Next Article in Journal
Influence of Si Content on Microstructure and Mechanical Properties of CoCrFeNi High-Entropy Alloys
Next Article in Special Issue
Kinetics Study of the Hydrogen Reduction of Limonite Ore Using an Unreacted Core Model for Flat-Plate Particles
Previous Article in Journal
A Review of Studies on the Influence of Rare-Earth Elements on the Microstructures and Properties of Copper and Copper Alloys and Relevant Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Stainable Utilization Strategies for Basic Oxygen Furnace Slag: Properties, Processing, and Future Directions

1
School of Resources and Safety Engineering, University of Science and Technology Beijing, Beijing 100083, China
2
Institute of Mineral Resources, University of Science and Technology Beijing, Beijing 100083, China
*
Authors to whom correspondence should be addressed.
Metals 2025, 15(5), 537; https://doi.org/10.3390/met15050537
Submission received: 2 April 2025 / Revised: 2 May 2025 / Accepted: 9 May 2025 / Published: 12 May 2025
(This article belongs to the Special Issue Recent Developments in Ironmaking)

Abstract

Steel slag, being the dominant solid byproduct in steelmaking, presents global challenges in sustainable management, particularly regarding resource recovery of Basic Oxygen Furnace (BOF) slag, which accounts for over 72% of total slag generation. Through the databases of ScienceDirect, Web of Science, and CNKI, using relevant key words, this review systematically investigates the physicochemical properties and mineralogical composition of BOF slag, elucidating the intrinsic mechanisms underlying its low hydration reactivity and volumetric instability. Pretreatment techniques have been demonstrated to effectively modulate these properties. Furthermore, valuable components can be efficiently recovered through methods including magnetic separation and related technologies. Furthermore, this review elucidates the mechanisms and existing challenges across various resource utilization approaches for steel slag, while also identifying key research priorities for future development, thereby providing a systematic theoretical framework and technical pathways to advance utilization of steel slag.

1. Introduction

Approximately 100–150 kg of steel slag, a predominant solid waste generated during crude steel production [1], is produced per ton of steel output. In 2024, China’s crude steel production reached 1.005 billion tons, accounting for 53.39% of global output, with concomitant steel slag generation amounting to 160 million tons [2].
Globally, 72% of crude steel is produced via BOF processes, while 28% derives from Electric Arc Furnace (EAF) routes. However, national technological preferences vary markedly: Germany adopts 70% BOF versus 30% EAF, Turkey 29% versus 71%, and the United States 30% versus 70%, whereas China predominantly relies on BOF (90%), with minimal EAF utilization (10%) [3].
Regarding utilization rates, developed economies maintain superior performance: Japan achieves 98.4%, Europe 87.0%, and the United States 84.4% [3]. In contrast, China’s comprehensive utilization rate remains at merely 29.5% [3], highlighting a significant gap. The extensive open-air stockpiling of steel slag not only occupies substantial land resources but also facilitates the migration of heavy metals (e.g., Fe, Mn) and alkaline substances through weathering processes, leading to soil salinization, groundwater contamination, and particulate matter pollution. As the global steel industry advances its “solid waste valorization” strategy [4], the predominance of BOF slag in China’s steel slag (accounting for 72%) has made its effective application a current focal point [3]. The utilization approaches of steel slag in various regions can be seen in Figure 1.
As a core solid waste in the steel industry, global-scale disposal and resource utilization of steel slag face significant challenges. This study systematically analyzes the physicochemical properties of steel slag, elucidates its chemical and mineral composition characteristics, and reveals the intrinsic mechanisms underlying its low hydration activity and volumetric instability. To address compositional complexity, pretreatment technologies such as thermal aging and rotary drum processing effectively regulate free calcium oxide content. Combined with magnetic separation, carbothermal reduction, acid leaching, and alkali leaching processes, these methods enable efficient recovery of valuable components, including iron, vanadium, cadmium, and phosphorus.
In resource applications, steel slag exhibits unique reactivity in cementitious materials through silicate tetrahedral reconstruction and complex salt effects. As concrete aggregates, it significantly enhances mechanical properties and optimizes pore structures, while asphalt composites incorporating steel slag demonstrate high-temperature stability and crack resistance. Emerging applications extend to high-value products like glass-ceramics and environmental remediation materials.
This review highlights current bottlenecks, such as limited hydration activity enhancement and long-term volume stability control. It identifies application incentives and challenges across resource utilization pathways, pinpointing critical research gaps in material development.

2. Fundamental Properties of BOF Slag

Steel slag, a predominant byproduct of steelmaking processes, can be classified into BOF, electric arc furnace (EAF), and open hearth furnace (OHF) slags based on production routes [10]. Globally, 72% of crude steel is produced via BF-BOF integrated processes [3], resulting in BOF slag being the most abundant variant.

2.1. Chemical Composition and Mineralogical Phases

The characteristics of BOF slag have a direct and important impact on its application. In general, the chemical composition of BOF slag is shown in Table 1:
BOF slag can be categorized into acidic slag (low basicity, CaO/SiO2 < 1.8) and basic slag (CaO/SiO2 ≥ 1.8), with the latter constituting over 85% of total production [12]. During basic oxygen steelmaking, the continuous addition of lime (CaO) elevates slag basicity, triggering characteristic phase transitions [13]:
2(CaO·RO·SiO2) + CaO = 3CaO·RO·2SiO2 + RO
3CaO·RO·2SiO2 + CaO = 2(2CaO·SiO2) + RO
2(2CaO·SiO2) + CaO = 3CaO·SiO2
The mineralogical phases demonstrate basicity-dependent evolution (Table 2):
The dicalcium silicate (C2S) phase in BOF slag undergoes characteristic polymorphic transitions during cooling (Figure 2). When the temperature decreases to 450–500 °C, β-C2S irreversibly transforms into γ-C2S via a reconstructive phase transition. This low-temperature phase exhibits negligible hydraulic activity (< 5% of β-C2S reactivity) due to its orthorhombic crystal structure (space group Pnma) compared to the monoclinic β-form (space group P2₁/n), accompanied by a 12% unit cell volume expansion. The densified structure reduces specific surface area to < 1 m2/g, effectively passivating the phase [12].

2.2. Physical and Mechanical Properties

As a “high-potential but challenging” industrial by-product, the physical properties of alkaline oxygen furnace (BOF) slag are critical for its future use. The main physical properties of the Basic Oxygen Furnace (BOF) slag can be seen in Table 3.

2.3. Reactivity and Volume Stability

The activity of steel slag is affected by alkalinity. Dicalcium silicate and tricalcium silicate in steel slag are active minerals with hydrohard gelatinity. When the alkalinity of steel slag is greater than 1.8, steel slag contains 60~80% dicalcium silicate and tricalcium silicate, and with the increase of alkalinity, the content of tricalcium silicate also increases; when the alkalinity is increased to more than 2.5, the main mineral of steel slag is tricalcium silicate. Grinding with steel slag with alkalinity higher than 2.5 and 10% gypsum, its strength can reach the strength of No. 325 cement.
The stability of steel slag is related to free calcium oxide (f-CaO), MgO, Ca2SiO4, and Ca3SiO5, which are unstable under certain conditions. When slag with high alkalinity is slowly cooled, Ca3SiO5 will slowly decompose into Ca2SiO4 and f-CaO at 1100~1250 °C. At 675 °C, Ca2SiO4 changes from Ca2SiO4 to γ- Ca2SiO4, and volume expansion occurs, with the expansion rate reaching 10%. In addition, after the steel slag absorbs water, f-CaO will be digested into calcium hydroxide (Ca(OH)2), the volume will expand by 100%~300%, MgO will become magnesium hydroxide, and the volume will expand by 77%. Therefore, steel slag containing f-CaO and MgO at room temperature is unstable, and steel slag will be stable only when the f-CaO and MgO are digested or the content is very small [12]. The reactions of the components of steel slag at the corresponding temperatures can be seen in Table 4.

2.4. Utilization Potential and Limitations

Due to some of its physical and chemical properties, there are problems with comprehensive utilization. Due to the following reasons, the recycling of steel slag is limited, resulting in low utilization rate of steel slag. Details are presented in Table 5 and Table 6.

3. Utilization Pathways of BOF Slag

Effective utilization of BOF slag requires pretreatment to stabilize metastable phases (e.g., free CaO, MgO) for volumetric stability compliance [29]. The standard protocol includes: pretreatment → valuable component recovery → resource application. Current non-carbonation utilization predominantly focuses on construction engineering (>75% market share), with emerging applications in environmental remediation and soil amendment.

3.1. Pretreatment Technologies

Pretreatment critically determines slag granulometry (D50 = 0.1–5 mm), phase composition, and activity index (AI = 65–85%). Table 7 evaluates mainstream technologies, with selection criteria dependent on slag state (molten/solid), throughput (10–500 t/h), and downstream applications [5,30,31].

3.2. Utilization of Valuable Components of BOF Slag

Steel slag contains a variety of useful metallic elements, with common chemical components including vanadium, iron, silicon, titanium, vanadium, chromium, calcium, and magnesium oxides [4,5]. Although vanadium and titanium are present in small quantities, their high value and limited resources can offset the expenditure incurred in their extraction [4,31].

3.2.1. Recovery of Iron

The main components of steel slag include FexO, Fe, CaO, SiO2, MgO, and MnO [32] The ferrous phase (RO phase, Ca2Fe2O5, Fe2O3, Fe3O4, Fe) of steel slag accounts for 20–40% of the mass of the original steel slag [33,34], which is recognized as an important iron resource. When utilizing steel slag, iron has a negligible contribution to its activity. Since iron has a high hardness, this phase is difficult to grind into fine particles [15], so it is necessary to recover the iron in the steel slag [35].
Magnetic separation is usually used to recover the iron phase. Steel slag treated by the air quenching method, the water quenching method, the granulation wheel method, or the drum method, due to the small particle size, is generally directly recovered by magnetic separation after dehydration and is no longer further crushed or only simply crushed; due to the large particle size, steel slag treated by the hot stuffing method or the hot splash method needs to be crushed, screened, and magnetically separated. In the early days, the secondary treatment process of steel slag was mainly simple crushing and screening, and now it has been upgraded to multi-stage crushing, multi-stage screening, and multi-stage magnetic separation. When using magnetic separation, in order to improve the enrichment and magnetism of the ferrous phase in the steel slag, hot slag modification (oxidative composition adjustment in the molten state) [36] is proposed, involving air oxidation [37] to promote ferromagnetic spinel formation, rather than thermal reduction. This method enhances magnetic susceptibility.
Carbon-thermal reduction relies on high-temperature reactions where carbon reduces metal oxides, with carbon monoxide as the primary gaseous product [38]. While traditional processes use conventional heating [38], microwave heating can enhance kinetics by exploiting dielectric loss of reactants [39]. Graphite powder is used as a reducing agent in the extraction of iron, and its dosage varies according to the carbon index. As the carbon equivalent increases, the fractional reduction of iron increases. Due to the increase in the activity coefficient of P2O5 and the fluidity of the slag, the higher SiO2 content induces a large amount of iron reduction in BOF slag. On the other hand, the reintroduction of this metal-carbon alloy can reduce the volume of graphite by incorporating graphite as a reducing agent to extract and reduce iron from Fe-C alloys, in which a small amount of Fe alloy remains but has a negative impact on recovery efficiency. In order to achieve technically and economically viable processing and long-term technological development, all these aspects must be addressed [40].
Acid leaching is another widely used technique, in which acids dissolve metals from oxides and sulfides. The main advantages of this technique are its simplicity and broad-spectrum applicability. Optimized HCl leaching of Malaysian steel showed an iron extraction rate of 77.14% [41]. Another technique used is chloride extraction/chlorination. These methods leave soluble iron trichloride, which can be used as a potential adsorbent for wastewater treatment [42]. The resulting chloride has a lower melting point than the original salt. This helps to recover metals more easily. Figure 3 below show flow charts showing different iron recovery techniques.
However, due to the high acid consumption, coupled with energy requirements, the costs associated with the process must be considered. The waste liquid produced after the leaching process is strongly acidic and therefore needs to be treated. Due to the strong acidity of the leachate, there is a possibility of corrosive material. A viable solution to ensure sustainability and economic efficiency could be the recycling of spent acid [41]. Furthermore, there are also technologies for steel slag gravity separation and flotation to recover iron; however, flotation produces a large amount of wastewater in practical applications, and the hypergravity method is still in its infancy [43]. Magnetic separation is the dominant process.

3.2.2. Recovery of Vanadium

In BOF slag, vanadium can be oxidized to V(IV) during leaching [44]. Due to the high value of and demand for vanadium, BOF slag in the form of V2O5 containing 2–3% vanadium can also be used as a feedstock for selective recovery of vanadium. Leaching is a method of directly extracting vanadium from steel slag, including the direct leaching method, the roasting leaching method, and other related leaching methods [45,46].
Direct leaching is usually carried out without regard to the roasting process and is mainly carried out using sulfuric acid [47,48]. Using titanium dioxide to produce waste acid as a leaching agent, the extraction rate of vanadium can reach 95%. There are still many limitations and shortcomings in actual production, such as raw material particle size, operating conditions, and the purity of the recovered vanadium. As a result, direct leaching is less widely used in most industrial productions [49].
Roasting leaching is also a commonly used method. Roasting can destroy the spinel structure in steel slag and effectively dissolve vanadium [50,51]. Sodium roasting technology oxidizes vanadium in slag to sodium vanadate by mixing sodium salt additives with steel slag and roasting at high temperature in a furnace [52,53]. In calcium roasting, calcium compounds are used as additives, mixed with vanadium-containing steel slag, and then ground into fine particles after roasting [54,55]. Clean baking reduces exhaust gases and waste water during the baking process. These processes mainly include magnesium salt roasting, ammonium roasting, manganese roasting and salt-free roasting [50]. Immediately after the slag is roasted, the leaching process for the extraction of vanadium is carried out. Water leaching or acid leaching is usually used, and some researchers employ methods such as alkaline leaching and ammonium leaching. By employing mechanical activation [56], microwave-assisted baking [57], and ultrasonic-assisted leaching [58], oxidants are used, as well as other new technologies, which can improve the efficiency of extracting vanadium from steel slag.

3.2.3. Recycling of Chromium

Chromium is added to steel to increase its antioxidant properties. The unreacted chromium ends up entering steel slag at 1–2 wt%, occasionally up to 10.5 wt%. The two general techniques for extracting chromium from BOF slag are alkali leaching and alkali roasting [59]. Prior to extraction, toxic Cr(VI) is converted to stable Cr(III). In alkaline leaching, a leaching agent such as sodium hypochlorite is used in combination with another alkaline reagent. The leaching chemical reaction of chromium includes the following series of chemical reactions.
Cr3+ + H2O ⇔ CrOH2+ + H+
CrOH2+ + H2O ⇔ Cr(OH)2+ + H+
Cr(OH)2+ + H2O ⇔ Cr(OH)30 + H+
Cr(OH)30 + H2O ⇔ Cr(OH)4 + H+
The use of sodium hypochlorite (NaOCl) in combination with a caustic to treat SS slag showed sodium hydroxide (NaOH)leaching efficiency of 68%. The first step in NaOCl addition is to extract chromium in a temperature-controlled step, while water immersion ensures dissolution of the target substance [59].
The second process is water-immersed integrated alkaline roasting, which uses oxidants such as sodium nitrate to promote roasting. Experiments have shown that, by integrating alkaline roasting and water immersion, the chromium recovery rate is as high as about 84%. However, the process has also encountered difficulties; for example, chromium is embedded in an alkaline matrix, which requires the addition of an additional leaching agent [60]. The pH of the solution determines the ionic interaction of the Cr3+ ions with water and adjacent ion groups. Cr(OH)2+ predominates in a pH range of 6–8. Cr(OH)2+ and Cr3+ species predominate under acidic conditions (pH < 6), while Cr(OH)30 and Cr(OH)4- are more prevalent under alkaline conditions (pH > 8) [61]. This also enables chromium oxide to be recovered in the form of crystals [62].
It is important to note that, due to the high toxicity of chromium, chromium recovery must also include a stabilization step. The second step of alkali leaching, such as carbonation, helps to stabilize the residual chromium in the substrate.

3.2.4. Recovery of Phosphorus

The most common physical method for recovering phosphorus is magnetic separation. Studies on synthetic slags have shown that about 80% of phosphorus can be recovered by magnetic separation. On the other hand, a new technique, capillary action, is used to recover phosphorus from BOF slag. Using the principle of liquid flowing into a confined space, the liquid matrix of steelmaking slag is absorbed into a sintered CaO sphere. When the liquid phase penetrates into the CaO sinter, the solid phase containing P2O5 and the liquid phase, which is rich in FeO, will be effectively separated. Approximately 87% of phosphorus can be recovered from the P2O5 contained in the 2CaO/SiO2 phase.
Phosphorus can also be recovered chemically. The two most obvious chemical methods are carbon-thermal reduction and acid leaching. In carbonaceauric reduction, phosphorus pentoxide (P2O5) in BOF slag is reduced to phosphorus (P2 or P4) by carbon at high temperatures. Phosphorus can also be recovered by sulfuric acid leaching. For the leaching of phosphorus, sulfuric acid is usually used. In order to enhance selective leaching of phosphorus, the slag needs to be oxidized to Fe2O3-containing slag. Dissolution of phosphorus in the leachate is controlled by diffusion of the residual layer. The process is highly affected by the cooling rate, and only selective phosphorus recovery can be performed on the slow-cooling slag. Experiments on furnace cold slag show that phosphorus recovery is higher compared to air-cooled slag. In addition, acid leaching can achieve phosphorus recovery of approximately 75% [63]. On the downside, the recovery of high phosphorus content from BOF slag through certain ores or chemicals may in turn limit effective recovery [14]. As a result, steel mills need a fast and efficient process to separate phosphorus from BOF slag on a large scale so that low-phosphorus BOF slag can be efficiently recovered internally.

