Next Article in Journal
Effect of Sintering Temperature and Time on Microstructure and Mechanical Properties of CoCrFeNiMn High-Entropy Alloys
Previous Article in Journal
Investigating the Tensile Properties of 22MnB5 After Austenitization and Quenching with Different Initial Microstructures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Investigation and Molecular Dynamics Modeling of the Effects of K2O on the Structure and Viscosity of SiO2-CaO-Al2O3-MgO-K2O Slags at High Temperatures

1
State Key Laboratory of Advanced Metallurgy, University of Science and Technology Beijing, Beijing 100083, China
2
Energy and Building Environment Engineering Institute, Henan University of Urban Construction, Pingdingshan 467041, China
*
Author to whom correspondence should be addressed.
Metals 2025, 15(6), 590; https://doi.org/10.3390/met15060590
Submission received: 23 March 2025 / Revised: 22 May 2025 / Accepted: 23 May 2025 / Published: 25 May 2025

Abstract

Variations in slag properties critically influence smelting operations and product quality. The effects of K2O on the CaO-SiO2-MgO-Al2O3-K2O slag system at 1823 K were systematically analyzed through an integrated approach combining viscosity measurements, FTIR spectroscopy, and molecular dynamics simulations. The results revealed a rapid 52% decrease in slag viscosity and an 18.32 kJ/mol reduction in activation energy as K2O content increased from 0% to 3%. K2O releases O2− ions that depolymerize Si-O network structures. Within the 3% to 5% range, structural network formation is promoted by the K2O-SiO2 reaction, resulting in increased slag viscosity and elevated activation energy. Molecular dynamics simulations elucidate the depolymerization of complex Si-O networks, accompanied by a proliferation of smaller [AlO4] tetrahedral fragments. The diminished Si-O-Si bridging oxygen (BO) bonds contrast with the enhanced increase in Si-O-K non-bridging oxygen (NBO) linkages. When K2O exceeds 3%, the diffusion capacity of K atoms becomes constrained as K2O participates in structural network assembly, a phenomenon validated by FTIR spectroscopic analysis. Elevated K2O concentrations enhance slag network polymerization, leading to increased viscosity. Therefore, the precise control of K2O content is critical during smelting operations and by-product manufacturing (e.g., glass or mineral wool) to optimize material performance. These findings provide theoretical support for controlling the alkali metal content during the actual metallurgical process and thus further optimizing blast furnace operation.

1. Introduction

Slag is a critical component in ironmaking composed primarily of ore gangue, fuel ash, flux, and reaction products. Its major chemical constituents include SiO2 (35–45%), CaO (33–45%), Al2O3 (8–17%), MgO (5–10%), and alkali metals (K2O, Na2O). Although typically present in trace amounts, alkali metals accumulate under high-temperature conditions in blast furnaces, converters, and gasifiers, altering slag viscosity, reactivity, and structure [1]. Slag also functions as a secondary resource for manufacturing mineral wool, glass, cement, and floor tiles, driving substantial research interest in slag systems [2]. Slag consists predominantly of silicates, with compositional variations exerting pronounced effects on both smelting operations and glass/cement production, particularly in the presence of Na2O and K2O. The effects of Na2O presence on slag, mineral wool, glass, and other molten systems have been systematically established, with physicochemical property modification enhancing smelting efficiency and product quality. In mineral wool and glass production systems, Na2O functions as both a fluxing agent and modifier, demonstrating measurable enhancements in material performance and industrial applicability [3,4,5,6,7,8,9]. The presence of K2O in silicate systems remains controversial. Some studies suggest that K2O reduces viscosity and fluidity by forming low-melting phases, whereas others argue it increases viscosity and the slag network polymerization degree, thereby enhancing system stability [10,11]. These variations are not only related to changes in the composition of the molten slag system but also associated with K2O content and temperature [12,13,14,15]. Current investigations on K2O in silicate systems remain predominantly conducted as either experimental investigations or computational modeling approaches, with notably limited studies implementing combined methodologies with cross-verification mechanisms [16,17,18]. Therefore, integrating modeling and experimental approaches to investigate K2O effects on molten slag systems enables the microstructural elucidation of network modifications while providing experimentally validated findings with enhanced technical credibility and industrial applicability.
Microstructural transformations directly govern macroscopic property alterations in slag systems. Viscosity, as a critical parameter of molten slag, exhibits acute sensitivity to compositional and structural variations, as such constituting a primary research focus. K2O demonstrates multifaceted influence on viscosity and flow characteristics within silicate-based slag systems. Higo et al. revealed that silicate slag viscosity increases with K2O content at K2O/Al2O3 molar ratios below 0.7, attributed to K2O-induced [AlO4] tetrahedral formation and elevated oxygen ion concentrations [19]. Conversely, viscosity decreases with further K2O additions beyond this threshold due to network structure depolymerization by excess K2O, enhancing flowability. Kim et al. demonstrated that at elevated temperatures (e.g., 1773 K), K2O additions increase silicate slag viscosity, whereas Na2O typically induces viscosity reduction, exhibiting contrasting effects [20]. Liu et al. reported that Na2O exhibits a more pronounced effect in reducing the viscosity of silicate slags compared to K2O, with further reduction in slag viscosity occurring upon the substitution of K2O with Na2O [21]. He et al. proposed that SiO2 reacts with Na2O to reduce the activity of potassium and calcium, thereby increasing slag viscosity—an effect particularly evident in slags with high silicon content [22]. Xing et al. established that precise K2O regulation is critical for optimizing slag fluidity and viscosity. Reducing K2O content lowers slag viscosity, thereby improving operational efficiency [23]. K2O supplies free oxygen ions, which disrupt Si-O-Si bonds in silicate networks, effectively reducing flow activation energy. As an alkaline oxide, K2O disrupts silicate network structures in slag systems. Within Mg-Al-Si-Ca containing slags, K2O reacts with SiO2 to generate non-bridging oxygen (NBO), reducing the polymerization degree through structural loosening, consequently decreasing viscosity [15,24,25,26,27]. Adjusting K2O content can effectively regulate slag fluidity and reactivity, enhancing smelting efficiency and product quality. K2O exerts dual effects on the viscosity and fluidity of silicate slag systems, with the specific outcomes depending on its concentration, temperature, and interactions with other components. The composition of coke, a key raw material in the smelting process, influences slag composition through the direct input of ash oxides. Coke ash (typically 10–20 wt.%) is rich in acidic oxides (SiO2, Al2O3) and minor basic oxides (CaO, MgO, K2O/Na2O), serving as a primary slag source. Ash from low-rank coal or low-quality coke may contain 1–5 wt.% K2O/Na2O. The entry of these alkali metals into the slag may affect its composition and properties [28]. These ash-derived components collectively impact slag composition, influencing properties such as viscosity and reactivity. This study aims to investigate the single-variable effect of K2O (0–5 wt.%) on slag by controlling the use of external high-alkali metal coke in the smelting process.
In metallurgy, metallurgical slags play a critical role in ironmaking and steelmaking processes, including desulfurization, dephosphorization, the protection of metal melts, and the regulation of heat and mass transfer. These functions are closely linked to slag properties such as viscosity and fluidity. Additionally, slag fluidity influences blast furnace operation and refractory erosion. K₂O accumulates in blast furnaces, affecting slag properties, thus underscoring the necessity of studying K₂O’s effects on slag properties [1]. This study employs rotational viscometry to investigate the viscosity of CaO-SiO2-MgO-Al2O3-K2O slag systems with a controlled CaO/SiO2 mass ratio of 1.32. Molecular dynamics simulations employing SCIGRESS software (Version Number: FJ 2.8.1) systematically characterized the network structures, polymerization degrees, atomic diffusivity, and bridging/non-bridging oxygen (BO/NBO) evolution within the slag system with controlled thermal gradients. FTIR spectroscopy was employed to perform qualitative analysis of the structural evolution in quenched slag at 1823 K, aiming to investigate the effects of K2O on Si and Al structural units. This study not only validated the consistency between the simulation and experimental results, thereby enhancing persuasiveness, but also deepened the understanding of K2O’s effects on slag viscosity and structure. It provides theoretical and technical support for applications of K2O-containing slag systems. This study integrates interdisciplinary methods to overcome the limitations of single approaches in understanding the complex mechanisms of K2O action, providing direct technical parameters for metallurgical slag process optimization (e.g., changes in slag viscosity due to changes in K2O content) and new analytical dimensions for fundamental research on multicomponent silicate systems. It advances precision regulation research linking microstructural features to macroscale properties to new heights.

