Startseite Characterization of silicon acrylic resin containing silica nanoparticles as candidate materials for antifouling and anticorrosion properties in seawater
Artikel Öffentlich zugänglich

Characterization of silicon acrylic resin containing silica nanoparticles as candidate materials for antifouling and anticorrosion properties in seawater

  • Weibin Jiang , Quanliang Niu , Lin Cheng , Tao Zhou EMAIL logo und Huasheng Xie
Veröffentlicht/Copyright: 27. Juli 2020

Abstract

The damage caused by marine fouling organisms to ships and underwater artificial equipment is becoming increasingly serious issue, and the prevention and control of marine biofouling has always been a research hotspot in marine coatings. Aiming at the problems of poor adhesion, long curing time and high curing temperature of low-surface energy marine antifouling coatings of organosilicon, a hydrophobic low-surface energy nano-SiO2/silicon acrylic resin nanocomposite coating was synthesized. The anticorrosive property of the composite coatings was analyzed by simulated seawater periodic immersion experiments. The gel permeation chromatography analysis showed that polydimethyl-siloxanes (PDMS) is involved in cross-linking reactions. The dynamic thermomechanical analysis indicated that the glass transition temperature of resin is 58 °C. The contact angle (CA) test showed that the CA of nanocomposite coating is 109.99°. All the detection results can support the excellent antifouling and anticorrosion performance of the low surface energy nanocomposite coatings.

1 Introduction

Biofouling, namely adsorption and adhesion of marine organisms onto the ship hulls, poses a threat to the economic viability and health state of ships, by increasing the fuel consumption, accelerating corrosion and dysfunction of surface coatings (Yebra et al. 2004). This complex process usually involves undesirable accumulation of a multitude of species (up to 4000) on artificial surfaces when immersed in seawater (Zhang et al. 2016). Generally, the process is divided into two stages: micro and macrofouling. In the former stage, a biofilm is formed by micro-organism gathering. In the latter stage, larger organisms start to adhere (Buskens et al. 2012). The benefits and detriments of all measures used to avoid the presence of biofouling in seawater immersed structures (boats and pilons) are shown in Table 1.

Table 1:

The benefits and detriments of all measures used to avoid the presence of biofouling in seawater immersed structures.

No. Method Advantage Disadvantage
1 Mechanical removal Complete and thorough Inefficient
2 Electrolytic sea water Convenient High energy consumption
3 Electrolytic heavy metal Simple installation, less investment, low cost Consume heavy metals, environmental pollution
4 Antifoul paint High efficiency, low cost, easy to use poor adhesion construction difficulty

The formation process of the marine pollution is shown in Figure 1. Prevention of biofouling plays a key role in the marine related industries and there is an increasing demand to develop effective antifouling coatings (Al-Naamani et al. 2017; Gogoi et al. 2018; Sathya et al. 2016; Yesudass et al. 2017).

Figure 1: 
					The formation process of the marine pollution.
Figure 1:

The formation process of the marine pollution.

Among all remedy approaches, tributyltin self-polishing copolymer paints (TBT-SPC paints) have been considered as the most effective one. However, due to their ecotoxicological effects on non-target organisms it was banned by the International Maritime Organization (IMO) since 2008 (Callow and Callow 2011; Yee et al. 2017). Given this, great efforts have been made to find alternative antifouling materials. The non-toxic environmental coatings based on the low surface energy principle are very promising, which are currently widely used along with biocide-based self-polishing coatings (Callow and Callow 2011; Sathya et al. 2016). More specifically, the usage of antifouling paints based on organosilicon acrylic resin is of more attractive.

