Next Article in Journal
Effect of Gangue Minerals on Pulp Rheology and Flotation Behavior of Smithsonite
Next Article in Special Issue
Utilization of Sugar Mill Waste Ash as Pozzolanic Material in Structural Mortar
Previous Article in Journal
Rock Magnetic Results from the Early Ordovician Limestone Rocks in the Northern Qaidam Block, Tibetan Plateau
Previous Article in Special Issue
Challenges, Regulations, and Case Studies on Sustainable Management of Industrial Waste
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of High Temperature Reconstruction and Modification on Phase Composition and Structure of Steel Slag

School of Materials and Metallurgy, Inner Mongolia University of Science and Technology, Baotou 014010, China
*
Author to whom correspondence should be addressed.
Minerals 2023, 13(1), 67; https://doi.org/10.3390/min13010067
Submission received: 24 November 2022 / Revised: 22 December 2022 / Accepted: 28 December 2022 / Published: 30 December 2022
(This article belongs to the Special Issue Management, Recycling and Reuse of Industrial Waste)

Abstract

:
This study investigates the pattern of influence of blast furnace slag tempering on the composition and structure of steel slag. The chemical composition, equilibrium phase composition, microscopic morphological characteristics and elemental composition of microscopic regions of steel slag and blast furnace slag modified by high temperature reconstruction were analyzed using X-ray diffractometer (XRD), FactSage7.1 thermodynamic analysis software, mineral phase microscopy and field emission scanning electron microscopy. The results show that blast furnace slag blending can promote the generation of a low melting point phase in the slag, as well as reducing its melting temperature and improving its high temperature kinetic conditions. On the one hand, the incorporation of blast furnace slag was found to promote the generation of C2S in the steel slag and improve its gelling activity. Notably, at 1400 °C, the C2S content (mass fraction) of steel slag modified with 15% high temperature reconstruction reached 39.04%, while that of unmodified steel slag at this temperature was only 16.92%, i.e., only 1/4 of the C2S content in the modified slag. On the other hand, the incorporation of blast furnace slag inhibited the generation of a-C2S-C3P and calcium ferrate minerals, refined the grains of calcium–aluminum yellow feldspar, reduced the alkalinity and promoted the generation of silicate phases with high hydration activity in steel slag.

1. Introduction

Steel slag, as a by-product of the high temperature steelmaking process, has a chemical composition and mineralogical composition very similar to those of cement and concrete admixtures [1]. However, the low content of cementitious active phase [2], poor stability [3] and high crushing cost of steel slag have become major factors limiting the scale of its use in blending and construction [4]. Therefore, at present, most steel mills still dispose of converter slag in traditional ways, such as piling [5] and landfilling [6,7,8]. Indeed, the recycling rate is only about 30% [9], which not only presents significant environmental issues [10] but is also a huge waste of resources [11], contributing to the high cost of steelmaking.
The temperature of molten steel slag is 1500~1700 °C, and the waste heat quality is high, making it of great value for development and utilization. In the process of steel slag cooling, according to the specific modification requirements, purposely adding modifiers to steel slag to adjust its physical and chemical properties can improve the potential to use this resource. Wang Huigang et al. [12] analyzed the main components of fly ash and their roles in steel slag; the results showed that SiO2 and Al2O3 react with f-CaO to generate stable phases which improve the stability of steel slag while also generating calcium silicate and calcium aluminate. Zhang Xiong [13] studied the carbonation mechanism of zeolite-modified steel slag products. The results showed that after pre-hydration curing for 1 d and carbonation for 2 h, a steel slag test block with 5% zeolite (CSZ5-1 d) had the best compressive strength, while a steel slag test block with 15% zeolite (CSZ15-1 d) had the best carbonation rate, with increases of 14% and 10.2%, respectively, compared with a pure steel slag test block. Xiang Ruiheng [14] studied preparation and foaming modifications of reconstructed steel slag powder with medium and high activity. It was observed that 75% converter steel slag, 4% bauxite and 21% lime, fired at 1290 °C for 90 min, had the highest C2S and C4AF contents after rapid air-cooling, resulting in an increase in water activity to 90.4%, a 2.03% reduction in the f-CaO mass fraction, a dissipated RO phase, and increased ease of grinding. It was also demonstrated that the best performance of porous reconstituted steel slag was achieved with a high temperature foaming agent SiC doping content of 1.6%. The activity index of modified steel slag could be increased to 98.2%, and the compressive strength of the composite cement mortar could reach 44.8 MPa. Rao Lei [15] adjusted the alkalinity (2.0~3.0) and Fe2O3 content (to about 20%) of a steel slag using quartz sand and coal dust. The results showed that the modified steel slag f-CaO mass fraction was reduced by 39.6%, the ease of grinding index was increased by 11% and the 7 d and 28 d activity indices were increased by 3% and 4.8%, respectively. Wang, Chang-Long [16] investigated the high temperature modification of steel slag using reservoir bottom sludge and electroslag stone. The results showed that the 28 d activity index reached 82.4%, with f-CaO and f-MgO mass fractions of 1.21% and 1.98%, respectively, when the compound modifier was dosed at 20% (reservoir bottom sludge:electroslag stone ratio of 3:1) and the treatment temperature was 1150 °C, which met the production requirements. Li et al. [17] analyzed the effect of electric furnace slag on the properties of steel slag. The results showed that the former can promote the improvement of coagulation in the latter. Zhang Zuoshun et al. [18] investigated the effect of iron tailings on the properties of high temperature steel slag. The results showed that iron tailings can enhance steel slag cementation, while the f-CaO content in the steel slag was reduced and its stability was improved. Lei Yunbo et al. [19] studied the effect of fly ash on steel slag in f-CaO. The results showed that an admixture of fly ash could reduce the f-CaO content in steel slag. Liu Shiye et al. [20] studied the effect of blast furnace slag on the physical phase in steel slag. The results obtained using a 10% blast furnace slag modified steel slag showed that at 1550 °C, the mass fraction of C2S and C3S increased significantly, f-CaO decreased to 1.64% and the stability improved. In addition, coke reduced the iron in the slag and improved the ease of grinding.
Most of the above research applied pure reagents or reducing agents for steel slag reduction modification treatments. Although the findings were striking, most of the treatment methods would be expensive, precluding their use on an industrial scale. Blast furnace slag is an abundant solid waste resource. Its binary alkalinity (R=CaO/SiO2) is around 1.0, and it has a high content of Al2O3, which can be used as a modifier to effectively reduce the alkalinity and stimulate the activity of steel slag, thereby also achieving the goal of “treating waste with waste”. However, few experimental studies have been carried out on the modification of steel slag by high temperature melting using blast furnace slag addition, and the phase composition and structure of steel slag modified by blast furnace slag are not clear. In this paper, under the premise of the “comprehensive utilization of solid waste”, we propose the use of water-quenched blast furnace slag to adjust the alkalinity and physical composition, eliminate the unstable factor and improve the gelation activity of steel slag. This research provides a theoretical basis and experimental data for the modification of steel slag using blast furnace slag.

2. Materials and Methods

2.1. Experimental Materials

In this experiment, water-quenched blast furnace slag and hot-spoiled steel slag were used as raw materials; their chemical compositions are shown in Table 1.

2.2. Parameters of FactSage7.1

The Equilib module in FactSage7.1 was used to calculate the phase composition of 100 g slag at equilibrium under different temperature conditions and to analyze variations in its phase composition. The specific parameters applied are shown in Table 2.

2.3. Experimental Protocol

(1) The steel slag and blast furnace slag were crushed using a crusher and screened using a mesh sieve (0.074 mm). According to the chemical composition and ratio shown in Table 1, blast furnace slag modulated steel slag was made. This was then loaded into the mixing tank on a mixing machine and mixed for 2 h before being taken out for use.
(2) We then placed the slag sample in a molybdenum crucible and put it into a KTF-1700-VT high-temperature vertical furnace. The temperature was ramped up at 10 °C/min from room temperature to 1000 °C, and then from 1000 °C to 1550 °C at a rate of 5 °C/min, holding at that temperature for 1 h. Subsequently, the sample was taken out for emergency water cooling. The whole process took place under an atmosphere of argon gas.
(3) An appropriate amount of roasted specimen was ground to achieve particle sizes of below 0.074 mm, and a composition analysis was carried out using a D8-advanced X-ray diffractometer with a Cu-Kα target, with a scanning range of 20°~80° and a scanning speed of 2°/min.
(4) Block specimens were taken. Their surface was first roughly ground and then finely ground and polished. Next, the microscopic morphological characteristics and elemental composition of the microscopic regions were analyzed using a ZEISS ultra-high resolution thermal field emission scanning electron microscope, produced by Zeiss, Germany.

3. Experimental Results and Analysis

3.1. Macrostructure of Modified Slag

The effect of high temperature melting of blast furnace slag modified steel slag after roasting are shown in Figure 1. It can be seen that the surface of the sample became rough with an increase of the proportion of blast furnace slag after roasting, and no crack occurred on the surface. The results show that the fluidity of the modified slag was reduced with an increase of the ratio of blast furnace slag, and that the surface of the synthetic slag was uneven during the cooling process, because the binary basicity of the blast furnace slag was about 1.0 and the fluidity of the low basicity blast furnace slag was poor at high temperature. The surface did not display cracks, indicating that the synthesis did not produce strong internal stresses during the cooling process. It has been shown that the main internal stresses within the steel slag originated from the crystallographic transformation of C2S and the volume expansion resulting from the hydrolysis of free calcium oxide (f-CaO) and free magnesium oxide (f-MgO). On the one hand, more pores formed on the surface of the steel slag not containing blast furnace slag after rapid cooling, while on the other hand, the internal pores became increasingly small for the steel slag mixed with increasing quantities of blast furnace slag. Our analysis indicated that with an increase of the quantity of blast furnace slag, the slag viscosity increased under high temperature conditions and the liquid phase liquidity became poor. Such high viscosity and low liquidity are not conducive to the formation of thin-walled large pore structures, making the internal pores of the slag fine and uneven.

3.2. Microstructure and Mineral Composition of Modified Slag

Blast furnace slag high-temperature molten modified steel slag roasted specimens were ground and analyzed using an XRD diffractometer (Smartlab, neo-confucianism, Japan) to determine their composition; the results are shown in Figure 2. As shown, the main phases after roasting were C2AS, C3AS3, β-C2S, and C3S. The main diffraction peak in the XRD diffraction pattern was C3S, while secondary diffraction peaks were C3AS3 and β-C2S. With an increase of blast furnace slag doping, the diffraction of the C3S peak at 2θ = 35.98° showed an overall enhancement. This indicates that the doping of blast furnace slag can promote the generation of C3S, which does not decompose into C2S and CaO during rapid cooling; instead, it indirectly reduces the f-CaO content in the slag and improves its stability. At the same time, C3S has better gelling activity, which enhances the proportion of gelling minerals in steel slag. The C2AS diffraction peak at 2θ = 31.18° and the diffraction intensity were significantly enhanced in the 15% blast furnace slag doped sample. The intensity of the β-C2S diffraction peak at 2θ = 33.30° did not change significantly, indicating that the doping of blast furnace slag had a small effect on β-C2S. C3AS3, an island silicate mineral with a high degree of hardness, showed a diffraction peak at 2θ = 31.30°. The generation of large quantities of this compound will increase the cost of steel slag crushing. As observed in the XRD plots, an increase in the blast furnace slag doping ratio enhanced the intensity of the C3AS3 diffraction peak in the steel slag. Therefore, it is suggested that the addition of 12% blast furnace slag admixture to steel slag is optimal.
In summary, in the smelting process, the blast furnace slag high temperature molten modified steel slag was fully reacted to form a new phase composition. Blast furnace slag tempering slag can promote the generation of gelling minerals such as C3S, thereby improving the stability of the steel slag, while blast furnace slag doping resulted in a higher proportion of non-gelling C2AS and higher hardness C3AS3 minerals, indirectly increasing the cost of the subsequent crushing of steel slag. Therefore, in the modification experiments of tempered steel slag, the blast furnace slag admixture percentage was kept below 15%, as this was found to be more conducive to the enhancement of the slag gelling activity [21,22], while the generated C3S indirectly consumed the f-CaO present in the slag and enhanced its stability.
In order to further confirm the interphase structure relationship of the modified steel slag, the microscopic morphological characteristics and elemental composition of the microscopic regions of the roasted experimental slag specimens were analyzed using mineral microscopy and field emission scanning electron microscopy (sigma300, Zeiss, Germany). The results of the analysis of Sr-0, Sr-8 and Sr-12 into the form hitting points and face sweeping elements are shown in Figure 3, Figure 4 and Figure 5 and Table 3, Table 4 and Table 5. From the analysis of the energy spectrum and surface scan in Figure 3 and the elemental distribution in Table 3, the percent elemental content at point 1 in Table 3 was converted into a molar ratio of n(MgO):n(FeO):n(Al2O3) = 1:1:2. On these bases, and given the distribution pattern of Mg, Fe, and Al in the vicinity of point 1, as determined by surface sweeping, it is likely that point 1 was an iron-aluminum spinel (MgO·FeAl2O4, MFA). At point 2 n(CaO):n(SiO2) = 1.35:0.96:0.17, and the binding surface sweep was more dispersed in the region, so the silicate phase with phosphorus attached at this point was considered to be Ca3(PO4)2 and CaSiO3. The substance ratio of each oxide at point 3 was about n(CaO):n(SiO2):n(Al2O3):n(FeO) = 4:3:1:1; a combined surface sweep suggested that calcium-aluminum feldspar mixed with calcium ironate (Ca2Al2SiO7, CaFe2O4) was present at this point. The material quantity ratio of each oxide at point 4 was roughly as follows: n(CaO):n(SiO2):n(Al2O3):n(MgO):n(MnO) = 1.8:1.4:0.3:0.3:0.4. Combined with the location of the ore phase, it was suggested that point 4 was the slag phase, consisting of magnesium aluminum spinel (MgAl2O4), calcium silicate (CaSiO3), and manganese oxide.
From the surface scan and energy spectrum shown in Figure 4, combined with the elemental distribution shown in Table 4, it is suggested that point 5 had the same physical phase as point 1, i.e., MFA. At point 6, the n (CaO): n (Fe2O3) ratio was 2:1. Additionally, the Ti content in the experimental slag was reduced, while the percentage share in the test was higher, considering the identification problem; therefore Ti was not considered as a factor at this stage, and the material phase at point 6 was considered to be dicalcium ferrate (2CaO·Fe2O3, C2F). The physical phases at points 7 and 8 were the same as those at point 2, i.e., silicate phases with phosphorus attached. The C content in the point sweep elements was high, possibly due to the use of a polishing agent; the main components of the polishing agent used were diamond micropowder and grinding media. As such, polishing agent residue on the surface of the specimen likely resulted in the high carbon content observed in the distribution of elements, especially around the hole residue, such as at point 8, where the carbon content was 41.47%. The detection of carbon was not further explored at this stage of our research.
The surface scan and energy spectrum shown in Figure 5, combined with the analysis of elemental distribution shown in Table 5, suggest that points 9, 11, and 12 had the same phase as point 6, i.e., C2F; the main phase at point 10 was calcium aluminum yellow feldspar (Ca2Al2SiO7, C2AS).
In summary, the scanning electron microscopy data complemented the XRD results. The phases contained in the steel slag, as determined via SEM-EDS, were MFA, a magnesium-iron phase solid solution with a low melting point, Ca3(PO4)2 and CaSiO3 (in an elliptical shape), C2F (with a milky white color), and C2AS (with a light gray color and hexagonal shape). C2AS and C2F formed a molten structure, and with the increase of blast furnace slag doping, the percentage of C2AS in the physical phase increased, causing significant cracks to appear in Sr-12. This shows that with an increase of the blast furnace slag doping ratio, the alkalinity of tempered slag decreases, which is conducive to the generation of C2S and the crystalline transformation of some C2S from β-C2S to γ-C2S, which expands to about 11% in volume. Additionally, the internal stresses resulted in the appearance of fine cracks inside the slag. On the other hand, the proportion of Al2O3 in blast furnace slag is relatively high, and the increase of its content promotes the generation of C2AS in tempered slag.
In order to further determine its macroscopic structure, the tempered slag was photographed using 300× and 500× lenses on a mineralogical microscope; see Figure 6. As shown, Sr-0 and Sr-10 were selected as the C2AS material phase fraction. It was observed that with an increase of blast furnace slag doping, the quantity of C2AS with a hexagonal shape in the material phase was reduced and black triangular forms appeared in the middle part. This indicated that the doping of blast furnace slag had a grain refining effect on C2AS. The Sr-8 calcium ferrate and silicate became interwoven, forming more needle-like calcium ferrate phase. Because needle-like calcium carbonate has high strength, an increase of its content can increase the strength of slag, indirectly increasing the cost of steel slag crushing. As shown in Sr-12 and Sr-15, silicate, calcium ironate, and calcium alumina yellow feldspar were present in relatively similar abundances. From Figure Sr-12 and Sr-15, it can be seen that the proportion of hexagonal calcium alumina yellow feldspar had decreased significantly, forming a smaller volume of calcium alumina yellow feldspar and wrapped by silicate. Additionally, the content of needle-like calcium ironate decreased significantly while that of silicate increased significantly. This was in accordance with the XRD diffraction pattern, which showed that the intensity of silicate (C3S) diffraction was enhanced with an increase in the blast furnace slag doping ratio.
In summary, on the one hand, blast furnace slag high-temperature molten tempered steel slag can refine the grain of calcium aluminum yellow feldspar, reduce the content of needle-like calcium ferrate, and indirectly reduce the cost of steel slag crushing. On the other hand, blast furnace slag blending can reduce the alkalinity of steel slag and promote the generation of a silicate phase with high hydration activity, providing a reference for the blending of steel slag in cement concrete.

3.3. Modified Slag FactSage7.1 Calculation Results and Analysis

In order to further determine the phase composition of the modified steel slag, the equilibrium phase composition of 100 g of tempered slag was simulated using the equilibrium module in the FactSage7.1 thermodynamic calculation software; the results are shown in Figure 7. As shown, in the low temperature section (1000 °C~1250 °C), an increase of blast furnace slag doping results in greater inhibition of the generation of C2S, a-C2S-C3P, and C2F and promotes the generation of magnesia rose pyroxene (Ca3MgSi2O8, C3MS2) and dicalcium aluminate (Ca2Al2O5, C2A). Our analysis suggests that when the temperature is lower than 1250 °C, P2O5 is very easy to solid-solve in C2S to form a-C2S-C3P, which consumes a large amount of C2S and causes the C2S content to decrease sharply at 1000 °C~1250 °C. According to existing research, P2O5 solid solution in C2S can inhibit the crystalline transformation of C2S from β-C2S to γ-C2S and improve the stability of steel slag. The alkalinity of blast furnace slag is about 1.0; it usually contains 13.92% Al2O3 and 8.26% MgO. With an increase of the blast furnace slag doping ratio, the Al2O3 and MgO contents in the slag after tempering increase, and the binary alkalinity decreases. The generation of C3MS2 and C2A in the low temperature section consumes some of the Ca2+ ions and indirectly inhibits the generation of C2S and Ca2Fe2O5, as C3MS2 and C2A are low melting point minerals. When the temperature is higher than 1250 °C, Ca2Fe2O5 and C2A partially melt and C3MS2 completely melts. All compounds enter the liquid phase when the temperature is higher than 1350 °C, thereby enriching the Ca2+ and Si4- ions in the liquid phase. As shown in Figure 7a, the C2S content in the tempered slag mixed with blast furnace slag increased at a higher rate than that in the steel slag without blast furnace slag at temperatures from 1250 °C to 1450 °C, and the C2S content in the equilibrium phase increased gradually as the temperature increased. The C2S content reached a maximum at a temperature of 1400 °C. The steel slag was tempered with 15% blast furnace slag and the C2S content reached 39.04 g. The C2S content in the slag at this temperature was only 16.92 g, i.e., only 1/4 that of the Sr-15 tempered slag. From Figure 7c, it can be seen that Sr-8, Sr-10, and Sr-12 started to enter the liquid phase at 1350 °C, and the amount of liquid phase increased with an increase in the ration of blast furnace slag doping. Our analysis indicated that the steel slag which was not mixed with blast furnace slag did not generate C3MS2 in the low-temperature section, and as such, as the temperature increased to 1350 °C, no low-melting point material melted. Combined with the data shown in Figure 7b), it was found that Sr-0 did not melt at 1350 °C, and therefore, no liquid phase was generated in Sr-0 at that temperature. Sr-15 did not produce a liquid phase at 1350 °C. Our analysis suggests that the liquid phase produced by the melting of C3MS2 and Ca2Fe2O5 at 1350 °C rapidly generated C2S with SiO2 in the steel slag, but, given that the melting temperature of C2S is high, a liquid phase was not reached at 1350 °C. A liquid phase environment can promote exchanges among ions and accelerate reactions. However, a large amount of dopant was shown to produce more non-gelatinized C2AS. Therefore, it was concluded that 12% dopant content is optimal. As shown in Figure 7g, the doping of blast furnace slag has an inhibitory effect on the formation of Ca7P2Si2O16 at temperatures ranging from 1000 °C to 1045 °C. At temperatures above 1050 °C, Sr-0~Sr-12 melted into the liquid phase, while Sr-15 slowly entered the liquid phase as the temperature increased. As shown in Figure 7e), the doping of blast furnace slag can indirectly inhibit the generation of a-C2S-C3P. At a temperature of 1300 °C, the a-C2S-C3P content in the slag not containing blast furnace slag was maximal, i.e., 37.67 g, while the a-C2S-C3P content in the slag tempered with 12% blast furnace slag was only 18.82 g. With an increase in temperature, the a-C2S-C3P content decreased rapidly. A temperature higher than 1350 °C is required to reach the melting point of a-C2S-C3P. As such, experiments had to be carried out at above 1450 °C. A comparison of our analysis compared with the experimental scan results showed that the roasted tempered slag contained a significant amount of C2AS. We concluded that the thermodynamic calculations yielded results for the equilibrium phase in the ideal state, while the experimental environment could not reach an ideal state, and the C2A generated by the tempered slag reacted with the SiO2 in the slag to produce C2AS minerals. This also explains the presence of the silicate phase in the form of CaSiO3 in the roasted specimens. In order to achieve the state described in the thermodynamic calculations, it would be necessary to increase the roasting temperature, extend the holding time, improve the liquid phase fluidity of the slag, improve the high temperature kinetic conditions of the slag, and enhance the electromagnetic stirring, gas stirring, or other stirring methods.
In summary, the doping of blast furnace slag can promote the generation of low melting point minerals such as C3MS2 and C2A, provide a liquid phase for the low temperature section of the tempering slag, improve the low temperature kinetic characteristics of the tempering slag, and promote the generation of C2S, making it possible to calculate the amount of liquid phase in the high temperature section at a roasting temperature of around 1550 °C. Blast furnace slag tempering slag, on the one hand, promotes the generation of low melting point minerals, improves the low temperature kinetic characteristics of the slag, promotes the generation of C2S, inhibits the generation of high strength minerals such as calcium ferrate, and reduces the strength of the slag, thereby indirectly reducing the cost of crushing. On the other hand, blast furnace slag doping can inhibit the generation of a-C2S-C3P, and, according to existing research, can result in the enrichment of P2O5, which has a strong inhibitory effect on the crystalline transformation of C2S. Therefore, reducing the content of a-C2S-C3P in a self-powdering steel slag is advantageous.

4. Conclusions

(1)
In blast furnace slag tempering steel slag, the blast furnace slag doping mass fraction should be about 12% or less in order to promote the generation of minerals with gelling activity such as C3S, to indirectly consume the f-CaO present in the steel slag, to improve the stability, to limit the abundance of non-gelling C2AS, and to achieve higher hardness via C3AS3 mineral generation, thereby reducing the cost of steel slag crushing.
(2)
As shown in the SEM-EDS data, the phases contained in the steel slag were MFA, a magnesium-iron phase solid solution with a low melting point, Ca3(PO4)2 and CaSiO3 (with elliptical shape), C2F (with a milky white color), and C2AS (with a light gray color and hexagonal shape). Blast furnace slag doping, on the one hand, can refine the grain of calcium aluminum feldspar and reduce the content of acicular calcium ferrate in the tempered slag. On the other hand, it can reduce the alkalinity of the slag and promote the generation of a silicate phase with high hydration activity.
(3)
Thermodynamic calculations showed that doping blast furnace slag promotes the generation of low melting point minerals (notably C3MS2 and C2A), provides a liquid phase for the low temperature section (1000 °C~1250 °C), improves the low temperature kinetic characteristics of the tempered slag, promotes the generation of C2S in the high temperature section (1250 °C~1600 °C), inhibits the generation of a-C2S-C3P and calcium ferrate minerals, and reduces the strength of steel slag.

Author Contributions

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

Funding

National Key R&D Program funded project (2020YFC1909105); Inner Mongolia Autonomous Region Science and Technology Major Special Project (2021ZD0016-05-04).

Data Availability Statement

All authors can confirm that all data used in this article are available for publication.

Conflicts of Interest

No conflict of interest exists in the submission of this manuscript, and the manuscript is approved by all authors for publication.

References

  1. Ibrahim, S.; Meawad, A. Towards green concrete: Study the role of waste glass powder on cement/superplasticizer compatibility. J. Build. Eng. 2022, 47, 103751. [Google Scholar] [CrossRef]
  2. Anandaraj, S.; Karthik, S.; Vijaymohan, S.; Rampradheep, G.S.; Indhiradevi, P.; Anusha, G. Effects of using white flour, zinc oxide and zinc ash as an admixture in mortar and concrete. Mater. Today Proc. 2022, 52, 1788–1793. [Google Scholar] [CrossRef]
  3. Cheng, X.; Tian, W.; Gao, J.F.; Gao, P. Performance evaluation and lifetime prediction of steel slag coarse aggregate concrete under sulfate attack. Constr. Build. Mater. 2022, 344, 128203. [Google Scholar] [CrossRef]
  4. Du, X.Q.; Huang, Z.; Ding, Y.; Xu, W.; Zhang, M.; Wei, L.B.; Yang, H.R. Feasibility Study of Grinding Circulating Fluidized Bed Ash as Cement Admixture. Materials 2022, 15, 5610. [Google Scholar] [CrossRef]
  5. Long, H.; Huang, X.Y.; Liu, M.J.; Cui, C.H.; Li, L.; Liao, Y.; Yan, D.H. The fate of heavy metals in the co-processing of solid waste in converter steelmaking. J. Environ. Manag. 2022, 311, 114877. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, G.; Schollbach, K.; Li, P.P.; Brouwers, H.J.H. Valorization of converter steel slag into eco-friendly ultra-high performance concrete by ambient CO2 pre-treatment. Constr. Build. Mater. 2021, 280, 122580. [Google Scholar] [CrossRef]
  7. Yong, Y.; Shaopeng, W.; Chao, L.; Dezhi, K.; Benan, S. Morphological Discrepancy of Various Basic Oxygen Furnace Steel Slags and Road Performance of Corresponding Asphalt Mixtures. Materials 2019, 12, 2322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Yee, C.L. Reutilization of dredged harbor sediment and steel slag by sintering as lightweight aggregate. Process Saf. Environ. Prot. 2019, 126, 287–296. [Google Scholar]
  9. Wang, L.; Wu, L.B.; Tang, Y.; Wang, B.H.; Sun, K.; Ishimwe, A. Liquefaction resistance behaviours of gravel steel slag. Eur. J. Environ. Civ. Eng. 2022, 26, 4643–4663. [Google Scholar] [CrossRef]
  10. Wang, Q.; Long, J.F.; Xu, L.L.; Zhang, Z.; Lv, Y.; Yang, Z.H.; Wu, K. Experimental and modelling study on the deterioration of stabilized soft soil subjected to sulfate attack. Constr. Build. Mater. 2022, 346, 128436. [Google Scholar] [CrossRef]
  11. Ju, J.R.; Feng, Y.L.; Li, H.R.; Xu, C.L.; 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]
  12. Ju, J.R.; Feng, Y.L.; Li, H.R.; Xu, C.L.; Yang, Y. Research progress and prospect of steel slag modification. Environ. Eng. 2020, 39, 136–140. [Google Scholar]
  13. Zhang, X. Carbonation Mechanism of Steel Slag Products Modified by Zeolite; Dalian University of Technology: Dalian, China, 2021. [Google Scholar]
  14. Xiang, R.H. Study on Preparation and Foaming Modification of Reconstituted Steel Slag Powder with Medium and High Activity; Guilin University of Technology: Guilin, China, 2021. [Google Scholar]
  15. Rao, L. Study on Internal Law of Composition, Structure and Properties of Converter Steel Slag and Its Application; University of Science and Technology Beijing: Beijing, China, 2020. [Google Scholar]
  16. Wang, C.L.; Zhao, G.F.; Wang, Y.B.; Zhang, S.H.; Zheng, Y.C.; Huo, Z.K.; Wang, Z.X.; Ren, Z.Z.; Zhou, J.Y. High temperature modification of steel slag from reservoir sediment and calcium carbide slag. Mater. Rev. 2022, 36, 127–133. [Google Scholar]
  17. Zaibo, L.; Sanyin, Z.; Xuguang, Z.; Tusheng, H. Cementitious property modification of basic oxygen furnace steel slag. Constr. Build. Mater. 2013, 48, 575–579. [Google Scholar]
  18. Zhang, Z.S.; Lian, F.; Liao, H.Q.; Yang, Q.; Cao, W.B. Properties of steel slag modified by iron tailings at high temperature. J. Univ. Sci. Technol. Beijing 2012, 34, 1379–1384. [Google Scholar]
  19. Lei, Y.B.; Zhang, Y.Z.; Xing, H.W.; Long, Y.; Tian, T.L. High temperature Melting and digestion of free CaO from converter slag mixed with fly ash. Hebei Metall. 2011, 4, 11–14. [Google Scholar]
  20. Liu, S.Y.; Wang, Z.J.; Peng, B.; Yue, C.S.; Guo, M. Physical and chemical basis of steel slag modified by blast furnace slag. J. Eng. Sci. 2018, 40, 557–564. [Google Scholar]
  21. Zhang, W.; Hao, X.S.; We, C.; Liu, X.M.; Zhang, Z.Q. Activation of low-activity calcium silicate in converter steelmaking slag based on synergy of multiple solid wastes in cementitious material. Constr. Build. Mater. 2022, 351, 128925. [Google Scholar] [CrossRef]
  22. Niu, F.; An, Y.; Zhang, J.; Chen, W.; He, S. Synergistic Excitation Mechanism of CaO-SiO2-Al2O3-SO3 Quaternary Active Cementitious System. Front. Mater. 2021, 8, 792682. [Google Scholar] [CrossRef]
Figure 1. Roasted blast furnace tempered slag specimens.
Figure 1. Roasted blast furnace tempered slag specimens.
Minerals 13 00067 g001
Figure 2. XRD diffraction pattern of blast furnace slag tempered slag roasted specimens.
Figure 2. XRD diffraction pattern of blast furnace slag tempered slag roasted specimens.
Minerals 13 00067 g002
Figure 3. Sr-0 SEM morphology, surface sweep, and energy spectra of four points.
Figure 3. Sr-0 SEM morphology, surface sweep, and energy spectra of four points.
Minerals 13 00067 g003
Figure 4. Sr-8 SEM morphology, surface sweep, and point energy spectra.
Figure 4. Sr-8 SEM morphology, surface sweep, and point energy spectra.
Minerals 13 00067 g004
Figure 5. Sr-12 SEM morphology, surface sweep, and energy spectra of points.
Figure 5. Sr-12 SEM morphology, surface sweep, and energy spectra of points.
Minerals 13 00067 g005
Figure 6. Mineralogram of blast furnace slag tempered slag roasting specimen. (1) to (4) are 300× magnified mineral phases, (5) and (6) are 500× magnified mineral phases.
Figure 6. Mineralogram of blast furnace slag tempered slag roasting specimen. (1) to (4) are 300× magnified mineral phases, (5) and (6) are 500× magnified mineral phases.
Minerals 13 00067 g006
Figure 7. Equilibrium phase composition of tempered slag. (ag) are the masses of Ca2SiO4, Ca2Fe2O5, liquid quantity, Ca2Al2O5, a-C2S-C3P, Ca3MgSi2O8, and Ca7P2Si2O16 in the material phase at equilibrium, respectively.
Figure 7. Equilibrium phase composition of tempered slag. (ag) are the masses of Ca2SiO4, Ca2Fe2O5, liquid quantity, Ca2Al2O5, a-C2S-C3P, Ca3MgSi2O8, and Ca7P2Si2O16 in the material phase at equilibrium, respectively.
Minerals 13 00067 g007
Table 1. Slag proportion and chemical composition (mass fraction, %).
Table 1. Slag proportion and chemical composition (mass fraction, %).
MaterialAlkalinity (R)w(CaO)w(SiO2)w(Al2O3)w(MgO)w(Fe2O3)w(FeO)w(P2O5)w(MnO)
Steel slag3.2741.0512.563.116.8915.3116.441.433.22
Blast furnace slag1.3043.8733.6914.888.831.750.36
Note: Sr-0~Sr-15 are steel slag and modified slag blended with 8, 10, 12 and 15% blast furnace slag, respectively. Alkalinity: Binary alkalinity is used in this paper, i.e., R=CaO/SiO2.
Table 2. Parameters setting of FactSage7.1.
Table 2. Parameters setting of FactSage7.1.
DatabaseFToxid7.1、FactPS7.1
Compound typeMonoxide
Solid solutionFToxide–SLAGA、FToxide–SPANA、FToxide–MeO–A、FToxide–cPyrA、FToxide–oPyr、FToxide–pPyrA、FToxide–LcPy、FToxide–WOLLA、FToxide–bC2S、FToxide–aC2S、FToxide–Mel、FToxide–OlivA
Table 3. Sr-0 SEM point element distribution (mass fraction, %).
Table 3. Sr-0 SEM point element distribution (mass fraction, %).
Elementw(C)w(O)w(Mg)w(Al)w(Si)w(P)w(Ca)w(V)w(Ti)w(Mn)w(Fe)
Point 12.7940.1613.4414.540.1600.635.1423.13
Point 24.7538.960.060.2112.572.1238.841.110.331.06
Point 32.8734.440.4110.753.610.1130.262.752.4112.39
Point 47.6231.854.314.5510.411.628.510.851.428.89
Table 4. Sr-8 SEM point element distribution (mass fraction, %).
Table 4. Sr-8 SEM point element distribution (mass fraction, %).
Elementw(C)w(O)w(Mg)w(Al)w(Si)w(P)w(Ca)w(S)w(V)w(Cr)w(Ti)w(Mn)w(Fe)w(Br)
Point 52.6538.9313.2620.380.0200.441.13.2819.94
Point 63.0534.680.372.413.040.0427.88.891.9817.73
Point 79.5335.250.180.0011.861.6937.031.250.291.091.84
Point 841.4722.240.10.737.151.0724.190.360.650.490.281.27
Table 5. Sr-12 SEM point element distribution (mass fraction, %).
Table 5. Sr-12 SEM point element distribution (mass fraction, %).
Elementw(C)w(O)w(Mg)w(Al)w(Si)w(P)w(Ca)w(Ti)w(Mn)w(Fe)
Point 94.8143.940.243.113.390.0223.747.351.6211.8
Point 103.6741.640.2515.269.280.0625.040.164.65
Point 113.2033.020.251.782.740.0029.689.621.6718.02
Point 127.6731.893.423.079.060.6919.724.9019.57
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

Hao, S.; Luo, G.; Lu, Y.; An, S.; Chai, Y.; Song, W. Effect of High Temperature Reconstruction and Modification on Phase Composition and Structure of Steel Slag. Minerals 2023, 13, 67. https://doi.org/10.3390/min13010067

AMA Style

Hao S, Luo G, Lu Y, An S, Chai Y, Song W. Effect of High Temperature Reconstruction and Modification on Phase Composition and Structure of Steel Slag. Minerals. 2023; 13(1):67. https://doi.org/10.3390/min13010067

Chicago/Turabian Style

Hao, Shuai, Guoping Luo, Yuanyuan Lu, Shengli An, Yifan Chai, and Wei Song. 2023. "Effect of High Temperature Reconstruction and Modification on Phase Composition and Structure of Steel Slag" Minerals 13, no. 1: 67. https://doi.org/10.3390/min13010067

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