Next Article in Journal
Enhancing Flame Resistance of Cellulosic Fibers Using an Ecofriendly Coating
Previous Article in Journal
Computational Simulations of Nanoconfined Argon Film through Adsorption–Desorption in a Uniform Slit Pore
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

APTES Modification of Molybdenum Disulfide to Improve the Corrosion Resistance of Waterborne Epoxy Coating

1
State Key Lab of Oil and Gas Reservoir Geology and Exploitation, Southwest Petroleum University, Rd. 8, Xindu District, Chengdu 610500, China
2
College of Chemistry and Chemical Engineering, Southwest Petroleum University, 8 Xindu Avenue, Xindu District, Chengdu 610500, China
3
Chengdu Evermaterials Tec Co, Chengdu 610500, China
*
Authors to whom correspondence should be addressed.
Submission received: 7 January 2021 / Revised: 25 January 2021 / Accepted: 27 January 2021 / Published: 2 February 2021

Abstract

:
MoS2 has been regarded as a promising addition for the preparation of epoxy-based coatings with high anticorrosion ability. However, its dispersion and compatibility remain significant challenges. In the present work, an organic thin layer was well coated on lamellar molybdenum disulfide (MoS2) via a simple modification of 3-aminopropyltriethoxysilane (APTES). The modification of hydrolyzing APTES on lamellar MoS2 effectively improved the dispersity of MoS2 in water-borne epoxy (WEP) and successfully enhanced the compatibility and crosslinking density of MoS2 with WEP. The influence of introducing MoS2-APTES into WEP coating on anticorrosion property for N80 steel was tested by electrochemical impedance spectroscopy (EIS), potentiodynamic polarization and salt spray test. The results exhibited that the |Z|0.01Hz value of MoS2-APTES/WEP still reached 3.647 × 107 Ω·cm2 even after the immersion time of 50 days in 3.5 wt.% NaCl solution, showing an extraordinary performance of corrosion resistance. The enhanced anticorrosion performance of composite coating could be resulted from the apparently increased dispersibility and compatibility of MoS2 in WEP.

1. Introduction

Metal corrosion has become a worldwide problem causing enormous economic losses and many casualties [1,2]. Among the conventional anticorrosive strategies, organic coatings have generated tremendous attention from researchers and have been extensively used in industrial fields, mainly due to their superior physical barrier abilities and relatively low costs [3,4,5]. However, solvent-based coatings containing many volatile organic compounds (VOCs) will heavily damage the eco-environment and human health [6,7]. Waterborne epoxy coating (WEP), one of the environment-friendly coatings, has drawn an increasing amount of attention owing to its many virtues like low noxiousness and excellent chemical resistance [8,9,10]. Unfortunately, WEP coatings could not attain satisfactory results of protecting metal from corrosion due to the formation of micro-pores or micro-cracks during coating curing caused by solvent evaporation [11]. To solve this problem, some methods have been developed to reinforce the anticorrosive capability of the WEP coatings. Among these diverse methods, various fillers such as graphene and its derivates, C3N4, and h–BN, etc., have been introduced into the waterborne epoxy, which is an effective way to reduce coating defects and promote the anticorrosion performance of WEP coatings [12].
To date, graphene, one of the two-dimensional (2D) materials, has already generated an increasing amount of interest among researchers in the area of anticorrosion, because of its superior impermeability to molecules, excellent chemical inertness, hydrophobicity, outstanding thermal and mechanical properties, etc. [13]. However, it is also well known that graphene cannot function as a long-term anticorrosion material, and it even accelerates metal corrosion once the coating is scratched [14]. This is a tremendous limit to graphene applied to anticorrosion. Even though previously relevant “self-healing” strategies can extend the life of graphene-based coatings, they cannot basically hinder the “corrosion-promotion activity”. Consequently, it is particularly crucial to hinder the corrosion-promotion activity of graphene-based anticorrosion coatings or find low conductivity/insulation materials [9,15].
MoS2, as a member of the transition metal dichalcogenides, has attracted a significant interest due to its 2D ultrathin atomic layer structure similar to that of graphene [16,17]. It is made up of one plane of hexagonally packed metal atoms sandwiched by two planes of chalcogenide atoms. The sandwich layers stack vertically and are loosely bonded via weak van der Waals forces [18,19]. The special lamellar structure of MoS2 endows it with excellent anticorrosion performance-like graphene [20]. Meanwhile, MoS2 possesses a high band gap in contrast to graphene, which has no influence on the electrical conductivity of the coatings when added with the MoS2 [21,22]. As a result, MoS2-based anticorrosion coatings will not cause “micro-galvanic corrosion” if the defects occur on the surface of the compounded coatings [23]. However, MoS2 has a tendency to agglomerate in the polymer matrix, because it possesses a large surface area and is short of functional groups. To improve its dispersibility and compatibility with the polymer matrix, the most effective methods to prepare high-performance composites is the surface modification of MoS2 including the non-covalent and covalent types [24]. For example, Wang et al. [25] found a non-covalently functionalized method to modify MoS2 with chitosan (CS), and the CS–MoS2 could be homogeneously dispersed in tetrahydrofuran, which could greatly decrease hazards connected with EP nanocomposites. Zhou et al. [26] devised an approach of covalent functionalization to graft MoS2 nanosheets with 3-mercaptopropyl trimethoxysilane and the functionalized MoS2 could be well dispersed in PVA without obvious aggregation.
Moreover, it is generally acknowledged that organic silane coupling agents could be used to modify materials. In this regard, Sepideh et al. [27] used (3-aminopropyl) triethoxysilane (APTES)-modified graphene oxide (GO) as a nanofiller to add into the epoxy coating, and the results showed that corrosion protection performance greatly improved because of the excellent interfacial interaction of GO in coating through silane modification. Sepideh et al. [28] reported 3-aminopropyl triethoxysilane (APTES) and 3-glysidyloxypropyl trimethoxysilane (GPTMS) as silane agents to modify GO sheets, which were introduced to epoxy coatings. The results exposed the incorporation of APTES-GO including amine end-groups, which better increased the corrosion resistance of epoxy coating. As we all know, APTES, a kind of typical silane coupling agent, could hydrolyze and graft to the surface of fillers, and its amine-end groups could improve the crosslinking density of epoxy resins.
In this study, MoS2 was functionalized with APTES in order to develop its dispersion and interfacial interactions with waterborne epoxy. Then, the hybrids MoS2-APTES were introduced into waterborne epoxy resin coatings to strengthen the anticorrosion performance. To the best of our knowledge, there is no study reporting the anticorrosion performance of APTES-modified MoS2 in waterborne epoxy. Finally, the corrosion resistance of the composite coatings was investigated by electrochemical impedance spectra (EIS), potentiodynamic polarization and salt spray test.

