Facile Route for Preparing Refractory Materials from Ferronickel Slag

Feb 2, 2018 - The feasibility of a facile technological route to preparation of refractory materials from a ferronickel slag with the addition of sint...
35 downloads 16 Views 7MB Size
Research Article Cite This: ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

pubs.acs.org/journal/ascecg

Facile Route for Preparing Refractory Materials from Ferronickel Slag with Addition of Magnesia Foquan Gu,† Zhiwei Peng,*,† Yuanbo Zhang,† Huimin Tang,† Lei Ye,† Weiguang Tian,‡ Guoshen Liang,‡ Mingjun Rao,† Guanghui Li,† and Tao Jiang† †

School of Minerals Processing and Bioengineering, Central South University, Changsha, Hunan 410083, China Guangdong Guangqing Metal Technology Co. Ltd., Yangjiang, Guangdong 529500, China



S Supporting Information *

ABSTRACT: The feasibility of a facile technological route to preparation of refractory materials from a ferronickel slag with the addition of sintered magnesia was verified in this study based on the thermodynamics analysis and the experimental exploration of the effect of the sintered magnesia addition on the phase transformation of ferronickel slag during the sintering process. For the first time, the results of thermodynamics calculation, X-ray diffraction (XRD), and scanning electron microscopy-energy dispersive X-ray spectroscopy (SEM-EDS) analyses revealed that the original phase of the slag can be transformed to high melting point phases by addition of MgO during the sintering process at high temperatures (e.g., 1350 °C). Specifically, the olivine in ferronickel slag decomposed initially, generating a low-iron olivine phase and an enstatite phase. With increasing addition of sintered magnesia, the enstatite phase changed to forsterite, and the iron, aluminum, and chromium components in the ferronickel slag converted to high melting point spinel phases, including magnesium aluminate spinel and magnesium chromate spinel via a low-magnesium transient phase. The experimental results showed that a good refractory material with refractoriness of 1660 °C, bulk density of 2.92 g/cm3, apparent porosity of 1.82%, and compressive strength of 100.61 MPa could be obtained when the slag was sintered with addition of 20 wt % sintered magnesia at 1350 °C for 3 h. Due to the low production cost and property superiority of the prepared refractory material over commercial counterparts, the method proposed in this study is expected to have widespread applications in recycling of ferronickel slag. KEYWORDS: Ferronickel slag, Refractory material, Sintering, Magnesia, Resources utilization



cement (as additive),11−16 concrete (as substitute of aggregates),17−20 glass ceramics,21−23 fire-resistant bricks,24 antislip pavement tiles,24 and geopolymers.25−32 Although the above applications of the ferronickel slag are technically feasible, the low activity of ferronickel slag, high contents of hazardous elements, the limited usage of slag and high operation cost restrain their development. Therefore, the majority of the slag is still temporarily disposed. Disposing this ferronickel production waste occupies a lot of land resources, leading to potential soil and water contamination. Therefore, it is essential to seek a feasible and viable route for treatment of the ferronickel slag with generation of value-added products. The main components of the ferronickel slag are magnesia and silica, which exhibit a similar composition to forsterite refractory materials. It was reported that value-added forsterite refractory materials ($1100−1300/t)33 could be produced using high-magnesia and silica waste as raw materials, including

INTRODUCTION During the production of ferronickel alloy, ferronickel slag is generated as a byproduct, after cooling of molten slag with water or air, of smelting of laterite ore in a high-temperature electric arc furnace with addition of a reducing agent.1 In 2016, the annual global output of ferronickel slag increased to about 11.934 Mt.2,3 Due to its large quantity and hazardous components (e.g., chromium), appropriate handling such a huge quantity of slag is a huge challenge to the ferronickel industry. For treatment of ferronickel slag, many efforts have been spent over the past few decades. Some studies examined the feasibility of recovering valuable metals (Ni, Co, Cu, et al.) directly from ferronickel slag via pyrometallurgical4−6 or hydrometallurgical methods.7−10 However, due to the low contents of these metals in the slag, the pyrometallurgical method has poor elemental recovery and thus low practicability.6 For the hydrometallurgical method, high-content magnesia and silica in the ferronickel slag often consume a large quantity of leaching agent and cause difficult subsequent separation of valuable components in solution.7 At present, only a small part of the produced ferronickel slag is recycled and used in production of © XXXX American Chemical Society