2. Experimental Part

2.1. Slag Sample Preparation

The selected slag system for the experiment was CaO-SiO2-MgO-Al2O3-K2O. Slag samples were prepared using analytical-grade reagents (purity ≥ 99 wt%) of CaO, SiO2, Al2O3, MgO, and K2O at a fixed basicity CaO/SiO2 = 1.32 (Shanghai Macklin Biochemical Technology Co., Ltd., Shanghai, China). The compositional variations are summarized in Table 1. This study employs a binary basicity definition, calculated as the mass fraction ratio of the basic oxide (CaO) to the acidic oxide (SiO2): R = w(CaO)/w(SiO2). Although the slag system contains MgO (basic oxide) and Al2O3 (amphoteric oxide), a simplified binary basicity was chosen to highlight the fundamental properties of the CaO-SiO2-based slag. MgO content was fixed at 10 wt.% as a background parameter (not included in the basicity formula), while Al2O3 (15 wt.%, fixed) served as a supplementary acidic component to control variables and focus on the structural effects of K2O. A 200 g sample was weighed and thoroughly ground/mixed in an agate mortar. An 80 g portion of the homogenized sample was placed into a graphite crucible and heated to 1823 K in a high-temperature furnace under argon gas protection. After 2 h of isothermal holding, the crucible was rapidly removed and quenched in cold water. The quenched samples were subjected to tests such as XRD and FTIR spectroscopy for the qualitative analysis of the slag structure. The remaining 120 g of the samples was used for viscosity tests.

2.2. Molecular Dynamics Simulation

Molecular dynamics simulation is commonly used for simulating the network structures of metallurgical slag and silicate systems, based on the Born–Mayer–Huggins (BMH) potential function as shown in Equation (1) [29,30,31,32,33,34]:
E = Aij*exp(−Bij*r) − Cij/r**6 − Dij/r**8
In Equation (1), E represents the interatomic pair potential between atomic types i and j at distance r. The first term denotes the repulsive interaction, while the second and third terms correspond to van der Waals forces, with potential parameters tabulated in Table 2. The simulation system was set up as a cubic box with three-dimensional periodic boundary conditions. A total of 6000 particles were included in the simulation, with the number of particles for each element calculated based on their molar ratios. Table 3 lists the atomic numbers, box edge length, and density of the simulation system. The calculations were performed using the SCIGRESS FJ 2.8.1 (Large-scale Atomic/Molecular Massively Parallel Simulator) package, with the BMH potential function implemented via user-defined pair coefficients (Table 2). The simulation employed the NVT ensemble with the Gear algorithm for trajectory integration [35]. The temperature profile over time is shown in Figure 1. It is a temperature control curve that varies with the time step. First, a random model was generated at 5000 K over 10,000 fs. Then, it was cooled to the target temperature of 1823 K in the next 10,000 fs. Finally, the slag structure at 1823 K was obtained after another 10,000 fs. The radial distribution function and coordination number were calculated using the last 1000 fs trajectory for analyzing the slag network structure. The simulation employed a time step of 1 fs for trajectory integration in the NVT ensemble, maintaining constant temperature using a Nosé–Hoover thermostat. Post-simulation analysis for radial distribution functions and coordination numbers utilized in-house Python (version number 2.7) scripts leveraging the MD Analysis library to process trajectory data from the final equilibrium state.