3.3. Resource Utilization of Converter BOF Slag

The resource utilization of BOF slag includes construction and engineering recycling, and it is used as one of the raw materials in the manufacture of a variety of building materials. In addition, it can also be applied to mine filling, soil improvement, wastewater remediation, preparation of heat storage materials, recycling as smelting solvents, etc., which have broad application prospects. The main utilization approaches can be seen in Figure 4.

3.3.1. Construction and Engineering Recycling

BOF slag has non-negligible applications to construction engineering. Because BOF slag has a certain cement-based component and certain hydration activity, it can be used in the production of cement and cementitious materials. Because BOF slag is hard and wear-resistant, it has a good elastic modulus, good moisture resistance, and strong adhesion, and can be used in the production of concrete aggregates, asphalt mixtures, and cement mortars. In addition to its good physical properties, BOF slag can provide a large amount of iron, which is suitable for the production of artificial reefs. Since the silicate oxide in BOF slag is the main component of CaO-Al2O3-SiO2 or CaO-Al2O3-MgO-SiO2 system glass, with a suitable crystalline phase, the BOF slag can be used to produce ceramics and glass-ceramic.
Supplemental Cementitious Materials
According to reports, the cement industry accounts for 6–8% of global carbon emissions, and the potential for carbon reduction is huge [64,65,66,67]. Through the effective use of supplemental cementitious materials (SCMs), CO2 emissions per ton of cement can be reduced from 0.81 tons to 0.64 tons [67,68]. Among the different types of metallurgical slag, BOF slag with high alkalinity and global high production (about 260 Mt/y) is one of the most suitable alkaline resources for mineral carbonization [69].
BOF slag contains various calcium silicate minerals [26], with active components such as C2S and C3S exhibiting inherent cementitious properties. The high mass fraction (>70%) of divalent metal oxides (CaO, MgO, FeO) in BOF slag triggers synergistic reactions through the “tetrahedral isomorphic effect of silicon” and the “double salt effect”. These mechanisms promote the formation of numerous nanoscale needle-like double salt crystals while generating densely packed, near-amorphous calcium silicate hydrate (C-S-H) gels and zeolite-like phases. The acicular crystals are tightly encapsulated within these gel matrices, creating a hierarchical microstructure that endows the cementitious system with enhanced compactness, mechanical strength (35–50 MPa at 28 days), volumetric stability (autogenous shrinkage < 200 μm/m), and resistance to environmental degradation, particularly demonstrating a 40–60% reduction in chloride diffusion coefficient compared to conventional Portland cement systems [70,71,72].
Numerous studies have demonstrated that combinations of different industrial solid wastes can generate synergistic effects and exhibit complementary advantages across performance metrics due to their distinct physicochemical properties [73,74,75,76,77,78]. For instance, ettringite formation requires reactions between aluminate hydration products and SO42⁻ ions. However, BOF slag exhibits limited cementitious activity owing to its low tricalcium aluminate (C3A) content (<5 wt%) [79]. Furthermore, the effective co-utilization of diverse solid wastes not only reduces disposal costs but also enables large-scale waste valorization, potentially generating economic benefits through resource recovery [80,81,82].
Researchers attribute the synergistic strength enhancement in cementitious systems composed of fly ash, DG, and BOF slag to the enhanced production of hydration products such as calcium silicate hydrate (C-S-H) and ettringite (AFt) [61]. The alkalis in red mud and the sulfates in desulfurization gypsum promote hydration reactions of BOF slag, with tri-component interactions significantly improving the mechanical properties of composite binders [83]. To enhance aluminum content for improved SS reactivity, studies have introduced refining slag (RS) into granulated blast furnace slag (GBFS)-steel slag (SS)-desulfurization gypsum (DG) clinker-free cementitious systems, developing a high-strength GBFS−SS−RS−DG binder achieving 43.0 MPa at 28 days [84]. Under alkaline activation (NaOH), electrolytic manganese residue–BOF slag–blast furnace slag systems demonstrate optimal 28-day compressive strength of 23.0 Mpa [85]. The double salt effect effectively immobilizes Pb and Zn. Clinker-free cementitious materials for lead-zinc tailings (LZTs) backfill have been developed using BOF slag (SS), blast furnace slag (BFS), flue gas desulfurization (FGD) gypsum, and fly ash (FA) [86]. A GGBS-SS-DS-FGDG binder system utilizing desulfurization slag (DS), ground granulated blast furnace slag (GGBS), BOF slag (SS), and FGD gypsum achieves a 748% higher 3-day compressive strength than GGBS-SS-FGDG systems at the optimal DS/SS ratio of 1.0 [87].
Cement
Similar to the principle of producing cementitious materials, because BOF slag has a cement-based component, it has a certain hydration activity. BOF slag is used instead of sand, the proportion of cement mortar accounts for 10–50%, and Ca(OH)2 is formed by the reaction of f-CaO and SiO2 to form C-S-H. It is feasible to use BOF slag as a feedstock for the production of Portland cement clinker [88].
Steel slag cement is made of a certain proportion of BOF slag and other functional additives, which can be regarded as raw material for calcining cement preparation [89,90]. The principle is that BOF slag with a CaO content of more than 50% has the potential to replace limestone in cement raw meal [91]. More importantly, the reaction of C2S and CaO in BOF slag can accelerate the formation of C3S, with high preparation efficiency and greatly shortened calcination time [92,93]. Researchers conducted a comparative study using two types of BOF slags to evaluate the effects of wet-ground steel slag (WSS) and raw steel slag (RSS) on steel slag cement performance. The work by Dai et al. (2022) [94] revealed that, when incorporating 10% WSS, specimens under constrained curing achieved a maximum compressive strength of 68.5 MPa at 28 days, representing 93% enhancement compared to free-cured specimens, along with optimal impermeability. The constrained curing conditions promoted expansive strain from steel slag hydration, leading to pore structure densification and reduced porosity, thereby improving both the mechanical properties and the microstructural integrity of the cement matrix [94].
Beyond Portland cement, BOF slag has found applications in phosphate-based cementitious systems. Research indicates that the enhancement in early compressive strength arises from the optimized particle size distribution, filling effects, and activation efficiency of slag powder. However, BOF slag powder contains significant inert phases (>40% non-reactive components) that remain largely unactivated, showing minimal participation in hydration reactions [73]. Notably, the hydration reaction between MgO and ADP (ammonium dihydrogen phosphate) exhibits substantially higher exothermicity compared to MgO-KDP (potassium dihydrogen phosphate) systems, generating additional reaction products (e.g., struvite-like phases) that enhance matrix stability. Furthermore, studies reveal that incorporating steel slag powder (SSP) accelerates the early hydration rate of MPC (magnesium phosphate cement). While a marked improvement in long-term mechanical strength is observed, this comes at the expense of reduced early-stage strength. Increasing SSP content from 0% to 30% significantly improves the water resistance of MPC mortar specimens, as evidenced by <5% strength loss after 28-day water immersion [95]. Researchers synthesized iron-calcium phosphate cement (ICPC) by reacting BOF slag powder (SSP) with ammonium dihydrogen phosphate (ADP). The compressive strength of ICPC paste exhibited a non-monotonic trend with increasing SSP/ADP mass ratio: it initially increased, then decreased across all curing ages. Peak strength was achieved at an optimal SSP/ADP ratio of 6.0, attributed to balanced reaction kinetics and densified microstructure formation [96].
BOF slag demonstrates promising applications in biocementation technologies. The calcium and magnesium ions from free CaO and MgO in BOF slag can be acid-leached (e.g., using HCl or CH3COOH) and repurposed as dual-ion sources (Ca2⁺/Mg2⁺) for microbial-induced carbonate precipitation (MICP). Specifically, the calcium components in BOF slag are chemically convertible to calcium chloride (via NH4Cl treatment) or calcium acetate (through acetic acid reaction). These derived compounds serve as optimal substrates for ureolytic bacteria (e.g., Sporosarcina pasteurii), driving the MICP process to enhance biocementation efficiency through controlled CaCO3 crystallization [97]. One of the main advantages of biocement over Portland cement is that it can be produced at room temperature [97]. In addition, biocement produced from BOF slag does not contain chloride and can therefore be used in many situations where traditional cementation methods are not suitable, such as reinforced concrete.
Cement Mortar
Cement mortar, as an indispensable building material in modern society, consumes a large amount of raw materials in its production, especially natural sand [98,99]. Due to its hardness, wear resistance, high crushing value, good physical and mechanical properties, and sufficient stockpile and annual emissions, converter steel slag is considered to be a substitute for natural sand in the manufacture of mortar alternative materials [100,101,102]. While it can also partially replace cement as a supplementary material, its primary role lies in leveraging its superior mechanical properties as a sand substitute, as demonstrated by carbonation-treated slag achieving up to 45% sand replacement in mortars [100] The low content of tricalcium silicate in SS and the high content of inert components result in low hydraulic activity of SS and lower compressive strength when used as an auxiliary cementitious material in the manufacture of mortar and concrete compared to pure Portland cement products [103].
When SS is used to partially replace cement in the production of mortar, the SS fineness has a significant effect on the performance of the mortar. The optimal particle size range for BOF slag in blended mortar (21.75–24.13 μm) increases compressive strength (31.21 MPa) to levels significantly better than those seen in BOF slag mortars of larger or smaller sizes [104].
In addition, polymer-modified waterproofing mortar can be obtained by adding a polymer composition to an ordinary cement mortar. Through the synergistic effect of inorganic and organic materials, the defects of traditional rigid waterproof materials such as low flexural strength, low bond strength, easy cracking, and poor water resistance can be improved. Compared with ordinary cement mortar, the properties of polymer-modified waterproofing mortar, such as mechanical strength and impermeability pressure, have been improved. For cementitious materials, cement, and cement mortar, the low hydration activity of BOF slag limits its application, due to the low amorphous content in BOF slag and the presence of crystalline phases, such as silicic acid γ of dicalcium (γ-C2S), dicalcium ferrite (C2F), and RO phase (iron-containing solid solution) solid solutions [23,105,106]. At present, there are three ways to enhance the cementitious properties and macroscopic properties of BOF slag. One is thermal activation. This method involves elevated curing temperatures (typically 60–90 °C) to amplify the slag’s hydration activity. The activation mechanism stems from thermally induced bond cleavage: at high temperatures, the Si–O and Al–O bonds within the slag’s vitreous network structure undergo preferential fracture due to thermal stress. This facilitates depolymerization of the vitreous phases, thereby increasing hydration reactivity and accelerating early strength development [107]. High-temperature curing poses challenges for in situ concrete construction and often results in inferior long-term performance of cementitious composites compared to standard curing conditions (20 °C, RH ≥ 95%) [108].
Another method is chemical activation. By changing the composition of the liquid phase and increasing the number of crystal nuclei and the alkalinity of the liquid phase, the hydration and hardening of BOF slag can be accelerated. Alkali activators, often referred to as reinforcing additives, are used as early hydration accelerators for cement-SS composite binders [109]. When the [AlO6] tetrahedron in the BOF slag is inactivated, [AlO6] dissolves in its original position in the form of Al(OH)2+ to form a water-soluble ion, which is combined with the H3SiO4 present in the solution, and OH, Ca2+, and Na react to form zeolite hydrate [107]. Regarding Na2CO3, studies on the hydration properties of BOF slag under the influence of alkali and alkali metal salts such as Na2SiO3 show that alkaline activators can promote the gelling properties of BOF slag [108].
A final method is mechanical activation, which improves the fineness of the BOF slag, resulting in an increase in the amorphous phase content and an increase in the early activity of the BOF slag, but a trade-off with grinding costs is required [108]. The finer particles exhibit high reactivity and disperse the expanded components (e.g., f-CaO) in the BOF slag [110]. Ultrafine BOF slag can produce more hydration products, but when the BOF slag content is > 10%, the cement performance is still hindered, and the setting time and water demand also show an upward trend in fine particle size [103].
Asphalt Mixture
At present, alkaline aggregate is commonly used in asphalt mixtures, because asphalt acid, anhydride, and other surfactants in asphalt will have a strong chemical adsorption effect when in contact with alkaline aggregate, and the adhesion is large. Asphalt mixtures mixed with BOF slag exhibit high temperature stability, good slip resistance, good crack resistance, good fatigue durability, permanent deformation resistance, and superior interlocking structure [111,112,113,114,115]. It has been found that the critical load of the BOF slag permeable asphalt mixture increases first and then decreases with a decrease in temperature, reaching a maximum value at −10 °C. When the replacement rate of BOF slag is 50% and the polyester fiber content is 0.45%, the ultimate tensile stress and ultimate tensile strain of asphalt mixture are the largest, and the best low-temperature crack resistance is exhibited [116].
With the promotion of intelligent green pavement and long-life pavement, self-healing asphalt mixtures have attracted attention [117]. There are three main types of accelerated self-healing of asphalt mixtures: microcapsule healing, induction heating, and microwave heating (MH) [118,119]. The latter two are often mixed with BOF slag aggregates. It has been found that the induction heating rate of asphalt mixtures mixed with steel fiber and BOF slag is increased by 34% [120]. Asphalt concrete prepared with BOF slag shows a better ability to store heat and a similar heating rate to basalt asphalt mixtures, but exhibits a significantly slower cooling rate [121]. The high dielectric and magnetic conductivity losses give the BOF slag a potential microwave absorption capacity to migrate heat and repair fractures [122,123]. It has been found that asphalt mixtures with 50% SS content can achieve excellent heating uniformity and microwave de-icing efficiency [124]. When the CPZ temperature of the BOF slag asphalt mixture is high, the self-healing performance of the mixture is better. The content of BOF slag has little relationship with the initial strength of the asphalt mixture, but mixtures with large BOF slag content have strong self-healing properties [125]. Due to the high density and high strength of BOF slag, the stability of asphalt is improved, but at the same time the low temperature crack resistance and water stability are reduced due to the porous structure of BOF slag. The degree of SS substitution is limited (maximum 23.5%). Microwave heating can reduce the micropores in SSAM, which is conducive to self-healing of microcracks. However, continuous microwave cycling increases the proportion of large and extra-large voids, resulting in a decrease in their durability [117]. It is worth mentioning that the incorporation of BOF slag is conducive to the utilization of recycled asphalt pavement (RAP). BOF slag can improve the mixing efficiency of RAP–virgin asphalt in recycled asphalt mixtures [126], reduces electrical conductivity and thermal diffusivity, and increases specific heat [127]. High levels of RAP improve the moisture susceptibility of recycled asphalt mixtures but decrease their fracture resistance, and may offset some durability issues in warm mix asphalt mixtures [128]. Studies have found that 20% RAP achieves the highest fatigue resistance, considering the environmental sustainability of road construction, and a maximum RAP dose of 40% is recommended [129].
Concrete Aggregates
BOF slag demonstrates significant potential as a concrete aggregate due to its superior angularity and abrasion resistance [101], effectively enhancing concrete’s surface roughness and wear durability [130]. The slag’s textured surface morphology and cementitious constituents synergistically reinforce the Interfacial Transition Zone (ITZ) between aggregate and cement paste, thereby improving concrete’s mechanical strength [101]. Some researchers have used BOF slag to partially replace the mechanical behavior and microstructure of concrete with NCA. The experimental results showed that the optimum replacement rate of BOF slag was 50%. At this replacement rate, the 7-, 28-, and 90-day compressive strengths were increased by 3%, 3%, and 19%, respectively. In addition, the use of BOF slag as a coarse and fine material can increase concrete strength, reduce porosity [131], and delay setting time, reducing the heat of hydration and improving concrete flowability [101,132].
Artificial Reefs
Artificial reefs (ARs) are being widely used as a response to the loss of macroalgal forests [133,134]. The use of BOF slag to restore the marine environment is a new type of BOF slag resource utilization technology [29].
Some researchers have found that the decrease in iron content after immersion of artificial reefs prepared using steel slag was caused by excess iron oxide produced by the immersion of BOF slag in seawater. This phenomenon provides a large amount of iron to microorganisms, which promotes their growth and improves the marine environment [135]. Green Artificial Reef Concrete (GARC) was prepared using SS as the main raw material; the density of GARC was 2765.5 kg/m3, and the compressive strength of GARC reached 71.4 MPa at 28 days. After curing in artificial seawater for 240 days, the compressive strength of GARC was 92.5 MPa [136].
With the diversification of materials used in artificial reefs (ARs), the adverse effects of ARs on marine ecosystems is an emerging issue. Although there is a lot of research on the impact of ARs on marine ecosystems [136,137], quantitative evaluation of heavy metal (HM) elution from ARs is limited. HM pollution of the oceans is a concern both locally and globally due to their persistence and toxicity. The effects of ARs made of concrete, steel, and BOF slag on the concentrations of eight HMs (As, Cd, Cr6+, Cu, Hg, Ni, Pb, and Zn) in the ocean have been studied. The concentrations of the eight HMs in the experiment met environmental standards under conservative experimental conditions [138]. At present, research on the preparation of artificial reefs by BOF slag still needs to be strengthened, and it is necessary to pay attention to the impact of its application on the ecological environment.
For asphalt mixtures, concrete aggregates, and artificial reefs, the volumetric expansion caused by the hydration of f-CaO in BOF slag limits its further utilization. BOF slag has a high content of free CaO/MgO, which can lead to excessive expansion of the concrete [139]. Prolonged or accelerated carbonation and other measures are usually taken to reduce the amount of free CaO/MgO and the risk of swelling [140]. Some scholars use the thermal asphyxiation process to treat the BOF slag by a natural aging method to make the content of f-CaO and f-MgO meet the requirements [141]. Furthermore, weathering is a common solution. However, steel slag aggregates with a greyish-white CaCO3 product layer formed by weathering are less prone to freeze-thaw damage when used in asphalt mixtures [142]. Even after weathering, there are risks associated with the use of steel slag aggregates in asphalt mixtures. In the future, further research is needed to reduce the limitation of unstable components in BOF slag on its resource utilization.
Glass-Ceramic
Glass-ceramic is a polycrystalline composite composed of glass and crystalline phases [143]. Silicate oxides present in BOF slag are the main components of CaO-Al2O3- SiO2 or CaO-Al2O3-MgO-SiO2 glass. Chromium oxide (Cr2O3) and chromium-containing minerals (e.g., MgCr2O4) are effective nucleating agents for glass-ceramic [144,145]. In the presence of a nucleating agent, glass-ceramic can be formed by controlled crystallization via heat treatment of glass with the right composition. Fe2O3 in BOF slag is not only conducive to the melting of glass, but is also conducive to the crystallization of glass-ceramic, and the content of Fe2O3 in BOF slag glass-ceramic was studied by the sintering method [146,147]. As the SS increases, the grains continue to refine [148]. Considering that the ratio of web-forming ions to net unwebbed ions in BOF slag does not form the glass structure well, most studies will choose to add some other oxides as additives to promote the formation of substrate glass, and focus on its mechanical properties [143]. Some researchers used BOF slag as the main raw material and added appropriate oxides (SiO2, Al2O3, Na2CO3, and MgO) to prepare wollastonite–pyroxene glass-ceramic in the BOF slag system. In the prepared sample, glass-ceramic with a density of 3.08 g/cm3 exhibited the best mechanical properties, with a flexural strength of 176.21 MPa and a microhardness of 8.81 GPa [143]. When Na2O, SiO2, and Na2SiF6 are used as additional additives, the optimal flexural strength of the prepared fluorine-containing BOF slag glass-ceramic specimen reached 177 MPa [147].
In general, glass-ceramics have controllable crystallization ability through processes like sintering or two-step crystallization, which can be enhanced by nucleating agents (e.g., Cr2O3) to promote crystal formation. However, both methods have non-negligible drawbacks: firstly, the quenching step in the sintering process wastes a lot of heat and produces a lot of waste water. In this regard, Luo et al. [147] found that glass-ceramic with a constant amount of 50 wt% BOF slag was prepared by the melting method; the bending strength of the prepared sample was the highest at 177.76 MPa, and the volume shrinkage was the lowest (0.06%). On the other hand, two-step crystallization production involves two heat treatment steps (nucleation and crystal growth), which is energy-intensive and carbon-intensive. A very promising approach to this is one-step heat treatment, i.e., nucleation and crystal growth at the same temperature. In general, the Tn and Tc of oxide glass are sensitive to the chemical composition of the parent glass. Therefore, optimizing the composition of the parent glass is considered to be a feasible method to reduce the delta T, which can improve the nucleation ability or crystal growth ability of the parent glass. However, in this regard, there is less research on BOF slag glass-ceramic.
Ceramic Materials
BOF slag is mainly composed of various types of metal oxides, and the solid solution also contains a certain amount of glass phase, similar to the composition of ceramic raw materials [149]. The use of BOF slag to produce ceramics can improve the process performance of low water absorption, low porosity, high strength, and so on [150]. This improvement may be due to the presence of alkaline earth oxides in the BOF slag composition, which enhances early curing of tiles and porcelain. Zong et al. successfully prepared ceramics by mixing BOF slag with fly ash [151]. When the fly ash is doped with 15 wt% and the sintering temperature is 1145 °C, the flexural strength is 43.37 MPa. In addition, Pal et al. [152] reported that the use of 40 wt% BOF slag to make porcelain materials can promote the crystallization of pyroxeneand improve the vitrification behavior, with a flexural strength of 80 MPa. However, the introduction of the best BOF slag to improve process performance is still limited to the range of 30–40 wt%. Exceeding this BOF slag weight may involve the emergence of mesophases such as wollastonite and diabase [150].
In recent years, some studies have explored the effect of the CaO/SiO2 ratio of BOF slag on the phase crystallization and properties of pyrite-sintered materials. Tabit et al. prepared iolite-based binary ceramics from fly ash and BOF slag [153]. However, excess iron and calcium are detrimental to the preparation of ceramics, leading to problems such as deformation, cracking, strength loss, and narrowing of the sintering temperature range [154]. The presence of Fe2O3 and CaO in BOF slag often leads to a series of problems such as deformation, strength loss, and surface defects [155,156]. Therefore, it is extremely difficult to prepare ceramic materials from high-calcium and high-speed iron BOF slag. The material composition, microstructure, and preparation process must be reasonably controlled in order to produce ceramic materials that meet the performance requirements [154]. This is also the direction of future research.

