Next Article in Journal
Experimental and Numerical Study on Flexural Behavior of Fold-Fastened Multi-Cellular Steel Panels
Next Article in Special Issue
A Simplified Empirical Model for Predicting Residual Flexural Capacity of Corroded Prestressed Concrete Beams
Previous Article in Journal
HFE-YOLO: Hybrid Feature Enhancement with Multi-Attention Mechanisms for Construction Site Object Detection
Previous Article in Special Issue
Electrochemical and Gravimetric Assessment of Steel Rebar Corrosion in Chloride- and Carbonation-Induced Environments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Research on 2-Methylimidazole-Lanthanum Nickel-Based Sol-Gel Conversion Coating for Oxide Scale Reinforcement Bars

Marine Corrosion and Protection Team, School of Chemical Engineering and Technology, Sun Yat-sen University, Zhuhai 519082, China
*
Author to whom correspondence should be addressed.
Buildings 2025, 15(23), 4272; https://doi.org/10.3390/buildings15234272
Submission received: 29 October 2025 / Revised: 18 November 2025 / Accepted: 24 November 2025 / Published: 26 November 2025
(This article belongs to the Special Issue Research on Corrosion Resistance of Reinforced Concrete)

Abstract

Corrosion induced by defective oxide scales severely compromises the durability of concrete structures. This study develops a dual-mechanism sol-gel protection strategy based on La3+/Ni2+/2-Methylimidazole (2-MI). First, 2-Methylimidazole-catalyzed epoxy ring-opening constructs defect-minimized Si–O–Si/C–O–C networks through 60 °C low-temperature curing, reducing microcrack formation and curing energy consumption compared to conventional 130 °C processing. Second, utilizing 400 °C waste heat from hot-rolled steel triggers pH-modulated La2O3/NiO co-deposition within oxide scale defects, enhancing corrosion resistance. After a 40-day immersion in SCP + 0.1 M NaCl, the coated reinforcement exhibits a low-frequency impedance modulus of 25.6 kΩ·cm2, achieving a 10.4-fold increase over untreated steel. Specimens embedded in 3.5 wt% NaCl mortar demonstrate a 120-day impedance modulus of 74.63 kΩ·cm2, exceeding the control by 8.03-fold. This strategy integrates efficient industrial waste heat utilization with energy-saving low-temperature curing, providing long-term corrosion protection for marine concrete structures.

1. Introduction

Reinforced concrete structures are widely used in marine infrastructure due to their corrosion resistance. The highly alkaline concrete matrix (pH ≥ 12.5) facilitates protective passive film formation on steel reinforcement, ensuring durability [1,2,3,4,5,6]. However, in marine settings, chloride ions (Cl) readily penetrate the porous oxide scale formed during the hot-rolling of steel, disrupting the passive film and initiating localized corrosion, stress corrosion cracking, and concrete spalling, ultimately leading to catastrophic structural failure [7,8,9]. Consequently, enhancing the corrosion resistance of reinforced concrete in marine environments has emerged as a critical research focus.
Traditional strategies include the following: stainless steel reinforcement (enhanced by Cr/Ni/Mo passivating elements), though prohibitively costly for large-scale applications [10,11]; and protective coatings and cathodic protection, limited by complex implementation, poor interfacial adhesion, and high long-term maintenance [12,13,14,15]. Recent advances target chemical conversion coatings on steel oxide scales/rust layers [16,17]. Sol-gel technology offers a low-cost, facile route to form dense inorganic–organic hybrid films that penetrate the substrate, creating covalent Si–O–M bonds for enhanced adhesion [18,19,20,21] and corrosion resistance [22,23,24]. Previous work by our group demonstrated that rare-earth ion doping (e.g., Ce3+) repairs inherent oxide scale defects by generating nano-CeO2 films at defect sites [24,25]. Further studies revealed that semi-coherent epitaxial growth of nanocrystalline CeO2 on the oxide scale introduces grain boundary defects, which are mitigated by Ni-induced formation of a stable NiFe2O4 spinel phase, significantly improving defect remediation and corrosion resistance [26]. Lanthanum—a rare earth with high oxygen affinity—forms dense La(OH)3/La2O3 layers on carbon steel, suppressing corrosion [27]. Thus, La3+-doped sol-gel systems hold promise for enhancing oxide scale remediation.
The current industrial practice primarily recovers high-grade waste heat (>650 °C) and medium-grade waste heat (230–650 °C) from hot-rolled rebars via water cooling for steam generation, rarely considering concurrent corrosion resistance enhancement. To address this gap, we propose a field-applicable innovation that efficiently utilizes medium-grade waste heat after high-grade heat recovery to integrate corrosion protection [28,29,30,31]. Our prior research on sol-gel coatings for steel reinforcement required a 130 °C curing temperature, where solvent evaporation induced microcracks that severely compromised coating performance [24,25,26]. Simultaneously, high-temperature curing increased production energy consumption, constraining large-scale application. Herein, we innovatively implement 60 °C low-temperature curing to minimize microcracking and energy consumption.
Industrial context: after 800–1000 °C water cooling, rebars enter the medium-grade waste heat zone (230–650 °C) with ~400 °C residual surface temperatures. Field implementation: direct immersion in La/Ni/2-Methylimidazole-modified sol-gel precursor for 30 s at this stage, followed by 60 °C curing, significantly enhances corrosion resistance through defect remediation. Lab simulation: (1) preheat rebars at 400 °C for 1 h in a muffle furnace; (2) immerse in precursor where 2-Methylimidazole catalyzes GPTMS epoxy ring-opening [32] and chelates La3+/Ni2+ [33,34]; and (3) cure at 60 °C for 4 h.
This strategy simultaneously reduces microcracks via crosslinking enhancement and low-temperature curing while upgrading waste heat utilization through corrosion protection. The study develops an energy-efficient sol-gel method for remediating oxide scale defects, improving concrete durability. Key innovations include the following: waste-heat-driven corrosion protection; 2-Methylimidazole-mediated chelation-crosslinking; and pH-modulated La2O3/NiO/SiO2 co-deposition for defect repair. This work provides a novel solution for enhancing marine concrete durability.

2. Materials and Methods

2.1. Materials

Substrate: as-received HRB400 steel reinforcement with oxide scale (composition in Table 1).
Chemicals: ethanol (ETOH), sodium chloride (NaCl), tetraethyl orthosilicate (TEOS), Macklin Biochemical Technology Co., Ltd. (Shanghai, China); (3-glycidyloxypropyl) trimethoxysilane (GPTMS), ammonia solution (NH3·H2O), Meryer Chemical Technology Co., Ltd. (Shanghai, China); 2-Methylimidazole, calcium hydroxide (Ca(OH)2), lanthanum nitrate hexahydrate (La(NO3)3·6H2O), nickel nitrate hexahydrate (Ni(NO3)2·6H2O), Aladdin Biochemical Technology Co., Ltd. (Shanghai, China). All reagents were employed as received without further purification.

2.2. Electrode and Electrolyte Preparation

Cylindrical steel samples (Ø 12 mm × 20 mm) with intact oxide scales served as working electrodes. Electrical contact was established by connecting one end to a copper wire, followed by sealing both ends with rosin-paraffin wax to expose a lateral surface area of 7.536 cm2. The electrolyte was a simulated concrete pore SCP solution (saturated Ca(OH)2 containing 0.1 M NaCl).

2.3. Sol-Gel Coating Deposition

A precursor solution was prepared by mixing TEOS, GPTMS, and ETOH (1:3:2 v/v) under magnetic stirring (600 rpm, 120 min) and aging for 3 days [26]. After adding specified volumes of 0.1 M La(NO3)3 and 0.5 M Ni(NO3)2, the pH was adjusted to 10.0 with 25 wt% NH3·H2O. To ensure safety during 400 °C immersion, operators wore flame-resistant lab coats and high-temperature resistant gloves, using titanium-coated extended tweezers (length ≥ 30 cm) to handle steel samples preheated at 400 °C for 1 h. Within a fume hood, samples were immersed in precursor solution for 30 s, during which the temperature dropped from 400 °C to 80 °C within 30 s. Operators maintained a safe distance (>30 cm) throughout the process. Immediately after immersion, samples were transferred to a 60 °C oven for 4 h of drying. This procedure ensured safety through engineering controls (ventilation system) and personal protective equipment, with solvent vapor concentration consistently <10% LEL (Lower Explosive Limit).
Coating nomenclature: LS: 1.5 mL La(NO3)3 + 0 mL Ni(NO3)2 (per 1 mL TEOS); 0.8LNS: 1.2 mL La(NO3)3 + 0.3 mL Ni(NO3)2; 0.2LNS: 0.75 mL La(NO3)3+ 0.75 mL Ni(NO3)2; 0.05LNS: 0.3 mL La(NO3)3 + 1.2 mL Ni(NO3)2; NS: 0 mL La(NO3)3 + 1.5 mL Ni(NO3)2. For 2-Methylimidazole-modified coatings, 0.52 wt% 2-MI (relative to GPTMS mass) was added during precursor preparation.

2.4. Mortar Specimen Fabrication

Portland cement (composition in Table 2) and deionized water (water–cement ratio 0.35) were mixed for 10 min, cast into 70 × 70 × 70 mm molds containing steel reinforcement, and cured at room temperature for 24 h. After demolding, exposed steel was sealed with epoxy resin, and specimens were cured for 28 days (20 ± 1 °C, RH 95 ± 2%).

2.5. Electrochemical Measurements

Tests were performed on a Gamry Interface 1010E system using a three-electrode configuration.
SCP solution tests: working electrode: HRB400 steel (7.536 cm2); reference electrode: saturated calomel electrode (SCE); counter electrode: platinum foil (2.5 × 2.5 cm2). EIS: 10 mV amplitude, 105–10−2 Hz after OCP stabilization (≥1 h); potentiodynamic polarization: −0.3 to 1.2 V vs. OCP, 0.333 mV/s.
Mortar-embedded steel tests: electrolyte: 3.5 wt% NaCl; working electrode: steel reinforcement embedded in mortar specimens; reference electrode: saturated calomel electrode (SCE); counter electrode: titanium mesh (other parameters identical to above). Data were analyzed with ZsimpWin 3.6 software.

2.6. Material Characterization

Surface/cross-section morphology: SEM (Thermo Fisher Apreo 2S, Waltham, MA, USA); elemental distribution: EDS (Oxford Instruments, Oxfordshire, UK); chemical states: XPS (Shimadzu AXIS SUPRA+, Kyoto, Japan; Vantage™ 6.9.0 software); functional groups: FTIR (Nicolet 6700, 32 scans, 4 cm−1 resolution, 400–4000 cm−1, Waltham, MA, USA); crystalline structure: XRD (Rigaku Ultima IV, Tokyo, Japan).

