Next Article in Journal
A Theoretical Framework for Analysing Technology Transfer Processes Using Agent-Based Modelling: A Case Study on Massive Technology Adoption (AMTEC) Program on Rice Production
Next Article in Special Issue
Simulation of Slag–Matte/Metal Equilibria for Complex and Low-Grade Raw Materials
Previous Article in Journal
A Study on the Sustainable Structural Relations between Social Exchange Relationship Characteristics and Social Contagion Effect in Beauty-Related One-Person Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Recycling of Blast Furnace Slag and Fluorite Tailings into Diopside-Based Glass-Ceramics with Various Nucleating Agents’ Addition

School of Iron and Steel, Soochow University, Suzhou 215137, China
*
Authors to whom correspondence should be addressed.
Sustainability 2021, 13(20), 11144; https://doi.org/10.3390/su132011144
Submission received: 30 August 2021 / Revised: 23 September 2021 / Accepted: 3 October 2021 / Published: 9 October 2021
(This article belongs to the Special Issue Metallurgical Solid Wastes Treatment and Utilization)

Abstract

:
Diopside-based glass-ceramics are successfully produced by recycling blast furnace slag and fluorite tailing with the addition of supplementary elements such as TiO2, Fe2O3 and Cr2O3 as nucleation agents, using a conventional melting method. The effects of various nucleating agents on the phase components and structure of the prepared glass-ceramics were evaluated by a differential scanning calorimeter, X-ray diffraction and scanning electron microscope–energy disperse spectrometer methods to determine the optimal dosage of nucleating agents. The results show that, in the preparation of diopside-based glass-ceramics, the suitable percentages of blast furnace and fluorite tailing are 55% and 45%, and the recommended composite nucleating agents consist of 1.5% Cr2O3, 2% TiO2 and 3% Fe2O3. Heat treatment was conducted at a nucleation temperature of 720 °C and a crystallization temperature of 920 °C, and the nucleation and crystallization durations were 1.0 h and 1.5 h, respectively. Under the abovementioned parameters, the obtained diopside-based glass-ceramics displayed a Vickers hardness of 7.12 GPa, density of 2.95 g·cm−3, water absorption of 0.02%, acid resistance of 0.23% and alkali resistance of 0.02%.

1. Introduction

Blast furnace slags are the major by-products in the ironmaking and steelmaking process, and its output is approximately 0.3 tons per ton of pig iron. In China, approximately 200 billion tons of blast furnace slag are generated per year, 40% of which are subjected to landfilling without further utilization [1,2,3]. Fluorite ore is widely distributed in South Africa, Mexico, and China, and its total reserves exceed 240 million tons [4]. Fluorite tailing is a typical industrial solid waste generated from the flotation process of fluorite ore, and its improper disposal not only occupies a large area of land, but also results in a serious contaminative source. Therefore, the recycle and comprehensive utilization of blast furnace slags and fluorite tailings are significant in order to solve the considerable waste of resources and environmental pollution.
Due to the major components in metallurgy slag (SiO2, CaO, MgO and Al2O3) matching well with the requirements of glass-ceramics, recycling metallurgy slag, especially blast furnace slag, to prepare glass-ceramics through a melting or sintering method has been consistently explored since the 1960s [5,6,7,8,9]. However, considering the basicity of blast furnace slag, the addition of SiO2 is necessary to guarantee the formation of glass phases during the high-temperature treatment. Previous literature [10,11,12] has reported that quartz sand, clay, kaolin and pure silica could be selected as proper supplements due to the relatively higher content of SiO2 during the blast furnace slag glass-ceramics preparation. Besides, some researchers put forward that glass-ceramics could be prepared by blast furnace slag with the addition of other industrial solid waste. Yang et al. [13,14,15,16] investigated the preparation of glass ceramics from blast furnace slag and multi-solid wastes including fly ash, waste glass, red mud, potash feldspar, etc. These research results would be of great interest regarding the potential industrial application of the utilizing industrial solid wastes. With the addition of multi-solid wastes in glass-ceramics preparation, the addition of nucleating agents is also necessary to promote crystallization and improve the physical and chemical properties of the products. Wang et al. [17,18,19,20,21,22] investigated the effect of a single nucleating agent as well as composite nucleating agents on glass-ceramics and demonstrated that the addition of the proper nucleating agent could effectively improve the physical and chemical properties of glass-ceramics. Although the preparation of glass-ceramics from solid wastes indicates a promising future, the fluctuations in chemical composition and the large amounts of impurity in elements seriously limit the usage percentage of solid wastes in glass-ceramics production plants in China.
The diopside-based glass-ceramic in the CaO–MgO–Al2O3–SiO2 (CMAS) system is one of the most widely applied glass-ceramics due to its high strength, high corrosion and high wear resistance [23,24,25]. The main components of blast furnace slag are CaO, MgO, Al2O3 and SiO2, which favor diopside formation by adjusting the CaO/SiO2 ratio in the glass network [26]. The main components of fluorite tailing are SiO2 and CaF2, which have been proven to be effective in crystallization by reducing the viscosity of the crystallizing glass [27,28]. Therefore, mixtures of blast furnace slag and fluorite tailing are suitable for diopside-based glass-ceramics. Our previous research demonstrated that fluorapatite-based glass-ceramics could be obtained from blast furnace slag and fluorite tailing, which exhibit excellent bioactive and mechanical properties [29].
In this study, blast furnace slag and fluorite tailing are employed to prepare diopside glass-ceramics with the addition of different nucleating agents. Firstly, glass-ceramics without the addition of a nucleation agent were investigated to determine the appropriate blast furnace slag and fluorite tailing percentages. Next, the effects of a single nucleation agent, namely TiO2, Fe2O3 and Cr2O3, on glass-ceramics were investigated to obtain the suitable dosage of the nucleation agent. Then, the addition of composite nucleating agents was explored by conducting orthogonal experiments to optimize the properties of diopside-based glass-ceramics. This research aims to elucidate a low-cost preparation method for diopside-based glass-ceramics from typical solid wastes and provide theoretical support for the efficient utilization of blast furnace slag and fluorite tailing.

