Next Article in Journal
Enhancing Tribological Performance of Micro-Arc Oxidation Coatings on 6061 Aluminum Alloy with h-BN Incorporation
Previous Article in Journal
Failure Causes Analysis of Circumferential Cracking on Gathering Pipeline in M Oil and Gas Field
Previous Article in Special Issue
Biocompatibility and Corrosion of Microplasma-Sprayed Titanium and Tantalum Coatings versus Titanium Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Size Effect of Graphite Nanosheet-Induced Anti-Corrosion of Hydrophobic Epoxy Coatings

1
State Key Laboratory of Green Building Materials, China Building Materials Academy Co., Ltd., Beijing 100024, China
2
School of Chemistry, Beihang University, Beijing 100191, China
*
Authors to whom correspondence should be addressed.
Coatings 2024, 14(6), 769; https://doi.org/10.3390/coatings14060769
Submission received: 28 May 2024 / Revised: 13 June 2024 / Accepted: 13 June 2024 / Published: 18 June 2024

Abstract

:
In order to broaden the selectivity of graphite nanosheet additives on epoxy resin-based coatings and verify the size effect, this work aims to dope graphite nanosheets of different sizes into the three-dimensional structure produced by cross-linking and curing epoxy resin and polyamide resin. In addition, a micro-nano level secondary structure and a surface with special roughness are constructed to obtain the composite epoxy hydrophobic coating. The influence of the size effect of graphite nanosheets on the hydrophobic performance and corrosion resistance of the coating is summarized as well. Among them, the optimized doping size (2.2 μm) of graphite nanosheets in the epoxy coating showed the largest impedance arc of 2.58 × 108 Ω cm2, which could form an excellent nano-network covering the micropores to impede the diffusion of corrosive medium. Through simulation calculation analysis, we also found that the edge site of graphene is more effective in capturing H2O and O2; therefore, a smaller size of graphene with a large edge can be more favorable. This work will be used as a reference for the industrial application of graphite anti-corrosive coating.

1. Introduction

Metal corrosion, induced by an electrochemical or chemical reaction at the surface of a metal, will significantly destroy its inherent properties (such as toughness, electrical, plasticity, and optical), thus posing a serious threat to construction, marine environments, transportation, and daily life [1,2,3]. The mechanism of corrosion phenomena is involved in a series of electrochemical reactions at the interface between corrosive media (such as Cl, H2O, and O2) and substrates [4,5]. Thus, it is necessary to design a surface coating to prevent the metal substrate from contacting these corrosive media [6,7,8]. Engineering a hydrophobic nanostructured coating is a significant method for anti-corrosive and self-cleaning applications for marine environments or transportation [9,10]. The hydrophobic surface of the substrate results in a reduced rolling angle, elevated water contact angle, and self-cleaning and anti-corrosion properties [11,12,13,14,15]. Hydrophobic nanostructured coatings act as a corrosion barrier by (1) preventing blister formation and establishing difficult pathways for Cl, H2O, and O2 and (2) prohibiting adhesive failure and enhancing corrosion resistance. However, most hydrophobic anti-corrosive coatings suffer from inadequate surface durability [16].
An organic–inorganic nanostructured coating has the features required to develop hydrophobic and anti-corrosive materials with long-term surface stability [17,18]. Developing carbonaceous nanomaterials, particularly graphene/graphite nanosheets, is favorable to employing durable and outstanding coatings for anti-corrosive protection because of their excellent electrical conductivity, physical properties, and honeycomb lattice structure, which can be used to obtain coatings with excellent barrier properties [19,20]. For example, Monetta et al. reported that 0.5 wt% and 1 wt% of graphene nanosheets were added to an additive-free waterborne epoxy resin applied to aluminum alloy. Electrochemical analysis revealed an improvement in the protective properties of the coating and a lesser amount of absorbed water than the unloaded film [21]. Kopsidas et al. also reported that the low content of graphene nanosheets (≤0.5 wt%) for EP coatings demonstrated 0% damage [22]. The doping of graphene nanosheets is only effective at low loadings, as higher contents result in accelerated corrosion that facilitates the conduction of corrosion currents. Li et al. fabricated sulfonated graphene to fill in a coating by introducing sulfonic acid groups on graphene sheets, and the most remarkable improvement was achieved by adding 1.0 wt% of graphene, which was effective for blocking the penetration of corrosive molecules [23]. In our previous work, the anti-corrosion and mechanical properties of Gr0.04-ZnO0.4 coatings were improved after adding the appropriate 0.04 wt% content of graphite nanosheets. Moreover, the parallel assembly of graphene occurs spontaneously, which leads to a remarkable improvement in corrosion protection [24,25]. Generally speaking, these graphene-based polymer coatings with thinner thickness and lighter weight of doping graphite nanosheets exhibit a more superior anti-corrosion effect than noumenal coating [26]. However, do the different sizes of graphite nanosheets as additives also have an impact on corrosion resistance? There are almost no relevant reports on the influence of the size effect of graphite nanosheets on the corrosion protection of coatings, and further exploration of the reaction mechanism is still needed.
The influence of the size effect of graphite nanosheets and the level of interface roughness on the hydrophobic performance of coatings was explored. In addition, the influence on stability, wear resistance, and corrosion resistance of the coating was summarized as well [27,28,29]. The operation mechanism of micro-nanoscale structure and size effect on the anti-corrosion process was determined by DFT simulation calculation [30,31]. Such a reaction mechanism can be used to adjust the size and morphology of two-dimensional materials doped in the coating and filtered into stable and high-efficiency anti-corrosion coatings. This method is capable of providing new design ideas for producing two-dimensional materials and epoxy matrices. We innovatively propose the design concept of regulating the size effect of two-dimensional materials in terms of preventing the corrosion of super-hydrophobics. This work will be used as a reference for industrial applications.
Herein, micro-kaolin, PTFE, and different sizes of graphite nanosheets were doped into WEP coatings (denoted as EGx/C1P0.2EP, X represents centrifugal rate). They were synthesized using a cost-effective stirring process in the presence of EP and alcohol. During the stirring process, different sizes of EGx were extensively introduced to EP coatings, and simultaneously, micro-kaolin and PTFE were uniformly created on the surface to form hydrophobic long chains. Among them, different sizes of EGx were prepared using a low-cost ball-milling and gradient centrifugation approach. Then, the EGx samples were characterized by XRD, Raman, SEM, and TEM to verify and successfully prepare the different sizes of EGx. The impedance arc of EG5000/C1P0.2EP in these coatings was the largest, with an impedance of 2.58 × 108 Ωcm2 after 7d immersing, indicating that the corrosion resistance of EG5000/C1P0.2EP is higher than that of EG1000/C1P0.2EP, EG3000/C1P0.2EP, and EG8000/C1P0.2EP. Even after 300 h NSS time, the EG5000/C1P0.2EP coating did not exhibit obvious corrosion features; it presented more corrosion resistance than the other sizes of EGx coatings. In addition, the theoretical calculation also showed that the chemisorption of H2O will cause the natural decomposition of H-O, producing the OH and H species on the edge of s-Gra and d-Gra. The smaller Gra can attract H2O and O2, reducing their attack on the coating, which can prohibit the corrosion of EGx/C1P0.2EP by oxidation. These results suggest that doping the optimized size (2.21 μm) of EG5000 could form a nano-network covering the micropores to impede the diffusion of the corrosive medium.