3. Results

3.1. Effect of La3+/Ni2+ Ratios and 2-Methylimidazole Content on Coating Morphology

3.1.1. Surface Morphology Analysis

Surface morphologies of Blank HRB400 steel and sol-gel conversion coatings under varied conditions are shown in Figure 1. The native oxide scale on the Blank steel exhibits abundant micropores and microcracks (Figure 1a), which facilitate corrosive species penetration, accelerating substrate corrosion and compromising concrete durability. Coatings derived from sol-gel precursors with different La3+/Ni2+ ratios (pH 10) effectively seal inherent defects in the oxide scale (Figure 1b–g). Notably, precursors without metal ions yield coatings with pronounced defects (Figure 1b), whereas metal-ion-containing formulations reduce defect density. This is attributed to chemical crosslinking between inorganic ions and organic moieties, forming M–O–M bonds that enhance three-dimensional network formation [35,36]. La3+-only coatings demonstrate superior smoothness and reduced colloidal particles versus Ni2+-only counterparts (Figure 1c,d), consistent with reports that La3+ improves film-forming capability in alkaline sol-gels [37].
Optimal performance occurs at La3+:Ni2+ = 0.8 (concentration ratio), generating coatings with minimal surface defects (Figure 1e). Deviations from this ratio degrade coating quality (Figure 1f,g), corroborating findings that dual-ion incorporation optimizes microstructure and functionality [38].
Building on the optimal 0.8 La3+/Ni2+ concentration ratio, 2-Methylimidazole (0.5–2 wt% relative to GPTMS) was introduced (Figure 1h–j). 2-Methylimidazole catalyzes epoxy ring-opening in GPTMS, accelerating hydrolysis/condensation and increasing crosslinking density [30,31,32]. Consequently, coatings with 1 wt% 2-MI exhibit the smoothest morphology with negligible particle aggregation (Figure 1i), attributable to enhanced sol stability and controlled metal ion hydrolysis.

3.1.2. Cross-Sectional Morphology Analysis

Backscattered electron (BSE) imaging reveals distinct differences in oxide scale morphology (Figure 2a–c). Blank HRB400 steel exhibits extensive defects/cracks (Figure 2a), facilitating Cl ingress. The 0.8LNS treatment forms a dense hybrid film that reduces defects but retains microcracks/pores (Figure 2b) [28,29]. Incorporation of 1 wt% 2-Methylimidazole eliminates defects via a crosslinking-enhanced 3D network, forming a homogeneous barrier (Figure 2c) [34,35,36]. Oxide scale thickness metrics demonstrate the following: Blank HRB400 steel average 8.54 µm (variance 3.21), and 0.8LNS averages 8.01 µm (variance 3.73), while 1 wt% 2-MI—0.8LNS shows 9.25 µm with significantly reduced variance (0.0958), indicating remarkable uniformity improvement. Sol-gel coating thickness metrics show the following: 0.8LNS averages 49.38 µm (variance 2.93) and 1 wt% 2-MI—0.8LNS averages 31.88 µm (variance 2.77).

3.2. Compositional Analysis

3.2.1. XRD Characterization

XRD patterns of Blank HRB400 and sol-gel coatings are shown in Figure 3a. Key observations are as follows: peaks at 2θ = 44.51°, 49.15°, and 56.85° are assigned to Fe [25,39]. Iron oxides are identified by peaks at 33.21°, 35.57°, 46.48°, 53.79°, 57.52°, 62.45°, 63.99° (Fe2O3); 36.14° (FeO); and 29.34°, 35.32°, 43.07°, 56.94°, 62.81° (magnetite, Fe3O4) [40,41]. Additional phases (e.g., FeOOH) are present. Coatings exhibit a broad peak near 2θ = 21°, confirming amorphous SiO2 formation [25].

3.2.2. FTIR Spectroscopy

FTIR spectra of precursor solutions and resultant coatings (Figure 3(b1,b2,d1,d2)) reveal the following:
Precursor solutions (Figure 3(b1)): peaks at 1170 and 965 cm−1 correspond to CH3 rocking vibrations; 1045 and 784 cm−1 to Si–O–Si bending; 878 cm−1 to Si–O–H stretching [42,43,44,45]; and 1258 and 911 cm−1 to epoxy group signatures (GPTMS) [25,43,46]. To determine whether La3+/Ni2+ and 2-Methylimidazole incorporation promote epoxy ring-opening, quantitative FTIR peak area ratio analysis of epoxy group/siloxane bonds was performed using OMNIC 9.2 software. The normalized peak areas yielded ratios of 1 wt% 2-MI—0.8LNS (0.0290), 0.8LNS (0.0403), and Undoped sol-gel (0.0749). This confirms metal ions and 2-Methylimidazole synergistically catalyze epoxy ring-opening, enhancing crosslinking density [30,31,32]. In the expanded view (400–700 cm−1, Figure 3(b2)), the absence of Ni–O (434–454, 558–570 cm−1) or La–O (486–499 cm−1) bands [47,48,49] confirms metal hydroxide predominance post-pH adjustment.
Coatings (Figure 3(d1)): similar profiles to precursors but with diminished O–H bands (3500–3150 cm−1), indicating that high-temperature condensation enhanced Si–O–Si network formation [50]. Epoxy peak attenuation parallels solution trends: quantitative FTIR peak area ratio analysis yields values of 0.0224, 0.0616, and 0.0655, consistent with precursor solution trends. In the expanded view (400–700 cm−1, Figure 3(d2)), the emergence of Ni–O (434–454/558–570 cm−1) and La–O (486–499 cm−1) bands [48,49] confirms La2O3/NiO formation during high-temperature deposition.

3.2.3. XPS Analysis

XPS spectra of Blank HRB400 steel and the optimized 1 wt% 2-MI—0.8LNS coating are presented in Figure 4. High-resolution spectra were deconvoluted using mixed Gaussian–Lorentzian functions (70:30 ratio) with Shirley background subtraction. All binding energies were referenced to adventitious carbon (C 1s = 284.80 eV).
For the Blank HRB400 steel (Figure 4(a1–a3)), the oxide scale composition is dominated by FeOOH (712.56 eV), Fe2O3 (710.53 eV), and FeO (709.28 eV) based on O 1s and Fe 2p deconvolution [51,52]. Synchronous La 3d and Ni 2p analysis confirms the absence of lanthanum and nickel signatures (Figure 4(a3)).
For the 1 wt% 2-MI—0.8LNS coating (Figure 4(b1–b6)), the survey spectra detected C, N, O, Si, La, and Ni, with attenuated Fe 2p signals (Figure 4(b3)). Key bonding states confirm successful coating formation: C 1s: O=C–O (287.35 eV), C–O–C (286.60 eV), C–O–Si (286.04 eV); O 1s: O–Si (532.55 eV); Si 2p: Si–O–C (103.30 eV), Si–OH (102.60 eV), Si–O–Si (101.92 eV) [49,50,51,52,53,54,55]; and N 1s spectra show nitrate-derived NO3 (407.31 eV), N–C (401.35 eV), and 2-Methylimidazole-associated N=C (399.73 eV) [56]. Metal oxide formation is evidenced by Ni 2p: NiO signatures at 852.81 eV (2p3/2) and 873.78 eV (2p1/2) with satellites (862.63, 880.91 eV) [57,58]; and La 3d: La2O3 peaks at 856.30 eV (3d3/2) and 837.32 eV (3d5/2) with satellites (852.81, 833.95 eV) [57,59]. These findings align with FTIR data confirming La2O3/NiO generation (Figure 3).
For the post-coating removal surface (Figure 4(c1–c3)), the exposed oxide scale retains FeOOH, Fe2O3, and FeO dominance [51,52], while residual O–Si bonding (Figure 4(c1)) indicates persistent Si–O–M covalent interfaces [54]. Crucially, La2O3 signatures persist at 856.30 eV (La 3d3/2) with satellites (851.99, 835.26 eV) [57,59], confirming in situ defect remediation via lanthanum oxide formation.

3.3. Corrosion Behavior of Sol-Gel Coatings in SCP + 0.1 M NaCl

3.3.1. Sol-Gel Conversion Coating Prepared Without 2-Methylimidazole

Electrochemical impedance spectroscopy (EIS) assessed the corrosion resistance of Blank HRB400 steel and coated steel in SCP + 0.1 M NaCl solution (Figure 5, Table 3) [16,17,60,61].
Nyquist analysis (Figure 5a): All sol-gel coatings exhibit larger capacitive loop diameters than untreated steel, indicating enhanced barrier properties against Cl; ingress. The 0.8LNS sol-gel coating (La3+:Ni2+ = 0.8) shows the maximum loop diameter, signifying optimal corrosion resistance. For the low-frequency impedance |Z|0.01Hz (Figure 5b), |Z|0.01Hz follows the order: Blank HRB400 < NS < LS < 0.05LNS < Undoped sol-gel < 0.2LNS < 0.8LNS. The 0.8LNS sol-gel coating achieves the highest |Z|0.01Hz (8.73 × 104 Ω·cm2), confirming superior protective performance.
Equivalent Circuit Modeling Rationale: The inherent microdefects and porosity within the oxide scale of steel reinforcement impart distinct electrical resistance and interfacial capacitance characteristics. These defects facilitate the permeation of corrosive species (e.g., Cl) to the metallic substrate, triggering electrochemical reactions. Consequently, the EIS data for untreated steel were modeled using the following equivalent circuit: (Rs(Q1(Ros(Q2Rct)))). (Figure 5(g1)) Following sol-gel modification, the coated steel exhibits an additional organic-inorganic hybrid layer with intrinsic coating resistance (Rf) and capacitance. This necessitates the following expanded circuit [25,26]: (Rs(Q0(Rf(Q1(Ros(Q2Rct)))))). (Figure 5(g2)) The following are circuit element definitions: Rs: solution resistance; Ros: oxide scale resistance; Rf: sol-gel coating resistance; Rct: charge transfer resistance; Q0/Q1/Q2: constant phase elements (CPEs), accounting for non-ideal capacitive behavior arising from interfacial heterogeneity. CPE Mathematical Formulation: The impedance of each CPE is defined as [62,63,64]:
Z   =   Y 0 1 j w n
where Y0: CPE admittance coefficient (Ω−1sncm−2); j: Imaginary unit; ω: angular frequency; and n: CPE exponent (0 ≤ n ≤ 1), where n = 0 is the pure resistor, and n = 1 is the ideal capacitor.
Fitted electrochemical parameters are presented in Table 2, with the total resistance Rtot (Rtot = Ros + Rf + Rct) compared across all samples. This parameter collectively represents the barrier properties against charge transfer and ionic penetration.
Electrochemical Performance Analysis: the 0.8LNS sol-gel coating exhibits the highest total resistance (Rtot) among all specimens (Figure 5d), demonstrating superior chloride diffusion barrier properties. Critical correlations emerge from constant phase element analysis. Oxide scale defect density correlates with CPE exponent n1 [26]. Blank HRB400 steel shows low n1 (0.42), indicating high defect concentration. Sol-gel modified specimens display significantly increased n1 values, confirming defect remediation. The 0.8LNS coating achieves maximum n1 (0.89), reflecting optimal defect sealing. Charge transfer resistance (Rct) of 0.8LNS coating exceeds Blank HRB400 steel by 35.5-fold, and oxide scale resistance (Ros) shows a 57.0-fold enhancement; these improvements confirm reduced exposed metallic area at coating-substrate interfaces and effective corrosion suppression.
Forty-day immersion tests in SCP + 0.1 M NaCl solution with daily electrochemical impedance spectroscopy (EIS) monitoring were conducted to assess the long-term corrosion resistance of sol-gel conversion coatings [60,61]. Figure 5e depicts the daily evolution of low-frequency impedance modulus |Z|0.01Hz for all specimens during the 40-day immersion in SCP + 0.1 M NaCl. Figure 5f compares the day 40 |Z|0.01Hz values. Critical observations are as follows: the Undoped sol-gel coating exhibited rapid degradation, reaching the lowest impedance among all sol-gel film samples. This confirms its poor long-term corrosion resistance and early failure in chloride environments. In contrast, the 0.8LNS coating maintained consistently high impedance throughout immersion. At day 40, it retained a |Z|0.01Hz value 6.0-fold higher than the Blank HRB400 steel (Figure 5f), demonstrating superior durability.