2. Raw Materials and Experimental Design

The raw materials used in this study include blast furnace slags and fluorite tailings, which are taken from an iron and steel company and ore treatment plant in China. In this article, the selected nucleating agents are TiO2, Fe2O3 and Cr2O3. Considering the purity of the added nucleating agents, the reagent-grade TiO2 (≥99.0%, average particle size ranging 0.1~0.3 μm), Fe2O3 (≥96.0%, particle size less than 5 μm) and Cr2O3 (AR) powders were purchased from the Aladdin company and used in the preparation of glass-ceramic samples via the melting method. The chemical compositions of the blast furnace slags and fluorite tailings are listed in Table 1.
The XRD patterns of the blast furnace slag and fluorite tailing are shown in Figure 1. As typical water-quenched slag, there are no obvious diffraction peaks in the XRD patterns of blast furnace slag due to its amorphous characteristics. In fluorite tailing, the major phases are SiO2, CaF2 and K2O·Al2O3·6SiO2.
In the CaO–MgO–Al2O3–SiO2 (CMAS) system, the diopside-based glass-ceramic exhibits high strength, excellent corrosion and wear resistance. Considering the major components in the blast furnace slag and fluorite tailing, the molar ratios of CaO, SiO2, MgO and Al2O3 were controlled in area 11 to guarantee the precipitation of diopside crystals, as shown in Figure 2. The detailed experimental scheme, according to the main crystalline phase designed in Figure 2, is listed in Table 2. The blast furnace slag percentage ranges from 45% to 65%. Correspondingly, the fluorite tailing percentage ranges from 55% to 35%. The nucleating agents selected in this study are TiO2, Fe2O3 and Cr2O3, and the addition percentages are listed in Table 2.

3. Preparation Process of Glass-Ceramics

The conventional melting method was used to prepare the glass-ceramics from blast furnace slag and fluorite tailing. The detailed preparation process is shown in Figure 3. Firstly, the blast furnace slag, fluorite tailing and nucleating agent powders were mixed homogeneously through ball-milling. During the ball-milling process, the ball-milling duration was 30 min and the ball-milling speed was 400 r/min. The ball-to-material mass ratio was 3.0. Next, the mixtures were charged into a platinum crucible, heated in a muffle furnace to 200 °C with a heating rate of 5 °C/min and kept constant at this temperature for 1 h to remove the moisture. Then, the mixtures were heated to 1500 °C at a heating rate of 5 °C/min and left at this temperature for 2 h until the material completely melted and all bubbles disappeared. After that, a small part of the molten material was quenched with water, and the nucleation and crystallization temperatures were determined by a differential scanning calorimeter (DSC). The remaining samples were poured into a preheating mold, and then annealed at 600 °C. Then, the heat treatment was conducted, wherein the nucleation and crystallization durations were 1.0 h and 1.5 h, respectively. After heat treatment, the glass-ceramic samples were cooled down to room temperature naturally in the furnace to release internal stress.

4. Characterization Methods and Performance Testing

The heat treatment temperatures significantly affect the nucleation and crystallization of glass-ceramics. In this study, the DSC technique was used to obtain the nucleation and crystallization temperatures with a heating rate of 10 °C/min in a nitrogen atmosphere (NETZSCH STA 449F5, NETZSCH, Waldkraiburg, Germany).
The crystalline phases were identified using X-ray diffraction (XRD, Ultima IV, Rigaku, Japan, Cu Kα, 10–90°) with a scanning speed of 2°/min. The fractured surfaces of the glass-ceramic samples were chemically etched for 3–5 s in a 4 vol % HF solution to observe the microstructures using scanning electron microscopy (SEM, SU5000, Hitachi, Japan). Meanwhile, the elemental compositions were determined using energy-dispersive spectroscopy (EDS, X-MAX 20, Oxford, UK).
The properties of the obtained glass-ceramics include Vickers hardness, density, water absorption, and acid and alkali resistance. The Vickers hardness of the glass-ceramics was measured using an automatic turret microhardness tester (HV-1000IS, Shanghai Institute of Optics and Fine Mechanics). Firstly, the glass-ceramics sample was ground and polished. Then, Vickers diamond indenters were used to press the test sample. The load was 500 g and the loading time was 10 s. After the test force was removed, the diagonal length of the indentation was measured. The Vickers hardness was calculated according to Equation (1).
HV = 2 P sin α 2 d 2 = 1.8544 P d 2
where HV is the Vickers hardness (MPa), P is the loading (N), α is the angle between two opposite faces of Vickers diamond indenters and d is the average length of the diagonal indentation (mm).
The water absorption of the glass-ceramics sample was tested by the following methods. Firstly, the sample was dried to a constant weight. Then, the samples were soaked in distilled water at room temperature for 48 h and taken out. The moisture on the surface of the samples was wiped off with a wrung-out wet towel, and the mass was immediately weighed. The water absorption rate was calculated according to Equation (2).
W a = m 0 m 1 m 0 × 100
where Wa is the water absorption (%), m0 is the mass of the sample before immersion (g) and m1 is the mass of the sample after immersion (g).
In the testing process for the chemical-resistant properties, the sample was dried to a constant weight. Then, the sample was immersed in the prepared sulfuric acid solution (1 vol.% chemical pure concentrated sulfuric acid) or sodium hydroxide solution (1.0 wt% chemical pure sodium hydroxide), wherein the liquid level was 30 mm higher than the sample, and the container mouth was sealed. It was soaked for 650 h, then taken out and washed with deionization water until the pH value was neutral. After drying, it was weighed. The mass loss rate was calculated according to Equation (3) to evaluate the acid or alkali corrosion resistance.
K = w 0 w 1 w 0 × 100
where K is the mass loss rate (%), w0 is the mass of the sample before immersion (g) and w1 is the mass of the sample after immersion (g).