2. Experiment

2.1. Materials

Molybdenum disulfide (MoS2, 99.9%) was obtained from McLean Biotech (Shanghai, China), and 3-aminopropyltriethoxysilane (APTES, 99%) and ethanol (≥99.7%) were purchased from Kelong Chemical Reagent Factory (Chengdu, China). Waterborne epoxy (BC2060H) and curing agent (BC916) provided by LianGu New Material Technology Co., Ltd. (Guangdong, China) were used to prepare the coatings. Distilled water was produced by the Milli-Q system (Millipore, NY, USA). All the reagents used in the experiment were directly utilized without other purification steps.

2.2. Preparation of MoS2-APTES

APTES was employed to modify MoS2 as follows: 0.3 g MoS2 was dispersed in 150 mL ethanol including 3 g APTES by sonification for 40 min, and then 5 mL of distilled water was dropped in the prepared suspension under sonification. The suspension was magnetically stirred at 500 rpm for 24 h at 80 °C. Then, modified MoS2 was centrifuged, washed several times with pure water and ethanol to insure the complete removal of unreacted APTES. In the end, the acquired products were dried by freeze-drying. The schematic of the synthesis procedure of MoS2-APTES was displayed in Figure 1.

2.3. Fabrication of MoS2-APTES Composite Coating

Firstly, a certain amount of MoS2 (0.3 wt.%) was dispersed in pure water, stirred and ultrasonicated for 5 min to obtain a homogeneous solution. Subsequently, the MoS2-APTES suspension was added into waterborne epoxy resin containing curing agent and mechanically stirring for 10 min to produce a homogeneous mixture. Finally, the mixture was sprayed on the prepared N80 steel substrates degreased by ethanol and acetone. The composite coating was placed at 25 °C for 7 days to promote curing. The average thickness of the coating was measured by the coating thickness gauge (QNIX 4500, Qnix, Bonn, Germany), which was about 40.0 ± 5 μm. Similarly, The WEP and MoS2/WEP were also prepared via the same method.

2.4. Characterization

Fourier transform infrared (FT-IR, WQF-520, Rui Li, Beijing, China) was used to initially characterize the chemical structure of the composite in the range of 500–4000 cm−1. To investigate the crystal structure of materials, X-ray diffraction (XRD, Bruker D8 ADVANCE A25X, Beijing, China) was carried out using copper Ka radiation source at a scan rate of 1°·s−1 from 5° to 80° and X-ray diffraction measurement was in Bragg–Brentano geometry. The morphology of MoS2-APTES was characterized by transmission electron microscopy (TEM, JEM-2100F, Japan Electron Optics Laboratory Co, Tokyo, Japan). X-ray photoelectron spectrometry (XPS, XSAM800, Al Kα ray, KRATOS, Manchester, UK) was employed to explore the surface chemistry. Thermogravimetric analysis (TGA) was conducted by TGA (SDTA851, METTLER, Zurich, Switzerland) on the condition that N2 atmosphere flowed over the open crucibles and a heating temperature increased from 40–800 °C with a rate of 10 °C/min. In order to study the dispersion ability of additions in WEP coatings, scanning electron microscopy (SEM, JSM-6700, JEOL, Tokyo, Japan) was utilized to survey the fracture surface of coatings after breaking the coatings in liquid nitrogen.

2.5. Corrosion Measurements

2.5.1. Electrochemical Test

The corrosion protection ability of the neat WEP, MoS2/WEP, MoS2-APTES/WEP was investigated by electrochemical impedance spectra (EIS) measurements which were from the CorrTest CS350 in 3.5 wt.% NaCl solution. Three traditional electrode systems including a working electrode (specimen coatings), reference electrode (saturated calomel electrode) and counter electrode (foil electrode) were applied in the measurement process. During the EIS test, the parameter of the frequency range was 10−2~105 Hz and the open circuit potential (OCP) always remains stable. In addition, the test data were fitted and evaluated by the ZsimpWin software. Additionally, the potentiodynamic polarization test was employed to further study the anticorrosive performance of the composite coating, which was conducted with a scan rate of 1 mV·s−1 from −200 mV (the cathodic direction) to +200 mV the anodic (direction).

2.5.2. Salt Spray Test

The long-term anticorrosion performance was estimated by conducting salt spray testing. According to the ASTM B117 standard, samples with a 4 cm scratch were exposed to 5 wt.% NaCl salt spray.