Received: November 20, 2017 Revised: January 27, 2018 Published: February 2, 2018 A

DOI: 10.1021/acssuschemeng.7b04336 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering ferrochromium slag,34 iron ore tailing35 and amorphous rice husk,36 after properly altering their chemical compositions at 1500−1650 °C for 3−6 h. Hence it may be feasible to prepare forsterite refractory materials by a similar approach from ferronickel slag despite the lack of relevant report in literature. Furthermore, the relatively high chromium content in the ferronickel slag may serve as an advantageous factor for preparation of refractory material as the Cr(III) contained in ferronickel slag can be stabilized in the high melting point spinel phase without leachability during the refractory material sintering process.37−40 This strategy will not only solve the potential chromium contamination of ferronickel slag in the preparation but also improve the refractoriness of the material. In the present study, the feasibility of a new and efficient technological route for preparing value-added refractory materials from a ferronickel slag was verified by altering its chemical and phase composition in the presence of low-cost sintered magnesia based on the thermodynamics analysis and experimental exploration of the effects of magnesia addition on the phase transformation of ferronickel slag during the sintering process. The findings are expected to provide a novel and useful guide for treatment of ferronickel slag.



Initially, the mixture of grounded ferronickel slag (86 wt % particles passing 74 μm sieve) and sintered magnesia (89 wt % particles passing 74 μm sieve) was carefully mixed in a planetary ball mill for 20 min. Briquettes of 20 mm in diameter and 20 mm in height were produced by manual hydraulic pressing at 100 MPa using 5 wt % of magnesium chloride solution (1.3 g/cm3) as the binder. The briquettes were then dried in a vacuum drying oven at 105 °C for 4 h. After drying, the briquettes were placed in a muffle furnace before sintering. After sintering at a designated temperature for a given period of time, the briquettes were cooled to room temperature and then taken out for the subsequent characterization tests. Instrumental Analyses. The chemical compositions of the samples were examined using an X-ray fluorescence spectrometer (XRF, PANalytical, Axios mAX, Almelo, The Netherlands). The phase constituents of the samples were identified by an X-ray diffraction spectrometer (XRD, D/max 2550PC, Japan Rigaku Co., Ltd.) using a Cu-anode target with the wavelength of 1.54056 Å, scan mode for the step scan, and step length of 0.02°. The morphological change of the sample during sintering was detected by using an electron scanning electron microscope (ESEM; FEI QUANTA 200; FEI, Eindhoven, The Netherlands) equipped with an EDAX energy dispersive X-ray spectroscopy (EDS) detector (EDAX Inc., Mahwah, NJ). Testing Standards. The values of refractoriness, bulk density and apparent porosity, and compressive strength of the refractory materials were measured according to the Chinese National Standard Test Methods (GBT 7322-2007, GBT 2997-2000, and GBT 5072-2008, respectively).



EXPERIMENTAL SECTION

RESULTS AND DISCUSSION Thermodynamic Analysis. The melting and crystallization behaviors of ferronickel slag can be deduced from the phase diagram of the MgO−SiO2−FeO system. Figure 1 shows the MgO−SiO2−FeO ternary phase diagram at 1350 °C. The red point in Figure 1 indicates the composition of ferronickel slag, and the dotted arrow indicates the possibility of altering its chemical composition by MgO addition. It is obvious that the refractoriness of the ferronickel slag can be increased by addition of MgO. Actually, for the ferronickel slag sintered in air atmosphere, the olivine will undergo the following reactions:43−45