The siloxane has been widely applied in various fields because of its good thermal and oxidative stability, high gas permeability, excellent dielectric properties and physiological inertness or biocompatibility (Giudice and Benitez 1995; Hussain et al. 2019; Jiang et al. 2019; Yilgör and Yilgör 2014). In addition, polydimethyl-siloxanes (PDMS) also display very low surface tension values (around 21–22 mN/m) (Wu 2003), paving a new way for the development of low surface energy antifouling coatings. Recently, nanoparticles are introduced as antifouling biocides in polymer matrices. An amphiphilic membrane surface, derived from block copolymers bearing hydrophilic poly (ethylene oxide) (PEO) and low surface energy PDMS segments, is synthesized via surface segregation process (Zhao et al. 2014), which representing a new approach to design antifouling surfaces with controllable composition and properties. Perfluorodecyl trichlorosilane based poly (dimethylsiloxane)-ZnO (FDTS-based PDMS-ZnO) nanocomposite coating with anticorrosion and antifouling capabilities is prepared by using a one-step fabrication technique (Arukalam et al. 2016), demonstrating low adhesion strength, surface energies and outstanding anticorrosive properties. Among many systems of antifouling coatings, acrylic dispersion and nano-materials are widely available and became popular systems to satisfy the various users’ demands. The low-surface energy material PDMS was introduced into the acrylic resin to improve the hydrophobic and stability properties (Ammar et al. 2016a). The silicon dioxide (SiO2) nanoparticle or WO3 was added to raise the antifouling, anticorrosion and mechanical characters (Ammar et al. 2016b; Jiang et al. 2020; He et al. 2020; Qiao et al. 2016). However, little attention has been paid to the anticorrosion capability, thereby we aim to synthesize an antifouling coating having excellent antifouling and anticorrosion merits.

In this article, the silicone acrylic resin was synthesized by free radical solution polymerization firstly. Then nano-SiO2 was modified by KH-550. After that, the nanocomposite coating was prepared by adding modified nano-SiO2 and assistant into the silicone acrylic resin copolymer under ultrasonic condition. The organosilicon acrylic polymer (P(MA-MMA–HEMA–AA–ST–PDMS)) synthesized was characterized. In addition, the antifouling properties of the SiO2/silicon acrylic nanocomposite coating are investigated.

2 Materials and methods

2.1 Reagents

Methyl acrylate (MA) (98%) andacrylicacid (AA) (99%) were obtained from Tianjin Guangfu Fine chemical research institute (Tianjin, China). Methyl methacrylate (MMA) (98%), styrene (ST) (99.5%) and hydroxyethyl methyl acrylate (HEMA) (99%) were from Sinopharm Chemical Reagent Co., Ltd. (Shanghai China). PDMS (99%) from Dow Corning Chemicals, USA and hexamethylenediisocyanate biuret (HDI biuret) was purchased from MYAY Reagent (Zhejiang, China) and Bayer Material Science Co., Ltd. (China), respectively. While ethyl acetate (99.5%) and butyl acetate (99%) which were used as mixed solvents was provided by Tianjin Fuyu Fine Chemical Reagent Co., Ltd. (Tianjin, China). An organic flatting agent BYK-331 (polyether-modified dimethylpolysiloxane copolymer) and defoamer BYK-052 (polyvinyl butyl ether in Stoddard solvent) were supplied by BYK Chemie (Wesel, Germany). Benzoperoxide (BPO) and (3-Aminopropyl) triethoxysilane (APTES, 99%) which served as free radical initiator and modifying agent were purchased from Aladdin Chemistry Co., Ltd. (China). All the reagents were employed as received without further purification.

2.2 Synthesis of silicone-acrylic resin

Silicone-acrylic resin copolymer was prepared by free radical polymerization in a 250 mL four neck round glass reactor equipped with a reflux condenser, mechanical stirrer, nitrogen gas inlet and thermometer. Initially, 42.82 g of ethyl acetate, butyl acetate and PDMS mixture were added into the reactor and heated to 75 °C with continuous stirring at the rate of 300 rpm under nitrogen atmosphere. 30 min later, the reactor was heated to 105 °C, then the acrylic monomer mixture (34.92 g) mixed with evocating agent (0.377 g) were added into the reactor at a constant speed using a peristaltic pump. The polymerization reaction was conducted at 105 °C under azeotropic distillation until complete conversion of the acrylic monomer to yield a light opalescent transparent viscous solution. The composition of the medium resin and its physical characteristics are shown in Table 2.

Table 2:

Recipe and properties of the silicone acrylic resin.