3. Results and Discussion

3.1. Characterization

To confirm the successful modification of APTES on MoS2, FTIR spectroscopy of MoS2, MoS2-APTES was displayed in Figure 2a. From the figure, there is no obvious absorption peak in the spectrum of original MoS2, because of the lack of functional groups on the surface of MoS2 [29]. In the case of MoS2-APTES, some peaks at 2917 and 2850 cm−1 are associated with the symmetric and asymmetric stretching of C–H [27]. Moreover, the broad absorbance between 3436 and 3308 cm−1 could be related to O–H stretching of water and N–H stretching vibrations. These prove the resultful modification of MoS2 with APTES. Furthermore, new bands at 1624, 1117, 1042, 912, 755, cm−1 are connected with the deformation (scissoring) of N–H, perpendicular Si–O stretching vibrations, in-plane Si–O stretching, O–H deformation of hydroxyl groups and perpendicular Si–O stretching vibrations, respectively [30]. All the above indicate that APTES has been successfully modified to the surface of MoS2.
XRD spectra were employed to investigate the crystalline structure of MoS2 and MoS2-APTES. As shown in Figure 2b, the crystalline diffraction peaks of MoS2 appear at 14.36°, 32.88° and 58.02°, which correspond to (0:0:2), (1:0:0) and (1:1:0) planes, respectively [31]. After modification with APTES, the hybrids exhibit no peak shift in contrast to MoS2, which manifests that there is no influence on the crystalline phrase of MoS2 via surface modification. Moreover, it is clear that the intensity of the (0:0:2) diffraction peaks of the MoS2-APTES hybrids obviously decreased compared to that of MoS2, which can probably be attributed to the thin layer of APTES hindering the aggregation of MoS2. Overall, these results demonstrate that APTES is successfully grown on the surface of MoS2.
In order to explore the morphology of MoS2 and MoS2-APTES, TEM was carried out, and the images are displayed in Figure 3. From Figure 3a, it could be clearly spotted that the average lateral size of MoS2 is close to 500 nm. According to the high magnification TEM image of MoS2-APTES (Figure 3b), a thin film of approximately 5 nm enveloping the surface of the particle can be seen, which proves that APTES successfully coats on the surface of MoS2 via hydrolysis reaction.
To investigate the chemical element composition of MoS2 modified with APTES, the XPS spectra analysis was performed. The wide-scan spectra of MoS2 and MoS2-APTES are exhibited in Figure 4a. The existences of Mo, S, C and O elements are spotted in both MoS2 and MoS2-APTE samples, and Si and N elements are detected from MoS2-APTES, revealing the successful incorporation of APTES on the surface of MoS2. The peak at 395.6 eV may be overlap between N1s and Mo 3p. The C and O elements of pristine MoS2 may be from the ambient atmosphere. Additionally, the apparent raised C1s and O1s intensity compared with the MoS2 mainly derives from the APTES, which indicates that APTES was also successfully introduced to MoS2. As shown in Figure 4b, the high-resolution spectrum of C1s for MoS2-APTES can be deconvoluted into four components situated at 283.60, 284.4, 285.3 and 286.4 eV, which correspond to C–Si, C–C, C–N and C–O bonds, respectively [32]. In addition, the presence of C–N and C–C derives from APTES, while C–Si is probably created by the dehydration condensation of Silanol [33]. Figure 4c displays the high-resolution Si 2p spectrum which can be split into three peaks. The two peaks at 102.3 and 103.2 eV are observed, which correspond to the Si–O–C and Si–O–Si, respectively [11]. The third peak located at 101.8 eV results from the Si–C bond. From the high-resolution Si 2p spectrum, it could be concluded that the silane coupling agent APTES hydrolyzes during the process of modification, and the produced silanol could undergo dehydration condensation by itself. In addition, the Mo 3d spectrum (Figure 4c) of MoS2-APTES shows five characteristic peaks. The two peaks located at Mo 3d5/2 (229.2 eV) and Mo 3d3/2 (232.4 eV) are attributed to characteristics of Mo4+, which prove the signs of MoS2. In the meantime, the peak appeared at 226.3 eV, which is the result of the strong overlap of the S2s and Mo 3d region. The peaks at 233.3 eV of Mo 3d5/2 and at 235.7 eV of Mo 3d3/2 are characteristic of Mo6+ species, which signify the crosslinking systems between MoS2 and APTES [34]. All of these XPS features reveal the successful modification of APTES on the surface of MoS2.
The thermal stability of MoS2 and modified MoS2 was studied via TGA from 40 °C to 800 °C under N2 atmosphere. The TGA curves of the initial MoS2 and MoS2-APTES hybrids are exhibited in Figure 5. There is no obvious loss of weight during the thermal decomposition procedure of the original MoS2, manifesting its excellent thermally stability in a nitrogen atmosphere [24]. For MoS2-APTES, an enormous amount of decline in weight is attributed to the thermal decomposition of APTES on the surface of MoS2 and the evaporation of the residual water. At the same time, the TGA curves also prove that the amount of APTES in MoS2-APTES hybrids is about 19.4 wt.%.
The fracture surface of different coatings was investigated by SEM and the images are displayed in Figure 6. For the neat WEP coating (Figure 6a), the fracture surface of it is relatively smooth [35]. After adding MoS2 into coating, it could be clearly spotted that the fracture surface of composite coating (Figure 6b) gets tough. However, some serious agglomeration of MoS2 can be observed, because the pristine MoS2 cannot effectively disperse in WEP resin and lacks interfacial bonding between MoS2 and resin [36]. In contrast with these, there is no obvious agglomeration in the MoS2-APTES/WEP composite coating (Figure 6c) and the MoS2 after the modification of APTES is well dispersed in waterborne epoxy resin, which highlights the better compatibility between MoS2-APTES and WEP. This still suggests that the MoS2-APTES/WEP composite coating can provide steel with superior corrosion resistance performance to corrosive media.