2. Materials and Methods

2.1. Materials

Graphite powders (CAS: 7782-42-5) were supplied by Sigma-Aldrich (Shanghai, China). Kaolin (CAS:1332-58-7), Poly(tetrafluoroethylene) (PTFE, CAS:9002-84-0), NaCl (99.9%), and 1H,1H,2H,2H-Perfluorodecyltriethoxysilane were purchased from Aladdin Holdings Group Co. Ltd. (Beijing, China). Waterborne epoxy dispersion and cosolvent were purchased from Shanghai Macklin Biochemical Co. Ltd. (Shanghai, China). The Q235 steel substrates (size: 20 mm × 20 mm × 0.3 mm) for electrochemical corrosion were firstly polished with 180-, 600-, and 1500-mesh abrasive paper, then ultrasonically cleaned in alcohol, and finally dried at room temperature. All reagents were used without further treatment.

2.2. Characterizations of Samples

The surface morphologies and physical size of samples were observed under a Scanning Electron Microscope (SEM, S-4800, Hitachi, Tokyo, Japan). The microstructure and lattice structure of the samples were observed by high-resolution transmission electron microscopy (HRTEM, JEM-2100F, Hitachi, Tokyo, Japan). Raman spectra were analyzed using 532 nm laser excitation by a Raman Spectroscope (Via Reflex, Renishaw, Wotton-under-Edge, UK). The XRD patterns were scanned at a rate of 5°/min from the 2θ angle of 5° to 90° with a characteristic wavelength using X-ray powder diffraction (XRD Terra, Innov-X, Woburn, MA, USA). Fourier transform infrared spectroscopy was taken by a spectrometer with He-Ne radiation (IR Prestige-21, Shimadzu, Kyoto, Japan). Static contact angle measurements were performed via a contact angle meter (OCAH200, Dataphysics, Filderstadt, Germany) using a sessile drop technique.

2.3. Synthesis of the Different Sizes of Graphite Sheets

First, 4 g of graphite powder (500 mesh), 2 mm, and 0.2 mm ZrO2 balls (weight ratio: 1:1) were added to a 250 mL jar along with 100 mL of N-vinyl pyrrolidone (NVP) and subjected to grinding for 200 h. After grinding, the obtained dispersion was collected and then centrifuged at 500 r/min for 20 min to remove any unexfoliated graphite. The resulting dispersion was then centrifuged at 1000 r/min for 20 min; the sediments, termed EG1000, were collected using a PTFE membrane and subsequently dried at 80 °C. The remaining supernatant, designated as A, was subjected to centrifugation at 3000 r/min for 20 min to yield sediments named EG3000; the leftover supernatant was designated as B. Similarly, exfoliated graphite sediments obtained via an identical centrifugation procedure conducted at 5000 and 8000 r/min were named EG5000 and EG8000, respectively.