Composition Weight (g) Function
MA 8.6 Cohesiveness, flexibility
MMA 10 Hardness
ST 10.4 Water resistance
HEMA 5.2 Viscosity
AA 0.72 Stickiness
PDMS 2.82 Hydrophobicity

2.3 Synthesis of modified SiO2 nanoparticles

Many nanoparticles are used in coating, such as nano SiO2, nano TiO2, nano ZnO, nano CaCO3, and so on. The SiO2 nanoparticles are widely used in polymer composites coating because the industrial production of nano SiO2 is currently the largest in the world. Moreover, adding nano SiO2 to the coating can form a shielding effect, resulting in enhance the anti-UV aging performance of the coating. Also, nano SiO2 has a three-dimensional mesh structure, which can form a mesh structure when the coating is dry, increasing the strength and gloss of the coating.

The amino-functionalized SiO2 was obtained through surface modification by using APTES ((3-Aminopropyl) triethoxysilane) as the silane coupling agent. 1 g of nanoparticles (10–80 nm) were dried in a vacuum drying oven under 100 °C for 12 h in advance and then dissolved in 22 mL of nitrogen methylpyrrolidone (NMP) under ultrasonic dispersion for 0.5 h. Then 5 mL of APTES and NMP mixed liquors were added into the beaker by dripping slowly under ultrasonic condition. After that, all the reactants were transferred into a three round bottom flask for 2 h under nitrogen atmosphere. After the reaction progress was completed, the modified SiO2 nanoparticle was separated by a centrifuge. After the modification, the as prepared SiO2 suspension exhibits unique properties such as higher molecular weight, less polar components and better hydrophobicity. Therefore, the suspension exhibits enhanced performance when dispersed in the organic polymers.

2.4 Preparation of the SiO2/silicon acrylic nanocomposite coating

The preparation of the SiO2/silicon acrylic nanocomposite coating was carried out in a 250 mL beaker at room temperature under ultrasonic condition. Initially the organosilicon acrylic resin and the curing agent HDI biuret were added with a ratio of 6:1. After 15 min, 0.5 wt% BYK-052 and 0.2 wt% BYK-331 were introduced into the reaction system to improve surface leveling, thus organosilicon acrylate was acquired. At the same time, 4.0 wt% modified nano-SiO2 nanoparticle was dispersed into the mixture. After the mixing process was carried out under ultrasonic condition for 0.5 h, the nanocomposite coating with wonderful dispersion was obtained.

2.5 Preparation of sample plates

Three rectangular aluminum plates (5 × 3 cm × 1 mm) were polished with abrasive sandpaper, followed by rinsing with deionized water and degreased with methanol. Then paint free plate was placed as blank sample and was named S1. Another two plates with coating on one side were deposited in a drying cabinet for 2 h before usage. The sample with silicone acrylic coating was named S2, while the nano-SiO2 (4.0 wt%) silicone acrylic coating was named S3. Each sample was painted with a brush and dried at room temperature for about one week to form a stable film with approximate 1 mm thickness.

2.6 Accelerated corrosion test in simulated seawater

To test the anticorrosion properties of the nanocomposite coating, all the samples were periodically immersed in the simulative marine environment. The simulated seawater is composed of 2.5% NaCl, 1.1% MgCl2, 0.40% Na2SO4 and 0.16% CaCl2 (Yamada et al. 2013). Before the simulated seawater periodic immersion experiment, the SEM analysis of the three samples was conducted to observe the surface morphology. Then, all samples were immerged into the simulated seawater for 8 h and exposed in the air for 16 h for a cycle and the experiments were carried out for 14 cycles (Cai and Liu 2006). After the experiment, three samples were tested again with scanning electron microscope. The corrosion resistance of the coating was detected by the surface morphology of the three samples before and after the experiments.