3.2. EIS Test

EIS was employed to investigate the corrosion resistance performance of the waterborne epoxy coatings at the steady open circuit potentials after an increasing amount of immersion time. The achieved studying data were drawn as Nyquist and Bode diagrams through the immersion time for 5, 15, 30, and 50 days. From Figure 7a,c,e, it could be obviously spotted that with the extending immersion time, the impedance diameters of all the epoxy coatings expose a decline in the Nyquist plots, which means that the anticorrosion property of the coatings is dwindling. As shown in Figure 7a,b, the pure WEP coating only has a semi-arc and one time constant when the immersion time is at 5 days. However, as the time extends to 15 days, the second time constant and two arcs appear, demonstrating that pure WEP coating cannot maintain the decent ability of corrosion resistance for N80 steel. Even though the arcs diameter of the MoS2/WEP coating (Figure 7b) was much larger than that of the neat WEP coating at the original immersion in 3.5 wt.% NaCl solution (5 days), it dramatically plunges and the second time constant appears similarly when the immersion time merely increases to 15 days. In this situation, the bad dispersibility and compatibility of MoS2 in the WEP coating are the main factors. In obvious contrast to the neat WEP and MoS2/WEP, MoS2-APTES/WEP coating (Figure 7e), exhibits the largest impedance arcs after the same immersion time and has only one semi-arcs. Additionally, this compound coating keeps one time constant from 5 to 50 days immersion, which proves that MoS2-APTES coating possesses an outstanding barrier ability and an excellent anticorrosion property.
In general, the lowest impedance moduli (|Z|0.01Hz) can be used as an effective method to assess the anticorrosive performance of the coating, which means that the higher |Z|0.01Hz of the coating, the better its protective ability [37]. For Figure 7b, it is clearly displayed that the neat WEP coating owns a low original value of |Z|0.01Hz about 3.5 × 107 Ω·cm2. Then, the impedance value dives heavily with the increase in the immersion period because of the entering of corrosive medium (O2, H2O, Cl) through the coating’s micropores or cracks. The variation of the |Z|0.01Hz value of MoS2/WEP (Figure 7d) coating has a similar tendency to the WEP coating, which is derived from the agglomeration of MoS2 in WEP resin. The MoS2-APTES/WEP coating (Figure 7f) shows the largest |Z|0.01Hz value (2.738 × 109 Ω·cm2). Above all, its value also has 3.647 × 107 Ω·cm2 even after the immersion time of 50 days in 3.5 wt.% NaCl solution, which can be ascribed to the homogeneous dispersion of MoS2-APTES in the WEP matrix and the interactions with them. All of the results prove the addition of MoS2-APTES into WEP coating and can offer carbon steel a long-period anticorrosion protection capacity.
The electrical equivalent circuit of ZSimpWin was used to fit the EIS data and was exhibited in Figure 8. The electrochemical parameters including CPEc and CPEdl represent the coating capacitance and double-layer capacitance, respectively. Rc, Rs, and Rct denote the coating resistance, solution resistance (resistance between the specimen coating electrode and saturated calomel electrode), and charge transfer resistance, respectively [38]. All these electrochemical parameters derived from the EIS charts are summed up in Table 1. The Rc values of neat WEP, and MoS2/WEP decline constantly following the extension of the immersion period. When the immersion time arrives at 15 days, the Rct fitted with the second equivalent circuit model appears, demonstrating that the erosive media reach the interface between the coating and steel going through the coating defects and the protective coating loses the ability of corrosion resistance for metal. In contrast, the MoS2-APTES/WEP coating exposes the higher Rc value and no Rct value even longing for an immersion time of 50 days, which indicates that the MoS2-APTES/WEP coating is endowed with the superior capability of blocking the penetration of corrosive media. These are another powerful evidence that the addition of MoS2-APTES could enhance the anticorrosion of WEP coating.
Potentiodynamic polarization test of the coatings was performed after immersing in 3.5 wt.% aqueous NaCl solution for 50 days and polarization diagrams were displayed in Figure 9. Generally speaking, if the protection coating has a higher corrosion potential (Ecorr) and a lower corrosion current density (icorr), it will be endowed with stronger anticorrosion property [39]. According to Figure 9, it can be apparently spotted that the MoS2-APTES/WEP coating shows more positive Ecorr in contrast to MoS2/WEP and a neat WEP coating. In addition, the icorr of MoS2-APTES/WEP coating is transparently lower than the others. The obtained results can be clarified via two explanations. One is that well distributed MoS2 modified by APTES could prolong the corrosive media’s routes to the substrate; another is that amine groups from APTES could lessen flaws in the WEP coating by increasing the crosslink density of the coating. Thus, MoS2-APTES/WEP coating displays the highest anticorrosion result among the three types of coatings.

3.3. Salt Spray Test

A salt spray test was employed to further evaluate the long-term anticorrosion property of coatings. The images of coatings, with 4 cm scratches on their surfaces after being made, put in salt spray of 5 wt.% NaCl for 5 and 15 days are displayed in Figure 10. For a neat WEP coating (Figure 10a,d), there are clear bubbles and corrosion products in the vicinity of the scratch after the exposure time reaches 5 days, exposing the inferior anticorrosion property of the neat waterborne epoxy coating. As the time extends to 15 days, the corrosion phenomenon increases heavily and much more corrosion products can be easily observed around the scratch, which are caused by more corrosion medium, which reaches the metal surface through coating flaws. After the addition of MoS2 (Figure 10b,e), the MoS2/WEP coating exhibits a similar corrosion protection result to the neat WEP coating during the salt spray test, because the bad dispersion of MoS2 in WEP and could not act as an effective barrier for the corrosive medium. In contrast to the neat WEP coating and MoS2/WEP coating, there are less corrosion products on the surface of MoS2-APTES/WEP composite coating (Figure 10c,f) even after 15 days of exposure time. The obtained results above prove that the MoS2-APTES/WEP coating has superior and long-term corrosion protection because of the physical barrier of MoS2-APTES.