2.4. Preparation of EGx/C1P0.2EP Composite Epoxy Coatings

Initially, a homogeneous solution was prepared by combining 1 g of micro-kaolin, 0.33 g of PTFE, 0.33 g of epoxy resin, and 0.05 wt% EG5000 in 3.3 mL of ethanol, followed by stirring for 1 h. Subsequently, 0.2 mL of 1H, 1H, 2H, and 2H-perfluorodecyltriethoxysilane was added to this solution and stirred for 2 h to form component A. Component B comprised 0.33 g of a waterborne epoxy curing agent. These two components were mixed and stirred at 1000 r/min for 30 min. The mixture was allowed to solidify at 30 °C until the surface was dry and then left to cure naturally for 48 h to form the hydrophobic coating, designated as EG5000/C1P0.2EP. The EG5000/C1P0.2EP nanocomposite coatings were applied to Q235 steel substrates using a 120# wire-wound rod and cured at 60 °C for 4 h. For comparative analysis, coatings EG1000/C1P0.2EP, EG3000/C1P0.2EP, and EG8000/C1P0.2EP were also fabricated using the same procedures by varying the additional amount of dimensional graphite sheets. In addition, hydrophobic epoxy coatings with 0 wt% C1P0.2EP, as well as 0.025, 0.04, and 0.075 wt% EG5000, were prepared to assess the effect of varying the EG5000 content.

3. Results

3.1. Characterization of Samples and Coatings

3.1.1. Morphology Characterization of Samples and Coatings

EGx (where X denotes the centrifugation rate) samples were designed and fabricated using a simple, cost-effective ball-milling and gradient centrifugation method, incorporating graphite nanosheets and an NVP solvent, as described in our previous study [25]. Typical surface morphologies of these samples and coatings were characterized via SEM and HRTEM. The SEM images (Figure 1) reveal that the sizes of EG1000, EG3000, EG5000, and EG8000 decrease progressively, illustrating the effectiveness of the gradient centrifugation method. Figure 2 demonstrates that the exfoliated EGx samples, processed through ball-milling and centrifugation, exhibit no fractures or agglomeration, maintaining uniform size and length. The average diameters of EG1000, EG3000, EG5000, and EG8000, measured across 100 samples, were found to be 4.91, 4.00, 2.21, and 1.04 μm, respectively. Figure S3 shows the actual samples, while Figure S1 provides a close-up view of various microholes and microcracks on the surface of the EP coating. These imperfections are likely due to poor structural density and substantial solvent evaporation during the curing process. Microscopic analysis at different sites revealed the average length of these microholes and microcracks to be between 2 and 4 μm. To address these defects, different sizes of graphite were introduced into the coating to decrease porosity and lengthen the penetration path for corrosive media, thereby enhancing corrosion resistance. The SEM images presented in Figure S2 show that the surface of EG5000/C1P0.2EP contains few micropores, which hampers the diffusion of O2, H2O, and Cl through these microchannels to the substrate and thereby considerably enhances its anti-corrosion properties.

3.1.2. Structural Characterization of Samples and Coatings

The molecular structure of the EGx samples was comprehensively analyzed via X-ray diffraction (XRD), Raman spectroscopy, and Fourier transform infrared (FTIR) spectroscopy. The XRD patterns of the EGx nanosheets (Figure S4) show diffraction peaks at 25.7° and 52.1°, which correspond to the (002) and (004) diffraction reflections of the graphite structure (PDF#75-1621). Notably, EG1000 and EG3000 exhibit a sharp (002) diffraction peak, indicating a highly regular spatial arrangement between lamellae [25]. However, the (002) diffraction peaks for EG5000 and EG8000 are broader, suggesting fewer layers and a reduced size of graphite nanosheets. The Raman spectroscopy analysis (Figure 3a) reveals that bulk graphite powders and EGx, excited by a 532 nm laser, display approximate resonance peaks at 1383 cm−1 and 1588 cm−1, representing the D and G bands, respectively. The intensity ratio of the D band to the G band (ID/IG) slightly increases from 0.03 for bulk graphite, 0.06 for EG1000, 0.11 for EG3000, and 0.17 for EG5000 to 0.23 for EG8000, indicating the introduction of a few defects during ball-milling [25]. Furthermore, Figure 3b demonstrates that the ID/IG ratio increases with increasing centrifugation speed, indicating that the defects are related to the edge structure and are influenced by the centrifugation process. Moreover, a blue shift is observed in the 2D peak of exfoliated EGx, signifying a reduction in the size of graphite nanosheets. In addition, the molecular structure of bulk graphite and EGx was examined using FTIR, as depicted in Figure 3c. The characteristic absorption peaks at 3410 cm−1, 2928 cm−1, 1630 cm−1, and 1080 cm−1 correspond to –OH, C–O, C–H, and C–O–C functional groups, respectively. The infrared spectra reveal that the functional groups in the EGx samples align closely with those of the original graphite, indicating that the chemical integrity of the graphite is maintained after ball milling.