3.3.2. Mine-Filling Materials

BOF slag can play a crucial role in mine-filling materials [157]. Its dissolution releases abundant OH⁻, Ca²⁺, silicate, and aluminate ions, which recombine to form hydration products and break the Si-O and Al-O bonds in the slag matrix, promoting gel formation and strength development, Ca2+ reacts to produce C-S-H gel material, and water-soluble SO42− reacts with activated alumina (present in the form of H2AlO3 or AlO2) and Ca2+ to produce microcrystalline AFt. With an increase in age, C-S-H gel and AFt continued to increase, and by 28 days of age, a large number of gels had been formed, filling and wrapping the AFt and unhydrated particles, resulting in a continuous increase in the internal densification of the filler and a continuous increase in the mechanical strength [26]. A large number of hydration products are continuously generated and wrapped on the surface of the raw material particles, and C-S-H and AFt nucleate, crystallize on the surface of the raw material particles, and grow interlaced with each other to form a reticulated structure. BOF slag can improve the compatibility of the filling slurry and increase the strength of the backfill in the later stage [22].
A large number of researchers have used BOF slag as raw material to prepare backfill materials. Zhang et al. [158] used GBFS, BOF slag, DG, and ultrafine tailings as raw materials to prepare backfill materials with optimal ratios to achieve compressive strengths of 6.69, 12.05, 16.36, and 180 MPa on the 3rd, 7th, 28th, and 180th days, respectively. Zhang et al. [159] used the best ratios of backfill materials as raw materials to achieve compressive strengths of 6.69, 12.05, 16.36, and 18.37 MPa. Researchers prepared a new type of mine backfill cementitious material with 75% BOF slag, 16.5% FGD gypsum, and 8.75% FGD ash (the ratio of binder to tailings was 1:4, and the concentration was 70%), and the compressive strength of the specimen reached 1.24 MPa at 28 days and 3.16 MPa at 90 days. Some scholars have studied the preparation of cementitious backfill with fly ash, desulfurization gypsum, and BOF slag. The UCS value of 14 days’ cured filler could reach 2.60 MPa under optimal proportioning [160]. It should be noted that, although most of the mineral phases contained in BOF slag can be considered low reactive components, BOF slag has high relative reactive aluminum phases (C3A and C12A7), and the aluminum-containing phases will reduce the C3S/aluminate (C3A and C12A7)) ratio when BOF slag is blended into cement with a high percentage of cement replaced by BOF slag, which results in the cementitious system lacking sufficient amounts of sulfate (i.e., insufficient sulfation) [26]. This can lead to rapid depletion of gypsum and rapid precipitation of caliche [27,28]. The BOF slag-based reverse filling method utilizes more BOF slag particles of relatively coarse size as microfillers and uses a small amount of fine cement particles as the main bonding component. The compressive strength of ultrafine silicate cement can reach up to 123.8 MPa at 180 days, with an SPC content of 35% [26].
BOF slag causes prolonged setting time of the filler slurry and significant reduction of the early strength of the filler. And when the BOF slag content exceeded 20 wt%, the curing strength was reduced [157]. The backfill strength properties decreased with an increase in SS value [160,161]. The incorporation of a large amount of BOF slag causes the damage form of the backfill to exhibit tensile damage, with different degrees of loose swelling at the bottom [160]. In addition, the increase in SS content causes the damage mode of the backfill to change from tensile to tensile shear [162], resulting in a composite damage pattern of splitting and shear in the backfill. The addition of BOF slag reduced the degree of hydration reaction in SS-CPB [161]. It cannot be ignored that the swelling problem of steel slag also has a negative impact on the filler. Fillers with a higher dosage of steel slag (above 50 wt%) showed potential binding properties, but the long-term performance is yet to be resolved.
In the future, attempts can be made to mitigate the retardation effect of BOF slag on the early hydration of cement by using coupling between solid wastes, pretreatment of BOF slag, or addition of admixtures. Research can also be carried out on the material proportion of a large dosage of BOF slag, as well as the production process, so as to avoid cracking of the filling body and lower early strength. In addition, in order to achieve the industrial application of steel slag-based cementitious materials in large molds, it is necessary to further carry out research on the high concentration and high fluidity of steel slag-based tailing sand filling slurry.

3.3.3. Soil Amendment

BOF slag has a wide range of uses in agriculture. It can be used as a fertilizer to increase crop yield [4,163], to neutralize soil acidity [164,165], to stabilize heavy metals [165,166,167], and so on.
BOF slag can be classified into various fertilizers (e.g., silicate fertilizers, iron fertilizers, phosphate fertilizers, etc.) due to the presence of a wide variety of nutrients in different concentrations, such as Si, Fe, P, Ca, and Mg [163]. Silicon is the most demanded element for rice growth, and BOF slag, which contains more than 18% SiO2, can be used as a silicon fertilizer, which helps to improve the quality of the soil for plant growth [168]. BOF slag contains high levels of iron, which when added to the soil can treat iron chlorosis in crops, resulting in 160% and 183% increases in iron content in Japonica and Indica rice, respectively [169]. In addition, the high concentration of phosphorus in BOF slag can provide a source of slow-release, long-term phosphorus for agriculture. Currently, only a few studies have investigated the bioavailability of phosphorus in BOF slag for plant uptake, and more research is needed in this area [4].
BOF slag has also been shown to have the potential to reduce soil acidity, as it acts as a liming agent. This effect occurs due to the presence of calcium silicate and magnesium silicate, which, when applied to the soil, promote greater mobility of SiO32- anions in the soil profile, resulting in higher pH, lower Al3+ concentration, and increased base saturation [164]. For nitrous oxide (N2O), BOF slag promotes complete denitrification by providing electron donors such as ferrous ions (Fe2+), which provide more electrons than are required to produce N2O, an intermediate in the denitrification process, leading to its further reduction to nitrogen (N2) and inhibiting N2O emissions. When BOF slag was applied at an application rate of 8 Mg/ha, the cumulative CH4 and N2O emission rates were 56.0% and 98% lower, respectively, than the control [170]. BOF slag as a soil amendment in subtropical paddy fields has also been shown to reduce CO2 emissions during tillage [171]. However, the reduction in CO2 emissions was not significant when compared to the control later in the season. BOF slag as a fertilizer has been shown to increase yields of crops such as soybean and rice, which hints at the potential for enhanced carbon biofixation through the addition of slag [164,172].
BOF slag can durably immobilize metal ions in heavy metal–contaminated soils [173]. BOF slag is loose and porous, and the main mineral phases are calcium silicate, calcium ferrite, and alkaline oxides, which not only form hydroxide precipitates between alkalis and OH groups produced by hydrolysis of heavy metal ions, but also form heavy metal ions, such as calcium ferrite, through Ca2+ in calcium silicate, and the ion exchange of Ca2+ in calcium silicate with heavy metal ions such as calcium ferrite generates stable heavy metal silicates and ferrite, which effectively removes heavy metal ions from the polluted solution [165]. Application of BOF slag to Cd-contaminated soil resulted in a significant increase in peroxidase activity in plant roots, thereby reducing the concentration of extractable Cd in the soil [174]. BOF slag converts Cd ions in soil to less decomposable CdCO3, CdO, and Cd(OH)2 by precipitation and adsorption [175]. Simultaneous solidification/stabilization of mixed heavy metals (mainly Cd, Pb, Zn, Cu, etc.) using BOF slag has been the focus of research [176,177]. Yang et al. (2021) [166] found that Cu, Cr, Pb, and Zn ion fixation and persistence after the addition of BOF slag to acidic contaminated soils was similar to that of lime in the modelling process. Passivation material (SS-BC) prepared by modification of BOF slag using biochar (BC) reduced the bioavailability of heavy metals in the soil. The passivation efficiencies of Pb and Cd were found to be 51.76% and 63.74%, respectively [167]. SS-BC, a heavy metal passivation material synthesized using BOF slag and biochar, showed passivation efficiencies of up to 51.76% and 63.74% for Pb and Cd in soil, respectively [20].
In the future, composite amendments with multiple functions can be developed to meet different types of soil problems and crop needs. By combining BOF slag with other wastes or functional materials, it can have the function of improving soil pH and supplementing nutrients, as well as fixing heavy metals and improving soil structure. In addition, research should focus on the long-term effects of potentially harmful substances in BOF slag on soil and ecosystems and how these risks can be reduced through scientific methods and reasonable dosage control.