3.4. The Corrosion Protection Mechanism

It can be concluded that the MoS2-APTES/WEP coating is endowed with a long-term and superior corrosion protection property to M80 steel, considering the results from the relevant EIS and salt spray test above. The corrosion protection mechanism of the MoS2-APTES/WEP coating is shown in Figure 11. In terms of neat WEP coating (Figure 11a), corrosive electrolytes such as H2O, O2, Cl can arrive at the interface between the coating and steel through diffusion channels, leading to metal corrosion [40]. Though MoS2 is added into the WEP coating, the MoS2/WEP coating (Figure 11b) cannot function effectively as a barrier impact on corrosive medium, because of its inferior dispersion in WEP. From Figure 11c, MoS2-APTES exposes homogeneous dispersion in WEP coating, giving rise to much more better barrier performance and more superior corrosion resistance properties, which is ascribed to the excellent compatibility of MoS2 and WEP. Moreover, amine-end groups of APTES could improve the crosslinking density between MoS2-APTES and WEP matrix.

4. Conclusions

In this work: a simple form of the anticorrosive addition MoS2-APTES was developed, which linked the virtues between MoS2 and APTES. Then, an MoS2-APTES hybrid, serving as a pigment, was added into the waterborne epoxy, which shows an apparently enhanced anticorrosion performance which effectively protects the N80 steel matrix. The successful modification of APTES on MoS2 is proved by FT-IR, XRD, TEM, XPS, and TGA. Moreover, the results of the SEM, EIS, potentiodynamic polarization and salt spray test reveal a long-term corrosion resistance of MoS2-APTES/WEP. The extraordinary performance of corrosion resistance is derived from the good dispersity of MoS2-APTES which can prolong the diffusion path of corrosive media going through the coating and APTES can strengthen the crosslinking density of waterborne epoxy resin.

Author Contributions

Conceptualization, Y.L. and S.Z.; methodology, Y.L.; software, P.X.; validation, Z.L., J.C. and F.Z.; formal analysis, C.Z.; investigation, Y.L.; resources, Y.H.; data curation, H.L.; writing-original draft preparation, Y.L.; writing-review and editing, C.C.; visualization, C.Z.; supervision, S.Z.; project administration, Z.L.; funding acquisition, Y.H. All authors have read and agreed to the published version of the manuscript.