3.1.3. Wettability Characterization of Samples and Coatings

The influence of different sizes of EGx on the wettability of coatings was studied by measuring the contact angles of four types of EGx/C1P0.2EP coatings. As shown inFigure 4a and Figure S5, the water contact angles of all four types of EGx/C1P0.2EP coatings are greater than 140° within 60 days, indicating that these coatings have low surface energy and excellent hydrophobicity. The EGx/C1P0.2EP coating effectively prevents the penetration of corrosive ions, making it difficult for corrosive media to adhere to the coating surface, thereby slowing down the corrosion process of the substrate (Table S1). After immersion in water for 60 days, the contact angles of all four types of EGx/C1P0.2EP decreased, but the water contact angle of the EG5000/C1P0.2EP coating remained the highest, indicating that EG5000/C1P0.2EP had the highest hydrophobicity and the lowest porosity. The significant decrease in the water contact angle for EG8000/C1P0.2EP coatings appears to be related to the size of the nanofiller (graphite nanosheets) in the resin; in fact, the smaller size of the nanofiller is more likely to move with increasing immersion time, slightly reducing the surface’s hydrophobicity. Figure 4b shows the contact angles of the EGx/C1P0.2EP coatings with different EG5000 contents as the soaking time increases. At the initial stage of immersion, the contact angles of the four coatings with different EG5000 contents were all greater than 145°, indicating that these coatings had excellent hydrophobic properties. From Figure 4b and Table S2, it can be observed that with increasing EG5000 content, the contact angle also gradually increases, with the maximum angle of the EG5000/C1P0.2EP coating with 0.075 wt% EG5000 exceeding 151°, indicating that the 0.075 wt% addition of EG5000 better filled the pores on the surface of the epoxy resin, reducing its porosity.

3.2. Anti-Corrosion Properties of EGx/C1P0.2EP Coatings

3.2.1. Electrochemical Characteristics

In this study, the impedance spectra of the EGx/C1P0.2EP coatings were measured in a 3.5 wt.% NaCl solution at room temperature. The corrosion protection properties of the materials correlate with the radius of the capacitive arc [14], and it was observed that the corrosion resistance of the EGx/C1P0.2EP coatings decreased over time, from 7 days to 60 days (Figure 5). After 7 days of immersion (Figure 5a), the impedance arc of the EG5000/C1P0.2EP coating was the largest, with an impedance of 2.58 × 108 Ω·cm2, indicating superior corrosion resistance compared with EG1000/C1P0.2EP, EG3000/C1P0.2EP, and EG8000/C1P0.2EP. This is attributed to the size of EG5000, which closely matches the pore size of the epoxy resin, effectively filling the holes and crevices to enhance corrosion resistance. After 30 days of immersion (Figure 5b), the impedance values of all coatings decreased by an order of magnitude from 108 Ω·cm2 to 107 Ω·cm2 because of the penetration of corrosive electrolytes. However, after 60 days of immersion (Figure 5c), EG5000/C1P0.2EP still exhibited the largest impedance arc among the coatings, with an impedance of 3.18 × 107 Ω·cm2. The EIS data were analyzed using ZSimpWin software version 3.30. The equivalent circuits in Figure S7 fit the EIS data with one time constant, where Rc represents coating resistance and C represents coating capacitance. As depicted in Table S3 and Figure 5, the Rc value of EG5000/C1P0.2EP is the largest, suggesting that this coating can more effectively prevent the penetration of the corrosion medium into the metal matrix when the size of the graphite nanosheet matches the pore size of the epoxy resin. Comparatively, the impedance spectra of EG5000/C1P0.2EP coatings with different doped EG5000 contents are shown in Figure 6. After soaking for 7 days, the data in Table S6 indicate that the coating with 0.075 wt.% EG5000 exhibited a greater electrochemical impedance (2.39 × 108 Ω·cm2) compared with those with lower contents of EG5000—0.025 wt.% (1.47 × 108 Ω·cm2), 0.04 wt.% (1.96 × 108 Ω·cm2), and 0.05 wt.% (2.18 × 108 Ω·cm2). From 7 to 40 days, the permeation of electrolytes resulted in a decrease in resistance values for coatings with varying EG5000 contents. However, after 40 days, the electrochemical impedance of the coating with 0.075 wt.% EG5000 remained at 5.95 × 107 Ω·cm2. As EG5000 content increases, the pores on the surface of the epoxy are fully filled, which reduces the porosity of the coating, increases the electrical impedance, and effectively compensates for the defects, thus protecting the steel matrix. Thus, the anti-corrosion properties of the hydrophobic epoxy resin coating were enhanced by the appropriate addition of graphite.
To study the state of the coating at different stages in detail, Bode modulus experiments were conducted. Similar to the impedance arc in the Nyquist diagram, the Bode modulus (f = 0.01 Hz) is an important electrochemical parameter for evaluating the corrosion resistance of coatings; organically coated samples with higher Bode moduli tend to exhibit better anti-corrosive properties. The impedance modulus of the EG5000/C1P0.2EP coating decreased with the extension of the corrosion time (Figure 7a), indicating a transition from capacitance to resistance in electrochemical behavior. Despite this, the coating resistance value of the EG5000/C1P0.2EP coating remained the highest after 60 days of immersion, maintaining the most stable anti-corrosive coating. This confirms that the appropriate size of graphite (EG5000) effectively suppresses the corrosive medium. Furthermore, it is straightforward to observe the number of time constants in Bode plots (phase angle), and the phase angle at high frequency can also be considered a useful parameter for evaluating a coating’s protective performance. Because of high resistance, this results in higher phase angles between the current and the voltage. At the initial stage of immersion (Figure S10), the phase angle curves show that the phase angles of all coated samples were greater than 60°, except for those coated with EG1000/C1P0.2EP. Notably, the phase angle of the EG5000/C1P0.2EP coating approached 85°. These results demonstrate that EG5000/C1P0.2EP hydrophobic films exhibit distinct corrosion behaviors with higher phase angles at high frequencies and significantly improved corrosion resistance, aligning with the conclusions drawn from Nyquist plots. Additionally, considering the literature discussed in the Introduction, the performance and yield were compared with the other materials listed in Table S9 [17,18,28,29,30]. Although different graphene derivatives were used with varying filler contents, our materials demonstrated the highest anti-corrosion properties, the longest immersion duration, and the smallest filler content.
Tafel polarization curves for four types of coatings are presented in Figure 7b. Specific data from a series of electrochemical measurements, including corrosion current (Icorr), corrosion potential (Ecorr), and corrosion rate (η), are listed in Table S4. Information about Icorr and Ecorr was derived from the intersection points of anodic and cathodic polarization curves. Generally, lower Icorr values or higher Ecorr values indicate better corrosion resistance [18]. For the pristine Q235, the corrosive current and corrosive potential were recorded at −631.36 mV and 9.40 × 10−6 A·cm−2, respectively. These results suggest the presence of numerous micropores or microcracks on the surface or interior of Q235, allowing O2, H2O, and Cl to diffuse through these microchannels to the substrate surface, thus reducing its corrosion resistance. In comparison, the corrosion potential of the EG5000/C1P0.2EP coating shifted in a positive direction, indicating that its anti-corrosion effect primarily reflected the inhibition of the anodic corrosion reaction. The optimized size of nanographite in the coating could form a nano-network that covers the micropores or microcracks of the Q235 substrate, impeding the diffusion of the corrosive medium. As the nanographite addition increased, Ecorr increased and Icorr decreased (Figure S6c and Table S7). With the addition of 0.075 wt.% EG5000, Ecorr improved as the nanographite content increased, reaching a peak of −442.76 mV on EG5000/C1P0.2EP, thereby exhibiting exceptional anti-corrosion properties. This effect is attributed to the moderate nanographite being adsorbed into the gaps between nano-kaolin and the epoxy resin, which reduces the likelihood of the corrosive medium penetrating the coating interior and prolongs the corrosion pathway.