3.3.2. Sol-Gel Coatings with 2-Methylimidazole Incorporation at Optimal La3+/Ni2+ Ratio

To enhance coating performance, sol-gel formulations incorporating 0.5–2 wt% 2-methylimidazole (relative to GPTMS mass) were applied to steel specimens at the optimal La3+:Ni2+ ratio of 0.8 (Figure 6). Following immersion in SCP + 0.1 M NaCl with daily EIS monitoring (Figure 6e, Table 4), the 1 wt% 2-MI—0.8LNS coating demonstrated significantly superior corrosion resistance, exhibiting dominant |Z|0.01Hz values beyond day 8. At day 40, it maintained 10.4-fold higher impedance than Blank HRB400 steel (Figure 6f).
Parameter analysis confirms the 1 wt% 2-MI—0.8LNS coating achieves optimal performance, exhibiting a maximized CPE exponent n1 = 0.93 (Table 4), surpassing the 0.8LNS coating (n1 = 0.89). This demonstrates superior defect remediation within the oxide scale, consistent with cross-sectional SEM observations (Figure 2). Concurrently, the oxide scale resistance (Ros) reaches 31.10 kΩ·cm2—a 70.68-fold increase over Blank HRB400 steel (0.44 kΩ·cm2)—validating exceptional electrochemical barrier integrity. Therefore, incorporating 1 wt% 2-Methylimidazole at the optimal La3+/Ni2+ ratio enhances the long-term corrosion resistance of sol-gel conversion coatings.
Electrochemical impedance analysis during 40-day immersion in SCP + 0.1 M NaCl, using equivalent circuits in Figure 5(g1) (Blank HRB400 steel) and Figure 5(g2) (1 wt% 2-MI—0.8LNS coating), reveals distinct corrosion resistance behaviors (Figure 7, Table 5). Both samples demonstrated an initial increase followed by a progressive decline in low-frequency impedance modulus |Z|0.01Hz. Critically, the Blank HRB400 steel exhibited rapid degradation (Figure 7b), while the 1 wt% 2-MI—0.8LNS coating maintained a gradual decline (Figure 7e), indicating superior barrier stability. At day 40, the 1 wt% 2-MI—0.8LNS coating demonstrated exceptional corrosion resistance, exhibiting a low-frequency impedance modulus of 25,600 Ω·cm2–10.4-fold higher than Blank HRB400 steel (2470 Ω·cm2). Critical electrochemical parameters confirm sustained barrier integrity: the coated sample maintained an oxide scale resistance of 17,200 Ω·cm2 (exceeding the blank sample’s 170·Ω cm2 by two orders of magnitude) and a charge transfer resistance of 2.41 × 105 Ω·cm2 (18.5-fold greater than the blank’s 0.13 × 105 Ω·cm2). These results validate the coating’s outstanding long-term protective capability against chloride ingress.

3.3.3. Potentiodynamic Polarization Analysis

Potentiodynamic polarization curves measured in SCP + 0.1 M NaCl (Figure 8a,b, Table 6 and Table 7) demonstrate that sol-gel coatings significantly enhance corrosion resistance. La3+/Ni2+-modified coatings exhibit positive shifts in corrosion potential (Ecorr) and reduced corrosion current density (icorr) compared to Blank HRB400 steel and undoped sol-gel. This improvement is attributed to La3+/Ni2+ forming dense passive films that block oxidant permeation to suppress anodic dissolution, while simultaneously precipitating as hydroxides/oxides at cathodic sites to cover active areas and inhibit oxygen reduction reactions.
Tafel slope analysis further quantifies these mechanisms: 0.8LNS coating demonstrates dominant anodic inhibition with a high βa value (196.12 mV/dec vs. 69.08 mV/dec for bare steel), reflecting La3+-induced passivation impeding metal dissolution. Its βc value (151.06 mV/dec) confirms cathodic suppression via hydroxide/oxide precipitation at cathodic sites. A total of 1 wt% 2-MI—0.8LNS achieves optimal cathodic control (βc = 200.94 mV/dec), attributed to enhanced coverage by La2O3/La(OH)3-based compounds and physical obstruction of oxygen reduction pathways by 2-MI coordination complexes. Its balanced βa (105.39 mV/dec) maintains anodic blocking capability.
Data in Table 6 and Table 7 confirm the 0.8 La3+/Ni2+ ratio as optimal, exhibiting the highest βa, lower icorr, and higher Ecorr. Incorporating 1 wt% 2-Methylimidazole further enhances performance, yielding the most positive Ecorr and the second-lowest icorr (0.11 µA/cm2), substantially outperforming the 0.8LNS film and achieving optimal cathodic control (βc = 200.94 mV/dec).

3.4. Post-Immersion Morphological and Compositional Analysis

3.4.1. Surface Morphology After 40-Day Immersion

After a 40-day immersion in SCP + 0.1 M NaCl, Blank HRB400 steel exhibited severe oxide scale spallation with extensive cracking and defect exposure (Figure 9a). Elemental mapping revealed calcium salt accumulation within defects, indicating cathodic site activation during corrosion (Figure 9(a1–a3)).
In contrast, the 1 wt% 2-MI—0.8LNS coating developed surface cracks due to alkaline-induced silicic acid dissociation (Figure 9b). Calcium ions permeated the sol-gel layer, forming calcium silicate precipitates that altered the silica network structure, evidenced by homogeneous Ca distribution within the coating (Figure 9(b1–b6)) [65].
Following ultrasonic removal of residual sol-gel, the underlying oxide scale of the coated sample remained intact without spallation or cracks (Figure 9c). Uniform elemental distribution of Si, Ca, La, and Ni throughout the oxide scale (Figure 9(c1–c6)) confirmed enhanced electrochemical homogeneity, eliminating localized cathodic/anodic zones. This contrasts sharply with Blank steel’s heterogeneous Ca accumulation (Figure 9(a3) vs. Figure 9(c3)). Macroscopic comparison (Figure 9(d1,d2)) validated the protective mechanism: Blank steel showed pronounced corrosion traces, while the coated specimen maintained pristine oxide integrity.

3.4.2. Cross-Sectional Morphology After 40-Day Immersion

Cross-sectional analysis (Figure 10) reveals distinct degradation patterns:
Blank HRB400 steel: Severe spallation of oxide scale with extensive cracking/pore formation (Figure 10(a–a3)), consistent with surface deterioration in Figure 9a. Residual oxide scale thickness (measured at thickest preserved regions) averages 2.55 µm (variance 5.465).
The 1 wt% 2-MI—0.8LNS: Sol-gel layer develops through-thickness cracks and delamination (Figure 10b), with residual thickness averaging 22.45 µm (variance 2.177), severely compromising barrier function. Modified oxide scale retains structural integrity with only minor microcracking, significantly outperforming blank samples. Elemental mapping (Figure 10(b3)) confirms calcium enrichment confined exclusively to the sol-gel layer, mirroring Figure 9b surface patterns. Oxide scale thickness averages 7.08 µm (variance 2.492), reflecting sustained barrier functionality despite sol-gel degradation in alkaline chloride environments.

3.4.3. Post-Immersion Surface Chemistry Analysis

XPS characterization after 40-day immersion in SCP + 0.1 M NaCl reveals critical interfacial transformations (Figure 11): Blank HRB400 steel exhibited substantial calcium carbonate deposition (C 1s: 290.17 eV; O 1s: O-Ca at 531.57 eV) from carbonation reactions [26,66]. The Fe 2p spectrum showed metallic iron signatures (706.11 eV) alongside FeOOH (712.16 eV), Fe2O3 (710.91 eV), and FeO (709.46 eV), confirming oxide scale spallation and substrate exposure—consistent with a macroscopic deterioration in Figure 9a [51,52]. Conversely, the 1 wt% 2-MI—0.8LNS coating demonstrated: calcium carbonate deposition within the sol-gel matrix, but no metallic Fe detection—validating intact oxide scale integrity (Figure 9c). Silica network preservation (Si 2p: 104.78 eV), indicating calcium silicate formation from alkaline hydrolysis [53]. Dominant La2O3 is present with minor NiO (Figure 11(b6)), mirroring pre-immersion elemental distribution patterns (Figure 4(c3)). These findings confirm the coating’s dual functionality: chemical stabilization of the oxide scale and barrier retention against chloride penetration [57,58,59].

3.5. Long-Term Corrosion Resistance in Mortar Systems

To validate practical performance, 1 wt% 2-MI—0.8LNS-coated steel was embedded in mortar blocks (cured at 20 ± 2 °C, 95 ± 2% RH for 28 days) and subjected to simulated seawater immersion (3.50 wt% NaCl) with 120-day EIS monitoring (Figure 12a–g). Equivalent circuit modeling (Figure 12h) identified three time constants: high-frequency: mortar matrix properties (RmQm); mid-frequency: oxide scale/coating behavior (RfQf); and low-frequency: charge transfer process (RctQdl) [67,68,69,70,71,72]. Critical performance metrics at Day 120 are as follows: |Z|0.01Hz: 74.63 kΩ·cm2 (1 wt% 2-MI—0.8LNS) vs. 9.29 kΩ·cm2 (Blank HRB400 steel), indicating that the impedance of the coated system is 8.0 times that of the blank steel.. This confirms the coated system’s exceptional barrier retention against chloride penetration in real-service conditions.

3.6. Corrosion Resistance Comparative Study of Oxide-Scaled Steel with Engineered Coatings

Compared to prior studies (Table 8), this work demonstrates distinct advantages:
Implementation of a low-temperature curing process, reducing energy consumption and enabling industrial-scale production; superior corrosion resistance in simulated concrete pore solution despite extended immersion; and development of a novel waste heat utilization strategy.

4. Discussion

4.1. Multifunctional Hierarchical Defense for Oxide-Scaled Steel

This study establishes a multifunctional defense system through synergistic integration of La3+, Ni2+, and 2-Methylimidazole within sol-gel matrices, achieving hierarchical corrosion protection on oxide-scaled steel via two core mechanisms (Figure 13):
1.
Sol-Gel Network Optimization
Chemical Design: La3+ acts as a Lewis acid catalyst (accelerating TEOS hydrolysis by polarizing Si–O–C2H5 bonds [73]), while Ni2+ stabilizes intermediate silanol groups (Si–OH). 2-Methylimidazole promotes epoxy ring-opening in GPTMS.
Microstructural Outcome: sol-gel films form hybrid Si–O–Si/C–O–C networks (FTIR, Figure 3). Coating defects are minimized by 1 wt% 2-Methylimidazole and 60 °C low-temperature curing (Figure 1 and Figure 2). 2-Methylimidazole chelation ensures uniform metal ion distribution [74].
Protective Performance: the sol-gel has enhanced barrier integrity through increased crosslinking density and defect reduction, improving physical barrier properties.
2.
Interfacial Bonding and Defect Remediation
Chemical Design: pre-oxidation at 400 °C generates reactive Fe-OH groups, forming covalent Si-O-Fe bonds with sol-gel silanol. At pH 10, La3+/Ni2+ form colloidal La(OH)3/Ni(OH)2 via sol-gel homogeneity [26]. Upon steel immersion, surface heat promotes La2O3/NiO formation while simultaneously ammonia evaporation lowers pH, triggering SiO2-Fe2O3reactions for defect transport. Additionally, Ni2+ reacts with hydroxylated iron oxides to form NiFe2O4 spinel (Figure 13b).
Microstructural Outcome: Defect Remediation and Sealing: adsorbed colloids dehydrate into La2O3(dominant)/NiO sealing layers at 400 °C at the interface, which deposit onto the oxide scale surface to remediate defects (Figure 13b). Validation is demonstrated with FTIR oxide peaks (Figure 3), XPS La2O3/NiO detection (Figure 4 and Figure 11), and surface/cross-section SEM (Figure 1, Figure 2, Figure 9 and Figure 10). Bonding networks correspond to Si-O-Fe covalent bonds (enhancing the coating/scale interfacial bonding strength) (XPS O 1s: 532.55 eV, Figure 4) and 2-Methylimidazole-coordinated Fe(MeIm)2/Fe(MeIm)3 protective complexes which adsorb on the steel surface, improving chloride resistance [75].
Protective Performance: Compared to Blank HRB400 steel, the 1wt%2-MI—0.8LNS film exhibits no spallation/cracking after 40-day immersion (Figure 9 and Figure 10), with persistent La2O3 signatures and absence of metallic Fe signals in XPS (Figure 11), confirming structural integrity and defect remediation efficacy. The impedance modulus |Z|0.01Hz = 25.6 kΩ·cm2 (10.4 × untreated) (Figure 5, Figure 6 and Figure 7).

4.2. Enhanced Corrosion Resistance in Concrete Structures

The 1 wt%-2MI—0.8LNS sol-gel-coated steel embedded in mortar blocks demonstrates exceptional corrosion resistance in simulated marine environments. After 120-day immersion in 3.50 wt% NaCl solution, the coated system maintains a low-frequency impedance modulus of 74.63 kΩ·cm2—8.03-fold higher than the uncoated control (9.29 kΩ·cm2). This performance aligns with our established corrosion inhibition mechanisms [26]:
C–S–H gel formation: sol-gel SiO2 reacts with cement-derived Ca(OH)2 to generate pore-filling calcium silicate hydrate (C–S–H), reducing chloride diffusion paths.
Pore refinement: silica particles and silanes physically seal micro-pores and capillaries within the mortar matrix, enhancing chloride barrier properties.
Stainless steel-like passivation: lanthanum incorporation promotes a chromium-equivalent passive film on the oxide scale [65,76,77], significantly improving anodic stability.

5. Conclusions

This study establishes a dual-mechanism strategy for marine concrete protection: catalytic crosslinking: 2-Methylimidazole-driven epoxy ring-opening constructs defect-minimized Si–O–Si/C–O–C networks via low-temperature curing (60 °C); and thermally activated repair: 400 °C steel interfaces enable pH-modulated La2O3/NiO co-deposition within oxide scale defects.
Quantitative outcomes: SCP + 0.1 M NaCl (40-day): |Z|0.01Hz = 25.6 kΩ·cm2 (10.4× untreated), and oxide scale integrity is preserved; and mortar in 3.5 wt% NaCl (120-day): |Z|0.01Hz = 74.63 kΩ·cm2 (8.03× control).
This design integrates low-temperature curing (60 °C vs. conventional 130 °C) with industrial waste heat utilization (400 °C), significantly reducing energy consumption during curing while efficiently harnessing medium-grade waste heat in industrial practice. The corrosion resistance of steel reinforcement is markedly enhanced, providing long-term protection for marine infrastructure. However, practical engineering applications still face the following challenges:
Scalability: requires synchronization with steel rolling line cooling rates to enable continuous production, alongside investigation into mortar-rebar interfacial behavior;
Long-term durability: necessitates > 24-month field exposure trials to validate performance degradation under tidal cycling and biofouling;
Cost-effectiveness: demands economic viability optimization through recycling process improvements.

Author Contributions

Conceptualization, Y.X. and X.W.; methodology, Y.Z.; software, Y.B.; validation, Y.X., X.W., and Y.B.; formal analysis, Y.X. and Y.Z.; investigation, X.W.; resources, Y.B.; data curation, Y.X.; writing—original draft preparation, Y.X. and Y.Z.; writing—review and editing, G.M.; visualization, G.M.; supervision, G.M.; project administration, G.M.; funding acquisition, G.M. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support from the National Natural Science Foundation of China (Nos. 52171093, 52371071), the National Key R&D Program of China (No. 2019YFE0111000).

Data Availability Statement

Data can be made available upon request.

Acknowledgments

The authors acknowledge the support from the mentioned funding sources and R&D initiatives.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Nóvoa, X.R. Electrochemical aspects of the steel-concrete system. A review. J. Solid State Electrochem. 2016, 20, 2113–2125. [Google Scholar] [CrossRef]
  2. Peng, G.; Yao, G.; Jia, H.; Zheng, S.; Yao, Y. Lifecycle Assessment of Seismic Resilience and Economic Losses for Continuous Girder Bridges in Chloride-Induced Corrosion. Buildings 2025, 15, 3315. [Google Scholar] [CrossRef]
  3. Chen, F.; Hua, L.; Zhang, J. The Deterioration of Low-Cycle Fatigue Properties and the Fatigue Life Model of Reinforcing Steel Bars Subjected to Corrosion. Buildings 2025, 15, 3313. [Google Scholar] [CrossRef]
  4. Wang, X.; Huang, P.; Yuan, Y.; Wang, D.; Yang, Y.; Liu, X. Chloride Diffusion and Corrosion Assessment in Cracked Marine Concrete Bridges Using Extracted Crack Morphologies. Buildings 2025, 15, 3214. [Google Scholar] [CrossRef]
  5. Cui, L.; Xiang, T.; Hu, B.; Lv, Y.; Rong, H.; Liu, D.E.; Zhang, S.; Guo, M.; Lv, Z.; Chen, D. Design of monolithic superhydrophobic concrete with excellent anti-corrosion and self-cleaning properties. Colloids Surf. A Physicochem. Eng. Asp. 2024, 685, 133345. [Google Scholar] [CrossRef]
  6. Wang, W.; Yu, H.; Yang, Z.; Wang, W.; Zhang, Q.; Jiang, Y.; Wang, P.; Marano, G.C. Photocathodic protection performance of CuInSe2/TiO2 n-n heterojunction films for HRB400 steel rebar under visible light. Colloids Surf. A Physicochem. Eng. Asp. 2025, 715, 136638. [Google Scholar] [CrossRef]
  7. Rodrigues, R.; Gaboreau, S.; Gance, J.; Ignatiadis, I.; Betelu, S. Reinforced concrete structures: A review of corrosion mechanisms and advances in electrical methods for corrosion monitoring. Constr. Build. Mater. 2021, 269, 121240. [Google Scholar] [CrossRef]
  8. Ding, L.; Dai, X.; Gan, Y.; Zeng, Y. Impact of Reinforcement Corrosion on Progressive Collapse Behavior of Multi-Story RC Frames. Buildings 2025, 15, 2534. [Google Scholar] [CrossRef]
  9. Caixinhas, A.; Tomé, J.; Melo, J.; Marreneca, G.; Furtado, A. Impact of Corrosion on the Behaviour of Reinforced Concrete Buildings. Buildings 2025, 15, 1267. [Google Scholar] [CrossRef]
  10. Gardner, L.; Bu, Y.; Francis, P.; Baddoo, N.R.; Cashell, K.A.; McCann, F. Elevated temperature material properties of stainless steel reinforcing bar. Constr. Build. Mater. 2016, 114, 977–997. [Google Scholar] [CrossRef]
  11. Rabi, M.; Shamass, R.; Cashell, K.A. Structural performance of stainless steel reinforced concrete members: A review. Constr. Build. Mater. 2022, 325, 126673. [Google Scholar] [CrossRef]
  12. Keßler, S.; Angst, U.; Zintel, M.; Gehlen, C. Defects in epoxy-coated reinforcement and their impact on the service life of a concrete structure. Struct. Concr. 2015, 16, 398–405. [Google Scholar] [CrossRef]
  13. Fares, S.; Meriggi, P.; De Santis, S.; de Felice, G. Experimental Investigation on the Tensile and Bond Durability of Galvanized Steel Reinforced Grout. Buildings 2025, 15, 3020. [Google Scholar] [CrossRef]
  14. Dong, Z.; Xu, K.; Chen, X.; Mao, Y.; Fu, C.; Zhang, Z.; Kessler, S. Unmasking the degradation of epoxy coating on the surface of steel subjected to stress in simulated concrete pore solution. Constr. Build. Mater. 2024, 452, 138931. [Google Scholar] [CrossRef]
  15. Xing, Y.; Zhang, Z.; You, Q.; Hu, J. Modified Aggregates for Mitigating Anodic Acidification in Impressed Current Cathodic Protection Systems Toward Infrastructure Modernization. Buildings 2025, 15, 1891. [Google Scholar] [CrossRef]
  16. Min, S.-H.; Lee, H.-S.; Singh, J.K. Effects of Different Inhibitors on the Corrosion Mitigation of Steel Rebar Immersed in NaCl-Contaminated Concrete Pore Solution. Buildings 2024, 14, 3559. [Google Scholar] [CrossRef]
  17. Tran, D.T.; Lee, H.-S.; Singh, J.K.; Yang, H.-M.; Jeong, M.-G.; Yan, S.; Ibrahim, I.S.; Ariffin, M.A.B.M.; Le, A.-T.; Singh, A.K. Effects of Hybrid Corrosion Inhibitor on Mechanical Characteristics, Corrosion Behavior, and Predictive Estimation of Lifespan of Reinforced Concrete Structures. Buildings 2025, 15, 1114. [Google Scholar] [CrossRef]
  18. Li, J.; Li, T.; Zeng, Y.; Chen, C.; Guo, H.; Lei, B.; Zhang, P.; Feng, Z.; Meng, G. A novel sol-gel coating via catechol/lysine polymerization for long-lasting corrosion protection of Mg alloy AZ31. Colloids Surf. A Physicochem. Eng. Asp. 2023, 656, 130361. [Google Scholar] [CrossRef]
  19. Figueira, R.B. Hybrid Sol–gel Coatings for Corrosion Mitigation: A Critical Review. Polymers 2020, 12, 689. [Google Scholar] [CrossRef] [PubMed]
  20. Li, J.; Li, S.; Chen, C.; Guo, H.; Lei, B.; Zhang, P.; Meng, G.; Feng, Z. Dopamine self-polymerized sol-gel coating for corrosion protection of AZ31 Mg Alloy. Colloids Surf. A Physicochem. Eng. Asp. 2023, 666, 131283. [Google Scholar] [CrossRef]
  21. Claire, L.; Marie, G.; Julien, G.; Jean-Michel, S.; Jean, R.; Marie-Joëlle, M.; Stefano, R.; Michele, F. New architectured hybrid sol-gel coatings for wear and corrosion protection of low-carbon steel. Prog. Org. Coat. 2016, 99, 337–345. [Google Scholar] [CrossRef]
  22. Chen, X.; Gao, Y.; Zhang, Y.; Hu, M.; Geng, Y.; Li, S.; Sui, S.; Liang, G. Properties evaluation of double silane system compound gel as a protective coating on concrete. J. Coat. Technol. Res. 2023, 21, 329–340. [Google Scholar] [CrossRef]
  23. Criado, M.; Sobrados, I.; Bastidas, J.M.; Sanz, J. Corrosion behaviour of coated steel rebars in carbonated and chloride-contaminated alkali-activated fly ash mortar. Prog. Org. Coat. 2016, 99, 11–22. [Google Scholar] [CrossRef]
  24. Zeng, Y.; Xu, P.; Liu, G.; Wang, T.; Lei, B.; Feng, Z.; Zhang, P.; Meng, G. ZIF-8 Modified Ce–Sol–gel Film on Rebar for Enhancing Corrosion Resistance. Acta Metall. Sin. (Engl. Lett.) 2024, 37, 2121–2135. [Google Scholar] [CrossRef]
  25. Zeng, Y.; Yang, L.; Liu, G.; Guo, Y.; Lei, B.; Feng, Z.; Guo, H.; Zhang, P.; Meng, G. Construction of Ce3+ modified conversion film on oxide scale of plain steel rebars to greatly enhance the corrosion resistance. J. Mater. Res. Technol. 2023, 27, 4403–4415. [Google Scholar] [CrossRef]
  26. Zeng, Y.; Xu, P.; Wang, T.; Xie, Y.; Liu, G.; Qian, H.; Feng, Z.; Lei, B.; Zhang, P.; Meng, G. Repairment of carbon steel oxide scale using Ce3+ and Ni2+ doped sol-gel technique. npj Mater. Degrad. 2024, 8, 99. [Google Scholar] [CrossRef]
  27. Cheng, X.; He, Y.; Song, R.; Li, H.; Liu, B.; Zhou, H.; Yan, L. Study of mechanical character and corrosion properties of La2O3 nanoparticle reinforced Ni-W composite coatings. Colloids Surf. A Physicochem. Eng. Asp. 2022, 652, 129799. [Google Scholar] [CrossRef]
  28. Koutsonikolas, D.; Kaldis, S.; Sakellaropoulos, G.P. A low-temperature CVI method for pore modification of sol–gel silica membranes. J. Membr. Sci. 2009, 342, 131–137. [Google Scholar] [CrossRef]
  29. Maia, F.; Yasakau, K.A.; Carneiro, J.; Kallip, S.; Tedim, J.; Henriques, T.; Cabral, A.; Venâncio, J.; Zheludkevich, M.L.; Ferreira, M.G.S. Corrosion protection of AA2024 by sol–gel coatings modified with MBT-loaded polyurea microcapsules. Chem. Eng. J. 2016, 283, 1108–1117. [Google Scholar] [CrossRef]
  30. Chen, Q.-A.; Liu, L.-Z.; Liu, X.; Liu, X.-H. Improvement to anti-rust property of hot rolled rebar by compact oxide scale. J. Cent. South Univ. 2017, 24, 2813–2818. [Google Scholar] [CrossRef]
  31. Zhou, Q.; Yin, Y.; Liu, Z.; Zhang, J. Optimising the corrosion performance of hot-rolled steel bar in concrete. Corros. Eng. Sci. Technol. 2022, 57, 568–579. [Google Scholar] [CrossRef]
  32. Chen, C.; Yu, M.; Zhan, Z.; Ge, Y.; Sun, Z.; Liu, J. Dual strategy of improved crosslinking degree and incorporated inhibitory fillers for high corrosion resistance sol-gel coatings. Corros. Sci. 2024, 234, 112145. [Google Scholar] [CrossRef]
  33. Goyal, S.; Jacob, J. 2,2′-Bipyridine containing chelating polymers for sequestration of heavy metal ions from organic solvents. J. Appl. Polym. Sci. 2022, 139, 52121. [Google Scholar] [CrossRef]
  34. Pang, K.; Xue, H.; Liu, H.; Sun, J.; Liu, T. Chelating Metal Ions in a Metal-Organic Framework for Constructing a Biomimetic Catalyst Through Post-modification. Chem. Res. Chin. Univ. 2022, 38, 1542–1546. [Google Scholar] [CrossRef]
  35. Zhang, Y.-L.; Chang, R.; Duan, H.-Z.; Chen, Y.-X. Metal ion and light sequentially induced sol–gel–sol transition of a responsive peptide-hydrogel. Soft Matter 2020, 16, 7652–7658. [Google Scholar] [CrossRef]
  36. Gutierres, A.; Pascual, S.; Fontaine, L.; Piogé, S.; Benyahia, L. The effect of metal ions on the viscoelastic properties of thermosensitive sol-to-gel reversible metallo-supramolecular hydrogels. Polym. Chem. 2018, 9, 2494–2504. [Google Scholar] [CrossRef]
  37. Zhu, X.; Li, S.; Ao, Q.; Yang, X. Deterioration and film forming abilities of the slurries used for preparing La0.8Sr0.2MnO3 films through composite sol–gel method. J. Alloys Compd. 2011, 509, 7093–7099. [Google Scholar] [CrossRef]
  38. Zhu, X.; Yin, X.; Zhang, Y.; Mei, Q.; Bai, J.; Yang, G.; Zhang, D.; Liang, X. Optimization via response surface methodology for Ni-La-V2O5 porous ion-storage film by sol–gel method. J. Sol-Gel Sci. Technol. 2025, 115, 84–97. [Google Scholar] [CrossRef]
  39. Chen, M.; Gai, J.-X.; Chen, C.-Y.; Li, J.-Q. Corrosion evolution of 15CrMn steel in sulfur-containing sodium aluminate solution. Mater. Lett. 2022, 310, 131464. [Google Scholar] [CrossRef]
  40. Both, J.; Szabó, G.; Katona, G.; Muresan, L.M. Tannic acid reinforced sol-gel silica coatings for corrosion protection of zinc substrates. Mater. Chem. Phys. 2022, 282, 125912. [Google Scholar] [CrossRef]
  41. Hamzah, E.; Hussain, M.F.; Ibrahim, Z.; Abdolahi, A. Corrosion Behaviour of Carbon Steel in Sea Water Medium in Presence of P. aeruginosa Bacteria. Arab. J. Sci. Eng. 2014, 39, 6863–6870. [Google Scholar] [CrossRef]
  42. Dehghani, A.; Sanaei, Z.; Fedel, M.; Ramezanzadeh, M.; Mahdavian, M.; Ramezanzade, B. Fabrication of an intelligent anti-corrosion silane film using a MoO42− loaded Micro/mesoporous ZIF67-MOF/multi-walled-CNT/APTES core-shell nano-container. Colloids Surf. A Physicochem. Eng. Asp. 2023, 656, 130511. [Google Scholar] [CrossRef]
  43. Mrad, M.; Dhouibi, L.; Montemor, M.F. Elaboration of γ-glycidoxypropyltrimethoxysilane coating on AA2024-T3 aluminum alloy: Influence of synthesis route on physicochemical and anticorrosion properties. Prog. Org. Coat. 2018, 121, 1–12. [Google Scholar] [CrossRef]
  44. Mrad, M.; Ben Amor, Y.; Dhouibi, L.; Montemor, M.F. Corrosion prevention of AA2024-T3 aluminum alloy with a polyaniline/poly(γ-glycidoxypropyltrimethoxysilane) bi-layer coating: Comparative study with polyaniline mono-layer feature. Surf. Coat. Technol. 2018, 337, 1–11. [Google Scholar] [CrossRef]
  45. Mohd Yusop, H.; Mohd Ismail, A.I.H.; Wan Ismail, W.N. Preparation and Characterization of New Sol–Gel Hybrid Inulin–TEOS Adsorbent. Polymers 2021, 13, 1295. [Google Scholar] [CrossRef]
  46. Jaramillo, A.F.; Baez-Cruz, R.; Montoya, L.F.; Medinam, C.; Pérez-Tijerina, E.; Salazar, F.; Rojas, D.; Melendrez, M.F. Estimation of the surface interaction mechanism of ZnO nanoparticles modified with organosilane groups by Raman Spectroscopy. Ceram. Int. 2017, 43, 11838–11847. [Google Scholar] [CrossRef]
  47. Zandi-zand, R.; Ershad-langroudi, A.; Rahimi, A. Silica based organic–inorganic hybrid nanocomposite coatings for corrosion protection. Prog. Org. Coat. 2005, 53, 286–291. [Google Scholar] [CrossRef]
  48. Ramesh, T.N. Isotopic effect on the hydration-dehydration studies of βbc-nickel hydroxide using infrared spectroscopy. Res. J. Chem. Environ. 2022, 26, 130–134. [Google Scholar]
  49. Salavati-Niasari, M.; Hosseinzadeh, G.; Davar, F. Synthesis of lanthanum carbonate nanoparticles via sonochemical method for preparation of lanthanum hydroxide and lanthanum oxide nanoparticles. J. Alloys Compd. 2011, 509, 134–140. [Google Scholar] [CrossRef]
  50. Romano, A.-P.; Fedel, M.; Deflorian, F.; Olivier, M.-G. Silane sol–gel film as pretreatment for improvement of barrier properties and filiform corrosion resistance of 6016 aluminium alloy covered by cataphoretic coating. Prog. Org. Coat. 2011, 72, 695–702. [Google Scholar] [CrossRef]
  51. Shang, B.; Ma, Y.; Meng, M.; Li, Y. The roles carbonate and bicarbonate ions play on pre-passive HRB400 rebars in simulated pore solutions of deteriorating concrete. Mater. Corros. 2018, 69, 1800–1810. [Google Scholar] [CrossRef]
  52. Yuan, X.; Wang, X.; Cao, Y.; Yang, H. Natural passivation behavior and its influence on chloride-induced corrosion resistance of stainless steel in simulated concrete pore solution. J. Mater. Res. Technol. 2020, 9, 12378–12390. [Google Scholar] [CrossRef]
  53. Sun, J.; Tang, Z.; Meng, G.; Gu, L. Silane functionalized plasma-treated boron nitride nanosheets for anticorrosive reinforcement of waterborne epoxy coatings. Prog. Org. Coat. 2022, 167, 106831. [Google Scholar] [CrossRef]
  54. He, S.; Sun, G.; Cheng, X.; Dai, H.; Chen, X. Nanoporous SiO2 grafted aramid fibers with low thermal conductivity. Compos. Sci. Technol. 2017, 146, 91–98. [Google Scholar] [CrossRef]
  55. Li, T.; Li, L.; Qi, J.; Chen, F. Corrosion protection of Ti6Al4V by a composite coating with a plasma electrolytic oxidation layer and sol-gel layer filled with graphene oxide. Prog. Org. Coat. 2020, 144, 105632. [Google Scholar] [CrossRef]
  56. Takiguchi, K.; Miura, K.; Ohtsu, N. XPS study of a laser-nitrided iron surface using a focused pulsed Nd:YAG laser under various conditions. Surf. Interface Anal. 2020, 52, 919–923. [Google Scholar] [CrossRef]
  57. Qiang, X.; Yao, Y.; Yin, J.; Da, P.; Mu, Z.; Shen, K.; Sun, Y.; Zhang, Y.; Li, P.; Li, Z.; et al. Activating La−O−Ni Bridge in Ordered Macroporous Interface for Electrochemical Urea Wastewater Purification. Angew. Chem. Int. Ed. 2025, 64, e202424014. [Google Scholar] [CrossRef] [PubMed]
  58. Huang, Z.; Zhou, W.; Hu, M.; Zhang, M.; Zhao, X.; Li, Y.; Hao, X.; Li, D.; Xu, J. Ultralong lifespan and high energy density soft-pack asymmetric supercapacitors based on electrochemically activated porous nickel oxide nanosheet array. Chem. Eng. J. 2024, 494, 152907. [Google Scholar] [CrossRef]
  59. Zhang, J.; Chen, J.; Luo, Y.; Chen, Y.; Zhang, C.; Luo, Y.; Xue, Y.; Liu, H.; Wang, G.; Wang, R. Engineering heterointerfaces coupled with oxygen vacancies in lanthanum–based hollow microspheres for synergistically enhanced oxygen electrocatalysis. J. Energy Chem. 2021, 60, 503–511. [Google Scholar] [CrossRef]
  60. An, H.; Xu, Z.; Meng, G.; Li, W.; Wang, Y.; Liu, B.; Wang, J.; Wang, F. High corrosion resistance film on rebar by cerium modification. J. Mater. Sci. Technol. 2021, 64, 73–84. [Google Scholar] [CrossRef]
  61. Ouassir, J.; Bennis, H.; Benqlilou, H.; Galai, M.; Hassani, Y.; Touhami, M.E.; Berrami, K.; Lemyasser, M. EIS study on erosion–corrosion behavior of BA35 and BA22 brasses in drinking water at various impingement angles. Colloids Surf. A Physicochem. Eng. Asp. 2020, 586, 124151. [Google Scholar] [CrossRef]
  62. Shi, X.; Nguyen, T.A.; Suo, Z.; Liu, Y.; Avci, R. Effect of nanoparticles on the anticorrosion and mechanical properties of epoxy coating. Surf. Coat. Technol. 2009, 204, 237–245. [Google Scholar] [CrossRef]
  63. Fan, B.; Zhu, H.; Li, H.; Tian, H.; Yang, B. Penetration of imidazoline derivatives through deposited scale for inhibiting the under-deposit corrosion of pipeline steel. Corros. Sci. 2024, 235, 112209. [Google Scholar] [CrossRef]
  64. Feng, Z. Corrosion Inhibition of AZ31 Mg Alloy by Aqueous Selenite (SeO32−). J. Electrochem. Soc. 2019, 166, C520, Erratum in J. Electrochem. Soc. 2020, 167, 089002. https://doi.org/10.1149/1945-7111/ab8928. [Google Scholar]
  65. Feng, H.; Le, H.T.N.; Wang, S.; Zhang, M.-H. Effects of silanes and silane derivatives on cement hydration and mechanical properties of mortars. Constr. Build. Mater. 2016, 129, 48–60. [Google Scholar] [CrossRef]
  66. Dong, Y.; Gao, Z.; Di, J.; Wang, D.; Yang, Z.; Wang, Y.; Guo, X.; Li, K. Experimental study on solidification and remediation of lead–zinc tailings based on microbially induced calcium carbonate precipitation (MICP). Constr. Build. Mater. 2023, 369, 130611. [Google Scholar] [CrossRef]
  67. Kim, C.; Goldsberry, R.; Karayan, A.I.; Milla, J.; Goehring, L.; Hassan, M.; Castaneda, H. Electrochemical evaluation of epoxy-coated-rebar containing pH-responsive nanocapsules in simulated carbonated concrete pore solution. Prog. Org. Coat. 2021, 161, 106549. [Google Scholar] [CrossRef]
  68. Shi, J.; Wu, M.; Ming, J. Long-term corrosion resistance of reinforcing steel in alkali-activated slag mortar after exposure to marine environments. Corros. Sci. 2021, 179, 109175. [Google Scholar] [CrossRef]
  69. Wang, Y.; Chen, R.; Hu, J.; Zhang, Z.; Huang, H.; Ma, Y.; Wei, J.; Zhang, Z.; Yin, S.; Wang, H.; et al. Surface characteristics and electrochemical behaviors of passive reinforcing steel in alkali-activated slag. Corros. Sci. 2021, 190, 109657. [Google Scholar] [CrossRef]
  70. Fan, Q.; Wang, Z.; Meng, X.; Zhang, K.; Ma, G.; Li, Z.; Meng, D. Multi-scale analysis of the strengthening mechanism of functionalized graphene as reinforcement in cement composites. Colloids Surf. A Physicochem. Eng. Asp. 2022, 651, 129729. [Google Scholar] [CrossRef]
  71. Li, G.; Wang, Z.; Leung, C.K.Y.; Tang, S.; Pan, J.; Huang, W.; Chen, E. Properties of rubberized concrete modified by using silane coupling agent and carboxylated SBR. J. Clean. Prod. 2016, 112, 797–807. [Google Scholar] [CrossRef]
  72. Nair, P.A.K.; Vasconcelos, W.L.; Paine, K.; Calabria-Holley, J. A review on applications of sol-gel science in cement. Constr. Build. Mater. 2021, 291, 123065. [Google Scholar] [CrossRef]
  73. Zhang, X.; Leddy, J.; Bard, A.J. Dependence of Rate Constants of Heterogeneous Electron Transfer Reactions on Viscosity. J. Am. Chem. Soc. 1985, 107, 3719–3721. [Google Scholar] [CrossRef]
  74. Zhang, X.; Yang, H.; Bard, A.J. Variation of the Heterogeneous Electron Transfer Rate Constant with Solution Viscosity: Reduction of Aqueous Solutions of [(EDTA) chromium (III)]-at a Mercury Electrode. J. Am. Chem. Soc. 1987, 109, 1916–1920. [Google Scholar] [CrossRef]
  75. Wang, Y.; Hu, J.; Ma, Y.; Zhang, Z.; Huang, H.; Wei, J.; Yin, S.; Yu, Q. A novel high-efficient MOFs-based corrosion inhibitor for the reinforcing steel in cement extract. Constr. Build. Mater. 2022, 317, 125946. [Google Scholar] [CrossRef]
  76. Rolandi, A.C.; Montes, F.; Morán Ayala, L.; Frontini, A.; Vázquez, M.; Valcarce, M.B. Pre-treatments in silicate solutions to mitigate reinforcing steel corrosion in chloride-contaminated environments. Corros. Eng. Sci. Technol. 2022, 57, 455–463. [Google Scholar] [CrossRef]
  77. Ming, J.; Zhou, X.; Zuo, H.; Jiang, L.; Zou, Y.; Shi, J. Effects of stray current and silicate ions on electrochemical behavior of a high-strength prestressing steel in simulated concrete pore solutions. Corros. Sci. 2022, 197, 110083. [Google Scholar] [CrossRef]
Figure 1. SEM images of HRB400 steel and HRB400 steel samples immersed in sol-gel solutions (pH = 10) with different metal ions: (a) Blank HRB400 steel; (b) Undoped sol-gel; (c) LS film; (d) NS film; (e) 0.8LNS film; (f) 0.2LNS film; (g) 0.05LNS film; (h) 0.5 wt% 2-MI—0.8LNS film; (i) 1 wt% 2-MI—0.8LNS film; (j) 2wt% 2-MI—0.8LNS film.
Figure 1. SEM images of HRB400 steel and HRB400 steel samples immersed in sol-gel solutions (pH = 10) with different metal ions: (a) Blank HRB400 steel; (b) Undoped sol-gel; (c) LS film; (d) NS film; (e) 0.8LNS film; (f) 0.2LNS film; (g) 0.05LNS film; (h) 0.5 wt% 2-MI—0.8LNS film; (i) 1 wt% 2-MI—0.8LNS film; (j) 2wt% 2-MI—0.8LNS film.
Buildings 15 04272 g001
Figure 2. Cross-sectional BSE images and EDS: (a,a1a3) Blank sample and distribution of O, Fe, C elements; (b,b1b3,c,c1c3) cross-sectional BSE images and distribution of O, Fe, Si elements for samples modified with 0.8LNS film and 1wt% 2-MI—0.8LNS film, respectively. The added red dashed line demarcates the interface boundary.
Figure 2. Cross-sectional BSE images and EDS: (a,a1a3) Blank sample and distribution of O, Fe, C elements; (b,b1b3,c,c1c3) cross-sectional BSE images and distribution of O, Fe, Si elements for samples modified with 0.8LNS film and 1wt% 2-MI—0.8LNS film, respectively. The added red dashed line demarcates the interface boundary.
Buildings 15 04272 g002
Figure 3. XRD spectra: (a) Blank HRB400, Undoped sol-gel, 0.8LNS film, 1 wt% 2MI—0.8LNS film; FTIR spectra: (b1,b2) Undoped sol-gel, 0.8LNS, 1 wt% 2-MI—0.8LNS solution FTIR spectra and magnified bands at 700 cm−1–400 cm−1; (c) sol-gel precursors GPTMS and TEOS; (d1,d2) FTIR spectra and magnified bands at 700 cm−1–400 cm−1 of Undoped sol-gel film, 0.8LNS film, 1 wt% 2-MI—0.8LNS film generated on high-temperature steel bar surface after immersion.
Figure 3. XRD spectra: (a) Blank HRB400, Undoped sol-gel, 0.8LNS film, 1 wt% 2MI—0.8LNS film; FTIR spectra: (b1,b2) Undoped sol-gel, 0.8LNS, 1 wt% 2-MI—0.8LNS solution FTIR spectra and magnified bands at 700 cm−1–400 cm−1; (c) sol-gel precursors GPTMS and TEOS; (d1,d2) FTIR spectra and magnified bands at 700 cm−1–400 cm−1 of Undoped sol-gel film, 0.8LNS film, 1 wt% 2-MI—0.8LNS film generated on high-temperature steel bar surface after immersion.
Buildings 15 04272 g003
Figure 4. (a1a3) Blank HRB400; (b1b6) 1 wt% 2-MI—0.8LNS film; XPS spectra of C 1s, O 1s, Si 2p, Fe 2p, Ni 2p, La 3d. (c1c3) exposed surface after removing the 1 wt% 2-MI—0.8LNS film from sample.
Figure 4. (a1a3) Blank HRB400; (b1b6) 1 wt% 2-MI—0.8LNS film; XPS spectra of C 1s, O 1s, Si 2p, Fe 2p, Ni 2p, La 3d. (c1c3) exposed surface after removing the 1 wt% 2-MI—0.8LNS film from sample.
Buildings 15 04272 g004
Figure 5. Electrochemical impedance plots of steel bars immersed at 400 °C in sol-gel solutions with pH = 10 and different Cla3+:CNi2+ ratios after treatment in 0.1 M NaCl + SCP solution: (a) Nyquist plots and fitting data; (b,c) Bode plots and fitting data; (d) comparison of fitted Rtot for different samples; (e) variation in low-frequency |Z|0.01Hz modulus for different samples during 40 days of immersion; (f) comparison of low-frequency |Z|0.01Hz modulus for different samples on the 40th day of immersion; equivalent circuit diagrams for different samples: (g1) represents the fitting circuit for unmodified mill-scale steel; (g2) represents the fitting circuit for mill-scale steel with sol-gel film.
Figure 5. Electrochemical impedance plots of steel bars immersed at 400 °C in sol-gel solutions with pH = 10 and different Cla3+:CNi2+ ratios after treatment in 0.1 M NaCl + SCP solution: (a) Nyquist plots and fitting data; (b,c) Bode plots and fitting data; (d) comparison of fitted Rtot for different samples; (e) variation in low-frequency |Z|0.01Hz modulus for different samples during 40 days of immersion; (f) comparison of low-frequency |Z|0.01Hz modulus for different samples on the 40th day of immersion; equivalent circuit diagrams for different samples: (g1) represents the fitting circuit for unmodified mill-scale steel; (g2) represents the fitting circuit for mill-scale steel with sol-gel film.
Buildings 15 04272 g005
Figure 6. Electrochemical impedance plots of steel bars immersed at 400 °C in pH = 10 solutions containing 0.8LNS film with different 2-Methylimidazole contents after treatment in 0.1 M NaCl + SCP solution: (a) Nyquist plots and fitting data; (b,c) Bode plots and fitting data; (d) domparison of fitted Rtot for different samples; (e) variation in low-frequency |Z|0.01Hz modulus for different samples during 40 days of immersion; (f) comparison of low-frequency |Z|0.01Hz modulus for different samples on the 40th day of immersion.
Figure 6. Electrochemical impedance plots of steel bars immersed at 400 °C in pH = 10 solutions containing 0.8LNS film with different 2-Methylimidazole contents after treatment in 0.1 M NaCl + SCP solution: (a) Nyquist plots and fitting data; (b,c) Bode plots and fitting data; (d) domparison of fitted Rtot for different samples; (e) variation in low-frequency |Z|0.01Hz modulus for different samples during 40 days of immersion; (f) comparison of low-frequency |Z|0.01Hz modulus for different samples on the 40th day of immersion.
Buildings 15 04272 g006
Figure 7. (a,d) Nyquist plots and fitting data; (b,e) and (c,f) Bode plots and fitting data; (g) low-frequency |Z|0.01Hz modulus variation during 40 days of immersion for Blank HRB400 steel and 1 wt% 2-MI—0.8LNS film in 0.1 M NaCl + SCP solution; (h) comparison of fitted Ros for both samples; (i) low-frequency Rct modulus variation for both samples on the 40th day of immersion.
Figure 7. (a,d) Nyquist plots and fitting data; (b,e) and (c,f) Bode plots and fitting data; (g) low-frequency |Z|0.01Hz modulus variation during 40 days of immersion for Blank HRB400 steel and 1 wt% 2-MI—0.8LNS film in 0.1 M NaCl + SCP solution; (h) comparison of fitted Ros for both samples; (i) low-frequency Rct modulus variation for both samples on the 40th day of immersion.
Buildings 15 04272 g007
Figure 8. (a) Potentiodynamic polarization curves of unmodified steel with mill-scale and steel with mill-scale coated with sol-gel films prepared at different La3+/Ni2+ concentration ratios in 0.1 M NaCl-containing simulated concrete pore solution; (b) potentiodynamic polarization curves of coated steel bars with varying 2-Methylimidazole content added to the base (La3+/Ni2+ concentration ratio = 0.8) in 0.1 M NaCl-containing simulated concrete pore solution.
Figure 8. (a) Potentiodynamic polarization curves of unmodified steel with mill-scale and steel with mill-scale coated with sol-gel films prepared at different La3+/Ni2+ concentration ratios in 0.1 M NaCl-containing simulated concrete pore solution; (b) potentiodynamic polarization curves of coated steel bars with varying 2-Methylimidazole content added to the base (La3+/Ni2+ concentration ratio = 0.8) in 0.1 M NaCl-containing simulated concrete pore solution.
Buildings 15 04272 g008
Figure 9. (ac) SEM images and corresponding elemental distribution maps of different sample surfaces after 40 days of immersion in 0.1 M NaCl + SCP solution. Specifically: (a,a1a3) Untreated steel bar sample with mill-scale (showing O, Fe, Ca element distribution); (b,b1b6) Sample coated with the 1 wt% 2-MI—0.8LNS film prepared under optimal conditions (showing O, Fe, Ca, Si, Ni, and La element distribution); (c,c1c6) Sample coated with the 1 wt% 2-MI—0.8LNS film after 40 days immersion and subsequent ultrasonic cleaning (showing O, Fe, Ca, Si, Ce, and Ni element distribution). (d1) Blank HRB400 sample (left) and the 1 wt% 2-MI—0.8LNS film coated sample (right) before immersion. (d2) Blank HRB400 sample (left) and the 1 wt% 2-MI—0.8LNS film coated sample (right) after 40 days immersion and subsequent ultrasonic cleaning.
Figure 9. (ac) SEM images and corresponding elemental distribution maps of different sample surfaces after 40 days of immersion in 0.1 M NaCl + SCP solution. Specifically: (a,a1a3) Untreated steel bar sample with mill-scale (showing O, Fe, Ca element distribution); (b,b1b6) Sample coated with the 1 wt% 2-MI—0.8LNS film prepared under optimal conditions (showing O, Fe, Ca, Si, Ni, and La element distribution); (c,c1c6) Sample coated with the 1 wt% 2-MI—0.8LNS film after 40 days immersion and subsequent ultrasonic cleaning (showing O, Fe, Ca, Si, Ce, and Ni element distribution). (d1) Blank HRB400 sample (left) and the 1 wt% 2-MI—0.8LNS film coated sample (right) before immersion. (d2) Blank HRB400 sample (left) and the 1 wt% 2-MI—0.8LNS film coated sample (right) after 40 days immersion and subsequent ultrasonic cleaning.
Buildings 15 04272 g009
Figure 10. SEM images and corresponding elemental distribution maps of the cross-sectional morphology of different samples after 40 days of immersion in 0.1 M NaCl + SCP solution. Specifically: (a,a1a3) Blank HRB400 (untreated steel bar sample with mill-scale) showing O, Fe, Ca element distribution; (b,b1b6) Sample coated with the 1 wt% 2-MI—0.8LNS film prepared under optimal conditions showing O, Fe, Ca, Si, Ni, and La element distribution. The added red dashed line demarcates the interface boundary.
Figure 10. SEM images and corresponding elemental distribution maps of the cross-sectional morphology of different samples after 40 days of immersion in 0.1 M NaCl + SCP solution. Specifically: (a,a1a3) Blank HRB400 (untreated steel bar sample with mill-scale) showing O, Fe, Ca element distribution; (b,b1b6) Sample coated with the 1 wt% 2-MI—0.8LNS film prepared under optimal conditions showing O, Fe, Ca, Si, Ni, and La element distribution. The added red dashed line demarcates the interface boundary.
Buildings 15 04272 g010
Figure 11. XPS and fitting spectra of (a1a3) Blank HRB400 and (b1b6) 1 wt% 2-MI—0.8LNS film after 40 days of immersion: C 1s, O 1s, Si 2p, Fe 2p, Ni 2p, La 3d.
Figure 11. XPS and fitting spectra of (a1a3) Blank HRB400 and (b1b6) 1 wt% 2-MI—0.8LNS film after 40 days of immersion: C 1s, O 1s, Si 2p, Fe 2p, Ni 2p, La 3d.
Buildings 15 04272 g011
Figure 12. (a,d) Nyquist plots and fitting data; (b,e) and (c,f) Bode plots and fitting data; (g) low-frequency |Z|0.01Hz modulus variation in mortar blocks prepared with Blank HRB400 and 1 wt% 2-MI—0.8LNS film immersed in simulated seawater for 120 days; (h) equivalent circuit diagram fitted for mortar blocks.
Figure 12. (a,d) Nyquist plots and fitting data; (b,e) and (c,f) Bode plots and fitting data; (g) low-frequency |Z|0.01Hz modulus variation in mortar blocks prepared with Blank HRB400 and 1 wt% 2-MI—0.8LNS film immersed in simulated seawater for 120 days; (h) equivalent circuit diagram fitted for mortar blocks.
Buildings 15 04272 g012
Figure 13. Corrosion and protection mechanisms of different samples in 0.1 M NaCl + SCP solution: (a) Blank HRB400; (b) 1 wt% 2-MI—0.8LNS film.
Figure 13. Corrosion and protection mechanisms of different samples in 0.1 M NaCl + SCP solution: (a) Blank HRB400; (b) 1 wt% 2-MI—0.8LNS film.
Buildings 15 04272 g013
Table 1. The chemical composition of HRB400 steel.
Table 1. The chemical composition of HRB400 steel.
ElementCSiMnVFe
Percent (wt%)0.230.371.570.07balance
Table 2. Mass fraction of main chemical composition of cement.
Table 2. Mass fraction of main chemical composition of cement.
ComponentSiO2MgOAl2O3Fe2O3CaOSO3Loss
Content (%)20.863.505.903.6162.542.431.16
Table 3. Electrochemical data from EIS fitting for sol-gel films constructed on mill-scale steel with different La3+/Ni2+ concentration ratios in simulated concrete pore solution containing 0.1 M NaCl.
Table 3. Electrochemical data from EIS fitting for sol-gel films constructed on mill-scale steel with different La3+/Ni2+ concentration ratios in simulated concrete pore solution containing 0.1 M NaCl.
SampleRs (Ω·cm2)Q1 (×10−5 Ω−1·cm−2·sn)n1Rf (×103 Ω·cm2)Q0 (×10−5 Ω−1·cm−2·sn)n0Ros (×103 Ω cm2)Q2 (×10−5 Ω−1·cm−2·sn)n2Rct (×105 Ω·cm2)Chi-Squared (×10−4)
Blank HRB40052.5645.000.56///0.4426.760.750.124.0
Undoped sol-gel111.100.890.761.4211.980.769.4022.290.541.326.2
LS film116.609.670.780.9634.850.5411.204.600.891.171.7
0.8LNS film77.094.350.891.5812.300.6925.100.050.854.269.0
0.2LNS film99.949.530.730.577.600.6422.593.300.923.937.3
0.05LNS film65.881.540.881.0932.220.6013.4213.350.521.260.8
NS film77.3714.370.621.4829.350.628.9732.000.680.530.6
Table 4. Electrochemical data from EIS fitting for sol-gel coatings constructed on mill-scale steel with varying 2-Methylimidazole content added to the base (La3+/Ni2+ concentration ratio = 0.8) in simulated concrete pore solution containing 0.1 M NaCl.
Table 4. Electrochemical data from EIS fitting for sol-gel coatings constructed on mill-scale steel with varying 2-Methylimidazole content added to the base (La3+/Ni2+ concentration ratio = 0.8) in simulated concrete pore solution containing 0.1 M NaCl.
SampleRs (Ω·cm2)Q1 (×10−5 Ω−1·cm−2·sn)n1Rf (×103 Ω·cm2)Q0 (×10−5 Ω−1·cm−2·sn)n0Ros (×103 Ω cm2)Q2 (×10−5 Ω−1·cm−2·sn)n2Rct (×105 Ω·cm2)Chi-Squared (×10−4)
Blank HRB40052.5645.000.56///0.4426.760.750.124.0
0.8LNS film77.094.350.891.5812.300.6925.100.050.854.269.0
0.5 wt% 2-MI—0.8LNS film66.304.370.901.0515.950.7727.800.110.920.527.2
1 wt% 2-MI—0.8LNS film59.753.050.932.1715.200.7331.100.040.891.779.3
2 wt% 2-MI—0.8LNS film79.321.860.911.7310.490.7526.550.330.921.497.3
Table 5. Electrochemical data from EIS fitting for Blank HRB400 and 1 wt% 2-MI—0.8LNS film in 0.1 M NaCl + SCP solution after 40 days of immersion.
Table 5. Electrochemical data from EIS fitting for Blank HRB400 and 1 wt% 2-MI—0.8LNS film in 0.1 M NaCl + SCP solution after 40 days of immersion.
SampleTime (Day)Rs (Ω·cm2)Q1 (×10−5 Ω−1·cm−2·sn)n1Rf (×103 Ω·cm2)Q0 (×10−5 Ω−1·cm−2·sn)n0Ros (×103 Ω cm2)Q2 (×10−5 Ω−1·cm−2·sn)n2Rct (×105 Ω·cm2)Chi-Squared (×10−4)
Blank HRB4001D52.5645.000.56///0.4426.760.750.124.0
5D43.4048.160.58///0.3441.220.740.486.9
10D49.9554.000.55///2.4511.670.861.129.4
15D77.1558.390.53///0.6641.850.680.372.8
20D52.88107.300.52///0.2181.610.700.133.0
25D69.3885.190.55///0.30103.600.730.159.6
30D105.2039.420.62///0.20142.000.610.329.3
35D75.6641.780.61///0.18157.700.560.279.3
40D99.1061.330.50///0.1789.080.500.131.0
1 wt% 2-MI—0.8LNS film1D59.753.050.932.1715.200.7331.103.810.891.779.3
5D64.423.310.922.0816.360.7330.603.530.902.809.7
10D47.953.080.901.8114.830.7430.1038.450.864.539.4
15D40.993.090.891.4213.160.7331.1092.000.877.354.9
20D22.522.900.911.1315.540.7329.5029.460.854.180.9
25D55.347.890.850.5011.710.7826.70865.000.603.216.5
30D64.577.350.860.6211.970.7824.30953.000.542.875.3
35D55.259.240.820.3910.410.7921.60125.000.542.848.9
40D76.998.460.850.6311.860.7917.20142.200.522.417.6
Table 6. Electrochemical parameters from potentiodynamic polarization curves for sol-gel films constructed on mill-scale steel with different La3+/Ni2+ concentration ratios in simulated concrete pore solution containing 0.1 M NaCl.
Table 6. Electrochemical parameters from potentiodynamic polarization curves for sol-gel films constructed on mill-scale steel with different La3+/Ni2+ concentration ratios in simulated concrete pore solution containing 0.1 M NaCl.
Sampleicorr (µA/cm2)Ecorr (V vs. SCE)βa (mV/dec)|βc| (mV/dec)
Blank HRB4000.82779−0.5404169.0887.44
Undoped sol-gel0.71633−0.4826189.3399.66
LS film0.19144−0.42969112.53136.46
0.8LNS film0.13085−0.43933196.12151.06
0.2LNS film0.18628−0.45639150.58122.77
0.05LNS film0.12585−0.48048102.37104.24
NS film0.19911−0.45290102.48100.86
Table 7. Electrochemical parameters from potentiodynamic polarization curves for sol-gel coatings constructed on mill-scale steel with varying 2-MI content added to the base (La3+/Ni2+ concentration ratio = 0.8) in simulated concrete pore solution containing 0.1 M NaCl.
Table 7. Electrochemical parameters from potentiodynamic polarization curves for sol-gel coatings constructed on mill-scale steel with varying 2-MI content added to the base (La3+/Ni2+ concentration ratio = 0.8) in simulated concrete pore solution containing 0.1 M NaCl.
Sampleicorr (µA/cm2)Ecorr (V vs. SCE)βa (mV/dec)|βc| (mV/dec)
Blank HRB4000.82779−0.5404169.0887.44
Undoped sol-gel0.71633−0.4826189.3399.66
0.8LNS film0.13085−0.43933196.12151.06
0.5 wt% 2−MI—0.8LNS film0.08300−0.4314294.11125.67
1 wt% 2−MI—0.8LNS film0.10870−0.38397105.39200.94
2 wt% 2−MI—0.8LNS film0.19502−0.4096795.18176.67
Table 8. Comparative analysis of engineered sol-gel coatings versus previous studies.
Table 8. Comparative analysis of engineered sol-gel coatings versus previous studies.
Research SystemSteel Bar Temperature (°C)Sol-Gel Solution Temperature (°C)Curing Temperature (°C)Immersion Time in SCP Solution (days)Impedance Gains (-Fold)
Ce Sol-gel film [25]2525130303.0
Ce-Ni Sol-gel film [26]2535130305.8
ZIF-8—Ce sol-gel film [24]2525130305.0
1 wt%-2MI—0.8LNS40025604010.4
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

Xie, Y.; Zeng, Y.; Wang, X.; Bai, Y.; Meng, G. Preparation and Research on 2-Methylimidazole-Lanthanum Nickel-Based Sol-Gel Conversion Coating for Oxide Scale Reinforcement Bars. Buildings 2025, 15, 4272. https://doi.org/10.3390/buildings15234272

AMA Style

Xie Y, Zeng Y, Wang X, Bai Y, Meng G. Preparation and Research on 2-Methylimidazole-Lanthanum Nickel-Based Sol-Gel Conversion Coating for Oxide Scale Reinforcement Bars. Buildings. 2025; 15(23):4272. https://doi.org/10.3390/buildings15234272

Chicago/Turabian Style

Xie, Yuhao, Yanwei Zeng, Xinwei Wang, Yuxin Bai, and Guozhe Meng. 2025. "Preparation and Research on 2-Methylimidazole-Lanthanum Nickel-Based Sol-Gel Conversion Coating for Oxide Scale Reinforcement Bars" Buildings 15, no. 23: 4272. https://doi.org/10.3390/buildings15234272

APA Style

Xie, Y., Zeng, Y., Wang, X., Bai, Y., & Meng, G. (2025). Preparation and Research on 2-Methylimidazole-Lanthanum Nickel-Based Sol-Gel Conversion Coating for Oxide Scale Reinforcement Bars. Buildings, 15(23), 4272. https://doi.org/10.3390/buildings15234272

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