2.3. Viscosity Test

The slag viscosity was measured via the rotational method, with the experimental setup illustrated in Figure 2. First, 120 g of pre-melted slag was placed into crucible and positioned in the constant temperature zone of a resistance furnace. The sample was heated to 1823 K in an argon atmosphere and held for 30 min. A molybdenum rotor was then immersed into the slag, with its lower end positioned approximately 10 mm above the crucible bottom and suspended within the crucible. The graphite crucible dimensions were Φ30 × 75 mm (inner diameter × height), and the molybdenum rotor dimensions were Φ15 × 20 mm (outer diameter × height). Viscosity measurements commenced once stable temperature and viscosity values were achieved. Three replicate measurements were taken per test group and averaged. Prior to each measurement, calibration was performed using standard castor oil.

2.4. Slag Structural Analysis

XRD measurements were performed using a Bruker D8 Advance diffractometer (Bruker, Billerica, MA, USA) with Cu Kα radiation (40 kV, 40 mA), a scanning range of 10–90° (2θ), and a step size of 0.02°. XRD was used to verify the amorphous state of all samples. Analysis using the PDF-5+ database and high-precision algorithms in JADE (version number 6.5) software ensured result reliability. FTIR spectroscopy was applied to analyze the structural characteristics of molten slags by identifying the vibration modes of functional groups (e.g., Si-O and Al-O bonds), enabling insights into the degree of polymerization and network connectivity in slag systems. FTIR spectroscopy was employed to analyze the structural evolution of the slag and corroborate the results of molecular dynamics simulations. Infrared spectroscopy (model: Thermo Fisher iS50, Waltham, MA, USA) was performed by homogenously grinding 2.0 mg of sample with 200 mg KBr in an agate mortar, followed by compressing the mixture into flakes. A wavenumber range of 400–4000 cm−1 was selected (with analysis focused on 400–1200 cm−1) with the aim of qualitatively analyzing structural changes in [SiO4] and [AlO4] groups [36].

3. Results and Discussion

3.1. XRD Analysis

Figure 3 shows the XRD patterns of the samples. Broad diffuse peaks (2θ ≈ 20°–35°) are a characteristic feature of the “short-range order, long-range disorder” structure in silicates, directly confirming that the bulk of the sample is an amorphous phase with minor crystalline phases of MgAl2O4 and SiC. Minor amounts of carbon (C) may be present due to the use of a graphite crucible during testing.

3.2. FTIR Spectroscopic Analysis

Figure 4 presents the FTIR spectra of samples with varying K2O contents. This technique enables the qualitative analysis of structural modifications in [SiO4] and [AlO4] tetrahedral units, as well as alterations in Si-O-Al bridging. The FTIR spectra analysis in Figure 4 indicates three distinct regions: 400–600 cm−1 [37] corresponding to Si-O-Al bending vibrations, 630–750 cm−1 [38] associated with asymmetric stretching vibrations of [AlO4], and 750–1200 cm−1 reflecting symmetric band variations of [SiO4] [39]. The spectral region between 750 and 1200 cm−1 can be further resolved into four silicon-associated structural units: Q0 (monomeric units), Q1 (dimeric structures), Q2 (chain-like configurations), and Q3 (layered network architectures) [36,40]. As evident from Figure 4, the characteristic bands associated with [SiO4] tetrahedra and the peak height h of Si-O-Al linkages exhibit a marked reduction as the K2O content increases from 1% to 3%. This trend signifies that free oxygen dissociates silicon-centered network structures, simplifying molecular complexity and lowering the polymerization degree of the slag system. Figure 4 reveals that at 5% K2O content, the observed spectral changes reverse previous trends, indicating that excessive K2O enhances the degree of polymerization. This leads to silicate structural complexity and strengthens linkages between [SiO4] and [AlO4] tetrahedra. Additionally, the [AlO4] tetrahedral peaks become slightly sharper with the increase in K2O content, indicating the enhanced structural ordering of these units. This observation aligns with the simulated reduction in Al-O coordination numbers, suggesting a proliferation of Al-associated molecular clusters within the system. However, compared to Si-O bonds, Al-O bonds exhibit greater bond lengths and reduced stability, rendering them more susceptible to O2−-induced depolymerization. This mechanism ultimately contributes to the observed viscosity reduction. In addition, when the K2O content exceeds 3%, the curves of Q0, Q1, and Q3 change from being flat to slightly convex [38]. This indicates that the number of chainlike and sheet-like structures increases, and the degree of polymerization rises. As a result, the viscosity increases when the K2O content exceeds 3%.