3 Results and discussion

3.1 Analysis of infrared spectra of silicone acrylic resin

The FT-IR spectra of silicon acrylic resin are showed in Figure 2. The peaks appearing around 3520 cm−1and 3025 cm−1 are the –OH and –CH telescopic vibration. The absorption bands at 2945 cm−1 can be assigned to the –CH2 symmetric stretching vibration (Smith 1998). The absorption bands at 1730 cm−1 is attributed to the CdbndO asymmetrical stretching vibration (Andrews 1992). The absorption band at 1451 cm−1 is associated with –CH3 bending vibration (Azémard et al. 2014). The characteristic peaks at 1261 and 988 cm−1 are ascribed to the –COOH stretching vibration (Kuanet al. 2005). The absorption bands at 1074 and 517 cm−1 are attributed to the Si–O asymmetrical stretching vibration (Ammar et al. 2016a). The absorption bands at 841 cm−1 is related to C–Si telescopic vibration (Ammar et al. 2016a). Finally, the absorption bands at 760 and 700 cm−1 are corresponds to the vibrations of aromatic groups (Su and Chang 2003).

Figure 2: 
						FT-IR spectra of silicon acrylic resin.
Figure 2:

FT-IR spectra of silicon acrylic resin.

To prove that PDMS is involved in cross-linking reactions, the comparative experiments are designed. Experimental procedure is the same as section 2.2. The resins were synthesized with adding PDMS (labeled as Resin 1) and without PDMS (labeled as Resin 2). From the gel permeation chromatography analysis to Resin 1 and Resin 2, it can be known from Table 3 that all molecular weight (including weight-average molecular weight, number-average molecular weight, higher average molecular weight and peak molecular weight) of Resin 1 is larger than those of Resin 2. Combined with FT-IR analysis, it can be concluded that PDMS participated in the cross-link reaction.

Table 3:

Molecular weight of Resin 1 and Resin 2.

Formula Weight-average molecular weight Number-average molecular weight Peak molecular weight Higher average molecular weight
Resin 1 21068 9456 11038 47410
Resin 2 17118 8991 10546 31130

3.2 FT-IR spectra of nano-SiO2 before and after modification with KH-550

The FT-IR spectra of nano-SiO2 before and after modification with KH-550 are displayed in Figure 3. In the IR spectra of the nano-SiO2 before modification with KH-550 (Figure 3a), the peaks around 3447 cm−1 can be assigned to the Si-OH telescopic vibration. The absorption bands at 1087 cm−1 are attributed to Si–O–Si antisymmetric stretching vibration. The absorption bands at 780 cm−1 can be assigned to the Si–O–Si symmetrical stretching vibration. There are also characteristic absorption bonds of Si–O–Si bending vibration at 487 cm−1. These groups are also seen in the FT-IR spectra of nano-SiO2 after modification with KH-550 (Figure 3b). However, the intensity of absorption bands at 3438 cm−1are obviously enhanced because of the simultaneous appearance of –NH2 and Si–OH absorption peaks. In addition, new characteristic peaks are also appeared on the spectral line b. The peak around 2982 cm−1 is –CH2 stretching vibration. The absorption bands at 1397 cm−1 are associated with C–N stretching vibration. The absorption bands at 1172 cm−1 are C–C stretching vibration. The absorption bands at 864 cm−1 are C–Si stretching vibration. All the presented data indicates the successful modification of nano-SiO2 by KH-550.

Figure 3: 
						FT-IR spectra of nano-SiO2 (a) before modification with KH-550, (b) after modification with KH-550.
Figure 3:

FT-IR spectra of nano-SiO2 (a) before modification with KH-550, (b) after modification with KH-550.

3.3 Test of glass transition temperature

The glass transition temperature (Tg) is one of the characteristics of polymer materials, and the transformation temperature of non-stereotyped polymer from glass state to high-bounce state is the minimum temperature of free movement of the non-stereotyped polymer macromolecule chain segment, which has an important influence on the performance and processing process of polymer materials. In general, Tg value for seawater applications should be controlled at 55 to 60 °C.