5. Results and Discussion

5.1. Effect of Blast Furnace Slag Percentage on Diopside-Based Glass-Ceramics

Figure 4 shows the DSC curves of the parent glass with different percentages of blast furnace slag and fluorite tailing. In the DSC testing, the parent glass samples A1–A5 were heated with a heating rate of 10 K/min under an N2 protective atmosphere. Generally, Tg is defined as the glass transition temperature, which is considered the nucleation starting temperature. Due to the indistinct exothermic peak in the DSC curves, the differential of the DSC curves was introduced to determine the location of the indistinct exothermic peak at Tg. Exothermic peaks associated with glass transition temperatures (Tg) were observed between 676 °C and 700 °C. Sample A5 exhibits the highest Tg, which may be attributed to the lowest fluorite tailing percentage among the five batches. The exothermic peaks of crystallization (Tc) for sample A1 appeared around 886 °C. For samples A2 to A5, there were two different exothermic peaks of crystallization that could be observed, implying a double crystallization process occurred and the crystallization was facilitated by fluorite tailing [29].
According to the literature [27], the nucleation and crystallization temperatures for heat treatment schedules should be set (20–100) °C higher than Tg and Tc. Considering the powder sample used in DSC testing performed at a lower crystallization boundary energy, the difference value was (10–40) °C. Therefore, the nucleation temperature selected was at 710 °C, and the double crystallization temperatures were set at 900 °C and 980 °C. Figure 5 presents the morphologies of samples A1 to A5 after heat treatment. The surface and cross-section appearances indicate that large amounts glass phases existed in the glass-ceramic samples.
Figure 6 shows the XRD patterns of samples A1 to A5 after heat treatment. Though the diopside phase (Ca(Mg,Al)(Si,Al)2O6) peaks could be observed in the five samples, the diffraction peak intensity was insufficient, which indicates that diopside glass-ceramics preparation is difficult when only using blast furnace slag and fluorite tailing without the addition of other nucleating agent. In samples A4 and A5, small amounts of a wollastonite phase (CaSiO3) appeared, which is mainly due to the higher blast furnace slag percentage decreasing the SiO2 content and facilitating the precipitation of crystals with a lower Si/O ratio. In sample A3, the diopside diffraction peak intensity was higher relative to the other samples. Therefore, sample A3, containing a blast furnace slag percentage of 55% and a fluorite tailing percentage of 45%, was selected as the basic case for the subsequent research considering the effect of different nucleating agents.

5.2. Effect of TiO2 Addition on Diopside-Based Glass-Ceramics

Figure 7 shows the DSC curves of the parent glass with different TiO2 content. In the DSC testing, the parent glass samples with different TiO2 content were heated at a heating rate of 10 K/min under an N2 protective atmosphere. With an increase in TiO2 addition from 0 to 5%, the parent glass transition temperature Tg clearly decreased from 683 °C to 664 °C. This is because Ti4+, exhibiting strong field intensity, would exist as [TiO6] or [TiO4] in the glass and destroy the glass network structure, thereby decreasing the stability of the glass phase. Compared with the A3 sample, the crystallization temperature Tc decreased first and then increased. The shape of the crystallization exothermic peak in the Ti3 sample was quite sharp, which indicates the appropriate TiO2 addition could improve crystallization ability. Based on the Tg and Tc obtained from the DSC curves, the nucleation and crystallization temperatures were set at 690 °C and 900 °C, respectively.
Figure 8 shows the XRD patterns, crystallinity and morphologies of the glass-ceramics with different TiO2 content after heat treatment. It should be pointed out that the crystallinity of the glass-ceramics in the present study was analyzed by the multi-peak separation method through Jade software. With the addition of TiO2, the main crystalline phases in the glass-ceramics were anorthite CaAl2Si2O8 and Ti-bearing diopside (Ca,Ti)(Mg,Al)(Si,Al)2O6. In sample Ti1, small amounts of anorthite and Ti-bearing diopside peaks appeared, and the crystallinity in this case was only 30.2%. When the TiO2 content was increased to 3% and 5%, the intensity of CaAl2Si2O8 and (Ca,Ti)(Mg,Al)(Si,Al)2O6 peaks noticeably increased, and the corresponding crystallinity also increased to 45.9% and 50.2%. Therefore, TiO2 addition could effectively facilitate crystallization. Besides, the glass-ceramics samples with TiO2 addition exhibited structural integrity without cracks.
Figure 9 shows the SEM-EDS analysis of the glass-ceramics with different TiO2 content. As shown in Figure 9a, large amounts of glass phases existed in sample Ti1 due to its low crystallinity. From the magnification image in Figure 9b, many crystal grains can be seen distributed in the glass phases. As shown in Figure 9b, large amounts of columnar crystals were precipitated when increasing the TiO2 content to 3%, indicating one-dimensional volumetric crystallization occurred. As shown in the elemental distribution images of sample Ti3, the Ti distribution overlapped with Mg and Al, which is consistent with the precipitation of Ti-bearing diopside (Ca,Ti)(Mg,Al)(Si,Al)2O6 from XRD analysis results. Therefore, it can be deduced that [TiO4] and [TiO6] may act as nucleation particles and promote crystallization. In sample Ti5, the size of columnar crystals increased noticeably with uneven distribution, as shown in Figure 9c. In view of the crystallinity and crystal distribution, the content of TiO2 added should be less than 3%.