3.3. Viscosity Change

Figure 5 shows the viscosity and activation energy changes caused by varying K2O content at 1823 K, with experimental data and modeled viscosity values displaying aligned trends. The simulated viscosity values represent shear viscosities calculated using the SCIGRESS molecular dynamics software. To calculate the activation energy, shear viscosities at 1853 K, 1823 K, and 1793 K were first computed using kinetic software (SCIGRESS FJ 2.8.1), and the activation energy was then determined via the Arrhenius equation. Namely, for each condition, shear viscosities at 1853 K, 1823 K, and 1793 K temperatures were calculated, and then one activation energy was derived using the Arrhenius equation, resulting in five activation energy values for five conditions. The Arrhenius-type equation (Equation (2)) is transformed into a linear form by taking the natural logarithm: lnη = lnη0 + (Eη/R)·(1/T), where lnη is the ordinate and 1/T is the abscissa [2]. By performing linear fitting, the slope k = Eη/R is obtained, and thus the activation energy Eη = k·R is calculated. The activation energy (Eη) for viscous flow in slag is directly related to its viscosity: a higher Eη indicates greater energy required for atomic/molecular movement, leading to higher resistance to flow (i.e., higher viscosity). The activation energy was calculated to characterize the kinetic barriers governing key processes in the slag system, such as viscous flow or ionic diffusion, which are critical for understanding temperature-dependent properties like viscosity. These data provide mechanistic insights into how structural modifications (e.g., network depolymerization/polymerization induced by K2O) influence atomic mobility, directly linking microstructural changes to macroscale thermophysical behavior.
η = η 0 e x p ( E η R T )
where η0 is the exponential factor; Eη denotes the activation energy for viscous flow; R represents the universal gas constant; and T is the absolute temperature.
Figure 5 results show that in the CaO-SiO2-MgO-Al2O3-K2O slag system, K2O at low concentrations (≤3%) significantly reduces slag viscosity by 52% and activation energy by 18.32 kJ/mol, indicating that K2O acts as a network modifier causing depolymerization of [SiO4] and [AlO4] tetrahedral structures. Additionally, the provision of O2− ions by K2O induces depolymerization of macromolecular structures, increasing the concentration of smaller molecular species and consequently reducing viscosity. K+ ions exhibit lower ionic potential and higher mobility compared to Ca2+ ions, enhancing O2− activity, which loosens the structural network, reduces the overall slag system dissociation energy, and lowers activation energy. When K2O exceeds 3%, the slag system exhibits increased viscosity and elevated activation energy. This phenomenon may be attributed to the K2O-SiO2 reaction [16], which generates high-melting-point phases that abruptly enhance the melting point. The resulting structural complexity drives concurrent rises in both viscosity and activation energy. The simulated values being higher than experimental ones in Figure 5 may stem from multiple factors. Simulations often rely on idealized models that might disregard impurities present in experiments that can reduce viscosity. Additionally, theoretical models may oversimplify molecular interactions or structural complexities (e.g., the actual dynamic behavior of the silicate network in slag), overestimating viscosity. We also measured the softening (Ts), melting (Tm), and flow temperatures (Tp) of the samples using an ash fusion tester, as shown in Figure 6. We observed that these temperatures decreased with the increase in K2O content but increased slightly when K2O reached 5 wt.%. As a network modifier, K2O reduces slag viscosity and melting difficulty by breaking siloxane bonds (disrupting [SiO4] tetrahedron polymerization), thus causing Ts, Tm, and Tp to decline with the rise in K2O—consistent with trends from viscosity and activation energy experiments, which validates the experimental accuracy.

3.4. Radial Distribution Functions (RDFs) and Coordination Numbers (CNs)

The radial distribution function (RDF) is typically employed to analyze structural evolution in slag systems, including ordering, bond lengths, and the polymerization degree. It establishes critical correlations between microscopic atomic arrangements and macroscopic properties. When the particle count within a radial shell from r to r + δr is defined as n(r), the RDF g(r) can be mathematically expressed by Equation (3) [41].
g ( r ) = 1 ρ 0 n ( r ) V = 1 ρ 0 n ( r ) 4 π r 2 δ r
Figure 7 presents the variations in radial distribution functions (RDFs) and coordination numbers (CNs) for Si-O, Al-O, Mg-O, Ca-O, and K-O pairs in slag systems with and without K2O. The average bond lengths of these atomic pairs correspond to first peak positions at 1.6 Å, 1.75 Å, 1.95 Å, 2.35 Å, and 2.65 Å, respectively. These values remain invariant across K2O concentrations, consistent with prior literature findings [2]. The K-O bonds exhibit the longest bond length with relatively weak bond strength, predisposing them to ionic character formation at elevated temperatures. This consequently disaggregates the macromolecular network of the slag structure into smaller molecular clusters. Longer bond lengths correlate with sharper radial distribution peaks, indicating enhanced atomic pair stability. Consequently, the bond stability hierarchy is as follows: Si-O > Al-O > Mg-O > Ca-O > K-O. Furthermore, the coordination number curves in Figure 7b,d reveal stable plateaus at 4.0 for Si-O and 4.2 for Al-O, demonstrating the predominance of tetrahedral structural units centered on silicon and aluminum within the slag matrix. The coordination curves of Mg-O, Ca-O, and K-O do not have stable plateaus. Therefore, they do not participate in the formation of the slag network and serve to provide charge balance. A wider plateau indicates a more stable structure. These findings are consistent with previous studies [27].
Figure 8 depicts the variations in radial distribution functions (RDFs) and coordination numbers (CNs) for Si-O, Al-O, Si-Si, and Si-K atomic pairs. As shown in Figure 8a, the first peak intensity of Si-O bonds decreases with the increase in K2O content, while the coordination number N rises from 3.95 to 4.05 (Figure 8b). This trend indicates that the silicon-associated network transitions from a compact to a more open configuration, with a reduced degree of polymerization. The stable [SiO4] tetrahedra undergo progressive depolymerization as the K2O concentration increases. When the K2O content reaches 5%, the Si-O peak intensity exhibits a subsequent increase. This suggests that structural reorganization in K+-Si interactions enhances Si-O bond stability, consistent with the viscosity experimental results. In contrast to the Si-O structure, the Al-O peak intensity also decreases (Figure 8c), while its coordination number (CN) diminishes (Figure 8d). This phenomenon indicates that cations generated during K2O-induced network depolymerization provide charge compensation to [AlO4] tetrahedra, causing the dissociation of larger Al-associated molecular aggregates into smaller, stabilized [AlO4] tetrahedral units. However, Al-O bonds exhibit greater bond lengths and lower stability compared to Si-O bonds, ultimately reducing slag system viscosity and activation energy. Furthermore, when K2O exceeds 3%, the coordination number (CN) of Al-O stabilizes, indicating that excessive K2O concentrations can trigger the reaggregation of free ionic species into high-melting-point phases. This structural reorganization compromises flow properties by reducing melt fluidity. Figure 8e,f demonstrate a concurrent reduction in the Si-Si peak intensity and coordination number with the decrease in K2O content. This trend correlates with the diminished bridging oxygen (BO) content within Si-O-Si linkages, further evidencing the declining polymerization degree of silicon-associated network structures. Figure 8g,h reveal the increased Si-K peak intensity and coordination number, indicating a rise in non-bridging oxygen (NBO) within Si-O-K configurations. This structural evolution drives the disaggregation of lamellar network architectures into smaller molecular chains or monomeric units. Notably, even during subsequent reconfiguration into larger aggregates, K2O remains integral to the polymerization process.