Materials. Ferronickel Slag. The ferronickel slag sample was obtained from the Rotary Kiln-Electric Furnace process of a ferronickel smelting plant. Its chemical composition is shown in Table S1. The sample was characterized by very low content of nickel oxide (0.06 wt %), high contents of silica and magnesia (48.29 and 30.95 wt %, respectively), and low magnesia/silica ratio (0.64). The X-ray diffraction (XRD) pattern (Figure S1) shows that the ferronickel slag was mainly composed of olivine. The SEM-EDS analysis of ferronickel slag (Figure S2) shows that its microstructure was constituted by three different colored regions. The dark region (point 1 in Figure S2) contained a phase whose chemical composition was close to olivine, which was in agreement with the XRD results. The gray region (point 2 in Figure S2) had a complex chemical composition with silica as the main ingredient. Between the gray and dark regions, a darker gray region (point 3 in Figure S2), with chemical composition similar to that of the gray region, was observed. Sintered Magnesia. Sintered magnesia was used as a source of MgO in this study due to its relatively high MgO content, low cost ($200−300/t),41 and high availability.42 The chemical composition of the sintered magnesia is given in Table S2, which shows the content of MgO was 94.51 wt %. The XRD pattern (Figure S3) shows that the sintered magnesia mainly consisted of periclase. Magnesium Chloride. Magnesium chloride of analytical grade was used as a binder in the preparation of refractory materials from ferronickel slag. Methods. Thermodynamic Calculation. For analysis of melting and crystallization behaviors of the ferronickel slag, the ternary phase diagrams of MgO−SiO2−FeO and MgO−SiO2−Fe2O3 systems were calculated by the software FactSage 7.0. Note that the slag contained only 0.06 wt % NiO. Therefore, its effect on the phase transformation is not shown in the phase diagrams. To evaluate the effect of MgO addition on the evolution of phases in ferronickel slag, the variations of the contents of thermodynamic equilibrium phases of the slag with addition of magnesia during the sintering process were also determined by the software. Experimental Procedure. In theory, the suitable molar ratio of magnesia to silica in forsterite refractory material is 0.94−1.33.43 According to the chemical compositions of ferronickel slag and sintered magnesia, the theoretical addition of sintered magnesia should be in the range of 14.1−26.31 wt % to ensure forsterite as the main phase of the refractory material. In this study, the addition of sintered magnesia was varied from 0% to 35 wt %.

2[(Mg·Fe)O·SiO2 ] + 3/2O2 → 2MgO·SiO2 + SiO2 + Fe2O3 (1)

2MgO·SiO2 + SiO2 → 2(MgO·SiO2 )

(2)

MgO·SiO2 + MgO → 2MgO·SiO2

(3)

In this case, the melting and crystallization of ferronickel slag follow the transformation from the MgO−SiO2−FeO to MgO−SiO2−Fe2O3 system. Hence the MgO−SiO2−Fe2O3 system should be used to explain the possible change process. As shown in Figure 2, the potential balanced phases of ferronickel slag at 1350 °C are SiO2, Mg2SiO3, and Fe2O3. In other words, with the addition of MgO, the original phases will gradually change to spinel and Mg2SiO4, and then to spinel, monoxide, and Mg2SiO4. According to the above phase diagrams analysis, one can conclude that the system’s refractoriness can be improved through addition of MgO, irrespective of the form of iron in the ferronickel slag (ferric or ferrous oxide). Theoretically, the effect of MgO addition on the evolution of phases in ferronickel slag can be determined by evaluating the variations of contents of thermodynamic equilibrium phases of the slag in the presence of MgO during the sintering process, as shown in Figure 3. It is observed that ferronickel slag contains high-refractoriness phases, such as forsterite and spinel, during the sintering process. Individual ferronickel slag will generate a B

DOI: 10.1021/acssuschemeng.7b04336 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 1. Phase diagram of the MgO−SiO2−FeO system.

large amount of liquid phase (olivine), which reduces the refractoriness of ferronickel slag. However, with the increase of MgO addition, the amounts of high melting point phases, including forsterite, spinel, and monoxide, in the ferronickel slag system are increased. Meanwhile, the quantity of the liquid phase declines rapidly. With addition of 20 wt % MgO, the content of forsterite increases to a maximum, and the content of spinel has a slight change when the addition of MgO is further increased. Obviously, with the MgO addition of 20 wt %, the ferronickel slag may achieve a high refractoriness. Phase Transformation Mechanism of Ferronickel Slag System during the Sintering Process. The phase transformation of ferronickel slag in the presence of magnesia during the sintering process was experimentally determined by investigating the effect of addition of magnesia varying from 0 to 35 wt % at the sintering temperature of 1350 °C for 3 h. The obtained refractory materials were characterized by XRD and SEM-EDS techniques. X-ray Diffraction Analysis. Figure 4 shows the XRD patterns of the refractory materials sintered with different additions of magnesia. The main phases of the refractory materials in the absence of magnesia were olivine and enstatite, which were similar to those of the ferronickel slag. As the addition of magnesia increased to 5 wt %, two additional phases of the refractory materials, namely forsterite and magnesium chrome spinel, were observed. With the further addition of magnesia, the phases of