The differential scanning calorimetry was used to obtain the glass transition temperature of the silicone acrylic resin to determine the phase diagrams and mixtures of polymer systems. The test was carried out under nitrogen flow at atmospheric pressure in a temperature range of 20 to 500 °C. In this temperature range three characteristic temperatures can be detected. According to Figure 4a, the glass transition temperature of organosilicon acrylic resin is in the range of 50–66 °C (relatively flat exothermic section). This is due to the resin appeared transition from the glassy state to the high elastic state in the temperature range. It can be seen from TG-DTG curve that the resin does not appear weight loss within the temperature range, and the exothermic section analysis shows that the glass transition temperature of the resin is about 58 °C. The endothermic peak appears at 164 °C. Under this temperature, the silicone acrylic resin copolymers show a weight loss of 1.8% (Figure 4b), which is due to the solvent evaporation (Zhang et al. 2016). The sample weight drops from 91.5% to 7.5%, a drop of 84% occurs between 327 and 447 °C, among which a conspicuous endothermic peak appears at 390 °C. This is because most polymers begin to break down at this temperature range. It can be concluded that the resin copolymers exert some thermal instability and indicate an endothermic small shift around 58 °C, which is assigned to glass transition temperature (Tg).

Figure 4: 
						(a) Differential scanning calorimetry and (b) TG-DTG of silicon acrylic resin.
Figure 4:

(a) Differential scanning calorimetry and (b) TG-DTG of silicon acrylic resin.

3.4 Contact angle (CA) measurements of applied resins

The CA to water can be used to evaluate the wetting of the solid surface. The greater the CA, the stronger the hydrophobic. In this work, the water CA measurements were introduced to investigate the hydrophobic character of the developed coatings (Ammar et al. 2016b). The CAs of four droplets of distilled water at different points of each sample was recorded by using JC2000 CA instrument. 5 μL water droplet was gently deposited on the sample surface and capturing the image was occurred directly in terms of measuring the static CA. The reported CA values were the averages of four measurements with less than 2° as a measurement error. The multiple averaging method is adopted to reduce CA measurement error. The CA of silicon acrylic resin coatings and nanocomposite coatings are measured in four different positions respectively. Figure 5 depicts CA values of silicon acrylic resin coating plate (Figure 5a) and nanocomposite coating plate (Figure 5b). It is well known that nano-SiO2 is easy to agglomerate because of its high surface energy and strong surface polarity. In this experiment, nanometer SiO2 was modified to reduce its surface energy, prevent its agglomeration and improve its affinity of organic phase. The experimental results showed that the CA values of silicon acrylic resin coating plate and nanocomposite coating plate were 108.5° and 110°, respectively. According to the experimental results, the surface of the coating can be maintained to some extent by the introduction of modified nano-SiO2 in silicon acrylic resin. The similar observations were previously reported by researchers introducing SiO2 nanoparticles into polymeric matrix (Wang et al. 2011). It is known that the antifouling effect of the coating display only the CA at above 98°. According to the CA test results, both silicon acrylic resin coating and nanocomposite coating possess the characteristics of preventing marine biofouling.

Figure 5: 
						Contact angle (CA) of (a) silicon acrylic resin coating plate and (b) nanocomposite coating plate.
Figure 5:

Contact angle (CA) of (a) silicon acrylic resin coating plate and (b) nanocomposite coating plate.

3.5 SEM image of coatings against seawater corrosion

SEM micrographs of all prepared sample plates show the efficiency of coatings against seawater corrosion (Figure 6). It was found that the sample plate S1 was badly corroded by simulated seawater with a cavity over the surface (S1 in Figure 6), and the sample plate S3 was slightly corroded and the sample plate S2 was scarcely corroded. Obviously, paint plays a key role in the resistance of seawater corrosion (S3 comparing to S1). In addition, the anticorrosion properties of the coating were significantly enhanced with the addition of nano-SiO2 (S3 comparing to S2). According to the SEM images of all the sample plates, it can be concluded that the nanocomposite coatings have excellent corrosion resistance to seawater.

Figure 6: 
						SEM images of three sample plates before and after the accelerated corrosion test.
Figure 6:

SEM images of three sample plates before and after the accelerated corrosion test.