3.5. Mean Square Displacements (MSDs) and [SiO4] Tetrahedral Population Variations

Figure 9 compares the diffusion capacities of K, Ca, Mg, Si, and Al. Within the same system in 3% K2O (Figure 9a), K exhibits the highest mobility, followed by Ca and Mg, while Si and Al demonstrate negligible displacement due to structural confinement within the network. This hierarchy confirms that K atoms predominantly exist as ionic species, providing charge compensation to the system, while the accompanying O2− ions facilitate network depolymerization. Figure 9b reveals that the diffusion capacity of K initially increases but subsequently declines with the increase in K2O content, exhibiting constrained mobility. The data indicate maximal K diffusion at 3% K2O. At 5% K2O, the aggregation of molecular clusters likely occurs, where K participates in structural network formation as a network-forming species. This integration restricts K atomic diffusion and displacement. Macroscopically, these structural modifications manifest as increased viscosity, altered flow properties, and the generation of high-melting-point phases.
In Figure 10, the variation in the number of Si-centered [SiO4] tetrahedra, statistically analyzed using MATLAB (version number R2016a), reveals a decrease from 592 to 560 followed by an increase to 572. These results demonstrate that K2O initially depolymerizes the Si-O network, but beyond 3% content, K2O participates in network formation, leading to a rebound in the tetrahedral population. This finding corroborates the previous simulation results, validating the consistency of the research outcomes.

4. Conclusions

The integrated analysis of the experimental and computational results demonstrates the dual role of K2O in the specific CaO-SiO2-MgO-Al2O3-K2O slag system investigated. K2O simultaneously acts as a network modifier and charge compensator, exhibiting concentration-dependent structural modulation behavior. During the initial stages of K2O addition, structural depolymerization is promoted, reducing network connectivity. This process increases the population of low-molecular-weight species, driving a significant viscosity reduction. However, beyond a critical threshold (3% in this study), K2O actively participates in the slag network architecture, inducing structural complexity through increased chain-like and layered configurations. This elevates the polymerization degree and viscosity, as evidenced by radial distribution function analyses and coordination number variations. The analysis of atomic mobility reveals constrained K diffusion in 5% K2O systems. Thus, the proper control of the K2O concentration changes plays a crucial role in regulating slag structure. During slag operations, the K2O concentration can be maintained within specified thresholds to minimize its impact on the slag structure. From an application perspective, while K2O exhibits its dual role in the specific CaO-SiO2-MgO-Al2O3-K2O slag system investigated here, analogous slag systems suggest this work offers guidance for precise K2O concentration control in industries reliant on slag viscosity regulation. In metallurgical processes, controlling K₂O content (e.g., 2–4%) to regulate slag viscosity and fluidity can optimize blast furnace operations. Leveraging K₂O’s dual roles (network modification and charge compensation) modulates slag–metal interfacial reaction kinetics. In addition, higher K2O levels (4–5%) can stabilize slag films and prevent nozzle clogging. In the manufacturing of specialty glasses or glazes, they enable precise control over melt flow during forming and cooling. In emerging applications like molten salt batteries and thermal energy storage, 2–4% K2O-containing slags may enhance heat transfer efficiency. In SiO2-Al2O3-based slag systems, K2O typically demonstrates consistent structural effects: high concentrations may induce structural crosslinking with SiO2, influencing overall slag properties. By establishing the K2O concentration–viscosity–structure relationship, this study provides a scientific basis for the multipurpose utilization of slags. Furthermore, this study provides actionable insights for optimizing metallurgical processes through the targeted control of alkali oxides by establishing a structure–property–metallurgical application relationship.