Funding

Open Fund (PLN161, PLN201806) of State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation (Southwest Petroleum University), the Research Program of Science and Technology Department of Sichuan Province (2018JY0302, 18YYJC0018) and the National Natural Science Foundation of China (51774245).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ding, J.; Zhao, H.; Shao, Z.; Yu, H. Bioinspired Smart Anticorrosive Coatings with an Emergency-Response Closing Function. ACS Appl. Mater. Interfaces 2019, 11, 42646–42653. [Google Scholar] [CrossRef] [PubMed]
  2. Li, Y.; Yang, Z.; Qiu, H.; Dai, Y.; Zheng, Q.; Li, J.; Yang, J. Self-aligned graphene as anticorrosive barrier in waterborne polyurethane composite coatings. J. Mater. Chem. A 2014, 2, 14139–14145. [Google Scholar] [CrossRef]
  3. Ramezanzadeh, B.; Haeri, Z. A facile route of making silica nanoparticles-covered graphene oxide nanohybrids (SiO2-GO); Fabrication of SiO2-GO/epoxy composite coating with superior barrier and corrosion protection performance. Chem. Eng. J. 2016, 303, 511–528. [Google Scholar] [CrossRef]
  4. Ramezanzadeh, B.; Niroumandrad, S.; Ahmadi, A.; Mahdavian, M.; Moghadam, M.M. Enhancement of barrier and corrosion protection performance of an epoxy coating through wet transfer of amino functionalized graphene oxide. Corros. Sci. 2016, 103, 283–304. [Google Scholar] [CrossRef]
  5. Cui, M.J.; Ren, S.M.; Zhao, H.C.; Xue, Q.J.; Wang, L.P. Polydopamine coated graphene oxide for anticorrosive rein-forcement of water-borne epoxy coating. Chem. Eng. J. 2018, 335, 255–266. [Google Scholar] [CrossRef]
  6. Cui, M.J.; Ren, S.M.; Chen, J.; Liu, S.; Zhang, G.G.; Zhao, H.C.; Wang, L.P.; Xue, Q.J. Anticorrosive performance of wa-terborne epoxy coatings containing water-dispersible hexagonal boron nitride (h-BN) nanosheets. Appl. Surf. Sci. 2017, 397, 77–86. [Google Scholar] [CrossRef]
  7. Ding, J.; Rahman, O.U.; Peng, W.; Dou, H.; Yu, H. A novel hydroxyl epoxy phosphate monomer enhancing the anticorrosive performance of waterborne Graphene/Epoxy coatings. Appl. Surf. Sci. 2018, 427, 981–991. [Google Scholar] [CrossRef]
  8. Chen, C.L.; Xiao, G.Q.; He, Y.; Zhong, F.; Li, H.J.; Wu, Y.Q.; Chen, J.Y. Bio-inspired superior barrier self-healing coating: Self-assemble of graphene oxide and polydopamine-coated halloysite nanotubes for enhancing corrosion resistance of waterborne epoxy coating. Prog. Org. Coat. 2020, 139, 105462. [Google Scholar] [CrossRef]
  9. Wu, Y.; He, Y.; Zhou, T.; Chen, C.; Zhong, F.; Xia, Y.; Xie, P.; Zhang, C. Synergistic functionalization of h-BN by mechanical exfoliation and PEI chemical modification for enhancing the corrosion resistance of waterborne epoxy coating. Prog. Org. Coat. 2020, 142, 105541. [Google Scholar] [CrossRef]
  10. Zhong, F.; He, Y.; Wang, P.Q.; Chen, C.L.; Yu, H.; Li, H.J.; Chen, J.Y. Graphene/V2O5@polyaniline ternary composites enable waterborne epoxy coating with robust corrosion resistance. React. Funct. Polym. 2020, 151, 104567. [Google Scholar] [CrossRef]
  11. Xia, Y.; He, Y.; Chen, C.; Wu, Y.; Zhong, F.; Chend, J. Co-modification of polydopamine and KH560 on g-C3N4 nanosheets for enhancing the corrosion protection property of waterborne epoxy coating. React. Funct. Polym. 2020, 146, 104405. [Google Scholar] [CrossRef]
  12. Zhong, F.; He, Y.; Wang, P.; Chen, C.; Lin, Y.; Wu, Y.; Chen, J. Self-assembled graphene oxide-graphene hybrids for enhancing the corrosion resistance of waterborne epoxy coating. Appl. Surf. Sci. 2019, 488, 801–812. [Google Scholar] [CrossRef]
  13. Chang, C.-H.; Huang, T.-C.; Peng, C.-W.; Yeh, T.-C.; Lu, H.-I.; Hung, W.-I.; Weng, C.-J.; Yang, T.-I.; Yeh, J.-M. Novel anticorrosion coatings prepared from polyaniline/graphene composites. Carbon 2012, 50, 5044–5051. [Google Scholar] [CrossRef]
  14. Zhang, C.L.; He, Y.; Li, F.; Di, H.H.; Zhang, L.; Zhan, Y.Q. h-BN decorated with Fe3O4 nanoparticles through mus-sel-inspired chemistry of dopamine for reinforcing anticorrosion performance of epoxy coatings. J. Alloys Compd. 2016, 685, 743–751. [Google Scholar] [CrossRef]
  15. Shi, H.; Liu, W.; Liu, C.; Yang, M.; Xie, Y.; Wang, S.; Zhang, F.; Liang, L.; Pi, K. Polyethylenimine-assisted exfoliation of h-BN in aqueous media for anticorrosive reinforcement of waterborne epoxy coating. Prog. Org. Coat. 2020, 142, 105591. [Google Scholar] [CrossRef]
  16. Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. MoS2Nanoparticles Grown on Graphene: An Advanced Catalyst for the Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2011, 133, 7296–7299. [Google Scholar] [CrossRef] [Green Version]
  17. Xiang, Q.; Yu, J.; Jaroniec, M. Synergetic Effect of MoS2and Graphene as Cocatalysts for Enhanced Photocatalytic H2Production Activity of TiO2Nanoparticles. J. Am. Chem. Soc. 2012, 134, 6575–6578. [Google Scholar] [CrossRef]
  18. Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive photodetectors based on monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497–501. [Google Scholar] [CrossRef]