5.3. Effect of Fe2O3 Addition on Diopside-Based Glass-Ceramics

Figure 10 shows the DSC curves of the parent glass with different Fe2O3 content in the range of 1 wt% to 5 wt%. Compared with the A3 sample, the nucleation temperature Tg and crystallization temperature Tc increased first and then decreased. Simultaneously, the steepness of the crystallization peak first increased and then decreased with Fe2O3 addition. The influence of Fe2O3 on the glass system could be divided into two different parts [30]. On the one hand, Fe3+ entered the glass phase in the form of [FeO4]5−, which could enhance the viscosity of glass and suppress crystallization. On the other, with a further increased Fe2O3 content, a fraction of Fe3+ occupied the octahedral position in the form of [FeO4]5− as well as Fe2+, which would promote crystallization and decrease the nucleation temperature Tg and crystallization temperature Tc. Based on the Tg and Tc obtained from the DSC curves, the nucleation temperature and crystallization temperature were set at 720 °C and 910 °C, respectively.
Figure 11 shows the XRD patterns, crystallinity and morphologies of the glass-ceramics with different Fe2O3 contents after heat treatment. With Fe2O3 addition, the main crystalline phase in the glass-ceramics is Fe-bearing diopside Ca(Mg,Fe)(Si,Al)2O6. This indicates the Fe2+ occupied the octahedral position in the diopside crystals. The Fe2O3 addition facilitated the crystallization effectively. When the Fe2O3 content was increased, the Ca(Mg,Fe)(Si,Al)2O6 peak intensity gradually increased, and crystallinity was enhanced from 54.1% to 63.0%. The morphologies of the glass-ceramics samples with Fe2O3 addition exhibited a typical double-layer structure. The surface of the samples crystallized, but the glass phase still appeared in the cross section, indicating that Fe2O3 addition could not realize volumetric crystallization.
Figure 12 shows the SEM-EDS analysis of the glass-ceramics with different Fe2O3 contents. A double-layer structure of glass-ceramics was observed and analyzed, as shown in Figure 12a,b. In the outside area of the Fe1 sample, large amounts of diopside crystals appeared. Furthermore, in the inner-side area, the crystals gradually disappeared, and the glass phase still existed. As shown in Figure 12b, the interface between the crystal area and the glass phase area was obvious, indicating that surface crystallization occurred with Fe2O3 addition. When the Fe2O3 content was increased to 3%, both the crystal size and distribution were unform, indicating that 3% Fe2O3 addition is appropriate for preparing diopside glass-ceramics from blast furnace slag and fluorite tailing. In the Fe5 sample, the growth of columnar crystals was facilitated, and the agglomeration of columnar crystals gradually appeared, as shown in Figure 12d. In the elemental distribution images, Fe was distributed homogenously in the crystal and glass phases, which is consistent with the previous literature [31].

5.4. Effect of Cr2O3 Addition on Diopside-Based Glass-Ceramics

Figure 13 shows the DSC curves of the parent glass with different Cr2O3 content. When the Cr2O3 addition was increased from 0 to 4%, the parent glass transition temperature Tg noticeably increased from 683 °C to 715 °C. The crystallization temperature Tc noticeably decreased because crystallization was effectively facilitated by the Cr2O3 introduced. Further increasing the Cr2O3 from 2% to 4%, the crystallization temperature decreased, and the exothermic peak strength was gradually weakened due to the fact that the extra Cr2O3 addition could increase the viscosity of glass and block the atomic diffusion, which indicates that Cr2O3 addition is excessive in the sample Cr3 and Cr4. Previous research demonstrated that the prevention of Cr2O3 precipitating in the glass matrix as a form of chromium spinel crystallites during the cooling stage of glass is mainly responsible for the reduction of viscosity. In contrast, the extra Cr2O3 could promote the precipitation of chromium spinel crystals and increase the viscosity, thereby restraining the bulk nucleation and subsequent crystallization processes [22,32]. Based on the Tg and Tc obtained from the DSC curves, the nucleation temperature and crystallization temperature were set at 720 °C and 910 °C, respectively.
Figure 14a shows the XRD patterns of the glass-ceramics with different Cr2O3 content after heat treatment. In sample Cr1, no diffraction peaks appeared and the sample was still amorphous. When the Cr2O3 content was increased to 2% and 3%, the diopside phase Ca(Mg,Al)(Si,Al)2O6 noticeably precipitated. In sample Cr4, small amounts of NaFe(SiO3)2 and KFeSiO4 phase also precipitated as well as the diopside phase. Noticeably, the intensity of the diffraction peaks was weakened compared with Cr2 and Cr3, which is because higher Cr2O3 content significantly increased the viscosity of glass and suppressed the crystals’ precipitation. As shown in Figure 14b, crystallinity perceptibly increased with an increase in the Cr2O3 content from 1% to 2%, and then decreased slightly with the increase in Cr2O3 content from 2% to 4%, reaching the maximum value of 53% with a Cr2O3 content of 2%.
Figure 15 shows the SEM-EDS analysis of the glass-ceramics with different Cr2O3 contents. In sample Cr1, there were no crystals observed, and large amounts of glass phases existed, which is consistent with the XRD analysis results. In sample Cr2, large amounts of granular crystals precipitated with a particle size of 1–2 μm, indicating three-dimensional volumetric crystallization occurred. Significantly, the granular crystals’ size and distribution were uniform, resulting in a compact structure of diopside glass-ceramic. When the Cr2O3 content was increased to 3%, the viscosity and nucleation points increased, which facilitated the growth of granular crystals and also accelerated the agglomeration of granular crystals. In sample Cr4, the atomic diffusion was further limited by the higher viscosity of glass, and the chromium-rich phases noticeably gathered as shown in the elemental distribution images. As a result, the crystals’ growth mode transformed from three-dimensional volumetric crystallization to one-dimensional volumetric crystallization and the columnar crystals gradually precipitated in sample Cr4. In view of the crystallinity and microstructure, the recommended Cr2O3 addition content is 2%.

5.5. Optimization of Composite Nucleating Agent Addition

Based on the effect of a single nucleating agent on the diopside glass-ceramic, the addition of composite nucleating agents was investigated by conducting orthogonal experiments. The detailed selected levels of the composite nucleating agents are listed in Table 3.
Figure 16 shows the DSC curves of the parent glass with the addition of different composite nucleating agents. The parent glass transition temperature ranged from 659 °C to 678 °C, and the crystallization peaks were located in the temperature range of 868 °C to 888 °C. In case 2 and case 3, the crystallization temperatures were relatively lower, and the shape of the crystallization peaks was quite sharp, which indicates the two cases have a better potential crystallization ability. Based on the DSC analysis results, the nucleation temperature and crystallization temperature were set at 720 °C and 920 °C, respectively.
Figure 17a shows the XRD patterns and crystallinity of the glass-ceramics with composite nucleating agents. In the nine batch samples, the diopside phase precipitated as the main and single crystalline phase, which indicates the addition of composite nucleating agents would not alter the crystal species. The crystallinity of the nine batch samples, shown in Figure 17b, ranged from 60% to 65%. In case 3, the crystallinity reached 64% and 65%, respectively.
Figure 18 presents the SEM images of glass-ceramics with the addition of different composite nucleating agents. In case 1, case 6 and case 8, the bulky crystals appeared with different sizes, which tends to produce a porous structure of the obtained glass-ceramics. In case 3, case 4, case 5, case 7 and case 9, large amounts of granular crystals precipitated and agglomerated, which results in an uneven crystals distribution. In case 2, the crystals’ size decreased to 0.2~0.5 μm. Simultaneously, the granular distribution was more uniform, and the grain boundary cracks diminished gradually. In view of the structure, crystal size and crystal distribution homogeneity, the composite nucleating agents in case 2 are obviously better than the others.
To evaluate the glass-ceramics prepared from blast furnace slag and fluorite tailing, properties including Vickers hardness, density, water absorption, acid resistance and alkali resistance were tested and are listed in Table 4. The best properties of the glass-ceramics were obtained in case 2, which is consistent with the abovementioned XRD and SEM analysis results. Therefore, in the preparation of diopside-based glass-ceramics, the appropriate percentages of blast furnace and fluorite tailing are 55% and 45%, and the recommended composite nucleating agents consist of 1.5% Cr2O3, 2%TiO2 and 3% Fe2O3. Under this condition, the diopside glass-ceramics displayed a Vickers hardness of 7.12 GPa, density of 2.95 g·cm−3, water absorption of 0.02%, acid resistance of 0.23% and alkali resistance of 0.02%.
Cimdins et al. [1] tested the physical and chemical properties of the glass-ceramics prepared from industrial waste, such as metallurgical slag, fly ash, etching refuse, peat, coal ash, glass waste and so on. The reported data include water uptake of 0.34–3.23 wt%, a final density of 2.93–3.05 g/cm3, and bending strength of 80–96 MPa. The material containing only waste had a durability (mass loss) of 3.02% in 0.1 N HCl. It can be seen that the physical and chemical properties of the glass-ceramics in our study are close to the data reported in the literature. Specifically, the acid and alkali resistance of the glass-ceramics prepared from blast furnace slag and fluorite tailings is better than that in the reported data.

6. Conclusions

The effects of nucleating agents on the diopside-based glass-ceramics prepared from blast furnace slag and fluorite tailing were investigated in this paper. The following conclusions could be drawn from this study.
(1)
The crystallization characteristics and mechanisms correlate to the addition of various nucleating agents. With the addition of Fe2O3 (ranging between 1 wt% and 5 wt%) or TiO2 (ranging between 1 wt% and 5 wt%), the parent glass realized surface crystallization, while the addition of Cr2O3 (ranging between 1 wt% and 4 wt%) promoted the transformation of the crystallization mode to three-dimensional volumetric crystallization.
(2)
The diopside-based glass-ceramics could be prepared from 55% blast furnace and 45% fluorite tailing with the addition of composite nucleating agents consisting of 1.5% Cr2O3, 2% TiO2 and 3% Fe2O3. The heat treatment was conducted at a nucleation temperature of 720 °C and a crystallization temperature of 920 °C, and the nucleation and crystallization durations were 1.0 h and 1.5 h, respectively.
(3)
Under optimized components and heat treatment parameters, the obtained diopside-based glass-ceramics displayed a Vickers hardness of 7.12 GPa, density of 2.95 g·cm−3, water absorption of 0.02%, acid resistance of 0.23% and alkali resistance of 0.02%.

Author Contributions

Conceptualization, W.Z. and H.G.; methodology, W.Z. and X.H.; software, S.H.; formal analysis, B.Y.; investigation, X.H.; writing—original draft preparation, W.Z.; writing—review and editing, W.Z. and B.Y.; visualization, S.H.; project administration, D.C. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are especially thankful to the Project of Transformation of Scientific and Technological Achievements of Inner Mongolia Autonomous Region (2019CG073), the National Natural Science Foundation of China (51574169), the Natural Science Research Project of Higher Education in Jiangsu Province (20KJB45002) and the Postdoctoral Research Foundation of Jiangsu Province (7114451120).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shew, H. RPR Cement Technology; Wuhan University of Technology Press: Wuhan, China, 1991; pp. 35–54. [Google Scholar]
  2. He, D.F.; Gao, C.; Pan, J.T.; Xu, A.J. Preparation of glass-ceramics with diopside as the main crystalline phase from low and me-dium titanium-bearing blast furnace slag. Ceram. Int. 2018, 44, 1384–1393. [Google Scholar] [CrossRef]
  3. Deng, L.; Jia, R.; Yun, F.; Zhang, X.; Li, H.; Zhang, M.; Jia, X.; Ren, D.; Li, B. Influence of Cr2O3 on the viscosity and crystallization behavior of glass ceramics based on blast furnace slag. Mater. Chem. Phys. 2020, 240, 122212. [Google Scholar] [CrossRef]
  4. Helbig, C.; Baldauf, H.; Mahnke, J.; Stöckelhuber, K.W.; Schulze, H.J. Investigation of Langmuir monofilms and flotation experi-ments with anionic/cationic collector mixtures. Int. J. Miner. Process. 1998, 53, 135–144. [Google Scholar] [CrossRef]
  5. Strnad, Z. Glass-Ceramic Materials; Elsevier Science Publishers: Amsterdam, The Netherlands, 1986; pp. 76–88. [Google Scholar]
  6. Pioro, L.; Pioro, I. Reprocessing of metallurgical slag into materials for the building industry. Waste Manag. 2004, 24, 371–379. [Google Scholar] [CrossRef]
  7. Guo, J.; Bao, Y.; Wang, M. Steel slag in China: Treatment, recycling, and management. Waste Manag. 2018, 78, 318–330. [Google Scholar] [CrossRef]
  8. Boltakova, N.; Faseeva, G.; Kabirov, R.; Nafikov, R.; Zakharov, Y. Utilization of inorganic industrial wastes in producing construction ceramics. Review of Russian experience for the years 2000–2015. Waste Manag. 2017, 60, 230–246. [Google Scholar] [CrossRef]
  9. Teo, P.T.; Seman, A.A.; Basu, P.; Sharif, N.M. Recycling of Malaysia’s electric arc furnace (EAF) slag waste into heavy-duty green ceramic tile. Waste Manag. 2014, 34, 2697–2708. [Google Scholar] [CrossRef]
  10. Wang, Z.-J.; Ni, W.; Jia, Y.; Zhu, L.-P.; Huang, X.-Y. Crystallization behavior of glass ceramics prepared from the mixture of nickel slag, blast furnace slag and quartz sand. J. Non-Cryst. Solids 2010, 356, 1554–1558. [Google Scholar] [CrossRef]
  11. Zhao, Y.; Chen, D.; Bi, Y.; Long, M. Preparation of low cost glass–ceramics from molten blast furnace slag. Ceram. Int. 2012, 38, 2495–2500. [Google Scholar] [CrossRef]
  12. Mostafa, N.Y.; Shaltout, A.; Abdel-Aal, M.S.; El-Maghraby, A. Sintering mechanism of blast furnace slag–kaolin ceramics. Mater. Des. 2010, 31, 3677–3682. [Google Scholar] [CrossRef]
  13. Yang, S.M.; Zhang, W. Research on glass ceramics of multi-solid waste slag by blast furnace slag and fly ash. Bull. Chin. Ceram. Soc. 2015, 34, 487–491. [Google Scholar]
  14. Gong, X.C.; Chen, Q.W.; Guo, H.W.; Yan, B.J.; Li, H.W.; Pei, F.J. Preparation of sintered glass ceramics with blast furnace slag and waste glass. Metal World 2018, 3, 13–17. [Google Scholar]
  15. Deng, L.; Zhang, X.; Zhang, M.; Jia, X.; Zhang, Z.; Li, B. Structure and properties of in situ synthesized FeSi2-diopside glass ceramic composites from Bayan Obo tailings, blast furnace slag, and fly ash. J. Alloy. Compd. 2019, 785, 932–943. [Google Scholar] [CrossRef]
  16. Yoshikawa, T.; Hirano, S.; Hirai, N.; Tanaka, T. Preparation of Porous Ceramics by the Hydrothermal Reaction of Blast Furnace Slag for Use in a Water-retentive Material. ISIJ Int. 2008, 48, 1322–1324. [Google Scholar] [CrossRef]
  17. Wang, Y.-C.; Xin, W.-B.; Huo, X.-G.; Luo, G.-P.; Zhang, F. Preparation and Properties of Blast Furnace Slag Glass Ceramics Containing Cr2O3. High Temp. Mater. Process. 2019, 38, 726–732. [Google Scholar] [CrossRef]
  18. Karamanov, A.; Hamzawy, E.M.A.; Karamanov, E.; Jordanov, N.B.; Darwish, H. Sintered glass-ceramics and foams by metal-lurgical slag with addition of CaF2. Ceram. Int. 2020, 46, 6507–6516. [Google Scholar] [CrossRef]
  19. Sun, R.J.; He, F.; Wang, L.G.; Zhang, W.T.; Liu, X.Q.; Yang, H.; Xie, J.L. Influence of TiO2 content on structure and properties of blast furnace slag glass-ceramics. Bull. Chin. Ceram. Soc. 2019, 38, 2542–2548. [Google Scholar]
  20. Zhang, W.T.; He, F.; Xiao, Y.L.; Xie, M.Q.; Xie, J.L.; Li, F.X.; Yang, H.; Luo, Z.H. Structure, crystallization mechanism, and prop-erties of glass ceramics from molten blast furnace slag with different B2O3/Al2O3. Mater. Chem. Phys. 2020, 243, 122664. [Google Scholar] [CrossRef]
  21. Zhang, W.; He, F.; Xiao, Y.; Xie, M.; Xie, J.; Sun, R.; Yang, H.; Luo, Z. Effects of Al/Na and heat treatment on the structure and properties of glass ceramics from molten blast furnace slag. Ceram. Int. 2019, 45, 13692–13700. [Google Scholar] [CrossRef]
  22. Khater, G. Influence of Cr2O3, LiF, CaF2 and TiO2 nucleants on the crystallization behavior and microstructure of glass-ceramics based on blast-furnace slag. Ceram. Int. 2011, 37, 2193–2199. [Google Scholar] [CrossRef]
  23. Guo, X.Z.; Cai, X.B.; Song, J.; Yang, G.Y.; Yang, H. Crystallization and microstructure of CaO-MgO-Al2O3-SiO2 glass-ceramics containing complex nucleation agents. J. Non-Cryst. Solids 2014, 405, 63–67. [Google Scholar] [CrossRef]
  24. Hang, H.N.; Cheng, Y.; Yang, Q.Q.; Send, T. Mechanical and tribological properties of calcia-magnesia-alumina-silica-based glass-ceramics prepared by in situ crystallization. Mater. Sci. Eng. A-Struct. 2006, 423, 170–174. [Google Scholar]
  25. Xiao, H.; Cheng, Y.; Yu, L.; Liu, H. A study on the preparation of CMAS glass–ceramics by in situ crystallization. Mater. Sci. Eng. A 2006, 431, 191–195. [Google Scholar] [CrossRef]
  26. Xie, C.S.; Gui, Y.L.; Song, C.Y.; Hu, B.S. Effect of CaO/SiO2; and heat treatment on the microstructure of glass-ceramics from blast furnace slag. Ceram. Silikaty 2016, 60, 146–151. [Google Scholar] [CrossRef] [Green Version]
  27. Likitvanichkul, S.; LaCourse, W.C. Effect of fluorine content on crystallization of canasite glass-ceramics. J. Mater. Sci. 1995, 30, 6151–6155. [Google Scholar] [CrossRef]
  28. Mukherjee, D.P.; Das, S.K. SiO2–Al2O3–CaO glass-ceramics: Effects of CaF2 on crystallization, microstructure and properties. Ceram. Int. 2013, 39, 571–578. [Google Scholar] [CrossRef]
  29. Chen, J.X.; Yan, B.J.; Li, H.W.; Li, P.; Guo, H.W. Vitrification of blast furnace slag and fluorite tailings for giving diop-side-fluorapatite glass-ceramics. Mater. Lett. 2018, 218, 309–312. [Google Scholar] [CrossRef]
  30. Yu, B.; Liang, K.; Gu, S. Effect of crystallization of CaO-P2O5-SiO2-MgO-F- glass-ceramics on its mechanical properties. Int. J. Miner. Metall. Mater. 2002, 9, 212–215. [Google Scholar]
  31. Liu, H.-P.; Huang, X.-F.; Ma, L.-P.; Chen, D.-L.; Shang, Z.-B.; Jiang, M. Effect of Fe2O3 on the crystallization behavior of glass-ceramics produced from naturally cooled yellow phosphorus furnace slag. Int. J. Miner. Metall. Mater. 2017, 24, 316–323. [Google Scholar] [CrossRef]
  32. Rezvani, M.; Eftekhari-Yekta, B.; Solati-Hashjin, M.; Marghussian, V.K. Effect of Cr2O3, Fe2O3 and TiO2 nucleants on the crystalli-zation behaviour of SiO2-Al2O3-CaO-MgO(R2O) glass-ceramics. Ceram. Int. 2005, 31, 75–80. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) blast furnace slag and (b) fluorite tailing.
Figure 1. XRD patterns of (a) blast furnace slag and (b) fluorite tailing.
Sustainability 13 11144 g001
Figure 2. CaO–SiO2–Al2O3 phase diagram with 10% mass MgO.
Figure 2. CaO–SiO2–Al2O3 phase diagram with 10% mass MgO.
Sustainability 13 11144 g002
Figure 3. Preparation process of diopside-based glass-ceramics from blast furnace slag and fluorite tailing.
Figure 3. Preparation process of diopside-based glass-ceramics from blast furnace slag and fluorite tailing.
Sustainability 13 11144 g003
Figure 4. DSC curves of the parent glass samples A1–A5.
Figure 4. DSC curves of the parent glass samples A1–A5.
Sustainability 13 11144 g004
Figure 5. Morphologies of glass-ceramics samples A1–A5 with different crystallization temperatures.
Figure 5. Morphologies of glass-ceramics samples A1–A5 with different crystallization temperatures.
Sustainability 13 11144 g005
Figure 6. XRD of A1 to A5 after heat treatment with a nucleation temperature of 710 °C and a crystallization temperature of 980 °C.
Figure 6. XRD of A1 to A5 after heat treatment with a nucleation temperature of 710 °C and a crystallization temperature of 980 °C.
Sustainability 13 11144 g006
Figure 7. DSC curves of the parent glass with different TiO2 content.
Figure 7. DSC curves of the parent glass with different TiO2 content.
Sustainability 13 11144 g007
Figure 8. XRD patterns and crystallinity of the glass-ceramics with different TiO2 content.
Figure 8. XRD patterns and crystallinity of the glass-ceramics with different TiO2 content.
Sustainability 13 11144 g008
Figure 9. SEM-EDS analysis of the glass-ceramics with different TiO2 content (a) SEM of sample Ti1; (b) high magnification SEM images of sample Ti1; (c) SEM of sample Ti3; (d) SEM of sample Ti5.
Figure 9. SEM-EDS analysis of the glass-ceramics with different TiO2 content (a) SEM of sample Ti1; (b) high magnification SEM images of sample Ti1; (c) SEM of sample Ti3; (d) SEM of sample Ti5.
Sustainability 13 11144 g009
Figure 10. DSC curves of the parent glass with different Fe2O3 content ranging between 1 wt% and 5 wt%.
Figure 10. DSC curves of the parent glass with different Fe2O3 content ranging between 1 wt% and 5 wt%.
Sustainability 13 11144 g010
Figure 11. XRD patterns, crystallinity and morphologies of the glass-ceramics with different Fe2O3 content (Fe1- with 1 wt% Fe2O3 addition, Fe3- with 3 wt% Fe2O3 addition, Fe5- with 5 wt% Fe2O3 addition).
Figure 11. XRD patterns, crystallinity and morphologies of the glass-ceramics with different Fe2O3 content (Fe1- with 1 wt% Fe2O3 addition, Fe3- with 3 wt% Fe2O3 addition, Fe5- with 5 wt% Fe2O3 addition).
Sustainability 13 11144 g011
Figure 12. SEM-EDS analysis of the glass-ceramics with different Fe2O3 content (a) SEM of outside area in sample Fe1; (b) SEM of the interface between crystal and glass phase area in sample Fe1; (c) SEM of sample Fe3; (d) SEM of sample Fe5.
Figure 12. SEM-EDS analysis of the glass-ceramics with different Fe2O3 content (a) SEM of outside area in sample Fe1; (b) SEM of the interface between crystal and glass phase area in sample Fe1; (c) SEM of sample Fe3; (d) SEM of sample Fe5.
Sustainability 13 11144 g012
Figure 13. DSC curves of the parent glass with different Cr2O3 content (A3- without Cr2O3 addition, Cr1- with 1 wt% Cr2O3 addition, Cr2- with 2 wt% Cr2O3 addition, Cr3- with 3 wt% Cr2O3 addition, Cr4- with 4 wt% Cr2O3 addition).
Figure 13. DSC curves of the parent glass with different Cr2O3 content (A3- without Cr2O3 addition, Cr1- with 1 wt% Cr2O3 addition, Cr2- with 2 wt% Cr2O3 addition, Cr3- with 3 wt% Cr2O3 addition, Cr4- with 4 wt% Cr2O3 addition).
Sustainability 13 11144 g013
Figure 14. XRD patterns (a) and crystallinity (b) of the glass-ceramics with different Cr2O3 content.
Figure 14. XRD patterns (a) and crystallinity (b) of the glass-ceramics with different Cr2O3 content.
Sustainability 13 11144 g014
Figure 15. SEM-EDS analysis of the glass-ceramics with different Cr2O3 content.
Figure 15. SEM-EDS analysis of the glass-ceramics with different Cr2O3 content.
Sustainability 13 11144 g015
Figure 16. DSC curves of the parent glass with the addition of different composite nucleating agents.
Figure 16. DSC curves of the parent glass with the addition of different composite nucleating agents.
Sustainability 13 11144 g016
Figure 17. XRD patterns and crystallinity of the glass-ceramics with the addition of different composite nucleating agents (a) XRD patterns; (b) crystallinity.
Figure 17. XRD patterns and crystallinity of the glass-ceramics with the addition of different composite nucleating agents (a) XRD patterns; (b) crystallinity.
Sustainability 13 11144 g017
Figure 18. SEM images of the glass-ceramics with the addition of different composite nucleating agents.
Figure 18. SEM images of the glass-ceramics with the addition of different composite nucleating agents.
Sustainability 13 11144 g018
Table 1. Chemical compositions of the blast furnace slag and fluorite tailing (wt%).
Table 1. Chemical compositions of the blast furnace slag and fluorite tailing (wt%).
ItemsSiO2CaOMgOAl2O3Fe2O3Na2OK2OTiO2F
Blast furnace slag33.9135.1310.0816.510.040.540.610.730.00
Fluorite tailing72.007.640.135.341.040.302.580.077.83
Table 2. Experimental scheme of glass-ceramic preparation.
Table 2. Experimental scheme of glass-ceramic preparation.
CaseBlast Furnace Slag (wt%)Fluorite Tailing (wt%)TiO2 (wt%)Fe2O3 (wt%)Cr2O3 (wt%)
A145.055.0///
A250.050.0///
A355.045.0///
A460.040.0///
A565.035.0///
Ti1mn1.0//
Ti3mn3.0//
Ti5mn5.0//
Fe1mn/1.0/
Fe3mn/3.0/
Fe5mn/5.0/
Cr1mn//1.0
Cr2mn//2.0
Cr3mn//3.0
Cr4mn//4.0
Note: The value of m and n is determined by the results of experiments A1 to A5.
Table 3. Levels of the selected composite nucleating agent components in the orthogonal experiments.
Table 3. Levels of the selected composite nucleating agent components in the orthogonal experiments.
CaseCr2O3TiO2Fe2O3
11.51.02.0
21.52.03.0
31.53.04.0
42.01.03.0
52.02.04.0
62.03.02.0
72.51.04.0
82.52.02.0
92.53.03.0
Table 4. Properties of the glass-ceramics with different composite nucleating agents.
Table 4. Properties of the glass-ceramics with different composite nucleating agents.
CaseVickers Hardness (GPa)Density
(g·cm−3)
Water Absorption (%)Acid Resistance (%)Alkali Resistance (%)
15.912.470.060.480.05
27.122.950.020.230.02
36.962.810.030.320.03
46.612.710.040.420.04
56.852.870.030.290.02
66.012.550.050.440.04
76.732.830.030.310.03
86.112.580.050.440.04
96.522.750.040.390.03
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, W.; Huang, X.; Yan, B.; Hu, S.; Guo, H.; Chen, D. Recycling of Blast Furnace Slag and Fluorite Tailings into Diopside-Based Glass-Ceramics with Various Nucleating Agents’ Addition. Sustainability 2021, 13, 11144. https://doi.org/10.3390/su132011144

AMA Style

Zhao W, Huang X, Yan B, Hu S, Guo H, Chen D. Recycling of Blast Furnace Slag and Fluorite Tailings into Diopside-Based Glass-Ceramics with Various Nucleating Agents’ Addition. Sustainability. 2021; 13(20):11144. https://doi.org/10.3390/su132011144

Chicago/Turabian Style

Zhao, Wei, Xiaofeng Huang, Bingji Yan, Shaoyan Hu, Hongwei Guo, and Dong Chen. 2021. "Recycling of Blast Furnace Slag and Fluorite Tailings into Diopside-Based Glass-Ceramics with Various Nucleating Agents’ Addition" Sustainability 13, no. 20: 11144. https://doi.org/10.3390/su132011144

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