Author Contributions

Conceptualization, Q.X. and H.Z.; Methodology, F.Y., Y.L. and J.W.; Software, F.Y. and Y.L.; Validation, F.Y. and Y.L.; Formal analysis, H.Z. and J.W.; Resources, Q.X.; Data curation, F.Y.; Writing—original draft, F.Y.; Writing—review and editing, Q.X. and J.W.; Supervision, H.Z. and J.W.; Project administration, H.Z.; Funding acquisition, Q.X. and J.W. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support of the National Natural Science Foundation of China (No. U1960205). Supported by China Baowu Low Carbon Metallurgical Innovation Foundation (BWLCF202206), China Minmetals Science and Technology Special Plan Foundation 2020ZXA01, Overseas Expertise Collaboration Base for Green and Intelligent Metallurgy, No: B21004.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Partyka, J.; Sitarz, M.; Leśniak, M.; Gasek, K.; Jeleń, P. The effect of SiO2/Al2O3 ratio on the structure and microstructure of the glazes from SiO2-Al2O3-CaO-MgO-Na2O-K2O system. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 134, 621–630. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, Z.; Meng, Z.; Liu, L.; Wang, H.; Sun, Y.; Wang, X. Structural and Viscous Insight into Impact of MoO3 on Molten Slags. Metall. Mater. Trans. B 2021, 52, 3730–3743. [Google Scholar] [CrossRef]
  3. Kim, H.; Kim, W.H.; Park, J.H.; Min, D.J. A Study on the Effect of Na₂O on the Viscosity for Ironmaking Slags. Steel Res. Int. 2010, 81, 17–24. [Google Scholar] [CrossRef]
  4. Blenau, L.W.; Sander, S.A.; Fuhrmann, S.; Charitos, A. Holistic valorization of fayalitic slag to pig iron and glass fibers. J. Clean. Prod. 2023, 396, 136575. [Google Scholar] [CrossRef]
  5. Li, Y.; Dai, W.-B. Modifying hot slag and converting it into value-added materials: A review. J. Clean. Prod. 2018, 175, 176–189. [Google Scholar] [CrossRef]
  6. Tesovnik, A.; Horvat, B. Rapid Immobilisation of Chemical Reactions in Alkali-Activated Materials Using Solely Microwave Irradiation. Minerals 2024, 14, 1219. [Google Scholar] [CrossRef]
  7. Elzeadani, M.; Bompa, D.V.; Elghazouli, A.Y. One part alkali activated materials: A state-of-the-art review. J. Build. Eng. 2022, 57, 104871. [Google Scholar] [CrossRef]
  8. Kupczewska-Dobecka, M.; Konieczko, K.; Czerczak, S. Occupational risk resulting from exposure to mineral wool when installing insulation in buildings. Int. J. Occup. Med. Environ. Health 2020, 33, 757–769. [Google Scholar] [CrossRef]
  9. Rawlings, R.D.; Wu, J.P.; Boccaccini, A.R. Glass-ceramics: Their production from wastes—A Review. J. Mater. Sci. 2006, 41, 729–744. [Google Scholar] [CrossRef]
  10. Jin, Z.N.; Lv, J.F.; Yang, H.Y. Effect of K₂O on Phase Relation and Viscosity of the CaO–SiO2—ZnO—FeO–Al2O3 System Slags. Russ. J. Non-Ferr. Met. 2020, 61, 153–161. [Google Scholar]
  11. Chang, Z.Y.; Jiao, K.X.; Ning, X.J.; Zhang, J.L. Novel Approach to Studying Influences of Na2O and K2O Additions on Viscosity and Thermodynamic Properties of BF Slags. Metall. Mater. Trans. B 2019, 50, 1399–1406. [Google Scholar] [CrossRef]
  12. Perez-Pariente, J.; Martens, J.A.; Jacobs, P.A. Crystallization mechanism of zeolite beta from (TEA)2O, Na2O and K2O containing aluminosilicate gels. Appl. Catal. 1987, 31, 35–64. [Google Scholar] [CrossRef]
  13. Michalski, J.R.; Kraft, M.D.; Sharp, T.G.; Williams, L.B. Mineralogical constraints on the high-silica martian surface component observed by TES. Icarus 2005, 174, 161–177. [Google Scholar] [CrossRef]
  14. Klima, K.M.; Luo, Y.; Brouwers, H.J.H.; Yu, Q. Effects of mineral wool waste in alkali activated-artificial aggregates for high-temperature applications. Constr. Build. Mater. 2023, 401, 132937. [Google Scholar] [CrossRef]
  15. Zhang, Y.; Zhang, S.; Chen, Y.; Çopuroğlu, O. The effect of slag chemistry on the reactivity of synthetic and commercial slags. Constr. Build. Mater. 2022, 335, 127493. [Google Scholar] [CrossRef]
  16. Koroleva, O.N.; Shtenberg, M.V.; Osipov, A.A. Structural Features of K2O-SiO2 Melts: Modeling and High-Temperature Experiments. Minerals 2023, 13, 94. [Google Scholar] [CrossRef]
  17. Xuan, W.; Guhl, S.; Zhang, Y.; Zhang, J.; Meyer, B. A Structural Viscosity Model for Silicate Slag Melts Based on MD Simulation. Metals 2022, 12, 1123. [Google Scholar] [CrossRef]
  18. Li, H.; Zhang, Q.; Wang, G. Influence of MgO on Viscosity and Structure of CaO-SiO2-MgO-Al2O3 Slags. Metals 2021, 11, 789–796. [Google Scholar]
  19. Higo, T.; Sukenaga, S.; Kanehashi, K.; Shibata, H.; Osugi, T.; Saito, N.; Nakashima, K. Effect of Potassium Oxide Addition on Viscosity of Calcium Aluminosilicate Melts at 1673–1873 K. ISIJ Int. 2014, 54, 2039–2044. [Google Scholar] [CrossRef]
  20. Kim, W.H.; Sohn, I.; Min, D.J. A Study on the Viscous Behaviour with K2O Additions in the CaO-SiO2-Al2O3-MgO-K2O Quinary Slag System. Steel Res. Int. 2010, 81, 735–741. [Google Scholar] [CrossRef]
  21. Liu, W.; Zuo, H. Mixed influence of Na2O and K2O on the viscous properties of aluminosilicate melt. Ceram. Int. 2025, in press. [CrossRef]
  22. He, C.; Fan, F.F.; Guo, J.; Yuan, M.; Qin, Y.; Wei, Y.; Yan, J. Viscosity determination of the biomass slag in the SiO2-CaO-K2O system based on the bond distribution. Fuel 2024, 356, 129642. [Google Scholar] [CrossRef]
  23. Xing, X.; Pang, Z.; Zheng, J.; Du, Y.; Ren, S.; Ju, J. Effect of MgO and K₂O on High-Al Silicon–Manganese Alloy Slag Viscosity and Structure. Minerals 2020, 10, 810. [Google Scholar] [CrossRef]
  24. Wu, G.; Yazhenskikh, E.; Hack, K.; Wosch, E.; Müller, M. Viscosity model for oxide melts relevant to fuel slags. Part 1: Pure oxides and binary systems in the system SiO2–Al2O3–CaO–MgO–Na2O–K2O. Fuel Process. Technol. 2015, 137, 93–103. [Google Scholar] [CrossRef]
  25. Fang, J.; Pang, Z.; Xing, X.; Xu, R. Thermodynamic Properties, Viscosity, and Structure of CaO-SiO2-MgO-Al2O3-TiO2-Based Slag. Materials 2020, 14, 124. [Google Scholar] [CrossRef]
  26. Jak, E. Viscosity Model for Slags in the Al2O3-CaO-FeO-K2O-Na2O-MgO-SiO2 System. In Proceedings of the VIII International Conference on Molten Slags, Fluxes and Salts, Santiago, Chile, 18–21 January 2009; pp. 1–10. [Google Scholar]
  27. Wu, G.; Yazhenskikh, E.; Hack, K.; Müller, M. Viscosity model for oxide melts relevant to fuel slags. Part 2: The system SiO2–Al2O3–CaO–MgO–Na2O–K2O. Fuel Process. Technol. 2015, 138, 520–533. [Google Scholar] [CrossRef]