3.2.2. Barrier Properties

To assess the corrosion resistance of the EP substrates doped with different sizes of EGx for over 300 h in an accelerated corrosion environment, coatings including C1P0.2EP, EG1000/C1P0.2EP, EG3000/C1P0.2EP, EG5000/C1P0.2EP, and EG8000/C1P0.2EP were subjected to a neutral salt spray test (NSS). As shown in Figure S9, all coatings experienced damage to varying degrees. However, the pristine C1P0.2EP exhibited the worst anti-corrosion performance, showing apparent rust at the cross-section and separation from the Q235 steel. In contrast, while the EG1000/C1P0.2EP, EG3000/C1P0.2EP, and EG8000/C1P0.2EP coatings developed brown rusty products and blisters after 300 h, the EG5000/C1P0.2EP coating did not exhibit conspicuous corrosion features such as rust or blisters. Thus, the NSS tests indicated that EG5000/C1P0.2EP exhibited the best anti-corrosion performance among these coatings. Notably, among the various nanographite-doped EG5000/C1P0.2EP coatings, the one with 0.075 wt% EG5000 maintained the best appearance after 400 h, demonstrating superior anti-corrosion effects (Figure S10).

3.3. Theoretical Calculation and Anti-Corrosion Mechanism of EGx/C1P0.2EP

To elucidate the mechanism by which graphite fragments enhance the corrosion resistance of EGx/C1P0.2EP, theoretical calculations were conducted using density functional theory with the PBE form of the generalized gradient approximation functional (GGA). The graphite (001) surface was modeled by slicing graphite along the [001] direction, choosing a single-layer and a double-layer slab for analysis. In the geometry optimizations, all atomic positions were allowed to relax. A vacuum layer of 18 Å was applied along the c and a axes to prevent periodic interactions.
Given that O2 and H2O are the primary oxidants in the EGx/C1P0.2EP environment, their adsorption was investigated. As depicted in Figure 8 and Figure S11, the adsorption sites on single-layer (s-Gra) and double-layer graphite (d-Gra) at the body center and edge locations were studied. Physisorption was observed for both H2O and O2 at body center sites, with adsorption energies close to zero. However, at edge sites, chemisorption occurred for both O2 and H2O on s-Gra and d-Gra. Figure 9 shows that the chemisorption of H2O leads to the natural decomposition of the H-O bond, producing OH and H species on the edges of s-Gra and d-Gra. Compared with the H2O (−0.95 eV) and O2 (−0.11 eV) adsorption on the EGx/C1P0.2EP body center, the edges of s-Gra/d-Gra were more favorable (Figure 8). These results suggest that smaller graphene particles, which have a large edge surface, are more effective in attracting H2O and O2, reducing their potential to damage the coating and thus inhibiting the corrosion of EGx/C1P0.2EP by oxidation. This theoretical insight aligns well with our experimental findings.