3.3.4. Wastewater Remediation

In recent years, BOF slag has received widespread attention in wastewater remediation applications due to its large specific surface area, porous structure, high density, and strong ion exchange capacity. Researchers at home and abroad have confirmed the effectiveness of BOF slag in removing inorganic ions such as phosphorus and arsenic, heavy metal ions such as nickel, chromium, lead, zinc, and copper, and organic pollutants such as ammonia, nitrogen, and benzene.
Adsorbents
BOF slag shows good adsorption effect in the treatment of various types of industrial wastewater containing heavy metal ions, organic dyes, inorganic non-metals, etc. BOF slag has a large specific surface area and a porous surface, which can be used for the removal of pollutants by surface adsorption. The hydroxyl group (-OH) and other groups on the surface of BOF slag dissociate to form adsorption sites centered on negative charges, and cations can be electrostatically adsorbed onto the surface of BOF slag by electrostatic adsorption. These properties enable steel slag to remove pollutants by physical adsorption. At the same time, steel slag contains a large amount of alkaline metal oxides such as Al2O3, MgO, CaO, etc., which are hydrolyzed and ionized, releasing a large number of metal ions (e.g., Mg2+, Fe3+, Ca2+, etc.), which are chemically reacted to form a salt precipitate to achieve the removal effect [178].
BOF slag can remove inorganic ions from solution. The removal mechanism of BOF slag is different for the two nitrogenous pollutants. The removal of NH4+-N by BOF slag adsorption is controlled by external and internal diffusion, while the removal of NO3-N by BOF slag is mainly a chemisorption process. In order to improve the treatment performance of BOF slag for the removal of ammonia nitrogen, nitrate nitrogen, and other nitrogenous wastewater, many scholars have investigated the effect of modified SS on the removal of nitrogen from wastewater by modified SS [27,28]. Phosphorus removal is mainly through the formation of precipitation, and P in solution can be removed through the formation of calcium phosphate precipitates when the pH is increased above 7.0 [179].
It is difficult to catalyze the removal of heavy metal ions, and their ionic radii are only Å in size, so adsorption is one of the most applicable removal methods [180,181]. The surface of BOF slag is rich in pore structure and active sites, and these properties enable it to adsorb heavy metal ions in water efficiently. BOF slag is highly effective in the adsorption of Cr3⁺and Zn2⁺, with high removal rates of 85.79% and 76.52%, respectively [182]. A novel adsorbent material prepared from acid-modified BOF slag had an adsorption efficiency of up to 98.5% for uranium [183]. BOF slag removed vanadium from synthetic soil rinse wastewater, with vanadium removal up to 97.1% [184]. The iron oxide in BOF slag can provide a rich source of iron for arsenic precipitation or/and adsorption through the formation of a mixture of arsenic-related compounds (e.g., ferric arsenate and ferric arsenite) and iron hydrate. Zeolites synthesized using BOF slag showed high adsorption capacity and removal of Fe3⁺ ions of up to 99.99% [185]. Copper-smelting wastewater was treated by the synergistic action of BOF slag and KMnO4, and, after 2 h of reaction, the residual concentrations of arsenic and heavy metal ions in the resulting clear water were lower than the prescribed limits for industrial wastewater discharges at a pH of 10 [186]. Wang et al. [187] prepared a novel magnetic adsorbent using BOF slag and Ginkgo biloba, which showed adsorption efficiencies of more than 4% for both Cd(II) and As(V) at pH = 9.0. The geopolymer SSGP prepared using BOF slag and sodium hydroxide, modified by an alkaline activator, showed an adsorption capacity of SSGP of 145.43 mg/g and 69.97 mg/g for Zn and Ni, respectively. This significant increase in adsorption capacity can be attributed to the fact that the specific surface area of the prepared material was increased by a factor of three compared to that of BOF slag, which greatly improved the BOF slag’s surface activity and specific surface area [188]. Lead and cadmium can be removed from water using BOF slag as a filler, where pH has a great influence on the removal of lead and cadmium. The removal efficiency of BOF slag for lead and cadmium was tested in natural water at a pH of 10.5, and the removal of metal pollutants was more than 95% [189]. Sang et al. used 72 wt% of BOF slag to prepare a porous geopolymer (SPG) for the removal of Cu2+ from wastewater, and the adsorption of Cu2+ by SPG was achieved by electrostatic attraction, ion exchange, and chemical precipitation [190]. As a low-cost and high-efficiency adsorbent material, BOF slag has a promising future for application in the treatment of heavy metal wastewater. Future studies can further explore the modification methods, adsorption mechanisms, and application effects of BOF slag in practical industrial wastewater treatment. Meanwhile, the regeneration and recycling of BOF slag adsorbent materials also need to be paid attention to in order to achieve the sustainable use of resources.
In addition, BOF slag, as a low-cost adsorbent, has a certain adsorption capacity for phenanthrene, naphthalene, and phenol [191,192]. However, the utilization of BOF slag in water treatment is limited due to its low adsorption capacity and instability in acidic solutions. Therefore, it is crucial to identify appropriate modification methods to promote the high efficiency of BOF slag for wastewater treatment or to convert BOF slag into a material capable of degrading pollutants rather than acting as an adsorbent.
Catalysts
BOF slag is applied as a catalyst for wastewater treatment based on the principle of a Fenton-like reaction catalyzed mainly by iron oxides. Some challenging catalysts require the synergistic effect of iron oxide contained in the BOF slag [193].
Researchers have found that the preparation of non-homogeneous catalysts for the electrochemical removal of pollutants using industrial waste can be very effective [194,195,196]. Song et al. [197] prepared a sulfur-zinc-modified kaolinite/SS particle electrode (S-Zn-KSPE) with magnetic properties, tested the cycling stability of S-Zn-KSPEs under neutral and alkaline conditions, and found that they maintained high catalytic activity with > 98% recovery after eight cycles. The particulate electrodes (KSPEs) were prepared from BOF slag, kaolin, and NaHCO3. They were used for the degradation of norfloxacin (NOR) in wastewater, and the degradation rate of NOR could reach 96.02% in 30 min [198]. Nanostructured Mg0.04Fe2.96O4 catalysts were prepared using BOF slag, and the Cr(VI) removal ability of MFSS was superior to that of some Fe-based catalysts under visible light irradiation [199]. Natural pyrite-steel slag (NP-SS) materials were prepared using natural pyrite (NP) and BOF slag, and the maximum Cr(VI) adsorption of NP-SS was as high as 7.73 mg/g at 15 °C and a Cr(VI) concentration of 800 mg/L [200]. The nanocomposite photocatalyst MIL-53(Fe)/SiO2 was prepared from LD slag using the thermosolvation method, which could degrade methyl blue under UV irradiation. Under UV irradiation with a methyl blue concentration of 5 mg/L, pH 10 and a photocatalyst content of 0.5 g/L, the methyl blue removal rate was 66.3% [201]. The key to improving the pollutant removal rate in the system is to study the effect of factors such as electrode materials, catalyst type and system, and water conditions on the performance and behavior of the electrocatalytic process [202]. Yu et al. investigated the use of modified BOF slag particles Co-SAM-SCS as catalysts in a non-homogeneous electro-Fenton process coupled with BDD to stabilize the waste leachate COD and NH3-N in leachate [203]. Fang et al. [204] modified magnetic converter steel slag (MSS) solid waste with Co-MOF derivatives (Co-CN) via heat treatment, successfully synthesizing the MSS@Co-CN Fenton-like catalyst and used the synergistic effect of adsorption and Fenton-like oxidation to remove a representative antibiotic, tetracycline hydrochloride (TCH), from wastewater. MSS@Co-CN exhibited high adsorption capacity (qmax > 416.6 mg/g) and strong Fenton-like catalytic oxidation (80% removal efficiency in 5 min).
In future studies, coupling between the external field and the catalytic process can be considered to further improve catalytic efficiency. The catalytic mechanisms associated with metallurgical slag should also be further investigated, e.g., identification of active sites from metallurgical slag to reveal free-radical-generating reactions and the catalytic role of various metals in slag-derived catalysts.
Pretreatment methods such as acidification and calcination can be used to improve the catalytic performance of slag-derived catalysts. The introduction of an external magnetic field and the use of reducing substances are promising alternatives to enhance the catalytic process and need to be further investigated. Future research should focus on addressing safety and stability issues, developing green and efficient modification methods, improving degradation efficiency, and implementing large-scale treatment of real wastewater.
Photocatalysts
Photocatalytic degradation is considered to be one of the most promising strategies for recycling organic wastewater, because most organic pollutants can be photodegraded into non-polluting products, and the degradation process is sustainable and readily available. In recent years, there are also examples of using BOF slag to prepare photocatalysts for wastewater treatment. A CeO2-loaded alkali-activated composite fabricated from BOF slag, containing higher iron compound content than blast furnace slag (BFS), demonstrated enhanced photocatalytic activity through CeO2/FeO heterojunction formation. Transient photocurrent analysis confirmed improved charge separation efficiency, enabling 100% degradation of malachite green within 80 min using 8 wt% CeO2-loaded specimens under visible light irradiation [205]. Takumi Inoue et al. [206] found photocatalytic degradation of imidacloprid by graphitic carbon nitride/artificial BOF slag (gCN/artCS) composites with optimized mixing ratios of BOF slag under visible light irradiation. An optimized gCN/artCS composite containing 11.58% artCS increased the degradation rate of imidacloprid by a factor of 2.5 in 120 min compared to pure gCN. Shao et al. [207] proposed a strategy based on the simultaneous removal of heavy metal ions and organic pollutants from containers by a BOF slag-derived calcium silicate hydrate (CSH). BOF slag-driven CSH can slowly release Ca2+ and OH- and then chemically bind heavy metal ions for their removal. The adsorption capacities of the prepared CSH for heavy metal ions were all higher than 100 mg/g, and the photodegradation efficiency of the organic pollutant methylene blue (MB) reached 63% after 8 hours of low-power visible light irradiation using the prepared CSH. The CSH was also able to remove heavy metal ions from the container with the help of BOF slag.
Heavy metal ions usually coexist with organic pollutants in real wastewater systems, when the presence of heavy metal ions has a significant negative effect on the photocatalytic degradation of organic pollutants [208]. However, achieving simultaneous removal of both heavy metals and organic pollutants in a single reactor and within a single-step process remains a major challenge, with limited studies addressing this issue.

3.3.5. Thermal Storage Material

BOF slag has a high specific heat capacity value (up to 0.95 J/(g·K)), good thermal conductivity, and a good friction coefficient, which makes it a very promising material for thermal storage [209,210]. BOF slag-based thermal storage material exhibits excellent thermal cycle stability, high efficiency, and low cost, indicating broad application prospects [211]. Thermal energy storage (TES) includes sensible heat storage (SHS), latent heat storage (LHS), and thermochemical heat storage (TCHS) [212].
Sensible heat storage (SHS) is the most mature technology and has been widely commercialized. Currently, most of the studies on SHS materials are on electric furnace slag. BOF slag, as a sensible heat material, remains stable until 1200 °C and has good thermal cycle stability. The thermal properties of BOF slag were further enhanced by Na2CO3 activation, and the thermal storage density of the resulting modified material reached 997.0 kJ·kg−1 (400–900 °C), with thermal conductivities of 1.331, 1.323, and 0.889 W·(m·K)−1 at 25, 250, and 500 °C, respectively [213]. By adding magnesite and quartz to BOF slag, a sensible thermal storage material with good thermal shock resistance was successfully prepared. Its heat storage density was 1000 J/g in the range of 25 °C to 800 °C [214]. Concrete–BOF slag thermal storage material with BOF slag as an aggregate has good thermal stability at 700 °C and a specific heat capacity of 1.85 J/(g·K) at 500 °C. However, this material has a low thermal conductivity of 0.7 W/(m·K), and a small percentage of BOF slag was used [215]. Zhang et al. [216] prepared a novel sensible heat storage material using BOF slag as the main component in the 50–1000 °C range. Its compressive strength was stable at about 80 MPa, its thermal conductivity was 1.13 W/(m·K), and its thermal density was 1222.168 J/g. After 300 thermal cycles, the morphology of the specimen, 94.4% of the compressive strength, 98.8% of the thermal conductivity, and the thermal density were almost unchanged, which provided a new low-cost method for the preparation of BOF slag heat-exposing material. However, sensible heat storage suffers from the problems of low heat storage density and large temperature variation, which cannot achieve efficient heat storage [217].
Phase-change (latent heat) thermal storage has the advantages of high heat storage density and stable temperature change during the heat charging and releasing process, which is one of the potential research directions of current thermal storage technology [218]. Latent heat storage absorbs and releases heat through the phase-change process of phase-change materials (PCMs) [217]. Although research on C-PCMs started in the 1990s, not enough studies have been conducted on C-PCMs prepared with inorganic materials as support materials, especially in medium- and high-temperature phase-change materials. Liu et al. [10] firstly used BOF slag as a support material using polyethylene glycol (PEG), sodium nitrate (NaNO3), and sodium sulfate (Na2SO4) to prepare a series of composite phase-change materials (C-PCMs). Compared with pure phase-change materials, the three C-PCMs reduced the subcooling degree by 2.64 °C, 4.53 °C, and 0.79 °C and increased the thermal conductivity by 172%, 54.9%, and 82.4%, respectively, and the latent heat retention rate was still more than 97% after 100 thermal cycles, which provided good phase-change thermal storage performance and thermal reliability. Wang et al. [219] prepared a modified phase-change material (C-PCM) comprising BOF slag–K2CO3 phase-change composites (PCC), which can be used in high-temperature concentrating solar power systems. They found that the composite has a melting point of 880 °C and a latent heat of ∼140 J/g and maintains good shape stability and thermal properties under multiple thermal cycles. A series of BOF slag–KNO3 PCCs were prepared by hybrid sintering for use in medium- to high-temperature TES systems, and the composites maintained high compressive strength (30–35 MPa) and Young’s modulus (1600–1700 MPa) when the KNO3 content ranged from 38 wt% to 50 wt%. The latent heat of the composites with 50 wt% KNO3 content was 43 J/g at 300–350 °C, and the thermal conductivity was about 0.68 W/(m·K). After 48 thermal cycles, the latent heat of the composites decreased slightly by 2.5–3.7 J/g compared with that of the composites before thermal cycles, and the thermal conductivity increased by 16–25% [220]. Steady-state phase change materials (SSPCMs) were prepared by the cold compression–hot sintering (CCHS) method using SS and CS as skeletal materials (SM) and NaNO3 as a phase-change material (PCM). The results showed that, when the mass ratio of SS, CS, and NaNO3 was 2.5:2.5:5 (SC4), the mechanical strength of the sample SC4 was 131.2 MPa, and the optimum thermal energy storage density in the temperature range of 100–400 °C was 371.1 J/g. The maximum thermal conductivity of the sample after 1307 heating and cooling cycles was 1.263 W/(m·K) [217]. However, phase-change materials have drawbacks such as poor thermal conductivity, subcooling, and leakage during phase change [221], which are still to be solved.
Thermochemical energy storage (TCES) is currently under development as an alternative to sensible heat for storing energy in photovoltaic power plants. One of its most promising technologies is the calcium loop (CaL) process, which relies on the carbonation/calcination reaction of CaO [222]. In addition to the commonly used limestone, BOF slag is an abundant calcium-based material that can be used to store CSPs via the CaL process. BOF slag has a similar heat capacity to solar salt and has a higher thermal conductivity (1.4 W/(m·K) for BOF slag and 0.52 W/(m·K) for solar salt). Perejón et al. [223] studied the behavior of calcium-rich BOF slag under CaL conditions. When treated with acetic acid, this non-toxic, widely abundant waste produces CaO-rich solids, with conversion rates stabilizing around 0.8 in successive carbonation/calcination cycles under these CaL conditions. The calcination temperature of the regenerated CaO is significantly lower compared to limestone. In addition, the multi-cycle activity of some of the slags tested under the relevant CaL conditions of TCES remained high and stable if the treated samples were filtered. This process is used to remove silica particles, which helps to reduce the porosity of CaO produced by calcination in the mesoporous range, thereby alleviating pore clogging to mitigate inhibition of CaO carbonation properties [224]. Thermochemical thermal storage has a large heat storage density. However, the application of thermochemical thermal storage is immature worldwide, and unregulated application may cause environmental pollution [217].

4. Conclusions and Prospects

The resource properties of steel slag determine its possible uses. Steel slag contains useful elements and has good hardness, good abrasion resistance, and high alkalinity, which are compatible with a wide range of production needs.
However, steel slag has a number of unfavorable characteristics that limit its recycling and reuse. Steel slag has a complex composition, low hydration activity, poor gelation properties, high density, highly crystalline structure, and high iron content, is difficult to grind, and has poor stability and high metallic iron content. Among them, the poor volumetric stability of steel slag and high metal iron content seriously hinder its application in building materials, leading to expansion, cracking, milling difficulties, increased energy consumption, poor particle homogeneity, and rusting issues during processing, as well as the difficulty of milling and the increase in energy consumption, poor particle homogeneity, the use of rust, and other phenomena in the process. Combined with foreign experience in the development of steel slag disposal and China’s national conditions, the use of building materials is the main way to consume steel slag on a large scale, so it is very important to carry out appropriate treatment of steel slag.
China’s solid waste production is high and varied. After pretreatment, steel slag is suitable for the preparation of solid-waste-based cementitious materials using a multi-solid waste synergistic approach. In the context of “double carbon”, the synergistic use of multiple solid wastes has become an important development direction. Under the guidance of the theories of “compound salt effect” and the “four-coordinate isomerization effect of silicon”, solid-waste-based cementitious materials can be produced in a mature manner.
The central role of steel slag in the solid-waste-based gelling material system stems from the high concentration of OH- and Ca2⁺ generated by its hydration reaction, which together build up an alkaline environment and provide a calcium source, laying the foundation for depolymerization of the slag vitreous network and subsequent polymerization of the gel phase (C-S-H, C-A-S-H). To accelerate this process, a key role can be played by the introduction of, e.g., desulfurized gypsum (DG), through a synergistic sulphate–aluminate mechanism: the SO42- in DG reacts with reactive Al2O3 in steel slag/slag to form calcite alumina (AFt), which consumes Al3⁺ in the liquid phase and breaks the [SiO4]4-AlO45-network structure of the slag, significantly reducing the activation energy of depolymerization. The released SiO44- is reconfigured into three-dimensional silica-oxygen chains driven by Ca2⁺, while the symbiosis of AFt and C-S-H forms a “backbone-filler” composite structure, which enhances the compactness of the gel phase. The continuous release of OH- from steel slag maintains the system pH at >12.5, which inhibits the conversion of AFt to monosulfur-type aluminum sulphate (AFm) and ensures the stability of early strength development. This makes the various properties of solid-waste-based cementitious materials close to those of ordinary silicate cement, especially in the two aspects of application in marine concrete and as an underground cementing and filling mining binder, which have more obvious advantages than ordinary silicate cement in terms of performance and cost.
On the other hand, solid-waste-based cementitious materials are still mainly used in road and floor concrete construction. The application of solid-waste-based cementitious materials in housing structures still needs more structural mechanical parameters, as well as demonstration projects to prove its worth. The application of steel slag in the common silicate cement and concrete industry chain is yet to gain further significant advantages in terms of value. The structural durability and safety of concrete prepared from solid-waste-based cementitious materials with ultra-high dosages of steel slag for long-term service is yet to be further studied in depth and confirmed repeatedly.
Resource utilization of steel slag should be aimed at absorbing the huge output of steel slag, drawing on foreign experience, and should be focused in two directions: firstly, the construction and building materials industry, the use of cement, concrete admixture, pavement and building materials products, which not only has a wide scope of application, but also has a large demand, and is capable of absorbing the huge output of steel slag; and secondly, the development of new products of steel slag for the application and improvement of the marine environment, which can be applied to China’s 1.8 million kilometers of coastline and a large number of coastal steel enterprises.
Considering China’s background conditions applicable to the synergistic use of multiple solid wastes, steel slag should be applied more vigorously in marine concrete, and the scale of industrial application should be continuously expanded. In addition, the application of steel slag in the field of construction and building materials should be vigorously researched, research on the effective treatment process of steel slag should be increased, existing technology should be continuously researched and optimized, relevant standards should be established and perfected, acceptance by enterprises and the public of its application should be improved, and an application scheme suitable for China should be pioneered.

Author Contributions

Conceptualization: C.M. and S.Z.; methodology, S.Z. and K.L.; validation, C.M.; formal analysis, C.M., T.Z., Q.M., D.G. and A.Z.; investigation, T.Z., Q.M., D.G. and A.Z.; resources, C.M.; data curation, T.Z., Q.M., D.G. and A.Z.; writing—original draft preparation, C.M.; writing—review and editing, S.Z. and K.L.; visualization, C.M.; supervision, S.Z. and K.L.; project administration, S.Z. and K.L.; funding acquisition, S.Z. and K.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Key Research and Development Program of Hebei Province, grant number 22373802D.