  28. Heikkilä, A.; Iljana, M.; Heikkinen, E.; Koskela, A.; Fabritius, T. Effect of Coal and Coke Ash on Blast Furnace Slag Properties: A Comparison Between Pulverized Coal, Charcoal, Fossil-Based Coke, and Biocoke. Steel Res. Int. 2022, 93, 2100188. [Google Scholar] [CrossRef]
  29. Mysen, B.O.; Virgo, D.; Harrison, W.J.; Scarfe, C.M. Solubility mechanisms of H₂O in silicate melts at high pressures and temperatures; a Raman spectroscopic study. Am. Mineral. 1980, 65, 900–914. [Google Scholar]
  30. Mysen, B.O.; Virgo, D.; Seifert, F.A. The structure of silicate melts: Implications for chemical and physical properties of natural magma. Rev. Geophys. 1982, 20, 353–383. [Google Scholar] [CrossRef]
  31. Belashchenko, D.; Ostrovski, O. Computer Simulation of Noncrystalline Ionic-Covalent Oxides in the SiO2-CaO-FeO System. Inorg. Mater. 2002, 38, 799–804. [Google Scholar] [CrossRef]
  32. Belashchenko, D.; Skvortsov, L. Molecular Dynamics Study of CaO-Al2O3 Melts. Inorg. Mater. 2001, 37, 476–481. [Google Scholar] [CrossRef]
  33. Bauchy, M. Structural, vibrational, and elastic properties of a calcium aluminosilicate glass from molecular dynamics simulations: The role of the potential. J. Chem. Phys. 2014, 141, 024507. [Google Scholar] [CrossRef] [PubMed]
  34. Bouhadja, M.; Jakse, N.; Pasturel, A. Stokes-Einstein violation and fragility in calcium aluminosilicate glass formers: A molecular dynamics study. Mol. Simul. 2014, 40, 251–259. [Google Scholar] [CrossRef]
  35. Sukenaga, S.; Saito, N.; Kawakami, K.; Nakashima, K. Viscosities of CaO-SiO2-Al2O3-(R2O or RO) Melts. ISIJ Int. 2006, 46, 352–358. [Google Scholar] [CrossRef]
  36. Liu, W.; Qin, J.; Xing, X.; Wang, J.; Zuo, H. Viscosity and Structure Evolution of CaO-SiO2-MgO-Al2O3-BaO Slag with the CaO/SiO2 Mass Ratio of 0.9. Ceram. Int. 2021, 47, 33337–33343. [Google Scholar] [CrossRef]
  37. Zhou, X.P.; Cheng, H.F.; Wang, Y. FTIR spectroscopic study of montmorillonite with different exchanged cations. J. Non-Cryst. Solids 2023, 598, 121543. [Google Scholar]
  38. Lodesani, F.; Menziani, M.C.; Hijiya, H.; Takato, Y.; Urata, S.; Pedone, A. Structural origins of the Mixed Alkali Effect in Alkali Aluminosilicate Glasses: Molecular Dynamics Study and its Assessment. Sci. Rep. 2020, 10, 2906. [Google Scholar] [CrossRef]
  39. Halász, I.; Agarwal, M.; Li, R.; Miller, N. What can vibrational spectroscopy tell about the structure of dissolved sodium silicates? Microporous Mesoporous Mater. 2010, 135, 74–81. [Google Scholar] [CrossRef]
  40. Dalby, K.; Topsøe, H.; King, P.L. A new approach to determine and quantify structural units in silicate glasses using micro Fourier-Transform Infrared spectroscopy. Am. Mineral. 2006, 91, 1783–1793. [Google Scholar] [CrossRef]
  41. Yang, F.; Han, Y.; Zhu, L. Analysis of basic network structure of continuous casting mould powder in electromagnetic field. Ironmak. Steelmak. 2019, 46, 827–834. [Google Scholar] [CrossRef]
Figure 1. Relationship between temperature and time step.
Figure 1. Relationship between temperature and time step.
Metals 15 00590 g001
Figure 2. Rotational viscometer setup and schematic diagram.
Figure 2. Rotational viscometer setup and schematic diagram.
Metals 15 00590 g002
Figure 3. XRD results of pre-melted slag.
Figure 3. XRD results of pre-melted slag.
Metals 15 00590 g003
Figure 4. FTIR spectra at different K2O contents.
Figure 4. FTIR spectra at different K2O contents.
Metals 15 00590 g004
Figure 5. Effect of K2O on viscosity and activation energy of CaO-SiO2-MgO-Al2O3-K2O slag system at 1823 K.
Figure 5. Effect of K2O on viscosity and activation energy of CaO-SiO2-MgO-Al2O3-K2O slag system at 1823 K.
Metals 15 00590 g005
Figure 6. Variations in ash fusion temperatures of samples.
Figure 6. Variations in ash fusion temperatures of samples.
Metals 15 00590 g006
Figure 7. The radial distribution functions (RDFs) and coordination numbers (CNs) for Si-O, Al-O, Mg-O, Ca-O, and K-O pairs in slag systems containing 0% and 1% K2O.
Figure 7. The radial distribution functions (RDFs) and coordination numbers (CNs) for Si-O, Al-O, Mg-O, Ca-O, and K-O pairs in slag systems containing 0% and 1% K2O.
Metals 15 00590 g007
Figure 8. Radial distribution functions and coordination numbers of different atomic pairs.
Figure 8. Radial distribution functions and coordination numbers of different atomic pairs.
Metals 15 00590 g008aMetals 15 00590 g008b
Figure 9. The atomic mean square displacements (MSDs) under different conditions. (a) Diffusion capacities of K, Ca, Mg, Si, and Al in slag containing 3% K2O; (b) K atomic diffusion behavior across systems with varying K2O concentrations.
Figure 9. The atomic mean square displacements (MSDs) under different conditions. (a) Diffusion capacities of K, Ca, Mg, Si, and Al in slag containing 3% K2O; (b) K atomic diffusion behavior across systems with varying K2O concentrations.
Metals 15 00590 g009
Figure 10. Variation in [SiO4] tetrahedra quantity with different K2O contents.
Figure 10. Variation in [SiO4] tetrahedra quantity with different K2O contents.
Metals 15 00590 g010
Table 1. Chemical composition design of slag (wt/%).
Table 1. Chemical composition design of slag (wt/%).
No.CaOSiO2MgOAl2O3K2OR
142.6332.29101501.32
241.7931.66101511.32
341.5331.46101521.32
440.6830.99101531.32
539.7930.14101551.32
Table 2. Numerical value of BMH potential function parameters.
Table 2. Numerical value of BMH potential function parameters.
IonicIonicAijBijCijDij
KK5.0665 × 10−206.251.5632 × 10−250
KCa1.6346 × 10−206.251.0421 × 10−250
CaCa5.2737 × 10−226.256.9467 × 10−260
MgK3.3626 × 10−216.252.0843 × 10−260
MgCa1.0849 × 10−216.251.3895 × 10−260
SiK1.3251 × 10−216.2500
MgMg2.2318 × 10−226.252.7791 × 10−270
SiCa4.2751 × 10−226.2500
AlK1.8339 × 10−216.2500
MgSi8.7947 × 10−236.2500
AlCa5.9169 × 10−206.2500
SiSi3.4656 × 10−236.2500
OK3.4457 × 10−206.062.0843 × 10−250
MgAl1.2172 × 10−226.2500
OCa1.1505 × 10−206.061.38950 × 10−250
AlSi4.7966 × 10−236.2500
OMg2.4829 × 10−216.062.7791 × 10−260
AlAl6.3869 × 10−236.2500
OSi1.0064 × 10−216.0600
OAl1.3792 × 10−216.0600
OO2.3993 × 10−205.882.7791 × 10−210
Table 3. Simulated particle number and box side length.
Table 3. Simulated particle number and box side length.
No.Particle CountBox Side Length/ADensity
SiAlCaKMgO
180246410740287337442.2913.052
2793465106134288336042.3533.047
3783465104967288334742.4163.041
47744661036101289333442.4843.035
57544671010172291330642.6143.024
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