4. Conclusions

To broaden the selectivity of graphite nanosheets as additives in epoxy resin and to verify their size effect, EGx was extensively incorporated into EP coatings. Micro-kaolin and PTFE were also uniformly integrated onto the surface to form hydrophobic long chains. EGx of various sizes was prepared using a cost-effective ball-milling and gradient centrifugation approach. Notably, the impedance arc of the EG5000/C1P0.2EP coating was the largest, registering an impedance of 2.58 × 108 Ω·cm2, indicating superior corrosion resistance compared with the EG1000/C1P0.2EP, EG3000/C1P0.2EP, and EG8000/C1P0.2EP coatings. Furthermore, even after 300 h of NSS testing, the EG5000/C1P0.2EP coating did not exhibit conspicuous corrosion features, demonstrating greater corrosion resistance than the other sizes of EGx coatings. The optimized size (2.21 μm) of EG5000 in the EG5000/C1P0.2EP coating forms a nano-network that covers micropores or microcracks, effectively impeding the diffusion of corrosive media. In addition, theoretical calculations revealed that the chemisorption of H2O leads to the natural decomposition of H–O, producing OH and H species on the edges of s-Gra and d-Gra. Compared with the adsorption energies for H2O (−0.95 eV) and O2 (−0.11 eV) on EGx/C1P0.2EP, the edges of s/d-Gra are more favorable. These findings suggest that smaller graphene particles are more effective in attracting H2O and O2, thereby reducing their potential to damage the coating and inhibit the corrosion of EGx/C1P0.2EP through oxidation.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/coatings14060769/s1. References [32,33,34,35,36] are cited in Supplementary Materials.

Author Contributions

Conceptualization, K.L. and D.W.; methodology, H.M. and X.L.; formal analysis, D.W.; investigation, Y.B. and Y.Z.; writing—original draft preparation, K.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China under Grant nos. 52102119, 52072356, and 52032011, the Young Elite Scientists Sponsorship Program by CAST (No. 2022QNRC001), and the talent project of Zao Zhuang (ZZYF-05).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are contained within this article and Supplementary Materials.