forsterite and magnesium chromate spinel remained, while the diffraction peaks of enstatite gradually decreased and finally vanished. When the addition of magnesia exceeded 15 wt %, the magnesium aluminate spinel formed. As the addition of magnesia increased continuously to 25 wt %, a periclase phase was observed. These conversions are summarized in Figure 5. It clearly demonstrates that the phase transformation of the original phase, olivine in the slag, undergoes four consecutive stages during the sintering process with increasing amount of magnesia (Note that iron oxide generated from decomposition of olivine entered into the spinels, as confirmed by Figures 6−8 and S4.) The addition of magnesia is expected to promote the generation of high-refractoriness phases. SEM-EDS Analysis. The morphologies of the refractory materials sintered were examined by the SEM-EDS technique. The results of samples 1 (0 wt % magnesia, 1350 °C, 3 h), 2 (20 wt % magnesia, 1350 °C, 3 h), 3 (35 wt % magnesia, 1350 °C, 3 h), and 4 (20 wt % magnesia, 1350 °C, 5 h) are shown in Figures 6−8 and S4, respectively. As shown in Figure 6, the morphologies of sample 1 had obviously changes compared with those of ferronickel slag. Sample 1 presented three different phases. As revealed by EDS results in Figure 6, spots 1−3 were constituted by an iron-rich phase, a silica-rich phase, and olivine, respectively. As shown in Figures 7 and 8, with increasing magnesia addition, the phases had slight changes. The iron contents in the iron-enriched and olivine phases were gradually C

DOI: 10.1021/acssuschemeng.7b04336 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 2. Phase diagram of the MgO−SiO2−Fe2O3 system.

confirmed by comparing the compositions of samples 1 (Figure 6) and 4 (Figure S4). With increasing sintering time and magnesia addition, the amount of phase of spot 2 decreased considerably (Figure 8). Properties of the Refractory Materials. The effect of magnesia addition on the refractoriness of the refractory material was studied when the sample was sintered at 1350 °C for 3 h. As shown in Figure 9, when the magnesia addition increased from 0 to 20 wt %, the refractoriness of the refractory material increased from 1438 to 1660 °C. With further increase of the magnesia addition, the refractoriness of refractory material increased slightly. According to the changing trend of refractoriness, one can divide it into three stages with increasing addition of magnesia. As shown in Figure 9, within the addition range of 0−10 wt % magnesia, olivine remained as the dominant phase, producing a refractory material with low refractoriness. In the addition range of 10−20 wt % magnesia, the amounts of forsterite and spinel increased continuously, significantly enhancing the refractoriness of the refractory material. When the addition of magnesia increased to 20−35 wt %, the forsterite and spinel phases became dominant, leading to a high refractoriness. These results demonstrated that the addition of magnesia facilitated the formation of high melting point phases, thereby substantially promoting refractory property of the target material, which agreed well with the XRD results in Figure 4. The effects of magnesia addition on the bulk density, apparent porosity, and compressive strength of refractory material was also evaluated under the same sintering conditions, as shown in Figure 10. When the magnesia addition increased

Figure 3. Calculated contents of thermodynamic equilibrium phases in the ferronickel slag system with addition of MgO at 1350 °C.

decreased, while the content of the magnesium in these three phases was increased. Based on the XRD results in Figure 4, it is inferred that the olivine in ferronickel slag will undergo decomposition, as shown by eqs 1−2, resulting in a low-iron olivine phase and an enstatite phase. During the sintering process, the enstatite reacted with the sintered magnesia, forming a forsterite phase, as shown by eq 3. The components of iron, aluminum, and chromium are enriched in the phase of spot 1, producing spinel. Compared with spots 1 and 3, spot 2 was actually composed of a low-magnesium transient phase, which would eventually change to the spinel phases. This transition can be D

DOI: 10.1021/acssuschemeng.7b04336 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 4. XRD patterns of refractory materials sintered with different additions of magnesia.