Yang, F.; Xue, Q.; Zuo, H.; Liu, Y.; Wang, J. Experimental Investigation and Molecular Dynamics Modeling of the Effects of K2O on the Structure and Viscosity of SiO2-CaO-Al2O3-MgO-K2O Slags at High Temperatures. Metals 2025, 15, 590. https://doi.org/10.3390/met15060590

AMA Style

Yang F, Xue Q, Zuo H, Liu Y, Wang J. Experimental Investigation and Molecular Dynamics Modeling of the Effects of K2O on the Structure and Viscosity of SiO2-CaO-Al2O3-MgO-K2O Slags at High Temperatures. Metals. 2025; 15(6):590. https://doi.org/10.3390/met15060590

Chicago/Turabian Style

Yang, Fan, Qingguo Xue, Haibin Zuo, Yu Liu, and Jingsong Wang. 2025. "Experimental Investigation and Molecular Dynamics Modeling of the Effects of K2O on the Structure and Viscosity of SiO2-CaO-Al2O3-MgO-K2O Slags at High Temperatures" Metals 15, no. 6: 590. https://doi.org/10.3390/met15060590

APA Style

Yang, F., Xue, Q., Zuo, H., Liu, Y., & Wang, J. (2025). Experimental Investigation and Molecular Dynamics Modeling of the Effects of K2O on the Structure and Viscosity of SiO2-CaO-Al2O3-MgO-K2O Slags at High Temperatures. Metals, 15(6), 590. https://doi.org/10.3390/met15060590

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