Conflicts of Interest

Authors Kuilin Lv, Yiwang Bao, Huachao Ma, Xiaogen Liu and Detian Wan were employed by the company China Building Materials Academy Co., Ltd. The remaining author Ying Zhu declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Bejinariu, C.; Burduhos-Nergis, D.-P.; Cimpoesu, N. Immersion behavior of carbon steel, phosphate carbon steel and phosphate and painted carbon steel in saltwater. Materials 2021, 14, 188. [Google Scholar] [CrossRef]
  2. Vovchenko, L.; Lazarenko, O.; Matzui, L.; Perets, Y.; Zhuravkov, A.; Fedorets, V.; Le Normand, F. Mechanical and electrical properties of the epoxy composites with graphite nanoplatelets and carbon nanotubes. Phys. Status Solidi 2014, 211, 336–441. [Google Scholar] [CrossRef]
  3. Ding, J.; Zhao, H.; Zheng, Y.; Zhao, X.; Yu, H. A long-term anticorrsive coating through graphene passivation. Carbon 2018, 138, 197–206. [Google Scholar] [CrossRef]
  4. Wu, F.; Chen, J.; Li, L.; Zhao, T.; Chen, R. Improvement of rate and cycle performence by rapid polyaniline coating of a MWCNT/Sulfur cathode. J. Phys. Chem. C 2011, 115, 24411–24417. [Google Scholar] [CrossRef]
  5. Perrin, F.; Phan, T.; Nguyen, D. Synthesis and characterization of polyaniline nanoparticles in phosphonic acid amphiphile aqueous micellar solutions for waterborne corrosion protection coatings. J. Polym. Sci. Polym. Chem. Ed. 2015, 53, 1606–1616. [Google Scholar] [CrossRef]
  6. Chen, F.; Liu, P. Conducting polyaniline nanoparticles and their dispersion for waterborne corrosion protection coatings. ACS Appl. Mater. Interfaces 2011, 3, 2694–2702. [Google Scholar] [CrossRef] [PubMed]
  7. Gu, L.; Zhao, X.; Tong, X.; Ma, J.; Chen, B.; Liu, S.; Zhao, H.; Yu, H.; Chen, J. Facile preparation of polyaniline nanoparticles and their dispersion for waterborne anticorrosion coatings. Int. J. Electrochem. Sci. 2016, 11, 1621–1631. [Google Scholar] [CrossRef]
  8. Qiu, S.; Chen, C.; Cui, M.; Li, W.; Zhao, H.; Wang, L. Corrosion protection performance of waterborne epoxy coatings containing self-doped polyaniline nanofiber. Appl. Surf. Sci. 2017, 407, 213–222. [Google Scholar] [CrossRef]
  9. Berber, H.; Sarac, A.; Yıldırım, H. A comparative study on water-based coatings prepared in the presence of oligomeric and conventional protective colloids. Prog. Org. Coat. 2011, 71, 225–233. [Google Scholar] [CrossRef]
  10. Wang, N.; Fu, W.; Sun, M.; Zhang, J.; Fang, Q. Effect of different structured TiO2 particle on anticorrosion properties of waterborne epoxy coatings. Corros. Eng. Sci. Technol. 2016, 51, 365–372. [Google Scholar] [CrossRef]
  11. Wang, N.; Diao, X.; Zhang, J.; Kang, P. Corrosion resistance of waterborne epoxy coatings by incorporation of dopamine treated Mesoporous-TiO2 particles. Coatings 2018, 8, 209. [Google Scholar] [CrossRef]
  12. Cui, M.; Ren, S.; Zhao, H.; Xue, Q.; Wang, L. Polydopamine coated graphene oxide for anticorrosive reinforcement of water borne epoxy coating. Chem. Eng. J. 2018, 335, 255–266. [Google Scholar] [CrossRef]
  13. Ding, R.; Zheng, Y.; Yu, H.; Li, W.; Wang, X.; Gui, T. Study of water permeation dynamics and anti-corrosion mechanism of Graphene/Zinc coatings. J. Alloys Compd. 2018, 748, 481–495. [Google Scholar] [CrossRef]
  14. Sheng, X.; Mo, R.; Ma, Y.; Zhang, X.; Zhang, L.; Wu, H. Waterborne epoxy resin/polydopamine modified zirconium phosphate nanocomposite for anticorrosive coating. Ind. Eng. Chem. Res. 2019, 58, 16571–16580. [Google Scholar] [CrossRef]
  15. Huang, T.; Su, Y.; Yeh, T.; Huang, H.; Wu, C.; Huang, K.; Chou, Y.; Yeh, J.; Wei, Y. Advanced anticorrosive coatings prepared from electroactive Epoxy-SiO2 hybrid nanocomposite materials. Electrochim. Acta 2011, 56, 6142–6149. [Google Scholar] [CrossRef]
  16. Martín-Fabiani, I.; Lesage, D.; Schulz, M.; Liu, Y.; Lee, M.; Duffy, B.; Agosto, F.; Lansalot, M.; Keddie, J. Enhanced water barrier properties of surfactant-free polymer films obtained by MacroRAFT-Mediated emulsion polymerization. ACS Appl. Mater. Interfaces 2018, 10, 11221–11232. [Google Scholar] [CrossRef] [PubMed]
  17. Ding, J.; Rahman, O.; 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]
  18. Gu, L.; Liu, S.; Zhao, H.; Yu, H. Facile Preparation of Water Dispersible Graphene Sheets Stabilized by Carboxylated Oligoanilines and Their Anticorrosion Coatings. ACS Appl. Mater. Interfaces 2015, 7, 17641–17648. [Google Scholar] [CrossRef]
  19. Li, Z.; Li, J.; Cui, J.; Qiu, H.; Yang, G.; Zheng, S.; Yang, J. Dispersion and parallel assembly of sulfonated graphene in waterborne epoxy anticorrosion coatings. J. Mater. Chem. A 2019, 7, 17937–17946. [Google Scholar] [CrossRef]
  20. Yu, Z.; Di, H.; Ma, Y.; Lv, L.; Pan, Y.; Zhang, C.; He, Y. Fabrication of graphene oxidealumina hybrids to reinforce the anti-corrosion performance of composite epoxy coatings. Appl. Surf. Sci. 2015, 351, 986–996. [Google Scholar] [CrossRef]
  21. Monetta, T.; Acquesta, A.; Carangelo, A.; Bellucci, F. Considering the effect of graphene loading in water-based epoxy coatings. J. Coat. Technol. Res. 2018, 15, 923–931. [Google Scholar] [CrossRef]
  22. Kopsidas, S.; Olowojoba, G.; Kinloch, A.; Taylor, A. Examining the effect of graphene nanoplatelets on the corrosion resistance of epoxy coatings. Int. J. Adhes. Adhes. 2021, 104, 102723. [Google Scholar] [CrossRef]
  23. Huang, H.; Li, M.; Tian, Y.; Xie, Y.; Sheng, X.; Jiang, X.; Zhang, X. Exfoliation and functionalization of α-zirconium phosphate in one pot for waterborne epoxy coatings with enhanced anticorrosion performance. Prog. Org. Coat. 2020, 138, 105390. [Google Scholar] [CrossRef]
  24. Lv, K.; Pan, R.; Zhang, L.; Tian, Y.; Yan, S.; Wan, D. Synergistically assembled graphene/ZnO composite to enhance anticorrosion performance of waterborne epoxy coatings. RSC Adv. 2022, 15, 9069–9076. [Google Scholar] [CrossRef]