refractoriness was restrained. When the magnesia addition increased continuously, the amount of spinel increased. Because the recrystallization capacity of spinel is weak and its formation is accompanied by a large volume expansion,43,45 which is harmful to the densification of the material, excessive magnesia addition will lower the bulk density of the refractory material. As expected, the changing tendency of bulk density was opposite to the variation of the apparent porosity but was consistent with that of the compressive strength of the refractory material. From Figures 9 and 10, a good refractory material with refractoriness of 1660 °C, bulk density of 2.92 g/cm3, apparent porosity of 1.82%, and compressive strength of 100.61 MPa was obtained when the ferronickel slag was sintered with addition of 20 wt % sintered magnesia at 1350 °C for 3 h. According to the Chinese National Refractory Material Production Standard (GB/T 2275-2007), the refractory materials obtained in this study are superior to those of commercial refractory materials

from 0 to 5 wt %, the bulk density of refractory material decreased from 2.85 to 2.74 g/cm3. As the magnesia addition was further increased, the bulk density of refractory material increased gradually. With the addition of 20 wt % magnesia, the bulk density of refractory material reached the maximum. However, as the magnesia addition exceeded 20 wt %, the bulk density of refractory material decreased significantly. Obviously, the changes were closely associated with the phase transformations of the sample during sintering. There was a lot of liquid phase (olivine) formed in the refractory material in the absence of magnesia, contributing to the densification of the material. This was followed by a reduction in densification caused by a little addition of magnesia (up to 5 wt %). However, with increasing magnesia addition to 20 wt %, the phase transformation of the refractory materials became more complete, repromoting densification of the material with higher bulk density. Meanwhile, the negative effect of liquid phase on E

DOI: 10.1021/acssuschemeng.7b04336 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 5. Phase transformations of refractory materials sintered with different additions of magnesia.

Figure 6. SEM-EDS analysis of refractory material (sample 1). F

DOI: 10.1021/acssuschemeng.7b04336 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 7. SEM-EDS analysis of refractory material (sample 2).



Figure 8. SEM-EDS analysis of refractory material (sample 3).

CONCLUSIONS The feasibility of a new and facile technological route for preparing refractory materials from a ferronickel slag was verified by evaluating the effect of magnesia addition on the preparation of refractory material from ferronickel slag based on both thermodynamic and experimental results. The thermodynamic analysis determined the thermodynamic equilibrium phases in

(magnesite brick trademark M-93, M-91, M-89, and M-87). It should be mentioned that the proposed technological route is also energy conservative because of its much lower sintering temperature in comparison with commercial ones (1350 °C vs 1500−1650 °C).39,46−49 Evidently, this study offers a very promising guide for facile production of value-added refractory materials from metallurgical slag with low cost of raw materials. G

DOI: 10.1021/acssuschemeng.7b04336 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 9. Effect of magnesia addition on the refractoriness of refractory material.

Figure 10. Effect of magnesia addition on the bulk density, apparent porosity, and compressive strength of refractory material.

ferronickel slag with addition of magnesia at 1350 °C, showing that the contents of forsterite and high melting point spinel phases increased with addition of magnesia. The X-ray diffraction and SEM-EDS studies revealed that the major phase, olivine, in ferronickel slag underwent decomposition, producing a low-iron olivine phase and an enstatite phase. The newly generated enstatite then reacted with the sintered magnesia, forming a forsterite phase. Meanwhile, the reactions between iron, aluminum, and chromium components in the slag and the sintered magnesia produced high melting point spinel phases (magnesium chromate spinel and magnesium aluminate spinel) via a low-magnesium transient phase. The formation of forsterite and the spinel phases contributed to superior properties (e.g., higher refractoriness) of the prepared refractory materials in comparison with commercial counterparts. The experimental

results showed that under the conditions of addition of 20 wt % sintered magnesia, sintering temperature of 1350 °C, and sintering time of 3 h, a good refractory material with refractoriness of 1660 °C, bulk density of 2.92 g/cm3, apparent porosity of 1.82%, and compressive strength of 100.61 MPa could be obtained. Overall, the proposed technological route is featured by low production cost, easy operation, and energy conservation.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.7b04336. Main chemical composition of ferronickel slag (Table S1); main chemical composition of sintered magnesia H

DOI: 10.1021/acssuschemeng.7b04336 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering



(11) Katsiotis, N. S.; Tsakiridis, P. E.; Velissariou, D.; Katsiotis, M. S.; Alhassan, S. M.; Beazi, M. Utilization of ferronickel slag as additive in portland cement: a hydration leaching study. Waste Biomass Valorization 2015, 6 (2), 177−189. (12) Owolabi, O. B.; Adeosun, S. O.; Aduloju, S. C.; Metu, C. S.; Onyedum, O. Review on novel application of slag fluxes and salts in metallurgical industry. Am. J. Chem. Mater. Sci. 2016, 3 (1), 1−5. (13) Vangelatos, I.; Angelopoulos, G. N.; Boufounos, D. Utilization of ferroalumina as raw material in the production of ordinary Portland cement. J. Hazard. Mater. 2009, 168 (1), 473−478. (14) Rahman, M. A.; Sarker, P. K.; Shaikh, F. U. A.; Saha, A. K. Soundness and compressive strength of Portland cement blended with ground granulated ferronickel slag. Constr. Build Mater. 2017, 140, 194−202. (15) Saha, A. K.; Sarker, P. K. Expansion due to alkali-silica reaction of ferronickel slag fine aggregate in OPC and blended cement mortars. Constr. Build Mater. 2016, 123, 135−142. (16) Lemonis, N.; Tsakiridis, P. E.; Katsiotis, N. S.; Antiohos, S.; Papageorgiou, D.; Katsiotis, M. S.; Beazi-Katsioti, M. Hydration study of ternary blended cements containing ferronickel slag and natural pozzolan. Constr. Build Mater. 2015, 81, 130−139. (17) Saha, A. K.; Sarker, P. K. Sustainable use of ferronickel slag fine aggregate and fly ash in structural concrete: Mechanical properties and leaching study. J. Cleaner Prod. 2017, 162, 438−448. (18) Saha, A. K.; Sarker, P. K. Compressive strength of Mortar containing ferronickel slag as replacement of natural sand. Procedia Eng. 2017, 171, 689−694. (19) Wang, G.; Thompson, R.; Wang, Y. Hot-mix asphalt that contains nickel slag aggregate: laboratory evaluation of use in highway construction. Transp. Res. Rec. 2011, 2208, 1−8. (20) Tiwari, A.; Singh, S.; Nagar, R. Feasibility assessment for partial replacement of fine aggregate to attain cleaner production perspective in concrete: a review. J. Cleaner Prod. 2016, 135, 490−507. (21) Ljatifi, E.; Kamusheva, A.; Grozdanov, A.; Paunović, P.; Karamanov, A. Optimal thermal cycle for production of glass-ceramic based on wastes from ferronickel manufacture. Ceram. Int. 2015, 41 (9), 11379−11386. (22) Karamanov, A.; Paunović, P.; Ranguelov, B.; Ljatifi, E.; Kamusheva, A.; Nacevski, G.; Karamanova, E.; Grozdanov, A. Vitrification of hazardous Fe-Ni wastes into glass-ceramic with fine crystalline structure and elevated exploitation characteristics. J. Environ. Chem. Eng. 2017, 5 (1), 432−441. (23) Rawlings, R. D.; Wu, J. P.; Boccaccini, A. R. Glass-ceramics: their production from wastes-a review. J. Mater. Sci. 2006, 41, 733− 761. (24) Kirillidi, Y.; Frogoudakis, E. Electric arc furnace slag utilization. In Proceedings of the 9th International Conference on Environmental Science and Technology, Rhodes, Greece, 2005; pp 768−772. (25) Maragkos, I.; Giannopoulou, I. P.; Panias, D. Synthesis of ferronickel slag-based geopolymers. Miner. Eng. 2009, 22 (2), 196− 203. (26) Komnitsas, K.; Zaharaki, D.; Perdikatsis, V. Effect of synthesis parameters on the compressive strength of low-calcium ferronickel slag inorganic polymers. J. Hazard. Mater. 2009, 161 (2−3), 760−768. (27) Sakkas, K.; Nomikos, P.; Sofianos, A.; Panias, D. Utilisation of FeNi-slag for the production of inorganic polymeric materials for construction or for passive fire protection. Waste Biomass Valorization 2014, 5 (3), 403−410. (28) Zhang, Z.; Zhu, Y.; Yang, T.; Li, L.; Zhu, H.; Wang, H. Conversion of local industrial wastes into greener cement through geopolymer technology: a case study of high-magnesium nickel slag. J. Cleaner Prod. 2017, 141, 463−471. (29) Zaharaki, D.; Komnitsas, K. Long term behaviour of ferronickel slag inorganic polymers in various environments. Fresen. Environ. Bull. 2012, 21 (8C), 2436−2440. (30) Komnitsas, K.; Zaharaki, D.; Bartzas, G. Effect of sulphate and nitrate anions on heavy metal immobilisation in ferronickel slag geopolymers. Appl. Clay Sci. 2013, 73, 103−109.