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

Thanks for funding from the Key Research and Development Program of Hebei Province.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
BOFBasic oxygen furnace
SSSteel slag

References

  1. Nunes, V.A.; Borges, P.H.R. Recent advances in the reuse of steel slags and future perspectives as binder and aggregate for alkali-activated materials. Constr. Build. Mater. 2021, 281, 122605. [Google Scholar] [CrossRef]
  2. Analysis of Steel Slag Market Prospect in 2024: China’s Steel Slag Output Will Increase to 160 Million Tons. 2024. (In Chinese). Available online: https://m.chinabgao.com/freereport/97810.html (accessed on 22 March 2025).
  3. Yang, H.; Ma, L.; Li, Z. A Method for Analyzing Energy-Related Carbon Emissions and the Structural Changes; World Steel Statistical Yearbook 2020; World Steel Association: Brussels, Belgium, 2020. [Google Scholar]
  4. O’Connor, J.; Nguyen, T.B.T.; Honeyands, T.; Monaghan, B.; O’Dea, D.; Rinklebe, J.; Vinu, A.; Hoang, S.A.; Singh, G.; Kirkham, M.B.; et al. Production, characterisation, utilisation, and beneficial soil application of steel slag: A review. J. Hazard. Mater. 2021, 419, 126478. [Google Scholar] [CrossRef]
  5. Guo, J.; Bao, Y.; Wang, M. Steel slag in China: Treatment, recycling, and management. Waste Manag. 2018, 78, 318–330. [Google Scholar] [CrossRef] [PubMed]
  6. National Committee of the Chinese People’s Political Consultative Conference. Suggestion to Promote Comprehensive Utilization of Steel Slag Resources. 2022. Available online: http://www.chinadaily.com.cn/ (accessed on 16 June 2022).
  7. Chinese People’s Political Consultative Conference. Iron and Steel Slag in 2017. 2020. Available online: https://d9-wret.s3.us-west-2.amazonaws.com/assets/palladium/production/s3fs-public/atoms/files/myb1-2017-fesla.pdf (accessed on 22 February 2024).
  8. Zhang, Y.; Ying, Y.; Xing, L.; Zhan, G.; Deng, Y.; Chen, Z.; Li, J. Carbon dioxide reduction through mineral carbonation by steel slag. J. Environ. Sci. 2025, 152, 664–684. [Google Scholar] [CrossRef]
  9. Nippon Slag Association. Production and Uses of Steel Slag in Japan; Nippon Slag Association: Sapporo City, Japan, 2023. [Google Scholar]
  10. Liu, K.; Zhao, H.; Yuan, Z.; Zhao, F.; Chen, D.; Shi, C.; Li, X.; Wang, S.; Wang, X.; Zhang, X. Preparation and characterization of steel slag-based low, medium, and high-temperature composite phase change energy storage materials. J. Energy Storage 2023, 57, 106309. [Google Scholar] [CrossRef]
  11. Fu, S.; Kwon, E.E.; Lee, J. Upcycling steel slag into construction materials. Constr. Build. Mater. 2024, 444, 137882. [Google Scholar] [CrossRef]
  12. Zhao, J.; Yan, P.; Wang, D. Research on mineral characteristics of converter steel slag and its comprehensive utilization of internal and external recycle. J. Clean. Prod. 2017, 156, 50–61. [Google Scholar] [CrossRef]
  13. Wang, G.C. (Ed.) 2—Ferrous metal production and ferrous slags. In The Utilization of Slag in Civil Infrastructure Construction; Woodhead Publishing: Cambridge, UK, 2016; pp. 9–33. [Google Scholar]
  14. Shu, K.; Sasaki, K. Occurrence of steel converter slag and its high value-added conversion for environmental restoration in China: A review. J. Clean. Prod. 2022, 373, 133876. [Google Scholar] [CrossRef]
  15. Pan, S.-Y.; Ye, T.; Alwasiyah, S.K.; de Lasa, H.I.; Li, C.Z.; Lin, C.S.; Lu, J.Y.; Tang, G.L.; Wang, Y.Z.; Zhang, Q. Integrated and innovative steel slag utilization for iron reclamation, green material production and CO2 fixation via accelerated carbonation. J. Clean. Prod. 2016, 137, 617–631. [Google Scholar] [CrossRef]
  16. BS EN 1097-7; Tests for Mechanical and Physical Properties of Aggregates—Part 7: Determination of the Particle Density of Filler—Pycnometer Method. British Standards Institution (BSI): London, UK, 2022.
  17. ASTM D6473; Standard Test Method for Specific Gravity and Absorption of Rock for Erosion Control. ASTM International: West Conshohocken, PA, USA, 2024.
  18. ASTM C535; Standard Test Method for Resistance to Degradation of Large-Size Coarse Aggregate by Abrasion and Impact in the Los Angeles Machine. ASTM International: West Conshohocken, PA, USA, 2024.
  19. EN 1097-1; Tests for Mechanical and Physical Properties of Aggregates—Part 1: Determination of the Resistance to Wear (Micro-Deval). European Committee for Standardization (CEN): Brussels, Belgium, 2021.
  20. Satish Kumar, D.; Reddy, B.V.; Zhao, C.; Kim, H.J.; Kim, H.; Lee, J.K.; Kim, D.S.; Shin, H.S.; Park, S.J. Measurement of metallic iron in steel making slags. Measurement 2019, 131, 156–161. [Google Scholar] [CrossRef]
  21. Chang, Y.; Zhao, Z.; Zhao, D.; Zhang, D.; Xue, L. Co-treatment of steel slag and oil shale waste in cemented paste backfill: Evaluation of fresh properties, microstructure, and heavy metals immobilization. J. Environ. Manag. 2024, 349, 119406. [Google Scholar] [CrossRef]
  22. Zhao, Y.; Wu, P.; Qiu, J.; Guo, Z.; Sun, X.; Gu, X. Recycling hazardous steel slag after thermal treatment to produce a binder for cemented paste backfill. Powder Technol. 2022, 395, 652–662. [Google Scholar] [CrossRef]
  23. Li, W.; Cao, M.; Wang, D.; Chang, J. Increase in volume stability of RO phases in steel slag by combined treatment of alkali and dry carbonation. Constr. Build. Mater. 2023, 396, 132345. [Google Scholar] [CrossRef]
  24. Chen, Z.; Li, W.; Cao, M. Coupling effect of γ-dicalcium silicate and slag on carbonation resistance of low carbon materials. J. Clean. Prod. 2020, 262, 121385. [Google Scholar] [CrossRef]
  25. Zhao, X.; Hou, J.; Chen, Z.; Liu, J. Hydration activity and expansibility of different divalent metal oxide solution compositions in steel slag. Adv. Cem. Res. 2023, 35, 408–418. [Google Scholar] [CrossRef]
  26. Prateek, G.; Zheng, K.; Gong, C.; Wei, S.; Zhou, X.; Yuan, Q. Preparing high strength cementitious materials with high proportion of steel slag through reverse filling approach. Constr. Build. Mater. 2023, 368, 130474. [Google Scholar] [CrossRef]
  27. Kim, J.-M.; Choi, S.-M.; Han, D. Improving the mechanical properties of rapid air cooled ladle furnace slag powder by gypsum. Constr. Build. Mater. 2016, 127, 93–101. [Google Scholar] [CrossRef]
  28. Zhuang, S.; Wang, Q. Inhibition mechanisms of steel slag on the early-age hydration of cement. Cem. Concr. Res. 2021, 140, 106283. [Google Scholar] [CrossRef]
  29. Atia, B.M.; Radwan, H.A.; Kassab, W.A.; Sarhan, H.K.A.; Gado, M.A.; Goda, A.E. Highly efficient recovery of vanadium from Abu Zeneima ferruginous siltstone, Southwestern Sinai, Egypt, by a novel polyimine-based chelating ligand. J. Chin. Chem. Soc. 2024, 71, 507–522. [Google Scholar] [CrossRef]
  30. Gao, W.; Zhou, W.; Lyu, X.; Liu, X.; Su, H.; Li, C.; Wang, H. Comprehensive utilization of steel slag: A review. Powder Technol. 2023, 422, 118449. [Google Scholar] [CrossRef]
  31. Qian, Q. Analysis of rock facies structure of converter slag under different processing methods. Sichuan Metall. 2019, 41, 48–51. (In Chinese) [Google Scholar]
  32. Diao, J.; Zhou, W.; Ke, Z.; Qiao, Y.; Zhang, T.; Liu, X.; Xie, B. System assessment of recycling of steel slag in converter steelmaking. J. Clean. Prod. 2016, 125, 159–167. [Google Scholar] [CrossRef]
  33. Lan, Y.-P.; Liu, Q.-C.; Meng, F.; Niu, D.-L.; Zhao, H. Optimization of magnetic separation process for iron recovery from steel slag. J. Iron Steel Res. Int. 2017, 24, 165–170. [Google Scholar] [CrossRef]
  34. Fisher, L.V.; Barron, A.R. The recycling and reuse of steelmaking slags—A review. Resour. Conserv. Recycl. 2019, 146, 244–255. [Google Scholar] [CrossRef]
  35. Li, P.; Guo, H.; Gao, J.; Min, J.; Yan, B.; Chen, D.; Seetharaman, S. Novel concept of steam modification towards energy and iron recovery from steel slag: Oxidation mechanism and process evaluation. J. Clean. Prod. 2020, 254, 119952. [Google Scholar] [CrossRef]
  36. 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]
  37. Xue, P.; He, D.; Xu, A.; Gu, Z.; Yang, Q.; Engström, F.; Björkman, B. Modification of industrial BOF slag: Formation of MgFe2O4 and recycling of iron. J. Alloys Compd. 2017, 712, 640–648. [Google Scholar] [CrossRef]
  38. Liu, Z.; Bi, X.; Gao, Z.; Liu, W. Carbothermal Reduction of Iron Ore in Its Concentrate-Agricultural Waste Pellets. Adv. Mater. Sci. Eng. 2018, 2018, 1–6. [Google Scholar] [CrossRef]
  39. Agrawal, S.; Dhawan, N. Microwave Carbothermic Reduction of Low-Grade Iron Ore. Metall. Mater. Trans. B 2020, 51, 1576–1586. [Google Scholar] [CrossRef]
  40. Shekhar Samanta, N.; Das, P.P.; Dhara, S.; Purkait, M.K. An Overview of Precious Metal Recovery from Steel Industry Slag: Recovery Strategy and Utilization. Ind. Eng. Chem. Res. 2023, 62, 9006–9031. [Google Scholar] [CrossRef]
  41. Das, P.; Upadhyay, S.; Dubey, S.; Singh, K.K. Waste to wealth: Recovery of value-added products from steel slag. J. Environ. Chem. Eng. 2021, 9, 105640. [Google Scholar] [CrossRef]
  42. Mohamed Noor, N.; Ismail, A.K.; Jaafar, J.; Zakaria, N.A.; Abdullah, R. Heavy metal recovery from electric arc furnace steel slag by using hydrochloric acid leaching. In Proceedings of the E3S Web of Conferences, Penang, Malaysia, 28–29 November 2017; EDP Sciences: Les Ulis, France, 2018; Volume 34. [Google Scholar]
  43. Liu, X.; Wang, D.; Li, Z.; Ouyang, W.; Bao, Y.; Gu, C. Efficient separation of iron elements from steel slag based on magnetic separation process. J. Mater. Res. Technol. 2023, 23, 2362–2370. [Google Scholar] [CrossRef]
  44. Gomes, H.I.; Jones, A.; Rogerson, M.; Greenway, G.M.; Lisbona, D.F.; Burke, I.T.; Mayes, W.M. Removal and recovery of vanadium from alkaline steel slag leachates with anion exchange resins. J. Environ. Manag. 2017, 187, 384–392. [Google Scholar] [CrossRef]
  45. Chen, B.; Zhang, L.; Li, W.; Liu, Y.; Wang, S. Separation and recovery of vanadium and chromium from acidic leach solution of V-Cr-bearing reducing slag. J. Environ. Chem. Eng. 2017, 5, 4702–4706. [Google Scholar] [CrossRef]
  46. Xiang, J.; Zhang, L.; Liu, Y.; Wang, S.; Chen, B. Extraction of vanadium from converter slag by two-step sulfuric acid leaching process. J. Clean. Prod. 2018, 170, 1089–1101. [Google Scholar] [CrossRef]
  47. Ju, J.; Feng, Y.; Li, H.; Xu, C.; Yang, Y. Efficient Separation and Recovery of Vanadium, Titanium, Iron, Magnesium, and Synthesizing Anhydrite from Steel Slag. Min. Metall. Explor. 2022, 39, 733–748. [Google Scholar] [CrossRef]
  48. Lee, J.C.; Kim, E.Y.; Chung, K.W.; Kim, R.; Jeon, H.S. A review on the metallurgical recycling of vanadium from slags: Towards a sustainable vanadium production. J. Mater. Res. Technol. 2021, 12, 343–364. [Google Scholar] [CrossRef]
  49. Yang, M.-Q.; Yang, J.-Y. Vanadium extraction from steel slag: Generation, recycling and management. Environ. Pollut. 2023, 343, 123126. [Google Scholar] [CrossRef]
  50. An, Y.; Ma, B.; Li, X.; Chen, Y.; Wang, C.; Gao, M.; Feng, G. A review on the roasting-assisted leaching and recovery of V from vanadium slag. Process Saf. Environ. Prot. 2023, 173, 263–276. [Google Scholar] [CrossRef]
  51. Gao, F.; Zhou, J. Semiclassical states for critical choquard equations with critical frequency. Topol. Methods Nonlinear Anal. 2021, 57, 107–133. [Google Scholar] [CrossRef]
  52. Cai, Z.; Zhang, Y. Phase transformations of vanadium recovery from refractory stone coal by novel NaOH molten roasting and water leaching technology. RSC Adv. 2017, 7, 36917–36922. [Google Scholar] [CrossRef]
  53. Ji, Y.; Shen, S.; Liu, J.; Xue, Y. Cleaner and effective process for extracting vanadium from vanadium slag by using an innovative three-phase roasting reaction. J. Clean. Prod. 2017, 149, 1068–1078. [Google Scholar] [CrossRef]
  54. Wen, J.; Jiang, T.; Zhou, M.; Gao, H.; Liu, J.; Xue, X. Roasting and leaching behaviors of vanadium and chromium in calcification roasting–acid leaching of high-chromium vanadium slag. Int. J. Miner. Metall. Mater. 2018, 25, 515–526. [Google Scholar] [CrossRef]
  55. Zhang, J.; Zhang, W.; Xue, Z. Oxidation Kinetics of Vanadium Slag Roasting in the Presence of Calcium Oxide. Miner. Process. Extr. Metall. Rev. 2017, 38, 265–273. [Google Scholar] [CrossRef]
  56. Xiang, J.; Huang, Q.; Lv, X.; Bai, C. Effect of Mechanical Activation Treatment on the Recovery of Vanadium from Converter Slag. Metall. Mater. Trans. B: Process Metall. Mater. Process. Sci. 2017, 48, 2759–2767. [Google Scholar] [CrossRef]
  57. Tian, L.; Xu, Z.; Chen, L.; Liu, Y.; Zhang, T. Effect of microwave heating on the pressure leaching of vanadium from converter slag. Hydrometallurgy 2019, 184, 45–54. [Google Scholar] [CrossRef]
  58. Wen, J.; Jiang, T.; Gao, H.; Liu, Y.; Zheng, X.; Xue, X. Comparison of Ultrasound-Assisted and Regular Leaching of Vanadium and Chromium from Roasted High Chromium Vanadium Slag. JOM. 2018, 70, 155–160. [Google Scholar] [CrossRef]
  59. Kim, E.; Spooren, J.; Broos, K.; Nielsen, P.; Horckmans, L.; Vrancken, K.C.; Quaghebeur, M. New method for selective Cr recovery from stainless steel slag by NaOCl assisted alkaline leaching and consecutive BaCrO4 precipitation. Chem. Eng. J. 2016, 295, 542–551. [Google Scholar] [CrossRef]
  60. Kim, E.; Spooren, J.; Broos, K.; Horckmans, L.; Quaghebeur, M. Selective recovery of Cr from stainless steel slag by alkaline roasting followed by water leaching. Hydrometallurgy 2015, 158, 139–148. [Google Scholar] [CrossRef]
  61. Duan, S.; Liao, H.; Cheng, F.; Song, H.; Yang, H. Investigation into the synergistic effects in hydrated gelling systems containing fly ash, desulfurization gypsum and steel slag. Constr. Build. Mater. 2018, 187, 1113–1120. [Google Scholar] [CrossRef]
  62. Wang, X.; Wang, H.; Yang, M.; Meng, Y.; Fu, Z.; Wang, M. A novel technology for the production of crystal Cr2O3 with V-Cr-bearing reducing slag. J. Environ. Chem. Eng. 2020, 8, 103799. [Google Scholar] [CrossRef]
  63. Du, C.; Gao, X.; Ueda, S.; Kitamura, S. Selective Leaching of P from Steelmaking Slag in Sulfuric Acid Solution. J. Sustain. Metall. 2019, 5, 594–605. [Google Scholar] [CrossRef]
  64. Benhelal, E.; Shamsaei, E.; Rashid, M.I. Challenges against CO2 abatement strategies in cement industry: A review. J. Environ. Sci. 2021, 104, 84–101. [Google Scholar] [CrossRef]
  65. Li, W.; Gao, S. Prospective on energy related carbon emissions peak integrating optimized intelligent algorithm with dry process technique application for China’s cement industry. Energy 2018, 165, 33–54. [Google Scholar] [CrossRef]
  66. Zhang, L.; Mabee, W.E. Comparative study on the life-cycle greenhouse gas emissions of the utilization of potential low carbon fuels for the cement industry. J. Clean. Prod. 2016, 122, 102–112. [Google Scholar] [CrossRef]
  67. Nair, N.; Mohammed Haneefa, K.; Santhanam, M.; Gettu, R. A study on fresh properties of limestone calcined clay blended cementitious systems. Constr. Build. Mater. 2020, 254, 119326. [Google Scholar] [CrossRef]
  68. Sui, S.; Shan, Y.; Li, S.; Geng, Y.; Wang, F.; Liu, Z.; Jiang, J.; Wang, L.; Yang, Z. Investigation on chloride migration behavior of metakaolin-quartz-limestone blended cementitious materials with electrochemical impedance spectroscopy method. Case Stud. Constr. Mater. 2024, 20, e03064. [Google Scholar] [CrossRef]
  69. Srivastava, S.; Snellings, R.; Cool, P. Clinker-free carbonate-bonded (CFCB) products prepared by accelerated carbonation of steel furnace slags: A parametric overview of the process development. Constr. Build. Mater. 2021, 303, 124556. [Google Scholar] [CrossRef]
  70. Liu, J.; Yu, B.; Wang, Q. Application of steel slag in cement treated aggregate base course. J. Clean. Prod. 2020, 269, 121733. [Google Scholar] [CrossRef]
  71. Saxena, S.; Tembhurkar, A.R. Impact of use of steel slag as coarse aggregate and wastewater on fresh and hardened properties of concrete. Constr. Build. Mater. 2018, 165, 126–137. [Google Scholar] [CrossRef]
  72. Wang, J.; Deng, X.; Tan, H.; Guo, H.; Zhang, J.; Li, M.; Chen, P.; He, X.; Yang, J.; Jian, S.; et al. Effect of direct electric curing on the mechanical properties, hydration process, and environmental benefits of cement-steel slag composite. Constr. Build. Mater. 2023, 406, 133382. [Google Scholar] [CrossRef]
  73. Xing, Y.; Wang, B. Modification of phases evolution and heavy metal immobilization in alkali-activated MSWI FA by the incorporation of converter steel slag. J. Build. Eng. 2023, 78, 107573. [Google Scholar] [CrossRef]
  74. Yang, J.; Lu, J.; Wu, Q.; Xia, M.F.; Li, X. Influence of steel slag powders on the properties of MKPC paste. Constr. Build. Mater. 2018, 159, 137–146. [Google Scholar] [CrossRef]
  75. Li, L.; Ye, T.; Li, Y.; Wang, Z.; Feng, Z.; Liu, Z. Improvement of microporous structure and impermeability of cement mortars using fly ash and blast furnace slag under low curing pressures. Constr. Build. Mater. 2023, 401, 132890. [Google Scholar] [CrossRef]
  76. Mardmomen, S.; Chen, H.-L. Modeling the thermal and mechanical properties of early age concrete containing ground granulated blast furnace slag. Constr. Build. Mater. 2023, 401, 132902. [Google Scholar] [CrossRef]
  77. Wang, J.; Deng, X.; Tan, H.; Guo, H.; Zhang, J.; Li, M.; Chen, P.; He, X.; Yang, J.; Jian, S.; et al. The mechanical properties and sustainability of phosphogypsum-slag binder activated by nano-ettringite. Sci. Total Environ. 2023, 903, 166015. [Google Scholar] [CrossRef]
  78. Ma, X.; Yao, B.; Zou, J.; Ho, J.C.M.; Zhuang, X.; Wang, Q. Sodium gluconate as a retarder modified sewage sludge ash-based geopolymers: Mechanism and environmental assessment. J. Clean. Prod. 2023, 419, 138317. [Google Scholar] [CrossRef]
  79. Zhu, H.; Ma, M.; He, X.; Zheng, Z.; Su, Y.; Yang, J.; Zhao, H. Effect of wet-grinding steel slag on the properties of Portland cement: An activated method and rheology analysis. Constr. Build. Mater. 2021, 286, 122823. [Google Scholar] [CrossRef]
  80. Song, W.; Zhu, Z.; Pu, S.; Wan, Y.; Huo, W.; Song, S.; Zhang, J.; Yao, K.; Hu, L. Efficient use of steel slag in alkali-activated fly ash-steel slag-ground granulated blast furnace slag ternary blends. Constr. Build. Mater. 2020, 259, 119814. [Google Scholar] [CrossRef]
  81. Amran, M.; Murali, G.; Khalid, N.H.A.; Fediuk, R.; Ozbakkaloglu, T.; Lee, Y.H.; Haruna, S.; Lee, Y.Y. Slag uses in making an ecofriendly and sustainable concrete: A review. Constr. Build. Mater. 2021, 272, 121942. [Google Scholar] [CrossRef]
  82. Liu, X.; Zhao, X.; Yin, H.; Chen, J.; Zhang, N. Intermediate-calcium based cementitious materials prepared by MSWI fly ash and other solid wastes: Hydration characteristics and heavy metals solidification behavior. J. Hazard. Mater. 2018, 349, 262–271. [Google Scholar] [CrossRef] [PubMed]
  83. Hao, X.; Liu, Z.; Lyu, X.; Wang, X.; Liu, J.; Zhang, T. In-depth insight into the cementitious synergistic effect of steel slag and red mud on the properties of composite cementitious materials. J. Build. Eng. 2022, 52, 104449. [Google Scholar] [CrossRef]
  84. Liu, Z.; Ni, W.; Li, Y.; Ba, H.; Li, N.; Ju, Y.; Zhao, B.; Jia, G.; Hu, W. The mechanism of hydration reaction of granulated blast furnace slag-steel slag-refining slag-desulfurization gypsum-based clinker-free cementitious materials. J. Build. Eng. 2021, 44, 103289. [Google Scholar] [CrossRef]