  25. Teng, C.; Xie, D.; Wang, J.; Yang, Z.; Ren, G.; Zhu, Y. Ultrahigh Conductive Graphene Paper Based on Ball-Milling Exfoliated Graphene. Adv. Funct. Mater. 2017, 27, 1700240. [Google Scholar] [CrossRef]
  26. Wang, J.; Du, P.; Zhao, H.; Pu, J.; Yu, C. Novel nitrogen doped carbon dots enhancing the anticorrosive performance of waterborne epoxy coatings. Nanoscale Adv. 2019, 1, 3443–3451. [Google Scholar] [CrossRef]
  27. Irfan, M.; Iqbal, S.; Ahmad, S. Waterborne reduced graphene oxide dispersed sebacic acid modified soy epoxy nanocomposite: A green and sustainable approach for high performance mechanically robust anticorrosive coatings. Prog. Org. Coat. 2022, 170, 106984. [Google Scholar] [CrossRef]
  28. Liu, S.; Gu, L.; Zhao, H.; Chen, J.; Yu, H. Corrosion Resistance of Graphene-Reinforced Waterborne Epoxy Coatings. J. Mater. Sci. Technol. 2016, 32, 425–431. [Google Scholar] [CrossRef]
  29. Wang, S.; Hu, Z.; Shi, J.; Chen, G.; Zhang, Q.; Weng, Z.; Wu, K.; Lu, M. Green synthesis of graphene with the assistance of modified lignin and its application in anticorrosive waterborne epoxy coatings. Appl. Surf. Sci. 2019, 484, 759–770. [Google Scholar] [CrossRef]
  30. Huan, Y.; Dan, L.; Xu, H. Graphene-induced enhanced anticorrosion performance of waterborne epoxy resin coating. Front. Mater. Sci. 2020, 14, 211–220. [Google Scholar]
  31. Lv, K.; Wan, D.; Pan, R. Curvature of NCNTs induced selectivity of CO2 electroreduction into CO. Carbon Neutralization 2022, 1, 189–197. [Google Scholar] [CrossRef]
  32. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  33. Kresse, G.; Furthmüller, J. Efficiency of Ab-initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  34. Kresse, G.; Hafner, J. Ab initio Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  35. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  36. Grimme, S.; Antony, J.; Ehrlich, S.; Krief, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 elements H–Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The surface SEM of EG1000 (a), EG3000 (b), EG5000 (c), and EG8000 (d); the red frame represents the field of vision, and the blue frame represents the graphene nanosheet.
Figure 1. The surface SEM of EG1000 (a), EG3000 (b), EG5000 (c), and EG8000 (d); the red frame represents the field of vision, and the blue frame represents the graphene nanosheet.
Coatings 14 00769 g001
Figure 2. HRTEM image of EG1000 (a), EG3000 (b), EG5000 (c), and EG8000 (d). Size statistics of EG8000 (e), EG5000 (f), EG3000 (g), and EG1000 (h).
Figure 2. HRTEM image of EG1000 (a), EG3000 (b), EG5000 (c), and EG8000 (d). Size statistics of EG8000 (e), EG5000 (f), EG3000 (g), and EG1000 (h).
Coatings 14 00769 g002
Figure 3. (a,b) Raman spectra of G, EG1000, EG3000, EG5000, and EG8000, respectively. (c) FTIR spectra of G, EG1000, EG3000, EG5000, and EG8000, respectively.
Figure 3. (a,b) Raman spectra of G, EG1000, EG3000, EG5000, and EG8000, respectively. (c) FTIR spectra of G, EG1000, EG3000, EG5000, and EG8000, respectively.
Coatings 14 00769 g003
Figure 4. (a) Water contact angle of EGx/C1P0.2EP coatings with different storage times. (b) Water contact angle of EG5000/C1P0.2EP with different EG5000 contents.
Figure 4. (a) Water contact angle of EGx/C1P0.2EP coatings with different storage times. (b) Water contact angle of EG5000/C1P0.2EP with different EG5000 contents.
Coatings 14 00769 g004
Figure 5. Nyquist plots of the EG1000/C1P0.2EP, EG3000/C1P0.2EP, EG5000/C1P0.2EP, and EG8000/C1P0.2EP coatings after 7 d (a), 30 d (b), and 60 d (c) of immersion, respectively.
Figure 5. Nyquist plots of the EG1000/C1P0.2EP, EG3000/C1P0.2EP, EG5000/C1P0.2EP, and EG8000/C1P0.2EP coatings after 7 d (a), 30 d (b), and 60 d (c) of immersion, respectively.
Coatings 14 00769 g005
Figure 6. Nyquist plots of EG5000/C1P0.2EP with different EG5000 addition amounts after 7 d (a), 30 d (b), and 60 d (c) of immersion.
Figure 6. Nyquist plots of EG5000/C1P0.2EP with different EG5000 addition amounts after 7 d (a), 30 d (b), and 60 d (c) of immersion.
Coatings 14 00769 g006
Figure 7. (a) Bode modulus plots of EG5000/C1P0.2EP after 7 d, 14 d, 30 d, and 60 d of immersion. (b) The polarization curve of EG1000/C1P0.2EP, EG3000/C1P0.2EP, EG5000/C1P0.2EP and EG8000/C1P0.2EP after 60 d of neutral salt spray test.
Figure 7. (a) Bode modulus plots of EG5000/C1P0.2EP after 7 d, 14 d, 30 d, and 60 d of immersion. (b) The polarization curve of EG1000/C1P0.2EP, EG3000/C1P0.2EP, EG5000/C1P0.2EP and EG8000/C1P0.2EP after 60 d of neutral salt spray test.
Coatings 14 00769 g007
Figure 8. The adsorption structures and energies of O2 and H2O at varying sites on single layers of graphite sheet.
Figure 8. The adsorption structures and energies of O2 and H2O at varying sites on single layers of graphite sheet.
Coatings 14 00769 g008
Figure 9. The adsorption energies of O2 and H2O at varying sites on single and double layers of graphite sheet as well as EGx/C1P0.2EP.
Figure 9. The adsorption energies of O2 and H2O at varying sites on single and double layers of graphite sheet as well as EGx/C1P0.2EP.
Coatings 14 00769 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lv, K.; Bao, Y.; Ma, H.; Liu, X.; Zhu, Y.; Wan, D. Size Effect of Graphite Nanosheet-Induced Anti-Corrosion of Hydrophobic Epoxy Coatings. Coatings 2024, 14, 769. https://doi.org/10.3390/coatings14060769

AMA Style

Lv K, Bao Y, Ma H, Liu X, Zhu Y, Wan D. Size Effect of Graphite Nanosheet-Induced Anti-Corrosion of Hydrophobic Epoxy Coatings. Coatings. 2024; 14(6):769. https://doi.org/10.3390/coatings14060769

Chicago/Turabian Style

Lv, Kuilin, Yiwang Bao, Huachao Ma, Xiaogen Liu, Ying Zhu, and Detian Wan. 2024. "Size Effect of Graphite Nanosheet-Induced Anti-Corrosion of Hydrophobic Epoxy Coatings" Coatings 14, no. 6: 769. https://doi.org/10.3390/coatings14060769

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