  19. Lee, Y.-H.; Zhang, X.Q.; Zhang, W.; Chang, M.-T.; Lin, C.-T.; Chang, K.-D.; Yu, Y.-C.; Wang, J.T.-W.; Chang, C.-S.; Li, L.-J.; et al. Synthesis of Large-Area MoS2Atomic Layers with Chemical Vapor Deposition. Adv. Mater. 2012, 24, 2320–2325. [Google Scholar] [CrossRef] [Green Version]
  20. Zhao, X.; Zhang, B.; Jin, Z.; Chen, C.; Zhu, Q.; Hou, B. Epoxy coating modified by 2D MoS2/SDBS: Fabrication, anticorrosion behaviour and inhibition mechanism. RSC Adv. 2016, 6, 97512–97522. [Google Scholar] [CrossRef]
  21. Xia, Z.; Liu, G.; Dong, Y.; Zhang, Y. Anticorrosive epoxy coatings based on polydopamine modified molybdenum disulfide. Prog. Org. Coat. 2019, 133, 154–160. [Google Scholar] [CrossRef]
  22. Jing, Y.; Wang, P.; Yang, Q.; He, Y.; Bai, Y. Molybdenum disulfide with poly(dopamine) and epoxy groups as an efficiently anticorrosive reinforcers in epoxy coating. Synth. Met. 2020, 259, 116249. [Google Scholar] [CrossRef]
  23. Ding, J.; Zhao, H.; Zhao, X.; Xu, B.; Yu, H. How semiconductor transition metal dichalcogenides replaced graphene for enhancing anticorrosion. J. Mater. Chem. A 2019, 7, 13511–13521. [Google Scholar] [CrossRef]
  24. Zhou, K.; Liu, C.; Gao, R. Polyaniline: A novel bridge to reduce the fire hazards of epoxy composites. Compos. Part A Appl. Sci. Manuf. 2018, 112, 432–443. [Google Scholar] [CrossRef]
  25. Wang, D.; Song, L.; Zhou, K.Q.; Yu, X.J.; Hu, Y.; Wang, J. Anomalous nano-barrier effects of ultrathin molybdenum di-sulfide nanosheets for improving the flame retardance of polymer nanocomposites. J. Mater. Chem. A 2015, 3, 14307–14317. [Google Scholar] [CrossRef]
  26. Zhou, K.; Jiang, S.; Bao, C.; Song, L.; Wang, B.; Tang, G.; Hu, Y.; Gui, Z. Preparation of poly(vinyl alcohol) nanocomposites with molybdenum disulfide (MoS2): Structural characteristics and markedly enhanced properties. RSC Adv. 2012, 2, 11695–11703. [Google Scholar] [CrossRef]
  27. Pourhashem, S.; Rashidi, A.; Vaezi, M.R.; Bagherzadeh, M.R. Excellent corrosion protection performance of epoxy composite coatings filled with amino-silane functionalized graphene oxide. Surf. Coat. Technol. 2017, 317, 1–9. [Google Scholar] [CrossRef]
  28. Pourhashem, S.; Vaezi, M.R.; Rashidi, A.; Bagherzadeh, M.R. Distinctive roles of silane coupling agents on the corrosion inhibition performance of graphene oxide in epoxy coatings. Prog. Org. Coat. 2017, 111, 47–56. [Google Scholar] [CrossRef]
  29. Zhou, K.; Gao, R.; Qian, X. Self-assembly of exfoliated molybdenum disulfide (MoS 2 ) nanosheets and layered double hydroxide (LDH): Towards reducing fire hazards of epoxy. J. Hazard. Mater. 2017, 338, 343–355. [Google Scholar] [CrossRef]
  30. Asadi, N.; Naderi, R.; Mahdavian, M. Doping of zinc cations in chemically modified halloysite nanotubes to improve protection function of an epoxy ester coating. Corros. Sci. 2019, 151, 69–80. [Google Scholar] [CrossRef]
  31. Karekar, S.E.; Pinjari, D.V. Sonochemical synthesis and characterization of molybdenum sulphide nanoparticles: Effect of calcination temperature. Chem. Eng. Process. Process. Intensif. 2017, 120, 268–275. [Google Scholar] [CrossRef]
  32. Knorr, D.B.; Tran, N.T.; Gaskell, K.J.; Orlicki, J.A.; Woicik, J.C.; Jaye, C.; Fischer, D.A.; Lenhart, J.L. Synthesis and Char-acterization of Aminopropyltriethoxysilane-Polydopamine Coatings. Langmuir 2016, 32, 4370–4381. [Google Scholar] [CrossRef] [PubMed]
  33. Xu, G.M.; Xu, G.N.; Chang, W.G.; Wang, H.Y. Fabrication of microsphere-like nano-MoO2 modified with silane coupling agent KH550. Mater. Res. Express. 2019, 6. [Google Scholar] [CrossRef]
  34. Tanhaei, M.; Ren, Y.; Yang, M.; Bussolotti, F.; Cheng, J.J.W.; Pan, J.; Chiam, S.Y. Direct control of defects in molybdenum oxide and understanding their high CO2 sorption performance. J. Mater. Chem. A 2020, 8, 12576–12585. [Google Scholar] [CrossRef]
  35. Chen, C.; He, Y.; Xiao, G.; Zhong, F.; Xia, Y.; Wu, Y. Graphic C3N4-assisted dispersion of graphene to improve the corrosion resistance of waterborne epoxy coating. Prog. Org. Coat. 2020, 139, 105448. [Google Scholar] [CrossRef]
  36. Hong, M.-S.; Park, Y.; Kim, J.-G.; Kim, K. Effect of Incorporating MoS2 in Organic Coatings on the Corrosion Resistance of 316L Stainless Steel in a 3.5% NaCl Solution. Coatings 2019, 9, 45. [Google Scholar] [CrossRef] [Green Version]
  37. Zuo, S.X.; Chen, Y.; Liu, W.J.; Yao, C.; Li, Y.R.; Ma, J.Q.; Kong, Y.; Mao, H.H.; Li, Z.Y.; Fu, Y.S. Polyaniline/g-C3N4 com-posites as novel media for anticorrosion coatings. J Coat. Technol. Res. 2017, 14, 1307–1314. [Google Scholar] [CrossRef]
  38. Ramezanzadeh, B.; Moghadam, M.H.M.; Shohani, N.; Mandavian, M. Effects of highly crystalline and conductive poly-aniline/graphene oxide composites on the corrosion protection performance of a zinc-rich epoxy coating. Chem. Eng. J. 2017, 320, 363–375. [Google Scholar] [CrossRef]