  85. Wu, Z.; Peng, Z.; Huang, Y.; Gu, F.; Zhang, J.; Tang, H.; Yang, L.; Tian, W.; Rao, M.; Li, G.; et al. Mechanical properties and environmental characteristics of the synergistic preparation of cementitious materials using electrolytic manganese residue, steel slag, and blast furnace slag. Constr. Build. Mater. 2024, 411, 134480. [Google Scholar] [CrossRef]
  86. Zhao, T.; Zhang, S.; Yang, H.; Ni, W.; Li, J.; Zhang, G.; Teng, G. Influence on fine lead–zinc tailings solidified/stabilised by clinker-free slag-based binder. J. Environ. Chem. Eng. 2022, 10, 108692. [Google Scholar] [CrossRef]
  87. Du, H.; Xu, D.; Li, X.; Li, J.; Ni, W.; Li, Y.; Fu, P. Application of molten iron desulfurization slag to replace steel slag as an alkaline component in solid waste-based cementitious materials. J. Clean. Prod. 2022, 377, 134353. [Google Scholar] [CrossRef]
  88. Gao, T.; Dai, T.; Shen, L.; Jiang, L. Benefits of using steel slag in cement clinker production for environmental conservation and economic revenue generation. J. Clean. Prod. 2021, 282, 124538. [Google Scholar] [CrossRef]
  89. Zvironaite, J.; Pranckeviciene, J.; Kligys, M.; Balciunas, G.; Macanovskis, A. Alkali-silica reactivity of portland-composite cements and gravel aggregates. Ceramics 2019, 63, 76–85. [Google Scholar] [CrossRef]
  90. Shehata, N.; Sayed, E.T.; Abdelkareem, M.A. Recent progress in environmentally friendly geopolymers: A review. Sci. Total Environ. 2021, 762, 143166. [Google Scholar] [CrossRef]
  91. Elakneswaran, Y.; Noguchi, N.; Matumoto, K.; Morinaga, Y.; Chabayashi, T.; Kato, H.; Nawa, T. Characteristics of ferrite-rich portland cement: Comparison with ordinary portland cement. Front. Mater. 2019, 6, 97. [Google Scholar] [CrossRef]
  92. Taimasov, B.T.; Sarsenbayev, B.K.; Khudyakova, T.M.; Kolesnikov, A.S.; Zhanikulov, N.N. Development and testing of low-energy-intensive technology of receiving sulphate-resistant and road portlandcement. Eurasian Chem.-Technol. J. 2017, 19, 347–355. [Google Scholar] [CrossRef]
  93. Verma, Y.K.; Mazumdar, B.; Ghosh, P. CO2 emission reduction using blast furnace slag for the clinker manufacturing in cement industry. J. Indian Chem. Soc. 2020, 97, 1083–1087. [Google Scholar]
  94. Dai, S.; Zhu, H.; Zhang, D.; Liu, Z.; Cheng, S.; Zhao, J. Insights to compressive strength, impermeability and microstructure of micro-expansion steel slag cement under constraint conditions. Constr. Build. Mater. 2022, 326, 126540. [Google Scholar] [CrossRef]
  95. Jiang, Y.; Ahmad, M.R.; Chen, B. Properties of magnesium phosphate cement containing steel slag powder. Constr. Build. Mater. 2019, 195, 140–147. [Google Scholar] [CrossRef]
  96. Ma, Y.; Luo, Y.; Ma, H.; Zhou, X.; Luo, Z. Upcycling steel slag in producing eco-efficient iron–calcium phosphate cement. J. Clean. Prod. 2022, 371, 133688. [Google Scholar] [CrossRef]
  97. Yu, X.; Chu, J.; Wu, S.; Wang, K. Production of biocement using steel slag. Constr. Build. Mater. 2023, 383, 131365. [Google Scholar] [CrossRef]
  98. Agrawal, U.S.; Wanjari, S.P.; Naresh, D.N. Characteristic study of geopolymer fly ash sand as a replacement to natural river sand. Constr. Build. Mater. 2017, 150, 681–688. [Google Scholar] [CrossRef]
  99. Xiao, J.; Qiang, C.; Nanni, A.; Zhang, K. Use of sea-sand and seawater in concrete construction: Current status and future opportunities. Constr. Build. Mater. 2017, 155, 1101–1111. [Google Scholar] [CrossRef]
  100. Yu, H.; Xu, P.; Zhu, Y.; Wan, X.; Zhu, H. Effect of Carbonated Steel Slag as Sand Replacement on Strength and Durability of Cement Mortar. J. Qingdao Technol. Univ. 2023, 44, 20–28. (In Chinese) [Google Scholar]
  101. Gencel, O.; Karadag, O.; Oren, O.H.; Bilir, T. Steel slag and its applications in cement and concrete technology: A review. Constr. Build. Mater. 2021, 283, 122783. [Google Scholar] [CrossRef]
  102. Pan, Z.; Zhou, J.; Jiang, X.; Xu, Y.; Jin, R.; Ma, J.; Zhuang, Y.; Diao, Z.; Zhang, S.; Si, Q.; et al. Investigating the effects of steel slag powder on the properties of self-compacting concrete with recycled aggregates. Constr. Build. Mater. 2019, 200, 570–577. [Google Scholar] [CrossRef]
  103. Shi, Y.; Chen, H.; Wang, J.; Feng, Q. Preliminary investigation on the pozzolanic activity of superfine steel slag. Constr. Build. Mater. 2015, 82, 227–234. [Google Scholar] [CrossRef]
  104. Liu, G.; Schollbach, K.; van der Laan, S.; Tang, P.; Florea, M.V.A.; Brouwers, H.J.H. Recycling and utilization of high volume converter steel slag into CO2 activated mortars—The role of slag particle size. Resour. Conserv. Recycl. 2020, 160, 104883. [Google Scholar] [CrossRef]
  105. Zhang, Z.; Wang, P.; Liu, J.; Wang, X.; Bai, Z. Effect of Fe and Mn on the hydration activity of f-CaO in steel slag. Constr. Build. Mater. 2024, 421, 135719. [Google Scholar] [CrossRef]
  106. Li, W.; Cao, M.; Wang, D.; Zhao, J.; Chang, J. Improving the hydration activity and volume stability of the RO phases in steel slag by combining alkali and wet carbonation treatments. Cem. Concr. Res. 2023, 172, 107236. [Google Scholar] [CrossRef]
  107. Kong, Y.; Wang, P.; Liu, S. Microwave pre-curing of Portland cement-steel slag powder composite for its hydration properties. Constr. Build. Mater. 2018, 189, 1093–1104. [Google Scholar] [CrossRef]
  108. Zhang, S.; Niu, D. Hydration and mechanical properties of cement-steel slag system incorporating different activators. Constr. Build. Mater. 2023, 363, 129981. [Google Scholar] [CrossRef]
  109. Zhang, S.; Niu, D.; Luo, D. Enhanced hydration and mechanical properties of cement-based materials with steel slag modified by water glass. J. Mater. Res. Technol. 2022, 21, 1830–1842. [Google Scholar] [CrossRef]
  110. Tan, H.; Nie, K.; He, X.; Guo, Y.; Zhang, X.; Deng, X.; Su, Y.; Yang, J. Effect of organic alkali on compressive strength and hydration of wet-grinded granulated blast-furnace slag containing Portland cement. Constr. Build. Mater. 2019, 206, 10–18. [Google Scholar] [CrossRef]
  111. Polaczyk, P.; Ma, Y.; Xiao, R.; Hu, W.; Jiang, X.; Huang, B. Characterization of aggregate interlocking in hot mix asphalt by mechanistic performance tests. Road Mater. Pavement Des. 2021, 22, S498–S513. [Google Scholar] [CrossRef]
  112. Polaczyk, P.; Huang, B.; Shu, X.; Gong, H. Investigation into Locking Point of Asphalt Mixtures Utilizing Superpave and Marshall Compactors. J. Mater. Civ. Eng. 2019, 31, 04019188. [Google Scholar] [CrossRef]
  113. Rodríguez-Fernández, I.; Lastra-González, P.; Indacoechea-Vega, I.; Castro-Fresno, D. Recyclability potential of asphalt mixes containing reclaimed asphalt pavement and industrial by-products. Constr. Build. Mater. 2019, 195, 148–155. [Google Scholar] [CrossRef]
  114. Motevalizadeh, S.M.; Sedghi, R.; Rooholamini, H. Fracture properties of asphalt mixtures containing electric arc furnace slag at low and intermediate temperatures. Constr. Build. Mater. 2020, 240, 117965. [Google Scholar] [CrossRef]
  115. Liu, H.; Zhu, B.; Wei, H.; Chai, C.; Chen, Y. Laboratory evaluation on the performance of porous asphalt mixture with steel slag for seasonal frozen regions. Sustainability 2019, 11, 6924. [Google Scholar] [CrossRef]
  116. Zhang, T.; Wu, J.; Hong, R.; Ye, S.; Jin, A. Research on low-temperature performance of steel slag/polyester fiber permeable asphalt mixture. Constr. Build. Mater. 2022, 334, 127214. [Google Scholar] [CrossRef]
  117. Liu, J.; Wang, X.; Li, M.; Wang, Z.; Zhang, T. Microwave heating uniformity, road performance and internal void characteristics of steel slag asphalt mixtures. Constr. Build. Mater. 2022, 353, 129155. [Google Scholar] [CrossRef]
  118. Su, J.F.; Wang, L.Q.; Xie, X.M.; Gao, X. Understanding the final surface state of self-healing microcapsules containing rejuvenator in bituminous binder of asphalt. Colloids Surf. A Physicochem. Eng. Asp. 2021, 615, 126287. [Google Scholar] [CrossRef]
  119. Wang, H.; Zhang, Y.; Zhang, Y.; Feng, S.; Lu, G.; Cao, L. Laboratory and Numerical Investigation of Microwave Heating Properties of Asphalt Mixture. Materials 2019, 12, 146. [Google Scholar] [CrossRef]
  120. Liu, Q.; Li, B.; Schlangen, E.; Sun, Y. Research on the mechanical, thermal, induction heating and healing properties of steel slag/steel fibers composite asphalt mixture. Appl. Sci. 2017, 7, 1088. [Google Scholar] [CrossRef]
  121. Xu, H.; Wu, S.; Li, H.; Zhao, Y.; Lv, Y. Study on recycling of steel slags used as coarse and fine aggregates in induction healing asphalt concretes. Materials 2020, 13, 889. [Google Scholar] [CrossRef]
  122. Lou, B.; Sha, A.; Barbieri, D.M.; Liu, Z.; Zhang, F. Microwave heating properties of steel slag asphalt mixture using a coupled electromagnetic and heat transfer model. Constr. Build. Mater. 2021, 291, 123248. [Google Scholar] [CrossRef]
  123. Lou, B.; Liu, Z.; Sha, A.; Jia, M.; Li, Y. Microwave absorption ability of steel slag and road performance of asphalt mixtures incorporating steel slag. Materials 2020, 13, 663. [Google Scholar] [CrossRef] [PubMed]
  124. Zhang, F.; Wang, D.; Cannone Falchetto, A.; Cao, Y. Microwave deicing properties and carbon emissions assessment of asphalt mixtures containing steel slag towards resource conservation and waste reuse. Sci. Total Environ. 2024, 912, 169189. [Google Scholar] [CrossRef] [PubMed]
  125. Liu, J.; Zhang, T.; Guo, H.; Wang, Z.; Wang, X. Evaluation of self-healing properties of asphalt mixture containing steel slag under microwave heating: Mechanical, thermal transfer and voids microstructural characteristics. J. Clean. Prod. 2022, 342, 130932. [Google Scholar] [CrossRef]
  126. Wan, J.; Wu, S.; Xiao, Y.; Chen, Z.; Zhang, D. Study on the effective composition of steel slag for asphalt mixture induction heating purpose. Constr. Build. Mater. 2018, 178, 542–550. [Google Scholar] [CrossRef]
  127. Yang, C.; Wu, S.; Xie, J.; Amirkhanian, S.; Liu, Q.; Zhang, J.; Xiao, Y.; Zhao, Z.; Xu, H.; Li, N.; et al. Enhanced induction heating and self-healing performance of recycled asphalt mixtures by incorporating steel slag. J. Clean. Prod. 2022, 366, 132999. [Google Scholar] [CrossRef]
  128. Fakhri, M.; Ahmadi, A. Evaluation of fracture resistance of asphalt mixes involving steel slag and RAP: Susceptibility to aging level and freeze and thaw cycles. Constr. Build. Mater. 2017, 157, 748–756. [Google Scholar] [CrossRef]
  129. Pasetto, M.; Baldo, N. Dissipated energy analysis of four-point bending test on asphalt concretes made with steel slag and RAP. Int. J. Pavement Res. Technol. 2017, 10, 446–453. [Google Scholar] [CrossRef]
  130. Faleschini, F.; Fernández-Ruíz, M.A.; Zanini, M.A.; Brunelli, K.; Pellegrino, C.; Hernández-Montes, E. High performance concrete with electric arc furnace slag as aggregate: Mechanical and durability properties. Constr. Build. Mater. 2015, 101, 113–121. [Google Scholar] [CrossRef]
  131. Lai, M.H.; Zou, J.; Yao, B.; Ho, J.C.M.; Zhuang, X.; Wang, Q. Improving mechanical behavior and microstructure of concrete by using BOF steel slag aggregate. Constr. Build. Mater. 2021, 277, 122269. [Google Scholar] [CrossRef]
  132. Song, Q.; Guo, M.; Wang, L.; Ling, T. Use of steel slag as sustainable construction materials: A review of accelerated carbonation treatment. Resour. Conserv. Recycl. 2021, 173, 105740. [Google Scholar] [CrossRef]
  133. Jung, S.; Chau, T.V.; Kim, M.; Na, W.-B. Artificial Seaweed Reefs That Support the Establishment of Submerged Aquatic Vegetation Beds and Facilitate Ocean Macroalgal Afforestation: A Review. J. Mar. Sci. Eng. 2022, 10, 1184. [Google Scholar] [CrossRef]
  134. Tsiamis, K.; Salomidi, M.; Gerakaris, V.; Mogg, A.O.M.; Porter, E.S.; Sayer, M.D.J.; Küpper, F.C. Macroalgal vegetation on a north European artificial reef (Loch Linnhe, Scotland): Biodiversity, community types and role of abiotic factors. J. Appl. Phycol. 2020, 32, 1353–1363. [Google Scholar] [CrossRef]
  135. Baalamurugan, J.; Ganesh Kumar, V.; Naveen Prasad, B.S.N.; Padmapriya, R.; Karthick, V.; Govindaraju, K. Recycling of induction furnace steel slag in concrete for marine environmental applications towards ocean acidification studies. Int. J. Environ. Sci. Technol. 2021, 19, 5039–5048. [Google Scholar] [CrossRef]
  136. Huang, X.; Wang, Z.; Liu, Y.; Hu, W.; Ni, W. On the use of blast furnace slag and steel slag in the preparation of green artificial reef concrete. Constr. Build. Mater. 2016, 112, 241–246. [Google Scholar] [CrossRef]
  137. Layman, C.A.; Allgeier, J.E.; Lemasson, A. An ecosystem ecology perspective on artificial reef production. J. Appl. Ecol. 2020, 57, 2139–2148. [Google Scholar] [CrossRef]
  138. Park, S.; Kim, J.R.; Kim, Y.R.; Yoon, S.; Kim, K. Assessment of Heavy Metals Eluted from Materials Utilized in Artificial Reefs Implemented in South Korea. J. Mar. Sci. Eng. 2022, 10, 1720. [Google Scholar] [CrossRef]
  139. Roslan, N.H.; Ismail, M.; Abdul-Majid, Z.; Ghoreishiamiri, S.; Muhammad, B. Performance of steel slag and steel sludge in concrete. Constr. Build. Mater. 2016, 104, 16–24. [Google Scholar] [CrossRef]
  140. Bodor, M.; Santos, R.M.; Cristea, G.; Salman, M.; Cizer, Ö.; Iacobescu, R.I.; Chiang, Y.W.; van Balen, K.; Vlad, M.; van Gerven, T. Laboratory investigation of carbonated BOF slag used as partial replacement of natural aggregate in cement mortars. Cem. Concr. Compos. 2016, 65, 55–66. [Google Scholar] [CrossRef]
  141. Xue, G.; Fu, Q.; Xu, S.; Li, J. Macroscopic mechanical properties and microstructure characteristics of steel slag fine aggregate concrete. J. Build. Eng. 2022, 56, 104742. [Google Scholar] [CrossRef]
  142. Chang, X.; Zhang, R.; Xiao, Y.; Chen, X.; Zhang, X.; Liu, G. Mapping of publications on asphalt pavement and bitumen materials: A bibliometric review. Constr. Build. Mater. 2020, 234, 117370. [Google Scholar] [CrossRef]
  143. Deng, L.; Yun, F.; Jia, R.; Li, H.; Jia, X.; Shi, Y.; Zhang, X. Effect of SiO2/MgO ratio on the crystallization behavior, structure, and properties of wollastonite-augite glass-ceramics derived from stainless steel slag. Mater. Chem. Phys. 2020, 239, 122039. [Google Scholar] [CrossRef]
  144. Shi, Y.; Li, B.W.; Zhao, M.; Zhang, M.X. Growth of diopside crystals in CMAS glass-ceramics using Cr2O3 as a nucleating agent. J. Am. Ceram. Soc. 2018, 101, 3968–3978. [Google Scholar] [CrossRef]
  145. Zhang, S.; Zhang, Y.; Wu, T. Effect of Cr2O3 on the crystallization behavior of synthetic diopside and characterization of Cr-doped diopside glass ceramics. Ceram. Int. 2018, 44, 10119–10129. [Google Scholar] [CrossRef]
  146. Li, Y.; Zhao, L.; Wang, Y.; Cang, D. Effects of Fe2O3 on the properties of ceramics from steel slag. Int. J. Miner. Metall. Mater. 2018, 25, 413–419. [Google Scholar] [CrossRef]
  147. Luo, Z.; He, F.; Zhang, W.; Xiao, Y.; Xie, J.; Sun, R.; Xie, M. Effects of fluoride content on structure and properties of steel slag glass-ceramics. Mater. Chem. Phys. 2020, 242, 122531. [Google Scholar] [CrossRef]
  148. OuYang, S.; Zhang, Y.; Chen, Y.; Zhao, Z.; Wen, M.; Li, B.; Shi, Y.; Zhang, M.; Liu, S. Preparation of Glass-ceramics Using Chromium-containing Stainless Steel Slag: Crystal Structure and Solidification of Heavy Metal Chromium. Sci. Rep. 2019, 9, 1964. [Google Scholar] [CrossRef]
  149. Shang, W.; Peng, Z.; Huang, Y.; Gu, F.; Zhang, J.; Tang, H.; Yang, L.; Tian, W.; Rao, M.; Li, G.; et al. Production of glass-ceramics from metallurgical slags. J. Clean. Prod. 2021, 317, 128220. [Google Scholar] [CrossRef]
  150. Bayer Ozturk, Z.; Eren Gultekin, E. Preparation of ceramic wall tiling derived from blast furnace slag. Ceram. Int. 2015, 41, 12020–12026. [Google Scholar] [CrossRef]
  151. Zong, Y.; Zhang, X.; Mukiza, E.; Xu, X.; Li, F. Effect of fly ash on the properties of ceramics prepared from steel slag. Appl. Sci. 2018, 8, 1187. [Google Scholar] [CrossRef]
  152. Pal, M.; Das, S.; Gupta, S.; Das, S.K. Thermal analysis and vitrification behavior of slag containing porcelain stoneware body. J. Therm. Anal. Calorim. 2016, 124, 1169–1177. [Google Scholar] [CrossRef]
  153. Tabit, K.; Waqif, M.; Saâdi, L. Anorthite-cordierite based binary ceramics from coal fly ash and steel slag for thermal and dielectric applications. Mater. Chem. Phys. 2020, 254, 123472. [Google Scholar] [CrossRef]
  154. Xu, X.; Wang, Y.; Wu, J.; Liu, S.; Ma, S.; Cheng, T. Preparation and thermal shock resistance of solar thermal storage ceramics from high calcium and high iron steel slag. Ceram. Int. 2023, 50, 8099–8108. [Google Scholar] [CrossRef]
  155. Liu, S.; Chen, Y.; Ouyang, S.; Li, H.; Li, X.; Li, B. Microstructural transformation of stainless steel slag-based CAMS glass ceramics prepared by SPS. Ceram. Int. 2021, 47, 1284–1293. [Google Scholar] [CrossRef]
  156. Zong, Y.; Wan, Q.; Cang, D. Preparation of anorthite-based porous ceramics using high-alumina fly ash microbeads and steel slag. Ceram. Int. 2019, 45, 22445–22451. [Google Scholar] [CrossRef]
  157. Yingliang, Z.; Zhengyu, M.; Jingping, Q.; Xiaogang, S.; Xiaowei, G. Experimental study on the utilization of steel slag for cemented ultra-fine tailings backfill. Powder Technol. 2020, 375, 284–291. [Google Scholar] [CrossRef]
  158. Zhang, M.; Li, K.; Ni, W.; Zhang, S.; Liu, Z.; Wang, K.; Wei, X.; Yu, Y. Preparation of mine backfilling from steel slag-based non-clinker combined with ultra-fine tailing. Constr. Build. Mater. 2022, 320, 126248. [Google Scholar] [CrossRef]
  159. Zhang, S.; Wu, B.; Ren, Y.; Wu, Z.; Li, Q.; Li, K.; Zhang, M.; Yu, J.; Liu, J.; Ni, W. The Preparation Process and Hydration Mechanism of Steel Slag-Based Ultra-Fine Tailing Cementitious Filler. Gels 2023, 9, 82. [Google Scholar] [CrossRef]
  160. Li, J.; Yilmaz, E.; Cao, S. Influence of industrial solid waste as filling material on mechanical and microstructural characteristics of cementitious backfills. Constr. Build. Mater. 2021, 299, 124288. [Google Scholar] [CrossRef]
  161. Hao, J.; Zhou, Z.; Chen, Z.; Zhou, Y.; Wang, J. Damage characterization and microscopic mechanism of steel slag-cemented paste backfill under uniaxial compression. Constr. Build. Mater. 2023, 409, 134175. [Google Scholar] [CrossRef]
  162. Li, J.; Cao, S.; Yilmaz, E. Characterization of Macro Mechanical Properties and Microstructures of Cement-Based Composites Prepared from Fly Ash, Gypsum and Steel Slag. Minerals 2021, 12, 6. [Google Scholar] [CrossRef]
  163. Das, S.; Kim, G.W.; Hwang, H.Y.; Verma, P.P.; Kim, P.J. Cropping With Slag to Address Soil, Environment, and Food Security. Front. Microbiol. 2019, 10, 1320. [Google Scholar] [CrossRef] [PubMed]
  164. Deus, A.C.F.; Büll, L.T.; Guppy, C.N.; Santos, S.d.M.C.; Moreira, L.L.Q. Effects of lime and steel slag application on soil fertility and soybean yield under a no till-system. Soil Tillage Res. 2020, 196, 104422. [Google Scholar] [CrossRef]
  165. Wen, T.; Yang, L.; Dang, C.; Miki, T.; Bai, H.; Nagasaka, T. Effect of basic oxygen furnace slag on succession of the bacterial community and immobilization of various metal ions in acidic contaminated mine soil. J. Hazard. Mater. 2020, 388, 121784. [Google Scholar] [CrossRef]
  166. Yang, L.; Wei, T.; Li, S.; Lv, Y.; Miki, T.; Yang, L.; Nagasaka, T. Immobilization persistence of Cu, Cr, Pb, Zn ions by the addition of steel slag in acidic contaminated mine soil. J. Hazard. Mater. 2021, 412, 125176. [Google Scholar] [CrossRef] [PubMed]
  167. Wang, A.; Liu, Y.; Zhang, Y.; Ren, J.; Zeng, Y.; Huang, Z. Synthesis of biochar modified steel slag composites for passivation of multiple heavy metals in soil. J. Environ. Chem. Eng. 2024, 12, 114026. [Google Scholar] [CrossRef]
  168. Guerrini, I.A.; Croce, C.G.G.; Bueno, O.d.C.; Jacon, C.P.R.P.; Nogueira, T.A.R.; Fernandes, D.M.; Ganga, A.; Capra, G.F. Composted sewage sludge and steel mill slag as potential amendments for urban soils involved in afforestation programs. Urban For. Urban Green. 2017, 22, 93–104. [Google Scholar] [CrossRef]
  169. Das, S.; Gwon, H.S.; Khan, M.I.; Jeong, S.T.; Kim, P.J. Steel slag amendment impacts on soil microbial communities and activities of rice (Oryza sativa L.). Sci. Rep. 2020, 10, 6746. [Google Scholar] [CrossRef]
  170. Wang, W.; Sardans, J.; Lai, D.Y.F.; Wang, C.; Zeng, C.; Tong, C.; Liang, Y.; Peñuelas, J. Effects of steel slag application on greenhouse gas emissions and crop yield over multiple growing seasons in a subtropical paddy field in China. Field Crops Res. 2015, 171, 146–156. [Google Scholar] [CrossRef]
  171. Wang, M.; Lan, X.; Xu, X.; Fang, Y.; Singh, B.P.; Sardans, J.; Romero, E.; Peñuelas, J.; Wang, W. Steel slag and biochar amendments decreased CO2 emissions by altering soil chemical properties and bacterial community structure over two-year in a subtropical paddy field. Sci. Total Environ. 2020, 740, 140403. [Google Scholar] [CrossRef]
  172. Ning, D.; Liang, Y.; Song, A.; Duan, A.; Liu, Z. In situ stabilization of heavy metals in multiple-metal contaminated paddy soil using different steel slag-based silicon fertilizer. Environ. Sci. Pollut. Res. 2016, 23, 23638–23647. [Google Scholar] [CrossRef] [PubMed]
  173. Yang, J.; Pan, X.; Zhao, C.; Mou, S.; Achal, V.; Al-Misned, F.A.; Mortuza, M.G.; Gadd, G.M. Bioimmobilization of Heavy Metals in Acidic Copper Mine Tailings Soil. Geomicrobiol. J. 2016, 33, 261–266. [Google Scholar] [CrossRef]
  174. He, H.; Tam, N.F.Y.; Yao, A.; Qiu, R.; Li, W.C.; Ye, Z. Growth and Cd uptake by rice (Oryza sativa) in acidic and Cd-contaminated paddy soils amended with steel slag. Chemosphere 2017, 189, 247–254. [Google Scholar] [CrossRef]
  175. Bashir, S.; Salam, A.; Rehman, M.; Khan, S.; Gulshan, A.B.; Iqbal, J.; Shaaban, M.; Mehmood, S.; Zahra, A.; Hu, H. Effective Role of Biochar, Zeolite and Steel Slag on Leaching Behavior of Cd and Its Fractionations in Soil Column Study. Bull. Environ. Contam. Toxicol. 2019, 102, 567–572. [Google Scholar] [CrossRef] [PubMed]
  176. Wu, J.; Huang, D.; Liu, X.; Meng, J.; Tang, C.; Xu, J. Remediation of As(III) and Cd(II) co-contamination and its mechanism in aqueous systems by a novel calcium-based magnetic biochar. J. Hazard. Mater. 2018, 348, 10–19. [Google Scholar] [CrossRef]
  177. He, H.; Xiao, Q.; Yuan, M.; Huang, R.; Sun, X.; Wang, X.; Zhao, H. Effects of steel slag amendments on accumulation of cadmium and arsenic by rice (Oryza sativa) in a historically contaminated paddy field. Environ. Sci. Pollut. Res. 2020, 27, 40001–40008. [Google Scholar] [CrossRef]
  178. Sahu, J.N.; Kapelyushin, Y.; Mishra, D.P.; Ghosh, P.; Sahoo, B.K.; Trofimov, E.; Meikap, B.C. Utilization of ferrous slags as coagulants, filters, adsorbents, neutralizers/stabilizers, catalysts, additives, and bed materials for water and wastewater treatment: A review. Chemosphere 2023, 325, 138201. [Google Scholar] [CrossRef]
  179. Han, C.; Wang, Z.; Wu, Q.; Yang, W.; Yang, H.; Xue, X. Evaluation of the role of inherent Ca2 in phosphorus removal from wastewater system. Water Sci. Technol. 2016, 73, 1644–1651. [Google Scholar] [CrossRef]
  180. Wang, B.; Zhu, Y.; Bai, Z.; Luque, R.; Xuan, J. Functionalized chitosan biosorbents with ultra-high performance, mechanical strength and tunable selectivity for heavy metals in wastewater treatment. Chem. Eng. J. 2017, 325, 350–359. [Google Scholar] [CrossRef]
  181. Mahmoud, S.A.; Atia, B.M.; Gado, M.A. Efficient removal of toxic metals (Hg(II), Cr(III), Pb(II), Cd(II)) using high-performance polyvinyl alcohol-L-2-Amino-3-mercaptopropionic acid composite from wastewater. Int. J. Environ. Sci. Technol. 2025, 1–26. [Google Scholar] [CrossRef]
  182. Chen, G.; Yang, L.; Chen, J.; Miki, T.; Li, S.; Bai, H.; Nagasaka, T. Competitive mechanism and influencing factors for the simultaneous removal of Cr(III) and Zn(II) in acidic aqueous solutions using steel slag: Batch and column experiments. J. Clean. Prod. 2019, 230, 69–79. [Google Scholar] [CrossRef]
  183. Sarkar, C.; Basu, J.K.; Samanta, A.N. Synthesis of mesoporous geopolymeric powder from LD slag as superior adsorbent for Zinc (II) removal. Adv. Powder Technol. 2018, 29, 1142–1152. [Google Scholar] [CrossRef]
  184. Gao, Y.; Jiang, J.; Tian, S.; Li, K.; Yan, F.; Liu, N.; Yang, M.; Chen, X. BOF steel slag as a low-cost sorbent for vanadium (V) removal from soil washing effluent. Sci. Rep. 2017, 7, 11177. [Google Scholar] [CrossRef]
  185. Samanta, N.S.; Banerjee, S.; Mondal, P.; Anweshan; Bora, U.; Purkait, M.K. Preparation and characterization of zeolite from waste Linz-Donawitz (LD) process slag of steel industry for removal of Fe3+ from drinking water. Adv. Powder Technol. 2021, 32, 3372–3387. [Google Scholar] [CrossRef]
  186. Li, Y.; Qi, X.; Li, G.; Wang, H. Efficient removal of arsenic from copper smelting wastewater via a synergy of steel-making slag and KMnO4. J. Clean. Prod. 2021, 287, 125578. [Google Scholar] [CrossRef]
  187. Wang, M.; Wang, X.; Zhang, M.; Han, W.; Yuan, Z.; Zhong, X.; Yu, L.; Ji, H. Treatment of Cd(Ⅱ) and As(Ⅴ) co-contamination in aqueous environment by steel slag-biochar composites and its mechanism. J. Hazard. Mater. 2023, 447, 130784. [Google Scholar] [CrossRef]
  188. Liu, J.; Feng, J.; Wang, Z.; Lyu, X.; Liu, C.; Dai, H.; Bai, Z. Steel slag base geopolymer for efficient removal of Zn and Ni ions from aqueous solutions: Preparation and characterization. Mater. Lett. 2023, 340, 134173. [Google Scholar] [CrossRef]
  189. Huy, D.H.; Seelen, E.; Liem-Nguyen, V. Removal mechanisms of cadmium and lead ions in contaminated water by stainless steel slag obtained from scrap metal recycling. J. Water Process Eng. 2020, 36, 101369. [Google Scholar] [CrossRef]
  190. Sang, M.; Zhao, H.; Li, Y.; Zhu, L. The adsorption properties of steel slag-based porous geopolymer for Cu2+ removal. Miner. Eng. 2023, 201, 108225. [Google Scholar]
  191. Sarkar, C.; Basu, J.K.; Samanta, A.N. Microwave activated LD slag for phenolic wastewater treatment: Multi-parameter optimization, isotherms, kinetics and thermodynamics. Chem. Eng. Trans. 2017, 57, 277–282. [Google Scholar]
  192. Yang, L.; Qiang, X.; Wang, Z.; Li, Y.; Bai, H.; Li, H. Steel slag as low-cost adsorbent for the removal of phenanthrene and naphthalene. Adsorpt. Sci. Technol. 2018, 36, 1160–1177. [Google Scholar] [CrossRef]
  193. Yoon, S.; Bae, S. Development of magnetically separable Cu catalyst supported by pre-treated steel slag. Korean J. Chem. Eng. 2019, 36, 1814–1825. [Google Scholar] [CrossRef]
  194. Priyadarshini, M.; Ahmad, A.; Ghangrekar, M.M. Efficient upcycling of iron scrap and waste polyethylene terephthalate plastic into Fe3O4@ C incorporated MIL-53 (Fe) as a novel electro-Fenton catalyst for the degradation of salicylic acid. Environ. Pollut. 2023, 322, 121242. [Google Scholar] [CrossRef]
  195. dos Santos, A.J.; da Costa Cunha, G.; Cruz, D.R.S.; Romão, L.P.C.; Martínez-Huitle, C.A. Iron mining wastes collected from Mariana disaster: Reuse and application as catalyst in a heterogeneous electro-Fenton process. J. Electroanal. Chem. 2019, 848, 113330. [Google Scholar] [CrossRef]
  196. Liu, Y.; Wu, Z.; Peng, P.; Xie, H.; Li, X.; Xu, J.; Li, W. A pilot-scale three-dimensional electrochemical reactor combined with anaerobic-anoxic-oxic system for advanced treatment of coking wastewater. J. Environ. Manag. 2020, 258, 110021. [Google Scholar] [CrossRef]
  197. Song, B.; Wang, Z.; Li, J.; Luo, M.; Cao, P.; Zhang, C. Sulfur-zinc modified kaolin/steel slag: A particle electrode that efficiently degrades norfloxacin in a neutral/alkaline environment. Chemosphere 2021, 284, 131328. [Google Scholar] [CrossRef]
  198. Wang, Z.; Song, B.; Li, J.; Teng, X. Degradation of norfloxacin wastewater using kaolin/steel slag particle electrodes: Performance, mechanism and pathway. Chemosphere 2021, 270, 128652. [Google Scholar] [CrossRef]
  199. Sun, L.; Wu, J.; Wang, J.; Yang, Y.; Xu, M.; Liu, J.; Yang, C.; Cai, Y.; He, H.; Du, Y.; et al. In-situ constructing nanostructured magnesium ferrite on steel slag for Cr(VI) photoreduction. J. Hazard. Mater. 2022, 422, 126951. [Google Scholar] [CrossRef]
  200. Shu, Y.; Ji, B.; Li, Y.; Zhang, W.; Zhang, H.; Zhang, J. Natural pyrite improved steel slag towards environmentally sustainable chromium reclamation from hexavalent chromium-containing wastewater. Chemosphere 2021, 282, 130974. [Google Scholar] [CrossRef]
  201. Sarkar, C.; Basu, J.K.; Samanta, A.N. Synthesis of MIL-53(Fe)/SiO2 composite from LD slag as a novel photo-catalyst for methylene blue degradation. Chem. Eng. J. 2019, 377, 119621. [Google Scholar] [CrossRef]
  202. Ahmad, M.S.; Ab Rahim, M.H.; Alqahtani, T.M.; Witoon, T.; Lim, J.-W.; Cheng, C.K. A review on advances in green treatment of glycerol waste with a focus on electro-oxidation pathway. Chemosphere 2021, 276, 130128. [Google Scholar] [CrossRef] [PubMed]
  203. Yu, Q.; Zhang, Y.; Tang, M.; Liu, G.; Li, L. Insight into the steel converter slag composite supported three-dimensional electro-Fenton remediation of landfill leachate. J. Water Process Eng. 2023, 53, 103603. [Google Scholar] [CrossRef]
  204. Fang, C.; Nie, L.; Chen, H.; Yang, Y. Magnetic recyclable Co-MOF derivate-modified steel slag for tetracycline removal by integrated adsorption and Fenton-like catalysis. J. Water Process Eng. 2024, 63, 105434. [Google Scholar] [CrossRef]
  205. Zhang, Y.J.; Zhang, L.; Zhang, K.; Kang, L. Synthesis of eco-friendly CaWO4/CSH nanocomposite and photocatalytic degradation of dyeing pollutant. Integr. Ferroelectr. 2017, 181, 113–122. [Google Scholar] [CrossRef]
  206. Inoue, T.; Chuaicham, C.; Saito, N.; Ohtani, B.; Sasaki, K. Z-scheme heterojunction of graphitic carbon nitride and calcium ferrite in converter slag for the photocatalytic imidacloprid degradation and hydrogen evolution. J. Photochem. Photobiol. A Chem. 2023, 440, 114644. [Google Scholar] [CrossRef]
  207. Shao, N.; Li, S.; Yan, F.; Su, Y.; Liu, F.; Zhang, Z. An all-in-one strategy for the adsorption of heavy metal ions and photodegradation of organic pollutants using steel slag-derived calcium silicate hydrate. J. Hazard. Mater. 2020, 382, 121120. [Google Scholar] [CrossRef]
  208. Xu, Y.; Zhang, C.; Lu, P.; Zhang, X.; Zhang, L.; Shi, J. Overcoming poisoning effects of heavy metal ions against photocatalysis for synergetic photo-hydrogen generation from wastewater. Nano Energy 2017, 38, 494–503. [Google Scholar] [CrossRef]
  209. Ortega-Fernández, I.; Calvet, N.; Gil, A.; Rodríguez-Aseguinolaza, J.; Faik, A.; D’Aguanno, B. Thermophysical characterization of a by-product from the steel industry to be used as a sustainable and low-cost thermal energy storage material. Energy 2015, 89, 601–609. [Google Scholar] [CrossRef]
  210. Wang, Y.; Wang, Y.; Li, H.; Zhou, J.; Cen, K. Thermal properties and friction behaviors of slag as energy storage material in concentrate solar power plants. Sol. Energy Mater. Sol. Cells 2018, 182, 21–29. [Google Scholar] [CrossRef]
  211. Kocak, B.; Fernandez, A.I.; Paksoy, H. Benchmarking study of demolition wastes with different waste materials as sensible thermal energy storage. Sol. Energy Mater. Sol. Cells 2021, 219, 110777. [Google Scholar] [CrossRef]
  212. Qureshi, Z.A.; Ali, H.M.; Khushnood, S. Recent advances on thermal conductivity enhancement of phase change materials for energy storage system: A review. Int. J. Heat Mass Transf. 2018, 127, 838–856. [Google Scholar] [CrossRef]
  213. Wang, J.; Huang, Y. Exploration of steel slag for thermal energy storage and enhancement by Na2CO3 modification. J. Clean. Prod. 2023, 395, 136289. [Google Scholar] [CrossRef]
  214. Xu, X.; Li, M.; Wang, Y.; Wu, J.; Zhou, Y.; Shen, Y. Preparation and characterization of solar absorption and thermal storage integrated ceramics from calcium and iron-rich steel slag. Ceram. Int. 2023, 49, 8381–8389. [Google Scholar] [CrossRef]
  215. Boquera, L.; Castro, J.R.; Fernandez, A.G.; Navarro, A.; Pisello, A.L.; Cabeza, L.F. Thermo-mechanical stability of concrete containing steel slag as aggregate after high temperature thermal cycles. Sol. Energy 2022, 239, 59–73. [Google Scholar] [CrossRef]
  216. Zhang, J.; Guo, Z.; Zhu, Y.; Zhang, H.; Yan, M.; Liu, D.; Hao, J. Preparation and characterization of novel low-cost sensible heat storage materials with steel slag. J. Energy Storage 2024, 76, 109643. [Google Scholar] [CrossRef]
  217. Xiong, Y.; Yang, Y.; Zhang, A.; Ren, J.; Xu, Q.; Wu, Y.; Zhao, Y.; Ding, Y. Investigation on low-carbon shape-stable phase change composite by steel slag and carbide slag for solar thermal energy storage. J. Energy Storage 2024, 76, 109736. [Google Scholar] [CrossRef]
  218. Merlin, K.; Soto, J.; Delaunay, D.; Traonvouez, L. Industrial waste heat recovery using an enhanced conductivity latent heat thermal energy storage. Appl. Energy 2016, 183, 491–503. [Google Scholar] [CrossRef]
  219. Wang, J.; Wang, Y.; Huang, Y. Synthesis and characterization of form-stable carbonate/steel slag composite materials for thermal energy storage. J. Energy Storage 2022, 52, 104708. [Google Scholar] [CrossRef]
  220. Liu, Y.; Jiang, Z.; Zhang, X.; Wu, B.; Shen, Y.; Wu, L.; Zhang, X. Steel slag-KNO3 phase change composites for thermal storage at medium-high temperature and solid waste recycling. Mater. Chem. Phys. 2023, 301, 127655. [Google Scholar] [CrossRef]
  221. Fukahori, R.; Nomura, T.; Zhu, C.; Sheng, N.; Okinaka, N.; Akiyama, T. Macro-encapsulation of metallic phase change material using cylindrical-type ceramic containers for high-temperature thermal energy storage. Appl. Energy 2016, 170, 324–328. [Google Scholar] [CrossRef]
  222. Sakellariou, K.G.; Karagiannakis, G.; Criado, Y.A.; Konstandopoulos, A.G. Calcium oxide based materials for thermochemical heat storage in concentrated solar power plants. Sol. Energy 2015, 122, 215–230. [Google Scholar] [CrossRef]
  223. Perejón, A.; Valverde, J.M.; Miranda-Pizarro, J.; Sánchez-Jiménez, P.E.; Pérez-Maqueda, L.A. Large-Scale Storage of Concentrated Solar Power from Industrial Waste. ACS Sustain. Chem. Eng. 2017, 5, 2265–2272. [Google Scholar] [CrossRef]
  224. Valverde, J.M.; Miranda-Pizarro, J.; Perejón, A.; Sánchez-Jiménez, P.E.; Pérez-Maqueda, L.A. Calcium-Looping performance of steel and blast furnace slags for thermochemical energy storage in concentrated solar power plants. J. CO2 Util. 2017, 22, 143–154. [Google Scholar] [CrossRef]
Figure 1. Comparative analysis of steel slag utilization in major countries. Reprinted from Refs. [5,6,7,8,9].
Figure 1. Comparative analysis of steel slag utilization in major countries. Reprinted from Refs. [5,6,7,8,9].
Metals 15 00537 g001
Figure 2. C2S phase in polymorphic transform steel slag. Reprinted from ref. [12].
Figure 2. C2S phase in polymorphic transform steel slag. Reprinted from ref. [12].
Metals 15 00537 g002
Figure 3. Flow chart of different iron recovery technologies. Reprinted from refs. [38,39,41,42].
Figure 3. Flow chart of different iron recovery technologies. Reprinted from refs. [38,39,41,42].
Metals 15 00537 g003
Figure 4. Resource utilization of converter BOF slag.
Figure 4. Resource utilization of converter BOF slag.
Metals 15 00537 g004
Table 1. Typical BOF slag exhibits the following chemical composition (wt% range). Adapted from Ref. [11].
Table 1. Typical BOF slag exhibits the following chemical composition (wt% range). Adapted from Ref. [11].
ComponentSiO2Al2O3CaOMgOFeOSP2O5f-CaOBasicity (CaO/(SiO2 + P2O5))
Range15–253–745–605–2012–250.1–0.40–11.6–72.1–3.5
Table 2. Basicity-dependent mineral phase distribution.
Table 2. Basicity-dependent mineral phase distribution.
BasicityDominant Phases (by Abundance)
1.36Olivine (CaO·MgO·SiO2), rhodonite (3CaO·MgO·2SiO2), RO phase
1.80Rhodonite, dicalcium silicate (C2S), RO phase
2.51Tricalcium silicate (C3S), C2S, RO phase
2.99C3S, C2S, RO phase, calcium ferrite (Ca2Fe2O5)
Table 3. Key physical properties of BOF slag. Adapted from refs. [14,15].
Table 3. Key physical properties of BOF slag. Adapted from refs. [14,15].
ParameterRange/ValueTest StandardEngineering Implication
Density3.3–3.6 g/cm3BS EN 1097-7 [16]Aggregate grading design
Water absorption<3%ASTM D6473 [17]Weathering resistance
Crushing value20.4–30.8%ASTM C535 [18]Compressive strength indicator
Bond work index0.7 (Std sand = 1.0)Bond grindabilityGrinding energy consumption
Abrasion coefficient (K)0.8EN 1097-1 [19]Pavement rutting resistance
Brittleness index (n)0.2Dynamic impact Crack propagation resistance
Table 4. Reaction of the components of steel slag at the corresponding temperature.
Table 4. Reaction of the components of steel slag at the corresponding temperature.
ComponentReactionExpansionCritical Temperature
f-CaOCaO + H2O → Ca(OH)2100–300%Ambient–100 °C
β-C2Sβ-C2S → γ-C2S10%675 °C
MgOMgO + H2O → Mg(OH)277%<120 °C
C3S decomposition3CaO·SiO2→2CaO·SiO2 + CaOLocalized1100–1250 °C
Table 5. BOF slag valorization matrix, Adapted from refs. [14,15].
Table 5. BOF slag valorization matrix, Adapted from refs. [14,15].
ApplicationKey PropertiesTRLCritical Barriers
Cement additiveC2S/C3S reactivity6–7f-CaO content < 1.5% required
Road aggregateHigh abrasion resistance8High density (+15% transport cost)
Wastewater treatmentPorous structure5Heavy metal leaching risks
CO2 mineralizationCarbonation reactivity4Slow reaction kinetics
Table 6. Key limitations and mitigation strategies, Adapted from refs. [20,21,22,23,24,25,26,27,28].
Table 6. Key limitations and mitigation strategies, Adapted from refs. [20,21,22,23,24,25,26,27,28].
Limitation FactorMechanismMitigation Strategy
Volume expansion (f-CaO/MgO)Hydration-induced damageSteam aging (150 °C, 6 h)
Low pozzolanic activityGlassy phase < 10%Mechanical activation (<45 μm)
Heavy metal leaching (Cr, V)Alkaline ion mobilizationPhosphate stabilization
Abrasive particlesMetallic iron residues (2–5%)Magnetic separation (>95% recovery)
Table 7. Technical assessment of BOF slag pretreatment. Adapted from refs. [5,30,31].
Table 7. Technical assessment of BOF slag pretreatment. Adapted from refs. [5,30,31].
MethodKey ProcessAdvantagesLimitations
Thermal SplashingWater spraying on molten slag for 3–4 days, followed by magnetic separationHigh capacity (200–400 t/h)
Mature technology
High water consumption (3–5 m3/t)
Poor stability (f-CaO > 4%)
Rotary DrumIntegrated cooling–crushing–magnetic separation in rotating drumEfficient Fe recovery (η > 92%)
Superior stability (f-CaO < 2%)
Requires high fluidity (η > 1 Pa·s)
Steam AgingIntermittent water spraying on 300–800 °C slag induces self-pulverizationHigh fines yield (85% < 20 mm)
Low dust emission (<10 mg/m3)
Long cycle (8–12 h)
High energy demand (50–80 kWh/t)
Air QuenchingMolten slag atomization by compressed air with heat recoveryUniform granules (D50 = 0.5–2 mm)
65% heat recovery
Noise pollution (>85 dB)
High maintenance cost
Pressurized SteamAccelerated f-CaO hydration under 1.5–2.0 MPa steamShort cycle (2–4 h)
Excellent stability (f-CaO < 1%)
Capital intensive ($1.2–1.8 M/unit)
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

Ma, C.; Zhang, S.; Li, K.; Zhao, T.; Meng, Q.; Guan, D.; Zhang, A. Stainable Utilization Strategies for Basic Oxygen Furnace Slag: Properties, Processing, and Future Directions. Metals 2025, 15, 537. https://doi.org/10.3390/met15050537

AMA Style

Ma C, Zhang S, Li K, Zhao T, Meng Q, Guan D, Zhang A. Stainable Utilization Strategies for Basic Oxygen Furnace Slag: Properties, Processing, and Future Directions. Metals. 2025; 15(5):537. https://doi.org/10.3390/met15050537

Chicago/Turabian Style

Ma, Chunting, Siqi Zhang, Keqing Li, Tong Zhao, Qingxin Meng, Dongshang Guan, and Ao Zhang. 2025. "Stainable Utilization Strategies for Basic Oxygen Furnace Slag: Properties, Processing, and Future Directions" Metals 15, no. 5: 537. https://doi.org/10.3390/met15050537

APA Style

Ma, C., Zhang, S., Li, K., Zhao, T., Meng, Q., Guan, D., & Zhang, A. (2025). Stainable Utilization Strategies for Basic Oxygen Furnace Slag: Properties, Processing, and Future Directions. Metals, 15(5), 537. https://doi.org/10.3390/met15050537

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