461.1 cm −1 [26][27][28][29]. These two specimens also showed the -OH band at 3370 cm −1 . MG bands at 2918 cm −1 and 2849 cm −1 resulted from, respectively, asymmetric and symmetric stretching vibration of methylene [24,30], indicating successful loading of sodium stearate ions into the garnet. The bands at 1434, 1464, and 1558 cm −1 of sodium stearate were attributed to vibration absorption of methyl and methylene, and asymmetric stretching vibration of the carboxylate group, respectively [30][31][32]. When sodium stearate was grafted onto the garnet, those bands interacted with the band at 1445 cm −1 , leading to a band shift, demonstrating chemical adsorption of sodium stearate in garnet. Stearate radicals reacted with Ca 2+ and Fe 3+ of garnet and formed a hydrophobic structure on the garnet surface. Minerals 2018, 8, x FOR PEER REVIEW 6 of 15 [26][27][28][29]. These two specimens also showed the -OH band at 3370 cm −1 . MG bands at 2918 cm −1 and 2849 cm −1 resulted from, respectively, asymmetric and symmetric stretching vibration of methylene [24,30], indicating successful loading of sodium stearate ions into the garnet. The bands at 1434, 1464, and 1558 cm −1 of sodium stearate were attributed to vibration absorption of methyl and methylene, and asymmetric stretching vibration of the carboxylate group, respectively [30][31][32]. When sodium stearate was grafted onto the garnet, those bands interacted with the band at 1445 cm −1 , leading to a band shift, demonstrating chemical adsorption of sodium stearate in garnet. Stearate radicals reacted with Ca 2+ and Fe 3+ of garnet and formed a hydrophobic structure on the garnet surface. Figure 4. The FTIR spectra of the garnet before and after ...
Context 5
... showed a strong hydrophilic surface that led to poor compatibility between itself and the matrix. Therefore, the hydrophilic surface of garnet powder must be modified to become oleophilic before preparation of composite materials. MG surface became nonpolar, which contributed to extreme hydrophobicity and lipophilicity. As a consequence, the contact angle ( Figure 3) exhibited remarkable variations due to surface modification with sodium stearate. As shown in Figure 3, contact angles of unMG and MG in water were approximately 20° and 135°, respectively. At the same time, contact angles of unMG and MG in benzene containing 2 wt % epoxy were 104° and 23°, respectively. This result indicates that unMG exhibited strong hydrophilicity and oleophobicity, whereas MG was extremely hydrophobic and lipophilic. The water contact angle of MG is larger than the 82.03° angle of EP [23]. The significant transformation of the contact angle confirms the successful grafting of sodium stearate on the surface of the garnet. Sodium stearate, with its excellent surface activation effect, remarkably enhances the hydrophobic properties of garnet. Compared with modification of nanoparticles [24,25], MG microparticles had a higher water contact angle. Figure 4 shows the FTIR spectra of unMG and MG. The garnets before and after modification exhibited the same stretching vibration bands at 1463.3, 913.5, 849.0, 690.3, 526.8, and 461.1 cm −1 As shown in Figure 3, contact angles of unMG and MG in water were approximately 20 • and 135 • , respectively. At the same time, contact angles of unMG and MG in benzene containing 2 wt % epoxy were 104 • and 23 • , respectively. This result indicates that unMG exhibited strong hydrophilicity and oleophobicity, whereas MG was extremely hydrophobic and lipophilic. The water contact angle of MG is larger than the 82.03 • angle of EP [23]. The significant transformation of the contact angle confirms the successful grafting of sodium stearate on the surface of the garnet. Sodium stearate, with its excellent surface activation effect, remarkably enhances the hydrophobic properties of garnet. Compared with modification of nanoparticles [24,25], MG microparticles had a higher water contact angle. Figure 4 shows the FTIR spectra of unMG and MG. The garnets before and after modification exhibited the same stretching vibration bands at 1463.3, 913.5, 849.0, 690.3, 526.8, and 461.1 cm −1 [26][27][28][29]. These two specimens also showed the -OH band at 3370 cm −1 . MG bands at 2918 cm −1 and 2849 cm −1 resulted from, respectively, asymmetric and symmetric stretching vibration of methylene [24,30], indicating successful loading of sodium stearate ions into the garnet. The bands at 1434, 1464, and 1558 cm −1 of sodium stearate were attributed to vibration absorption of methyl and methylene, and asymmetric stretching vibration of the carboxylate group, respectively [30][31][32]. When sodium stearate was grafted onto the garnet, those bands interacted with the band at 1445 cm −1 , leading to a band shift, demonstrating chemical adsorption of sodium stearate in garnet. Stearate radicals reacted with Ca 2+ and Fe 3+ of garnet and formed a hydrophobic structure on the garnet surface. Minerals 2018, 8, x FOR PEER REVIEW 6 of 15 [26][27][28][29]. These two specimens also showed the -OH band at 3370 cm −1 . MG bands at 2918 cm −1 and 2849 cm −1 resulted from, respectively, asymmetric and symmetric stretching vibration of methylene [24,30], indicating successful loading of sodium stearate ions into the garnet. The bands at 1434, 1464, and 1558 cm −1 of sodium stearate were attributed to vibration absorption of methyl and methylene, and asymmetric stretching vibration of the carboxylate group, respectively [30][31][32]. When sodium stearate was grafted onto the garnet, those bands interacted with the band at 1445 cm −1 , leading to a band shift, demonstrating chemical adsorption of sodium stearate in garnet. Stearate radicals reacted with Ca 2+ and Fe 3+ of garnet and formed a hydrophobic structure on the garnet surface. Figure 4. The FTIR spectra of the garnet before and after ...
Context 6
... showed a strong hydrophilic surface that led to poor compatibility between itself and the matrix. Therefore, the hydrophilic surface of garnet powder must be modified to become oleophilic before preparation of composite materials. MG surface became nonpolar, which contributed to extreme hydrophobicity and lipophilicity. As a consequence, the contact angle ( Figure 3) exhibited remarkable variations due to surface modification with sodium stearate. As shown in Figure 3, contact angles of unMG and MG in water were approximately 20° and 135°, respectively. At the same time, contact angles of unMG and MG in benzene containing 2 wt % epoxy were 104° and 23°, respectively. This result indicates that unMG exhibited strong hydrophilicity and oleophobicity, whereas MG was extremely hydrophobic and lipophilic. The water contact angle of MG is larger than the 82.03° angle of EP [23]. The significant transformation of the contact angle confirms the successful grafting of sodium stearate on the surface of the garnet. Sodium stearate, with its excellent surface activation effect, remarkably enhances the hydrophobic properties of garnet. Compared with modification of nanoparticles [24,25], MG microparticles had a higher water contact angle. Figure 4 shows the FTIR spectra of unMG and MG. The garnets before and after modification exhibited the same stretching vibration bands at 1463.3, 913.5, 849.0, 690.3, 526.8, and 461.1 cm −1 As shown in Figure 3, contact angles of unMG and MG in water were approximately 20 • and 135 • , respectively. At the same time, contact angles of unMG and MG in benzene containing 2 wt % epoxy were 104 • and 23 • , respectively. This result indicates that unMG exhibited strong hydrophilicity and oleophobicity, whereas MG was extremely hydrophobic and lipophilic. The water contact angle of MG is larger than the 82.03 • angle of EP [23]. The significant transformation of the contact angle confirms the successful grafting of sodium stearate on the surface of the garnet. Sodium stearate, with its excellent surface activation effect, remarkably enhances the hydrophobic properties of garnet. Compared with modification of nanoparticles [24,25], MG microparticles had a higher water contact angle. Figure 4 shows the FTIR spectra of unMG and MG. The garnets before and after modification exhibited the same stretching vibration bands at 1463.3, 913.5, 849.0, 690.3, 526.8, and 461.1 cm −1 [26][27][28][29]. These two specimens also showed the -OH band at 3370 cm −1 . MG bands at 2918 cm −1 and 2849 cm −1 resulted from, respectively, asymmetric and symmetric stretching vibration of methylene [24,30], indicating successful loading of sodium stearate ions into the garnet. The bands at 1434, 1464, and 1558 cm −1 of sodium stearate were attributed to vibration absorption of methyl and methylene, and asymmetric stretching vibration of the carboxylate group, respectively [30][31][32]. When sodium stearate was grafted onto the garnet, those bands interacted with the band at 1445 cm −1 , leading to a band shift, demonstrating chemical adsorption of sodium stearate in garnet. Stearate radicals reacted with Ca 2+ and Fe 3+ of garnet and formed a hydrophobic structure on the garnet surface. Minerals 2018, 8, x FOR PEER REVIEW 6 of 15 [26][27][28][29]. These two specimens also showed the -OH band at 3370 cm −1 . MG bands at 2918 cm −1 and 2849 cm −1 resulted from, respectively, asymmetric and symmetric stretching vibration of methylene [24,30], indicating successful loading of sodium stearate ions into the garnet. The bands at 1434, 1464, and 1558 cm −1 of sodium stearate were attributed to vibration absorption of methyl and methylene, and asymmetric stretching vibration of the carboxylate group, respectively [30][31][32]. When sodium stearate was grafted onto the garnet, those bands interacted with the band at 1445 cm −1 , leading to a band shift, demonstrating chemical adsorption of sodium stearate in garnet. Stearate radicals reacted with Ca 2+ and Fe 3+ of garnet and formed a hydrophobic structure on the garnet surface. Figure 4. The FTIR spectra of the garnet before and after ...

Citations

... Similar findings were presented in films of cassava starch reinforced with thermally modified kaolin [36] and taro starch reinforced with bentonite [96]. Furthermore, the surface roughness of the biodegradable films was investigated using a 3D optical surface profiler (zeta 20) [97][98][99][100] (Fig. 14). The application of clay with the starch matrix reduced roughness on the film's surface. ...
Article
Full-text available
Fillers improve the thermal stability and mechanical properties of bio-plastics. This research aims to synthesize and characterize biodegradable film from mango seed kernel starch reinforced with micro-pottery clay using glycerol as a plasticizer. The mango seed kernel starch was extracted and characterized in terms of solubility (75.01 ± 0.35%), swelling power (22.3 ± 0.24%), moisture content (10.33 ± 0.21%), ash content (0.41 ± 0.13%), and yield of starch (19.97 ± 0.021%). Pottery clay to starch ratios of 0, 5, 7, and 10% w/w, and glycerol to starch ratios of 0, 25, 30, and 35% w/v were evaluated. The film was analyzed for its physicochemical properties such as tensile strength, thickness, transparency, and biodegradability. Furthermore, the surface functional groups, morphology, and thermal properties of the film were analyzed using FT-IR, SEM, 3D optical surface profiler, and TGA. The control film has a tensile strength of 3.33 ± 0.008Mpa whereas the tensile strength of the reinforced film is 7.66 ± 0.012 MPa. The maximum biodegradation rate of the film is 74.38 ± 0.012 at 35% glycerol concentration. Increasing the amount of glycerol increased the solubility of the biodegradable film due to its strong affinity towards water molecules. Glycerol increases film thickness and reduces its transparency. The microstructure results indicate that clay particles are homogeneously dispersed with reduced surface roughness of reinforced films compared to the control. The mango seed kernel starch film reinforced with 5% pottery clay and 25% glycerol was selected based on tensile strength, thermal stability, and biodegradability properties. In summary, the addition of glycerol and pottery clay significantly affects the physicomechanical and thermal properties of films.
Article
Multifunctional epoxy‐polydimethylsiloxane nanocomposite coatings with antifouling and anticorrosion characteristics have been developed via in situ polymerization method at different loading (1, 3, and 6.5 wt.%) of ZnO nanoparticles to cater marine applications. A detailed comparative analysis has been carried out between epoxy‐polydimethylsiloxane control (EPC) and ZnO‐reinforced coatings to determine the influence of ZnO loading on various properties. The incorporation of ZnO in EPC led to increase in root mean square (RMS) roughness to 126.75 nm and improved hydrophobicity showing maximum contact angle of 123.5° with low surface energy of 19.75 mN/m of nanocomposite coating as compared with control coating. The differential scanning calorimetry (DSC) result indicated improved glass transition temperature of nanocomposite coatings with highest Tg obtained at 83.69°C in case of 1 wt.% loading of ZnO. The increase in hydrophobicity of the system was accompanied by upgraded anticorrosion performance exhibiting 98.8% corrosion inhibition efficiency (CIE) as compared with control coating and lower corrosion rate of 0.12 × 10−3 mm/year. The Taber abrasion resistance and pull‐off adhesion strength results indicated an increment of 34.7% and 150.7%, respectively, in case of nanocomposite coating as compared with the control coating. The hardness of nanocomposite coatings was also improved, and maximum hardness was found to be 65.75 MPa for nanocomposite coating with 1 wt.% of ZnO. Our study showed that the nanocomposite coating was efficient in inhibiting accumulation of marine bacteria and preventing biofouling for more than 8 months. The developed environment‐friendly and efficient nanocomposite material has a promising future as a high‐performance anticorrosive and antifouling coating for marine applications.