3.6 Influence of silicon monomer content on film performance

The influence of the silicon monomer content on the wettability of the resulting coated surfaces was shown in Figure 7. The incorporation of silicone monomer into the acrylic resin results in an increase in the CA at 104° compared to the uncoating (CA at 89°), which shows that PDMS within the polymeric coating has the ability to enhance hydrophobicity. It is because the PDMS possesses a significant low surface energy and the ability to produce hydrophobic surfaces. This leads to changes in the chemical composition due to the presence of the (–Si–O–Si–) silicone. However, from the experimental results, it was found that the CA was not increasing further when the content of the PDMS exceeds a certain value. This consequence can be ascribed to the saturation of the grafting of PDMS on the main chain of the polymer. Finally, the mass of the PDMS selected was 2.82 g.

Figure 7: 
						The water CA values of different PDMS content in acrylic resin. (1) 0 g PDMS (2) 0.94 g PDMS (3) 1.88 g PDMS (4) 2.82 g PDMS (5) 3.76 g PDMS.
Figure 7:

The water CA values of different PDMS content in acrylic resin. (1) 0 g PDMS (2) 0.94 g PDMS (3) 1.88 g PDMS (4) 2.82 g PDMS (5) 3.76 g PDMS.

The antifouling tests were not performed yet, but we are planning to measure the antifouling properties in future.

4 Summary

The low-surface energy nanocomposite coating with high hydrophobic, anticorrosion and excellent adhesion for resisting marine biofouling was prepared by mixing a low surface energy silicone acrylic resin and the modified nano-SiO2 to improve poor adhesion and difficulty in application of the existing commercial coatings. The comparative experimental results showed that PDMS is involved in cross-linking reactions. The infrared spectra indicated the successful grafting of silicon monomers on the main chain of the acrylic resin copolymers and the successful modification of nano-SiO2. The coating surface measured by the CA tester indicated high hydrophobicity. In addition, enhanced anticorrosion and hydrophobicity were observed after nano-SiO2 was added.


Corresponding author: Tao Zhou, Hunan Provincial Key Laboratory of Efficient and Clean Utilization of Manganese Resources, College of Chemistry and Chemical Engineering, Central South University, Changsha, 410083, Hunan, China, E-mail:

Quanliang Niu: Joint first author.


Funding source: Hunan Provincial Science and Technology Department

Award Identifier / Grant number: 216TP1007

Funding source: National Natural Science Foundation of China

Award Identifier / Grant number: 21376269

  1. Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: The authors acknowledge with gratitude the financial support of the National Natural Science Foundation of China (no. 21376269) and the project supported by the Hunan Provincial Science and Technology Plan, China (no. 2016TP1007).

  3. Conflicts of interest: The authors declare no conflicts of interest regarding this article.

References

Al-Naamani, L., Dobretsov, S., Dutta, J., and Burgess, J.G. (2017). Chitosan-zinc oxide nanocomposite coatings for the prevention of marinebiofouling. Chemosphere 168: 408–417, https://doi.org/10.1016/j.chemosphere.2016.10.033.Suche in Google Scholar PubMed

Ammar, S., Ramesh, K., Vengadaesvaran, B., Ramesh, S., and Arof, A.K. (2016a). Amelioration of anticorrosion and hydrophobic properties of epoxy/PDMS composite coatings containing nano ZnO particles. Prog. Org. Coat. 92: 54–65, https://doi.org/10.1016/j.porgcoat.2015.12.007.Suche in Google Scholar

Ammar, S., Ramesh, K., Vengadaesvaran, B., Ramesh, S., and Arof, A.K. (2016b). A novel coating material that uses nano-sized SiO2 particles to intensify hydrophobicity and corrosion protection properties. Electrochim. Acta 220: 417–26, https://doi.org/10.1016/j.electacta.2016.10.099.Suche in Google Scholar

Andrews, D.H. (1992). Progress in low temperature physics. Phys. Today 22: 107–109, https://doi.org/10.1063/1.3035332.Suche in Google Scholar

Arukalam, I.O., Oguzie, E.E., and Li, Y. (2016). Fabrication of FDTS-modified PDMS-ZnO nanocomposite hydrophobic coating with anti-fouling capability for corrosion protection of Q235 steel. Colloid. Interf. Sci. 484: 220–228, https://doi.org/10.1016/j.jcis.2016.08.064.Suche in Google Scholar PubMed