  39. Zheng, H.; Shao, Y.; Wang, Y.; Meng, G.; Liu, B. Reinforcing the corrosion protection property of epoxy coating by using graphene oxide–poly(urea–formaldehyde) composites. Corros. Sci. 2017, 123, 267–277. [Google Scholar] [CrossRef]
  40. Wang, H.H.; Qin, S.D.; Yang, X.F.; Fei, G.Q.; Tian, M.; Shao, Y.M.; Zhu, K. A waterborne uniform gra-phene-poly(urethane-acrylate) complex with enhanced anticorrosive properties enabled by ionic interaction. Chem. Eng. J. 2018, 351, 939–951. [Google Scholar] [CrossRef]
Figure 1. Schematic of the synthesis procedure of MoS2-3-aminopropyltriethoxysilane (APTES) via APTES hydrolysis.
Figure 1. Schematic of the synthesis procedure of MoS2-3-aminopropyltriethoxysilane (APTES) via APTES hydrolysis.
Coatings 11 00178 g001
Figure 2. (a) FT-IR spectra and (b) XRD patterns of MoS2, and MoS2-APTES.
Figure 2. (a) FT-IR spectra and (b) XRD patterns of MoS2, and MoS2-APTES.
Coatings 11 00178 g002
Figure 3. TEM images of (a) MoS2-APTES and (b) high-resolution TEM morphology of MoS2-APTES.
Figure 3. TEM images of (a) MoS2-APTES and (b) high-resolution TEM morphology of MoS2-APTES.
Coatings 11 00178 g003
Figure 4. (a) Wide XPS survey of MoS2 and MoS2-APTES and high-resolution spectra in (b) C 1s; (c) Si 2p and (d) Mo 3d binding energy regions of MoS2-APTES.
Figure 4. (a) Wide XPS survey of MoS2 and MoS2-APTES and high-resolution spectra in (b) C 1s; (c) Si 2p and (d) Mo 3d binding energy regions of MoS2-APTES.
Coatings 11 00178 g004
Figure 5. TGA curves of MoS2 and MoS2-APTES.
Figure 5. TGA curves of MoS2 and MoS2-APTES.
Coatings 11 00178 g005
Figure 6. SEM images of the fracture surface of neat waterborne epoxy coating (WEP) (a); MoS2/WEP (b); MoS2-APTES/WEP (c).
Figure 6. SEM images of the fracture surface of neat waterborne epoxy coating (WEP) (a); MoS2/WEP (b); MoS2-APTES/WEP (c).
Coatings 11 00178 g006
Figure 7. Nyquist and Bode plots of the neat WEP (a,b), MoS2/WEP (c,d) and MoS2-APTES/WEP (e,f) in a 3.5 wt.% NaCl solution.
Figure 7. Nyquist and Bode plots of the neat WEP (a,b), MoS2/WEP (c,d) and MoS2-APTES/WEP (e,f) in a 3.5 wt.% NaCl solution.
Coatings 11 00178 g007
Figure 8. The equivalent circuit models of the first (a) and the second (b).
Figure 8. The equivalent circuit models of the first (a) and the second (b).
Coatings 11 00178 g008
Figure 9. The polarization curves of composite coatings after in 3.5 wt.% NaCl solution for 50 days.
Figure 9. The polarization curves of composite coatings after in 3.5 wt.% NaCl solution for 50 days.
Coatings 11 00178 g009
Figure 10. The visual picture of the neat WEP (a,d), MoS2/WEP (b,e) and MoS2-APTES/WEP (c,f) samples after 5 and 15 days.
Figure 10. The visual picture of the neat WEP (a,d), MoS2/WEP (b,e) and MoS2-APTES/WEP (c,f) samples after 5 and 15 days.
Coatings 11 00178 g010
Figure 11. Schematic illustration of a corrosion protection mechanism for neat WEP (a), MoS2/WEP (b) and MoS2-APTES/WEP (c) coating.
Figure 11. Schematic illustration of a corrosion protection mechanism for neat WEP (a), MoS2/WEP (b) and MoS2-APTES/WEP (c) coating.
Coatings 11 00178 g011
Table 1. Electrochemical impedance data of neat and composite waterborne epoxy coatings.
Table 1. Electrochemical impedance data of neat and composite waterborne epoxy coatings.
Sample.TimeCPEc Rc (Ω·cm2)CPEdl Rct (Ω·cm2)
Y0 (ohm−1 cm−2sn)ncoat Y0 (ohm−1 cm−2sn)ndl
Epoxy5 d1.736 × 10−90.9533.517 × 107---
15 d2.043 × 10−80.9454.355 × 1062.145 × 10−60.731.19 × 106
30 d1.4 × 10−80.7951.911 × 1056.552 × 10−60.685.572 × 105
50 d1.369 × 10−50.674.448 × 1051.779 × 10−50.622.543 × 105
MoS25 d1.808 × 10−90.945.476 × 107---
15 d3.036 × 10−80.924.05 × 1064.258 × 10−60.783.688 × 106
30 d2.155 × 10−80.781.763 × 1057.106 × 10−60.733.225 × 105
50 d1.614 × 10−50.6632.272 × 1046.049 × 10−50.642.929 × 105
MoS2-APTES5 d3.59 × 10−100.9682.738 × 109---
15 d3.757 × 10−90.966.387 × 108---
30 d3.565 × 10−90.9583.091 × 108---
50 d1.63 × 10−90.9253.647 × 107---
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Y.; Zhang, S.; He, Y.; Chen, C.; Zhang, C.; Xie, P.; Zhong, F.; Li, H.; Chen, J.; Li, Z. APTES Modification of Molybdenum Disulfide to Improve the Corrosion Resistance of Waterborne Epoxy Coating. Coatings 2021, 11, 178. https://0-doi-org.brum.beds.ac.uk/10.3390/coatings11020178

AMA Style

Liu Y, Zhang S, He Y, Chen C, Zhang C, Xie P, Zhong F, Li H, Chen J, Li Z. APTES Modification of Molybdenum Disulfide to Improve the Corrosion Resistance of Waterborne Epoxy Coating. Coatings. 2021; 11(2):178. https://0-doi-org.brum.beds.ac.uk/10.3390/coatings11020178

Chicago/Turabian Style

Liu, Yang, Shihong Zhang, Yi He, Chunlin Chen, Chen Zhang, Peng Xie, Fei Zhong, Hongjie Li, Jingyu Chen, and Zhenyu Li. 2021. "APTES Modification of Molybdenum Disulfide to Improve the Corrosion Resistance of Waterborne Epoxy Coating" Coatings 11, no. 2: 178. https://0-doi-org.brum.beds.ac.uk/10.3390/coatings11020178

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop