Next Article in Journal
Mechanical Characteristics of 26H2MF and St12T Steels Under Torsion at Elevated Temperatures
Previous Article in Journal
Thermal Degradation and Flame Resistance Mechanism of Phosphorous-Based Flame Retardant of ABS Composites Used in 3D Printing Technology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Silica Sol on the Preparation and Oxidation Resistance of MoSi2@SiO2

1
College of Chemistry and Chemical Engineering, Cangzhou Normal University, Cangzhou 061001, China
2
Key Lab of Advanced Ceramics and Machining Technology of the Ministry of Education, School of Materials Science and Engineering, Tianjin University, Tianjin 300350, China
3
College of Science, Civil Aviation University of China, Tianjin 300300, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(13), 3203; https://doi.org/10.3390/ma18133203
Submission received: 14 May 2025 / Revised: 20 June 2025 / Accepted: 3 July 2025 / Published: 7 July 2025
(This article belongs to the Section Advanced and Functional Ceramics and Glasses)

Abstract

The limited oxidation resistance of MoSi2 between 400 °C and 600 °C restricts its aerospace applications. This study develops a silica-sol derived core-shell MoSi2@SiO2 composite to enhance the low-temperature oxidation resistance of MoSi2. Acidic, neutral, and basic silica sols were systematically applied to coat MoSi2 powders through sol-adsorption encapsulation. Two pathways were used, one was ethanol-mediated dispersion, and the other was direct dispersion of MoSi2 particles in silica sol. Analysis demonstrated that ethanol-mediated dispersion significantly influenced the coating efficiency and oxidation resistance, exhibited significantly decreased coating weight gains (maximum 27%) and increased oxidation weight gains (10–20%) between 340 °C and 600 °C compared with direct dispersion of MoSi2 particles with silica sol, ascribe to the kinetic inhibition of hydroxyl group condensation and steric hindrance of MoSi2-silica sol interface interactions of ethanol. Systematic investigation of silica sol encapsulation of MoSi2 revealed critical correlations between colloid properties and oxidation resistance of MoSi2@SiO2. Basic silica sol coated MoSi2 (BS-MoSi2) exhibits the lowest coating efficiency (coating weight gain of 7.74 ± 0.06%) as well as lowest oxidation weight gain (18.45%) between 340 °C and 600 °C compared with those of acid and neutral silica sol coated MoSi2 (AS-MoSi2 and NS-MoSi2), arises from optimal gelation kinetics, enhanced surface coverage via reduced agglomeration, and suppressed premature nucleation through controlled charge interactions under alkaline conditions.

1. Introduction

MoSi2 has emerged as a promising high-temperature structural material with multifunctional applications in ceramic, metallic, and C/C composite-based coatings. Its exceptional infrared emissivity makes it indispensable as the primary emittance agent in radiative thermal protection coatings [1,2,3]. The material’s inherent oxidation resistance at high temperatures (>800 °C) positions MoSi2 as the principal constituent of oxidation-resistant coatings for metals and C/C composites [4,5]. Furthermore, MoSi2, serving as self-healing particles, can oxidize and form a glass-like silicon dioxide (SiO2) layer, which can fill the cracks in thermal barrier coatings (TBCs) [6,7]. However, critical limitations persist: conventional MoSi2-containing coatings need to be applied in high-temperature conditions, and the material exhibits pulverization oxidation characteristics due to volume expansion within 400–600 °C [8,9]. Current research efforts focus on developing composite modification strategies to simultaneously suppress oxidation-induced decohesion and maintain mechanical integrity under service conditions. Addressing these critical limitations is essential for advancing MoSi2-based thermal protection coatings in aerospace.
Current strategies for enhancing MoSi2 oxidation resistance primarily focus on two material engineering paradigms: (1) alloying design through the addition of refractory metals (e.g., Al, Nb) forming intermetallic compounds [10,11], and (2) composite reinforcement using secondary phases (e.g., SiC, ZrB2) [12,13]. While these approaches effectively improve oxidation resistance in bulk composites, their application in powder-based materials with porous and loose structures faces critical limitations. For MoSi2 materials, their high-temperature oxidation resistance comes from the formation of a dense SiO2 protective layer at high temperatures, which prevents oxygen from entering and further oxidation of the internal materials [14]. Therefore, generating a protective coating through pre-oxidation has become an important way of improving the oxidation resistance of MoSi2 materials [15,16]. However, for MoSi2 powder, strict conditions are required to generate an in situ protective coating through the pre-oxidation method [17]. Advances in surface modification engineering propose an alternative pathway: chemical synthesis of core-shell structured MoSi2@SiO2 particles [18]. This bottom-up approach enables independent control over shell thickness and crystallinity through sol-gel parameters (pH, precursor ratio) while avoiding thermal degradation of the core material [19].
The use of SiO2 as a protective shell for MoSi2 material mainly lies in its high stability and optical transparency, which can protect MoSi2 from oxidation without affecting its emissivity. Xiang’s research suggests that the emissivity of both layered structures with SiO2 covered on the surface of MoSi2 and composites with SiO2 mixed with MoSi2 is enhanced significantly and demonstrates that oxides play a dual role in determining the emissivity of MoSi2, i.e., anti-reflective agents and binders [1]. Perhydropolysilazane, tetraethyl orthosilicate (TEOS) and silica sol can be used as precursors for generating silica shells. In patent US6391383, the SiB4 is dispersed in perhydropolysilazane, and after heat treatment at 400 °C, a SiO2 film is formed on the surface of SiB4 to prevent oxidation and serve as a reaction barrier, effectively improving the emissivity of the coating [20]. Perhydropolysilazane can be converted to SiO2 at lower temperatures, but it is expensive. Chen et al. prepared MoSi2@SiO2 particles via the tetraethyl orthosilicate (TEOS) hydrolysis self-assembly method with tetrabutyl ammonium bromide (TBAB) as an electrostatic adsorbent after heat treatment at 1200 °C [21]. Strict control conditions are required for the hydrolysis and polymerization of TEOS. In comparison, silica sol is an ideal precursor for forming silica shells because of its low price and easy control of the gel process [22,23]. Silica sol is a highly dispersed colloid of poly silicic acid with water as the dispersing medium, which is a heterogeneous inorganic polymer with good adhesion and film-forming properties [24]. In our previous work, core-shell structured MoSi2@SiO2 was prepared with the sol-gel process followed by pre-oxidation at 800 °C, in which a dense coating layer formed as the shell. It showed less weight gains (<1 wt%) than those prepared with the direct preoxidation method (<1.6 wt%) after cyclic-isothermal oxidation at 400–600 °C for 12 h [17,19], indicating that the sol-adsorption coating process played an important role in improving the oxidation resistance of MoSi2@SiO2, and, therefore, the factors (such as dispersant and pH) affecting the properties of silica sol, as well as the gel coating MoSi2@SiO2, need to be carefully studied. In this paper, the influence of the sol-adsorption process (such as dispersant and pH) on shell formation, as well as the oxidation resistance of core-shell MoSi2@SiO2, was mainly focused on without high-temperature pre-oxidation treatment. To reveal the influence of the silica gel coating on the oxidation resistance of MoSi2@SiO2, three kinds of silica sol were used as the precursor to prepare MoSi2@SiO2 particles, and the effect of silica sols on the oxidation resistance of MoSi2@SiO2 was deeply investigated.

2. Materials and Methods

MoSi2 powder (Eno Material, Qinhuangdao, China) with a median particle size (D50) of 3.721 μm was used as raw material. Aqueous silica sol was commercially available (Jinghuo Technique Glass Co., Ltd., Dezhou, China) with detailed parameters in Table 1. The MoSi2@SiO2 composites were fabricated through a sol-adsorption process using acidic, neutral, and basic silica sols (pH values detailed in Table 1) as precursors. Two pathways were used: one was ethanol-mediated dispersion, and the other was direct dispersion of MoSi2 particles in silica sol. Precisely weighed MoSi2 powders (±0.0001 g accuracy) received a certain amount of respective silica sols with a mass ratio of 1:3 (MoSi2:SiO2) via pipetting, followed by ultrasonic dispersion (40 kHz, 25 °C) for 10 min to achieve homogeneous distribution. In the ethanol-mediated dispersion process, the amount of ethanol used is 2.5 mL/g, depending on the mass of MoSi2. The mixture underwent magnetic stirring (800 rpm, 25 ± 1 °C) for 4 h to form stable suspensions. Subsequent phase separation involved centrifugation (3000 rpm, 20 °C, 30 min) and 24 h sedimentation in sealed containers. The supernatant was decanted, and the residue was subjected to sequential drying: 30 min at 65 °C for solvent removal followed by 4 h at 105 °C for complete dehydration. The resultant xerogels were designated as AS-MoSi2 (acidic), NS-MoSi2 (neutral), and BS-MoSi2 (basic) based on colloid pH characteristics.
The coating efficiency was quantified through gravimetric analysis using Equation (1):
w / % = m 2 m 1 m 1 × 100 % ,
where w represents the coating weight gain rate (%), m1 indicates the pre-treatment MoSi2 mass (g), and m2 denotes the post-treatment composite mass (g). All mass measurements were conducted using the analytical balance (±0.1 mg precision).
The surface chemical groups of AS-MoSi2, NS-MoSi2, and BS-MoSi2 were characterized by Fourier transform infrared spectroscopy (FTIR, Nicolet iS5, Thermo Scientific, Waltham, MA, USA) employing the KBr pellet technique in transmission mode. Phase evolution during oxidation was analyzed through X-ray Diffraction (XRD, TD-3500 Tongda, Dandong, China) with Cu Kα radiation (λ = 1.5406 Å). Microstructural features and elemental distribution were examined using scanning electron microscopy (SEM, TM3030, Hitachi, Tokyo, Japan) coupled with energy-dispersive spectroscopy (EDS). The specific surface area and pore structure of the samples were determined using a specific surface area analyzer (Kobu X1000, Builder, Hangzou, China).
Thermogravimetric analysis (TG, HCT-4, Hengjiu, Beijing, China) was conducted to evaluate non-isothermal oxidation kinetics under an air atmosphere (heating rate 10 °C/min, 40–1100 °C). For isothermal cyclic oxidation evaluation, disc specimens (diameter 1.33 cm, thickness 0.25 ± 0.001 cm) were prepared by cold pressing 0.5000 ± 0.0010 g powder samples at 20 kPa for 10 s. The cyclic oxidation tests at 400 °C, 500 °C, 600 °C, and 700 °C followed a standardized procedure: samples were loaded into preheated furnaces, held at the target temperature for 1 h, then quenched to room temperature. Mass variation was quantified according to Equation (2) using an analytical balance (±0.1 mg precision) after each oxidation cycle.
w o / % = m h m 0 m 0 × 100 %
where wo is the oxidation weight change percent (%), mh is the weight of samples after isothermal cyclic oxidation for a certain time (g), and m0 is the initial sample weight (g).

3. Results

3.1. Effects of Ethanol-Mediated Process on Coating Efficiency and Oxidation Resistance

In order to analyze the effect of ethanol-mediated dispersant, a comparison was made between the xerogel complexes obtained from the two pathways. Samples obtained from ethanol-mediated dispersion are labeled as “with ethanol”, and direct mixing of MoSi2 with silica sol is labeled as “without ethanol”. As shown in Figure 1a, all AS-MoSi2, NS-MoSi2, and BS-MoSi2 samples fabricated via ethanol-mediated dispersion exhibited significantly reduced coating efficiency (CE) values when compared with direct mixing. Notably, this solvent-dependent CE discrepancy reached 27% maximum difference for NS-MoSi2 systems, suggesting that ethanol-mediated dispersion plays an important role in oxide layer formation.
When MoSi2 is directly mixed with silica sol, interfacial water molecules react with MoSi2 surfaces to form oxygen-containing functional groups (hydroxyl –OH and Si–O bonds). Surface hydroxyl species (Si–OH) undergo condensation polymerization through dehydration reaction (Reaction (1)):
Si–OH (MoSi2) + Si–OH → Si–O–Si (MoSi2) + H2O
During solvent evaporation, these Si–OH groups interconnect to construct a three-dimensional SiO2 network on MoSi2 surfaces.
In ethanol-mediated systems, particle interactions follow DLVO theory [25,26]. Ethanol’s weak chemical adsorption via ethoxy (–OC2H5) and trace hydroxyl groups exhibits lower reactivity compared with aqueous systems [25], resulting in suppressed premature Si–O–Si network formation. Ethanol’s dual role includes: kinetic inhibition of hydroxyl group condensation and steric hindrance of MoSi2-silica sol interface interactions. These effects of ethanol explain the reduced coating efficiency (CE) observed in samples fabricated via ethanol-mediated dispersion system, where slower reaction kinetics lead to 10–30% lower coating weight gain rate values versus direct mixing processing.
Thermogravimetric (TG) analysis revealed distinct oxidation resistance trends among AS-MoSi2, NS-MoSi2, and BS-MoSi2 prepared via ethanol-mediated dispersion versus direct mixing. All xerogel composites exhibited a similar trend but reduced mass gain compared to pristine MoSi2, with ethanol-free preparations showing 10–20% lower weight gains between 340 °C and 600 °C than their ethanol-dispersed counterparts. As shown in Figure 1b, the NS-MoSi2 system demonstrated delayed oxidation kinetics: oxidation initiation shifted from 340 °C (MoSi2) to 390 °C, with peak reaction rates occurring at 480 °C (vs. 450 °C for MoSi2).
MoSi2 undergoes an oxidation reaction as shown in Reaction (2), with the theoretical weight gain of the complete reaction being 73.63%.
2MoSi2(s) + 7O2(g) = 4SiO2(s) + 2MoO3(s)
As illustrated in Figure 1b, the TG curves reveal that there exists an upward trend of the weight change between 340 °C and 750 °C, but when the temperature exceeds the volatilization temperature of the oxidation product MoO3 (750 °C), the weight change begins to decline. Notably, the weight gains of MoSi2, NS-MoSi2 (with ethanol), and NS-MoSi2 (without ethanol) between 340 °C and 600 °C are 35.79%, 28.34%, and 24.44%, respectively. These data indicate that NS-MoSi2 experiences a postponed onset of oxidation and reduced oxidation weight gain compared with unmodified MoSi2, confirming that the gel coating enhances oxidation resistance of the coated composites to some degree. Furthermore, the NS-MoSi2 sample prepared through the direct mixing of MoSi2 with silica sol exhibits superior oxidation resistance compared with the sample where MoSi2 was first dispersed in ethanol prior to silica sol addition. This difference is likely attributable to the microstructural characteristics visible in Figure 1c,d.
While ethanol dispersion theoretically should reduce particle agglomeration and create more uniform particle exposure for silica sol polymerization, the slower polymerization kinetics under ethanol conditions result in a less dense protective film formation, as evidenced by the SEM images (Figure 1c). The direct mixing method appears to produce a more continuous and uniform coating layer (Figure 1d), leading to improved oxidation protection.
These findings underscore the significant influence of preparation methodology on protective coating characteristics and resulting oxidation resistance. For optimal antioxidant performance in practical applications, the direct mixing method of MoSi2 and silica sol is recommended for preparing MoSi2@SiO2 composites.

3.2. Effect of Silica Sol on the Oxidation Resistance of Gel Coating MoSi2

3.2.1. Effect of Silica Sol on Coating Formation

As illustrated in Figure 1a, there is a clear correlation between the coating weight gain rate and the type of silica sol used. The AS-MoSi2 sample exhibits the highest coating weight gain rate of 9.12 ± 0.26%, followed by NS-MoSi2 of 8.60 ± 0.41%, while BS-MoSi2 demonstrates the lowest value of 7.74 ± 0.06%.
The zeta potential of MoSi2 in aqueous solution reveals that the surface charge of MoSi2 progressively becomes more negative as the pH increases from 3 to 11 [27]. This negative surface charge characteristic plays a critical role in the coating formation process when MoSi2 is dispersed in silica sol. The surface charge generation on MoSi2 can be attributed to the surface reaction described by Reaction (3):
SiOH (MoSi2) + H2O = SiO (MoSi2) + H3O+
The isoelectric point of pure silica is typically around pH 3 [25]. Consequently, in acidic silica sol environments (pH < 3), the silicon hydroxyl groups undergo protonation according to Reaction (4), resulting in positively charged silica particles (Si–OH2+).
SiOH + H+ = SiOH2+
In neutral and alkaline silica sol conditions (pH > 3), deprotonation occurs as shown in Reactions (5) and (6). These reactions lead to negatively charged silica surfaces (Si–O).
SiOH + H2O = SiO + H3O+
SiOH + OH = SiO + H2O
During the sol-gel process, condensation reactions represent the critical step for coating formation on MoSi2 particles. The mechanisms of condensation are based on two steps: protonation/deprotonating (according to the pH), and then reaction of the protonated/deprotonated particle with the raw species [28]. These condensation mechanisms depend on the surface charge characteristics, which are pH-dependent:
In acidic conditions, the positively charged silica sol particles (Si–OH2+) are strongly attracted to the negatively charged MoSi2 surface (Si–O) through electrostatic interactions. This results in efficient adsorption and deposition of silica onto MoSi2 particles, yielding the highest coating weight gain rate observed. In neutral conditions, the silica sol particles exhibit reduced surface charge, leading to poorer colloidal stability and increased tendency for particle aggregation. This results in decreased coating efficiency and lower weight gain compared to acidic conditions. While in alkaline conditions, the silica sol particles carry negative surface charges (Si–O), which create electrostatic repulsion with the similarly charged MoSi2 surface (Si–O). This significantly reduces adsorption efficiency and produces the lowest coating weight gain rate.

3.2.2. Surface Characteristics and Thermogravimetric Analysis

To investigate the surface chemical characteristics of MoSi2 before and after silica coating treatments, Fourier transform infrared spectroscopy (FTIR) was employed, and the spectra are shown in Figure 2a. The unmodified MoSi2 sample exhibited negligible absorption features across the measured spectral range (4000–400 cm−1), indicating minimal surface functionalization. In contrast, the coated samples (AS-MoSi2, NS-MoSi2, and BS-MoSi2) all displayed distinct absorption bands at approximately 1100 cm−1, 800 cm−1, and 475 cm−1, which correspond to the asymmetric stretching, symmetric stretching, and bending vibrations of the Si–O–Si network structure [29,30]. These findings confirm the successful formation of silica-based coatings on the MoSi2 particle surfaces.
The thermogravimetric analysis (Figure 2b) demonstrates that silica coating substantially enhances the oxidation resistance of MoSi2. The oxidation onset temperature increased from approximately 340 °C for pristine MoSi2 to about 400 °C for all coated derivatives. Between 340 °C and 600 °C, the weight gains due to oxidation were quantified as 31.58%, 28.34%, and 18.45% for AS-MoSi2, NS-MoSi2, and BS-MoSi2, respectively. This trend (BS-MoSi2 < NS-MoSi2 < AS-MoSi2) indicates that BS-MoSi2 exhibits superior antioxidant capacity, likely due to differences in coating microstructure and density.
Scanning electron microscopy (SEM) images (Figure 2c–e) reveal pronounced morphological differences among the silica-coated MoSi2 variants, which correlate with their respective oxidation behaviors. The AS-MoSi2 sample shows relatively large aggregated particles with uneven surface coverage and some interparticle connectivity (Figure 2c). The NS-MoSi2 displays partial surface coating with observable silica aggregate formations (Figure 2d). In contrast, BS-MoSi2 exhibits the most uniform morphology (Figure 2e), with well-dispersed particles and a consistent gel-like coating layer that appears to fully encapsulate the MoSi2 core. These structural variations provide clear visual evidence explaining the differences in oxidation resistance observed in the thermal analysis.

3.2.3. Oxidation Behavior of the Silica-Coated MoSi2

The isothermal cyclic oxidation behavior of AS-MoSi2, NS-MoSi2, and BS-MoSi2 was systematically investigated at temperatures ranging from 400 °C to 700 °C. As illustrated in Figure 3, the weight change kinetics demonstrated distinct temperature-dependent patterns for all three coating types. XRD patterns of the AS-MoSi2, NS-MoSi2, and BS-MoSi2 after isothermal oxidation tests for 4 h are shown in Figure 4.
At 400 °C (Figure 3a), AS-MoSi2 and NS-MoSi2 exhibited relatively rapid high oxidation weight gains compared to that of BS-MoSi2, with weight gains exceeding 40% after 4 h of exposure. This significant weight increase was accompanied by the visible pulverization and volumetric expansion of the samples, indicating vigorous oxidation reactions at this relatively low temperature. Diffraction peaks corresponding to MoO3 (PDF#76-0003) derived from the oxidation of MoSi2 can be detected in all three samples tests at 400 °C, and the diffraction peaks corresponding to MoSi2 (PDF#80-0544) are relatively weak, especially for AS-MoSi2 (Figure 4a), resulting from the oxidation of MoSi2 at this temperature.
As the temperature increased to 500 °C (Figure 3b), the oxidation rate decreased noticeably for all samples compared to those at 400 °C, with weight gains lower than 20%. Notably, BS-MoSi2 consistently showed the lowest weight gain across all time points, while AS-MoSi2 displayed the most pronounced oxidative degradation, particularly in the later stages of exposure. Although the peak reaction rates occur near 500 °C according to the TG analysis (Figure 2b), diffraction peaks corresponding to MoSi2 after isothermal cyclic oxidation at 500 °C are stronger than those at 400 °C (Figure 4b). The samples were oxidized less intensely at 500 °C than those at 400 °C, which may be due to the less volume expansion at the initial stage of oxidation, avoiding more channels for oxygen to enter.
At higher temperatures of 600 °C and 700 °C, the oxidation kinetics exhibited a two-stage behavior: an initial rapid weight gain phase within the first hour, followed by a gradual stabilization period (Figure 3c,d). According to thermal analysis (Figure 2b), the TG curves at 600–700 °C are close to the plateau, indicating a slow oxidation reaction rate at this stage. Among the three samples, BS-MoSi2 maintained the lowest weight gain throughout the entire exposure period (<10%), with only minor surface cracking observed in its visual micrographs (Figure 3c,d). In contrast, AS-MoSi2 and NS-MoSi2 samples displayed increasingly severe surface cracking and structural degradation upon oxidation time, particularly in AS-MoSi2. In XRD patterns, weak diffraction peaks corresponding to MoO3 (PDF#76-0003) were detected in sample tests at 700 °C compared with those at 600 °C (Figure 4c,d), which may be attributed to the volatilization of MoO3 at this temperature.
The observed differences in oxidation behavior among AS-MoSi2, NS-MoSi2, and BS-MoSi2 can be attributed to variations in their surface adsorption and the consequent effects on protective oxide layer formation. The enhanced oxidation resistance of BS-MoSi2 is likely related to its unique compositional characteristics that promote the formation of more adherent and continuous protective oxide scales.
The surface elemental distributions of AS-MoSi2, NS-MoSi2, and BS-MoSi2 were investigated using EDS mapping in correlation with SEM morphology (Figure 5). The combined SEM with BSE imaging (Figure 5a,c,e) and EDS mapping (Figure 5b,d,f) data reveal distinct elemental distributions that correlate with microstructural evolution.
In AS-MoSi2 (Figure 5a,b), the SEM image reveals large particulate aggregates with rough surface texture, while EDS mapping shows intense red regions of oxygen, indicating localized oxygen enrichment. This corresponds to thick but uneven gel layer formation, likely due to heterogeneous nucleation during coating deposition. For NS-MoSi2 (Figure 5c,d), SEM imaging demonstrates finer particulate dispersion with reduced aggregation. The corresponding EDS mapping exhibits sparse red oxygen signals of oxygen, consistent with the lower surface charge density of MoSi2 in neutral conditions, which restricts effective SiO2 gel polymerization and leads to discontinuous coating formation. The BS-MoSi2 sample (Figure 5e,f) displays uniform surface morphology with a tightly packed surface in SEM imaging. EDS mapping confirms homogeneous elemental distribution, particularly notable for the significantly reduced molybdenum content (19 wt% vs. 30 wt% in AS-MoSi2). The notably high silicon content (80 wt%) suggests the effective suppression of oxidation through alkaline-induced Si–O–Si network densification, forming a conformal gel coating that minimizes MoSi2 exposure.
This systematic comparison demonstrates how pH-mediated sol-adsorption modulates both coating morphology and chemical composition, with alkaline conditions proving most effective for achieving homogeneous protective layers on MoSi2 surfaces.
To further investigate the pore structure of the samples, adsorption/desorption behaviors of MoSi2, AS-MoSi2, NS-MoSi2, and BS-MoSi2 were evaluated through the BET test. Figure 6 illustrates the nitrogen adsorption–desorption isotherms and pore size distributions of the samples. According to IUPAC classification, the curve of MoSi2 shows type III isotherm, and the curves of the AS-MoSi2, NS-MoSi2, and BS-MoSi2 show type IV isotherms with H2a hysteresis loops (Figure 6a). The IV curve indicates that the porous material has multiple adsorption forces on the surface, a pronounced pore structure, and an increase in adsorption rate at low relative pressures, indicating the presence of mesoporous structures within the samples (Figure 6b) [31].
Table 2 summarizes the detailed test data. BS-MoSi2 possesses the smallest pore size (2.86 nm) and the largest specific surface area (66.61 m2/g) among the three samples, resulting in poor oxidation resistance. NS-MoSi2 has similar surface areas but relatively larger pore sizes (6.11 nm) compared with BS-MoSi2 (4.72 nm), providing channels for oxygen to diffuse inward. Overall, BS-MoSi2 has the smallest specific surface area and middle pore size; therefore, it has the best oxidation resistance.

3.3. Mechanism Analysis of Coating Encapsulation of MoSi2

The coating formation mechanism can be elucidated through thermodynamic, kinetic, and surface interaction analyses. Thermodynamically, heterogeneous nucleation of silica sol particles on MoSi2 surfaces requires less surface energy than homogeneous nucleation in aqueous phases [32], establishing preferential surface coating. Three key factors govern this coating process: the dispersion and surface properties of MoSi2 itself, the colloidal stability of silica sol, and the reactivity of MoSi2 with silica sol.
Firstly, the isoelectric point of MoSi2 is between pH 1–2, and the zeta potential of MoSi2 becomes more negative with increasing pH, indicating that MoSi2 is negatively charged in all three types of silica sol. MoSi2 has poor dispersibility and is prone to aggregation in acidic environments, while in alkaline environments, the dispersibility is better and the particles are easier to be encapsulated [27].
Secondly, in terms of kinetics, the polymerization of silica sol involves competition between film-forming coating and nucleation coating [33]. That is, if the gel speed of silica sol is controlled, a uniform SiO2 film can be formed on the surface of MoSi2. If the polymerization of silica sol is too fast, it will lead to self-nucleation and the formation of nucleation coating. pH is a key factor affecting the gelation time of silica sol. When the concentration is the same, the curve of gelation time (t) with pH changes in an “N” shape [34]. The pH corresponding to the lgt of acidic silica sol and alkaline silica sol is located at the high point of the N-shaped curve, and the gelation speed of silica sol is slow, which is conducive to the formation of continuous and uniform film coating. The pH corresponding to the lgt of neutral silica sol is located at the low point of the N-shaped curve, with a fast gelation rate, which is not conducive to film coating and tends to self-nucleate.
Lastly, as mentioned earlier, in acidic environments, negatively charged MoSi2 and positively charged silica sol undergo electrostatic attraction, leading to rapid adsorption and deposition of silica sol on the surface of MoSi2 particles, resulting in high coating efficiency; in neutral environment, less surface groups lead to less condensation; and in basic environment, there is electrostatic repulsion between charges of the same type, but polymerization reactions can still occur.
Based on the above analysis, schematic diagrams of the film formation mechanism of AS-MoSi2, NS-MoSi2, and BS-MoSi2 are shown in Figure 7. Under acidic conditions (pH < 3), MoSi2 exhibits poor dispersibility due to electrostatic aggregation. Although rapid adsorption of silica sol particles (Si–OH2+) occurs via electrostatic attraction to the negatively charged MoSi2 surface (Si–O), the high gelation rate promotes premature nucleation rather than uniform film growth. This results in thick but porous gel films with irregular coverage and limited oxidation resistance. In neutral silica sol, MoSi2 has enhanced dispersibility but fewer surface functional groups, which is not conducive to the formation of adsorption layers. At the same time, the silica sol has a fast gelation rate, resulting in the formation of SiO2 aggregates between particles. In basic silica sol, the good dispersibility of MoSi2, the electrostatic repulsion between particles, and the slow gelling speed of silica sol are conducive to the polymerization to form a uniform gel film.
The coating quality and oxidation resistance of BS-MoSi2 arise from optimal gelation kinetics under alkaline conditions (slow, homogeneous growth), enhanced surface coverage via reduced agglomeration in high-pH media, and suppressed premature nucleation through controlled charge interactions. This pH-dependent behavior provides a framework for optimizing sol-gel derived coatings, where colloidal stability and reaction kinetics are precisely balanced to achieve conformal encapsulation.

4. Conclusions

In this study, acidic, neutral, and basic silica sols were systematically applied to coat MoSi2 powders through sol-adsorption and encapsulation. Two distinct fabrication routes were employed: (1) ethanol-mediated dispersion of MoSi2 followed by silica sol impregnation, and (2) direct mixing of MoSi2 with silica sol. The impact of preparation methodology and silica sol pH on protective film characteristics and resulting oxidation resistance of core-shell structured MoSi2@SiO2 powders was systematically investigated. The main results are as follows.
(1)
AS-MoSi2, NS-MoSi2, and BS-MoSi2 specimens exhibited significantly reduced coating efficiency values (maximum 27%) when fabricated via ethanol-mediated dispersion compared with direct mixing. All xerogel powder exhibited reduced mass gain compared to pristine MoSi2 in TG analysis, with ethanol-free preparations showing 10–20% lower weight gains between 340 °C and 600 °C than their ethanol-mediated dispersed counterparts. The NS-MoSi2 demonstrated delayed oxidation kinetics: oxidation initiation shifted from 340 °C (MoSi2) to 390 °C, with peak reaction rates occurring at 480 °C (vs. 450 °C for MoSi2).
(2)
There is a clear correlation between the coating weight gain rate and the type of silica sol used. The AS-MoSi2 sample exhibits the highest coating weight gain rate at 9.12 ± 0.26%, followed by NS-MoSi2 with 8.60 ± 0.41%, while BS-MoSi2 demonstrates the lowest value at 7.74 ± 0.06%. While the weight gains between 340 °C and 600 °C due to oxidation were quantified as 31.58%, 28.34%, and 18.45% for AS-MoSi2, NS-MoSi2, and BS-MoSi2, respectively, indicate that BS-MoSi2 exhibits superior oxidant resistance.
(3)
The oxidation resistance of BS-MoSi2 arises from the coating quality, which is determined by optimal gelation kinetics under alkaline conditions (slow, homogeneous growth); enhanced surface coverage via reduced agglomeration in high-pH media; and suppressed premature nucleation through controlled charge interactions.

Author Contributions

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

Funding

This research was funded by the Natural Science Foundation of Hebei Province, grant number E2022110002; Natural Science Foundation of Cangzhou City, grant number 221001005D and the Advanced Composite Materials and Coating Research and Innovation Team of Cangzhou Normal University, grant number cxtdl2302.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Huimin, X.; Fuzhi, D.; Yanchun, Z. Secrets of high thermal emission of transition metal disilicides TMSi2 (TM = Ta, Mo). J. Mater. Sci. Technol. 2021, 89, 114–121. [Google Scholar] [CrossRef]
  2. Tao, X.; Xu, X.; Guo, L.; Hong, W.; Guo, A.; Hou, F.; Liu, J. MoSi2-borosilicate glass coating on fibrous ceramics prepared by in-situ reaction method for infrared radiation. Mater. Des. 2016, 103, 144–151. [Google Scholar] [CrossRef]
  3. Yang, X.; Wan, Y.; Li, J.; Liu, J.; Wang, M.; Tao, X. High Emissivity MoSi2-SiC-Al2O3 Coating on Rigid Insulation Tiles with Enhanced Thermal Protection Performance. Materials 2024, 17, 220. [Google Scholar] [CrossRef] [PubMed]
  4. Cai, Z.; Zhang, D.; Chen, X.; Huang, Y.; Peng, Y.; Xu, C.; Huang, S.; Pu, R.; Liu, S.; Zhao, X.; et al. A novel ultra-high-temperature oxidation protective MoSi2-TaSi2 ceramic coating for tantalum substrate. J. Eur. Ceram. Soc. 2019, 39, 2277–2286. [Google Scholar] [CrossRef]
  5. Xuanru, R.; Junshuai, L.; Wei, L.; Yuwen, H.; Ke, S.; Can, M.; Hong’ao, C.; Weiguang, W.; Leihua, X.; Ziyu, L.; et al. Influence of MoSi2 on oxidation protective ability of TaB2-SiC coating in oxygen-containing environments within a broad temperature range. J. Adv. Ceram 2020, 9, 703–715. [Google Scholar] [CrossRef]
  6. Ying, C.; Xun, Z.; Zwaag, S.; Willem, G.; Sloof, P.X. Damage evolution in a self-healing air plasma sprayed thermal barrier coating containing self-shielding MoSi2 particles. J. Am. Ceram. Soc. 2019, 102, 4899–4910. [Google Scholar] [CrossRef]
  7. Denise, K.; Denial, E.; Robert, V. Degradation and lifetime of self-healing thermal barrier coatings containing MoSi2 as self-healing particles in thermo-cycling testing. Surf. Coat. Technol. 2022, 437, 128353. [Google Scholar] [CrossRef]
  8. Chou, T.; Nieh, T. Mechanism of MoSi2 pest during low temperature oxidation. J. Mater. Res. 1993, 8, 214–226. [Google Scholar] [CrossRef]
  9. Hansson, K.; Halvarsson, M.; Tang, J.; Pompe, R.; Sundberg, M.; Svensson, J. Oxidation behaviour of a MoSi2-based composite in different atmospheres in the low temperature range (400–550 °C). J. Eur. Ceram. Soc. 2004, 24, 3559–3573. [Google Scholar] [CrossRef]
  10. Zenk, C.; Gibson, K.; Maier-Kiener, V.; Neumeier, S.; Korte-Kerzel, S. Low temperature deformation of MoSi2 and the effect of Ta, Nb and Al as alloying elements. Acta Mater. 2019, 181, 385–398. [Google Scholar] [CrossRef]
  11. Niannian, L.; Jun, G.; Wanxia, W.; Sheng-Chi, C.; Kunlun, W.; Yong, W.; Chao-Kuang, W.; Hui, S. Oxidation resistance of Cr-modified MoSi2 composites at high temperature. Int. J. Refract. Met. Hard Mater. 2024, 119, 106497. [Google Scholar] [CrossRef]
  12. Sedighi, A.; Adeli, M.; Soltanieh, M. Investigation of the effect of SiC additions on the high-temperature oxidation behavior of combustion-synthesized MoSi2. J. Mater. Res. Technol. 2024, 30, 187–196. [Google Scholar] [CrossRef]
  13. Ji, X.; Chen, Y.; Yao, L.; Zhang, Y.; Ren, X.; Wang, P.; Korneev, P.; Levashov, E.; Shi, J.; Kang, X.; et al. Enhanced oxidation resistance of ZrB2-MoSi2 coating through MoSi2-TaSi2 double-silicide alloying modifying. Corro. Sci. 2024, 233, 112070. [Google Scholar] [CrossRef]
  14. Bundschuh, K.; Schüze, M.; Müller, C.; Greil, P.; Heider, W. Selection of materials for use at temperatures above 1500 °C in oxidizing atmospheres. J. Eur. Ceram. Soc. 1998, 18, 2389–2391. [Google Scholar] [CrossRef]
  15. Feng, P.; Wang, X.; He, Y.; Qiang, Y. Effect of high-temperature preoxidation treatment on the low-temperature oxidation behavior of a MoSi2-based composite at 500 °C. J. Alloys Compd. 2009, 473, 185–189. [Google Scholar] [CrossRef]
  16. Yakaboylu, G.; Yumak, T.; Sabolsky, K.; Sabolsky, E. Effect of high temperature preoxidation treatment on the oxidation behavior of MoSi2- and WSi2-Al2O3 composites. J. Alloys Compd. 2020, 816, 152499. [Google Scholar] [CrossRef]
  17. Guo, L.; Liu, B.; Zhang, J.; Zhang, X.; Miao, C.; Wu, Y.; Du, H. Low-temperature oxidation resistance of core-shell structure MoSi2@SiO2. J. Chin. Ceram. Soc. 2022, 50, 270–276. [Google Scholar] [CrossRef]
  18. Guerrero, M.; Pérez, J.; Liz, M. Recent progress on silica coating of nanoparticles and related nanomaterials. Adv. Mater. 2010, 22, 1182–1195. [Google Scholar] [CrossRef]
  19. Guo, L.; Hu, X.; Tao, X.; Du, H.; Guo, A.; Liu, J. Low-temperature oxidation resistance of the silica-coated MoSi2 powders prepared by sol-gel preoxidation method. Ceram. Int. 2020, 46, 23471–23478. [Google Scholar] [CrossRef]
  20. Matsunaga, H.; Iwata, M.; Shinji, K.; Kenji, Y.; Kazuhiro, H.; Yoshitsugu, K.; Takamitsu, I.; Masakazu, H.; Takamitsu, F. High Emittance Glass Coating Material, High Emittance Glass Coating, and Method of Produce High Emittance Glass Coating. U.S. Patent NO. 6391383, 21 May 2002. [Google Scholar]
  21. Chen, Y.; Zhu, S.; Ji, Y.; Ma, Z.; Wei, H. Oxidation resistance and infrared emissivity of MoSi2@SiO2 particles prepared via TEOS hydrolysis self-assembly method. J. Alloys Compd. 2019, 810, 151745. [Google Scholar] [CrossRef]
  22. Lei, X.; Sun, H.; Yuan, X.; Zhang, W.; Jiang, P.; Zhang, L. Exploration of silica sol-gel coating for the preparation of core-shell structure VO2(M)@SiO2. Inorg. Chem. Ind. 2024, 56, 46–54. [Google Scholar]
  23. Tao, X.; Li, J.; Liu, J.; Cai, G.; Zhang, J.; Wang, M. Improved oxidation resistance and infrared emissivity of tantalum disilicide particles with sol-gel derived SiO2 coatings. J. Alloys Compd. 2024, 971, 172765. [Google Scholar] [CrossRef]
  24. Zhang, R.; Li, J.; Wang, Z.; Qi, C.; Zhuo, J.; Wan, Y.; Zhang, Y.; Liu, H.; Xiao, B.; Wang, M. Preparation of porous mullite ceramics composed entirely of overlapping and interlocking mullite whiskers through whisker in-situ growth. J. Adv. Ceram. 2025, 14, 9221055. [Google Scholar] [CrossRef]
  25. Yoko, F.; Patrick, S. Relative colloidal stability in ethanol of powders with Si-O surface species. J. Ame. Ceram. Soc. 2004, 85, 2945–2948. [Google Scholar] [CrossRef]
  26. Yoko, F.; Patrick, S. The role of Si-O species in the colloidal stability of silicon-containing ceramic powders. J. Eur. Ceram. Soc. 2004, 24, 17–23. [Google Scholar] [CrossRef]
  27. Sandlin, M.; Butt, D.; Taylor, T.; Petrovic, J. Aqueous zeta potentials of molybdenum disilicide. J. Mater. Sci. Let. 1997, 16, 1336–1338. [Google Scholar] [CrossRef]
  28. Tamar, Y.; Sasson, Y. Examination of the regime controlling sol-gel based colloidal silica aggregation. J. Non-Cryst. Solids 2013, 380, 35–41. [Google Scholar] [CrossRef]
  29. Dyah, A.; Budi, H.; Neny, K.; Nazopatul, P.; Noviyan, D.; Irzaman, I. Functional groups, band gap energy, and morphology properties of annealed silicon dioxide (SiO2). Egypt. J. Chem. 2023, 66, 529–535. [Google Scholar] [CrossRef]
  30. Liu, J.; Wan, Y.; Xiao, B.; Li, J.; Hu, Z.; Zhang, R.; Hu, X.; Liu, J.; Cai, G.; Liu, H.; et al. The preparation and performance analysis of zirconium-modified aluminum phosphate-based high-temperature (RT-1500 °C) resistant adhesive for joining alumina in extreme environment. J. Adv. Ceram. 2024, 13, 911–932. [Google Scholar] [CrossRef]
  31. Yang, X.; Wan, Y.; Yang, N.; Hou, Y.; Chen, D.; Liu, J.; Cai, G.; Wang, M. The effect of different diluents and curing agents on the performance of epoxy resin-based intumescent flame-retardant coatings. Materials 2024, 17, 348. [Google Scholar] [CrossRef]
  32. Cui, A.; Wang, T.; Jin, Y.; Sun, M. Thermodynamic research on the coating process of silica nano film on titanate particles surface. Chem. J. Chin. Univ. 2001, 22, 1543–1545. [Google Scholar] [CrossRef]
  33. Cui, A.; Wang, T.; Jin, Y.; Xiao, S.; Ge, X. Mechanism and structure analysis of TiO2 surface coated with SiO2 and Al2O3. Chem. J. Chem. J. Chin. Univ. 1998, 19, 1727–1729. [Google Scholar] [CrossRef]
  34. Wenfu, Y.; Ruren, X. Chemical reactions in aqueous solutions with condensed liquid state. Prog. Chem. 2022, 34, 1454–1491. [Google Scholar]
Figure 1. (a) Coating efficiency of AS-MoSi2, NS-MoSi2, and BS-MoSi2 fabricated via ethanol-mediated dispersion vs. direct mixing. (b) Comparative thermogravimetric (TG) curves of MoSi2 and NS-MoSi2 (ethanol-mediated vs. direct mixing). (c,d) Scanning electron microscopy (SEM) microstructures of NS-MoSi2 under different processing routes: (c) ethanol-mediated and (d) direct mixing.
Figure 1. (a) Coating efficiency of AS-MoSi2, NS-MoSi2, and BS-MoSi2 fabricated via ethanol-mediated dispersion vs. direct mixing. (b) Comparative thermogravimetric (TG) curves of MoSi2 and NS-MoSi2 (ethanol-mediated vs. direct mixing). (c,d) Scanning electron microscopy (SEM) microstructures of NS-MoSi2 under different processing routes: (c) ethanol-mediated and (d) direct mixing.
Materials 18 03203 g001
Figure 2. FTIR spectrum with red lines indicating the vibration frequency locations of Si–O–Si (a) and TG (b) curves of MoSi2, AS-MoSi2, NS-MoSi2, and BS-MoSi2, and SEM images of AS-MoSi2 (c), NS-MoSi2 (d) and BS-MoSi2 (e).
Figure 2. FTIR spectrum with red lines indicating the vibration frequency locations of Si–O–Si (a) and TG (b) curves of MoSi2, AS-MoSi2, NS-MoSi2, and BS-MoSi2, and SEM images of AS-MoSi2 (c), NS-MoSi2 (d) and BS-MoSi2 (e).
Materials 18 03203 g002
Figure 3. The isothermal cyclic oxidation behavior of AS-MoSi2, NS-MoSi2, and BS-MoSi2 at 400 °C (a), 500 °C (b), 600 °C (c), and 700 °C (d), with physical images of the samples in (c,d).
Figure 3. The isothermal cyclic oxidation behavior of AS-MoSi2, NS-MoSi2, and BS-MoSi2 at 400 °C (a), 500 °C (b), 600 °C (c), and 700 °C (d), with physical images of the samples in (c,d).
Materials 18 03203 g003
Figure 4. XRD patterns of AS-MoSi2, NS-MoSi2, and BS-MoSi2 after isothermal cyclic oxidation at 400 °C (a), 500 °C (b), 600 °C (c), and 700 °C (d) for 4 h.
Figure 4. XRD patterns of AS-MoSi2, NS-MoSi2, and BS-MoSi2 after isothermal cyclic oxidation at 400 °C (a), 500 °C (b), 600 °C (c), and 700 °C (d) for 4 h.
Materials 18 03203 g004
Figure 5. SEM with BSE imaging and EDS mapping of the particles in red squares with element contents of AS-MoSi2 (a,b), NS-MoSi2 (c,d), and BS-MoSi2 (e,f).
Figure 5. SEM with BSE imaging and EDS mapping of the particles in red squares with element contents of AS-MoSi2 (a,b), NS-MoSi2 (c,d), and BS-MoSi2 (e,f).
Materials 18 03203 g005
Figure 6. Nitrogen adsorption-desorption isotherms (a) and pore size distributions (b) of MoSi2, AS-MoSi2, NS-MoSi2, and BS-MoSi2.
Figure 6. Nitrogen adsorption-desorption isotherms (a) and pore size distributions (b) of MoSi2, AS-MoSi2, NS-MoSi2, and BS-MoSi2.
Materials 18 03203 g006
Figure 7. Schematic diagram of film formation mechanism of AS-MoSi2, NS-MoSi2 and BS-MoSi2.
Figure 7. Schematic diagram of film formation mechanism of AS-MoSi2, NS-MoSi2 and BS-MoSi2.
Materials 18 03203 g007
Table 1. Parameters of the silica sol.
Table 1. Parameters of the silica sol.
TypepHSiO2 Content (%)Na2O Content (%)Density
(20 °C, g·cm−3)
Viscosity
(20 °C, mPas)
Acid silica sol2.92 ± 0.0130 ± 0.1≤0.061.17 ± 0.01≤7
Neutral silica sol7.31 ± 0.0132 ± 0.1≤0.11.18 ± 0.01≤7
Alkaline silica sol9.49 ± 0.0132 ± 0.1≤0.31.20 ± 0.01≤10
Table 2. Parameters of MoSi2, AS-MoSi2, NS-MoSi2, and BS-MoSi2.
Table 2. Parameters of MoSi2, AS-MoSi2, NS-MoSi2, and BS-MoSi2.
TypeMoSi2AS-MoSi2NS-MoSi2BS-MoSi2
BET surface area (m2/g)6.1566.6149.1748.75
Total pore volume of adsorption (cm3/g)0.0210.1020.1600.126
Pore size (nm)5.722.866.114.72
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

Guo, L.; Zhang, J.; Miao, C.; Feng, S.; Fan, X.; Du, H.; Liu, J.; Wang, M. Effect of Silica Sol on the Preparation and Oxidation Resistance of MoSi2@SiO2. Materials 2025, 18, 3203. https://doi.org/10.3390/ma18133203

AMA Style

Guo L, Zhang J, Miao C, Feng S, Fan X, Du H, Liu J, Wang M. Effect of Silica Sol on the Preparation and Oxidation Resistance of MoSi2@SiO2. Materials. 2025; 18(13):3203. https://doi.org/10.3390/ma18133203

Chicago/Turabian Style

Guo, Linlin, Jinjun Zhang, Chengpeng Miao, Shuang Feng, Xiaozhen Fan, Haiyan Du, Jiachen Liu, and Mingchao Wang. 2025. "Effect of Silica Sol on the Preparation and Oxidation Resistance of MoSi2@SiO2" Materials 18, no. 13: 3203. https://doi.org/10.3390/ma18133203

APA Style

Guo, L., Zhang, J., Miao, C., Feng, S., Fan, X., Du, H., Liu, J., & Wang, M. (2025). Effect of Silica Sol on the Preparation and Oxidation Resistance of MoSi2@SiO2. Materials, 18(13), 3203. https://doi.org/10.3390/ma18133203

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