Azémard, C., Vieillescazes, C., and Ménager, M. (2014). Effect of photodegradation on the identification of natural varnishes by FT-IR spectroscopy. Microchem. J. 112: 137–149, https://doi.org/10.1016/j.microc.2013.09.020.Suche in Google Scholar

Buskens, P., Wouters, M., Rentrop, C., and Vroon, Z. (2012). A brief review of environmentally benign antifoulingand foul-release coatings for marine applications. J. Coat. Technol. Res. 10: 29–36, https://doi.org/10.1007/s11998-012-9456-0.Suche in Google Scholar

Cai, J.P. and Liu, M. (2006). Novel accelerated corrosion test for LY12CZ and LC4CS aluminum alloys. Trans. Nonferr. Met. Soc. China. 16: s1415–s1418.Suche in Google Scholar

Callow, J.A. and Callow, M.E. (2011). Trends in the development of environmentally friendly fouling-resistant marine coatings. Nat. commun. 2: 244, https://doi.org/10.1038/ncomms1251.Suche in Google Scholar PubMed

Giudice, C.A., Benitez, J.C. (1995). Coatings for corrosion protection in seawater structures. Corros. Rev. 13: 81–190. https://doi.org/10.1515/corrrev.1995.13.2-4.81.Suche in Google Scholar

Gogoi, B., Barua, S., Sarmah, J.K., and Karak, N. (2018). In situ synthesis of a microbial foulingresistant, nanofibrillar cellulose-hyperbranched epoxy composite for advanced coating applications. Prog. Org. Coat. 124: 224–231, https://doi.org/10.1016/j.porgcoat.2018.04.025.Suche in Google Scholar

He, Y.X., Wu, D.Y., Zhou, M.Y., Liu, H., Zhang, L., Chen, Q.Y., Yao, B.H., Yao, D.H., Jiang, D.F., Liu, C.T., et al. (2020). Effect of MoO3/carbon nanotubes on friction and wear performance of glass fabric-reinforced epoxy composites under dry sliding. Appl. Surf. Sci. 506: 144946, https://doi.org/10.1016/j.apsusc.2019.144946.Suche in Google Scholar

Hussain, A.K., Sudin, I., Basheer, U.M., and Mohd Yusop, M.Z. (2019). A review on graphene-based polymer composite coatings for the corrosion protection of metals. Corros. Rev. 37: 343–363, https://doi.org/10.1515/corrrev-2018-0097.Suche in Google Scholar

Jiang, W.B., Dai, A.Q., Zhou, T., and Xie, H.S. (2019). Hybrid polysiloxane/polyacrylate/nano-SiO2 emulsion for waterborne polyurethane coatings. Polym. Test. 80: 106110, https://doi.org/10.1016/j.polymertesting.2019.106110.Suche in Google Scholar

Jiang, W.B., Jin, X.F., Li, H., Zhang, S.Y., Zhou, T., and Xie, H.S. (2020). Modification of nano-hybrid silicon acrylic resin with anticorrosion and hydrophobic properties. Polym. Test. 82: 106287, https://doi.org/10.1016/j.polymertesting.2019.106287.Suche in Google Scholar

Kuan, H.C., Ma, C.C.M., and Chang, W.P. (2005). Synthesis, thermal, mechanical and rheological properties of multiwall carbon nanotube/waterborne polyurethane nanocomposite. Compos. Sci. Technol. 65: 1703–1710, https://doi.org/10.1016/j.compscitech.2005.02.017.Suche in Google Scholar

Qiao, B., Liang, Y., Wang, T.J., and Jiang, Y. (2016). Surface modification to produce hydrophobic nano-silica particles using sodium dodecyl sulfate as a modifier. Appl. Surf. Sci. 364: 103–109, https://doi.org/10.1016/j.apsusc.2015.12.116.Suche in Google Scholar

Sathya, S., Murthy, P.S., Das, A., Gomathi Sankar, G., Venkatnarayanan, S., and Pandian, R. (2016). Marine antifouling property of PMMA nanocomposite films: results of laboratory and field assessment. Int. Biodeter. Biodegr. 114: 57–66, https://doi.org/10.1016/j.ibiod.2016.05.026.Suche in Google Scholar

Smith, B. (1998). Infrared spectral interpretation: a systematic approach. CRC Press, Boca Raton.Suche in Google Scholar

Su, Y.C. and Chang, F.C. (2003). Synthesis and characterization of fluorinated polybenzoxazine material with low dielectric constant. Polymer 44: 7989–7996, https://doi.org/10.1016/j.polymer.2003.10.026.Suche in Google Scholar

Wang, S., Li, Y., Fei, X., Sun, M., Zhang, C., and Li, Y. (2011). Preparation of a durable superhydrophobic membrane by electrospinning poly (vinylidene fluoride) (PVDF) mixed with epoxy-siloxane modified SiO2 nanoparticles: a possible route to superhydrophobic surfaces with low water sliding angle and high water contact angle. J. Colloid. Interf. Sci. 359: 380–388, https://doi.org/10.1016/j.jcis.2011.04.004.Suche in Google Scholar PubMed

Wu, S. (2003). Surface and interfacial tensions of polymers, oligomers, plasticizers, and organic pigments. In: Book: the Wiley database of polymer properties. John Wiley & Sons, pp. 125–138.Suche in Google Scholar

Yamada, N., Suzumura, M., Koiwa, F., and Negishi, N. (2013). Differences in elimination efficiencies of escherichia coli in freshwater and seawater as a result of TiO2 photocatalysis. Water Res. 47: 2770–2776, https://doi.org/10.1016/j.watres.2013.02.023.Suche in Google Scholar PubMed

Yebra, D.M., Kiil, S., and Dam-Johansen, K. (2004). Antifouling technology—past, present and future steps towards efficient and environmentally friendly antifouling coatings. Prog. Org. Coat. 50: 75–104, https://doi.org/10.1016/j.porgcoat.2003.06.001.Suche in Google Scholar

Yee, M.S.L., Khiew, P.S., Lim, S.S., Chiu, W.S., Tan, Y.F., and Kok, Y.Y. (2017). Enhanced marine antifouling performance of silver-titania nanotube composites from hydrothermal processing. Colloid. Surface. A. 20: 701–711, https://doi.org/10.1016/j.colsurfa.2017.02.034.Suche in Google Scholar

Yesudass, S.A., Mohanty, S., Nayak, S.K., and Rath, C.C. (2017). Zwitterionic–polyurethane coatingsfor non-specific marine bacterial inhibition: a nontoxic approach for marine application. Eur. Polym. J. 96: 304–315.10.1016/j.eurpolymj.2017.09.019Suche in Google Scholar

Yilgör, E. and Yilgör, I. (2014). Silicone containing copolymers: synthesis, properties and applications. Prog. Polym. Sci. 39: 1165–1195, https://doi.org/10.1016/j.progpolymsci.2013.11.003.Suche in Google Scholar

Zhang, J., Pan, M., Luo, C., Chen, X., Kong, J., and Zhou, T. (2016). A novel composite paint (TiO2/fluorinated acrylic nanocomposite) for antifouling application in marine environments. J. Environ. Chem. Eng. 4: 2545–2555, https://doi.org/10.1016/j.jece.2016.05.002.Suche in Google Scholar

Zhao, X., Su, Y., Li, Y., Zhang, R., Zhao, J., and Jiang, Z. (2014). Engineering amphiphilic membrane surfaces based on PEO and PDMS segments for improved antifouling performances. J. Membrane. Sci. 450: 111–123, https://doi.org/10.1016/j.memsci.2013.08.044.Suche in Google Scholar

Received: 2019-10-27
Accepted: 2020-05-31
Published Online: 2020-07-27
Published in Print: 2020-08-27

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 2.10.2025 von https://www.degruyterbrill.com/document/doi/10.1515/corrrev-2019-0091/html
Button zum nach oben scrollen