(Table S2); XRD pattern of ferronickel slag (Figure S1); SEM-EDS analysis of ferronickel slag (Figure S2); XRD pattern of sintered magnesia (Figure S3); and SEM-EDS analysis of refractory material (sample 4; Figure S4) (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: +86-731-88877656. Fax: +86-731-88830542 (Z.P.). ORCID

Zhiwei Peng: 0000-0003-1720-0749 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was partially supported by the National Natural Science Foundation of China under Grants 51774337 and 51504297, the Natural Science Foundation of Hunan Province, China, under Grant 2017JJ3383, the Key Laboratory for Solid Waste Management and Environment Safety (Tsinghua University) Open Fund under Grant SWMES2017-04, the Hunan Provincial Co-Innovation Center for Clean and Efficient Utilization of Strategic Metal Mineral Resources under Grant 2014-405, the Guangdong Guangqing Metal Technology Co. Ltd. under Grant 738010210, the Innovation-Driven Program of Central South University under Grant 2016CXS021, and the Shenghua Lieying Program of Central South University under Grant 502035001.



REFERENCES

(1) Dourdounis, E.; Stivanakis, V.; Angelopoulos, G. N.; Chaniotakis, E.; Frogoudakis, E.; Papanastasiou, D.; Papamantellos, D. C. Highalumina cement production from FeNi-ERF slag, limestone and diasporic bauxite. Cem. Concr. Res. 2004, 34 (6), 941−947. (2) http://www.insg.org/docs/INSG_Press_Release_April_2017. pdf (accessed on November 20, 2017). (3) Balomenos, E.; Dimitrios, P. Iron recovery and production of high added value products from the metallurgical by-products of primary aluminum and ferronickel industries. In Proceedings of the 3rd International Slag Valorisation Symposium; Belgium, Leuven, 2013; pp 161−72. (4) Pan, J.; Zheng, G. L.; Zhu, D. Q.; Zhou, X. L. Utilization of nickel slag using selective reduction followed by magnetic separation. Trans. Nonferrous Met. Soc. China 2013, 23 (11), 3421−3427. (5) Perederiy, I.; Papangelakis, V. G.; Buarzaiga, M.; Mihaylov, I. Cotreatment of converter slag and pyrrhotite tailings via high pressure oxidative leaching. J. Hazard. Mater. 2011, 194, 399−406. (6) Shen, Y. F.; Xue, W. F.; Li, W.; Tang, Y. L. Selective recovery of nickel and cobalt from cobalt-enriched Ni-Cu matte by two-stage counter-current leaching. Sep. Purif. Technol. 2008, 60, 113−119. (7) Shen, H. T.; Forssberg, E. An overview of recovery of metals from slags. Waste Manage. 2003, 23, 933−949. (8) Li, Y. J.; Papangelakis, V. G.; Perederiy, I. High pressure oxidative acid leaching of nickel smelter slag: Characterization of feed and residue. Hydrometallurgy 2009, 97, 185−193. (9) Huang, F. R.; Liao, Y. L.; Zhou, J.; Wang, Y. Y.; Li, H. Selective recovery of valuable metals from nickel converter slag at elevated temperature with sulfuric acid solution. Sep. Purif. Technol. 2015, 156, 572−581. (10) Ettler, V.; Kvapil, J.; Sebek, O.; Johan, Z.; Mihaljevic, M.; Ratie, G.; Garnier, J.; Quantin, C. Leaching behaviour of slag and fly ash from laterite nickel ore smelting (Niquelandia, Brazil). Appl. Geochem. 2016, 64, 118−127. I

DOI: 10.1021/acssuschemeng.7b04336 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering (31) Yang, T.; Yao, X.; Zhang, Z. H. Geopolymer prepared with highmagnesium nickel slag: Characterization of properties and microstructure. Constr. Build Mater. 2014, 59, 188−194. (32) Choi, Y.; Choi, S. Alkali-silica reactivity of cementitious materials using ferro-nickel slag fine aggregates produced in different cooling conditions. Constr. Build Mater. 2015, 99, 279−287. (33) http://www.nhclsq.com/cpsb.php?zycp=%C3%BE%D7%A9 (accessed on November 20, 2017). (34) Sahu, N.; Biswas, A.; Kapure, G. U. Development of refractory material from water quenched granulated ferrochromium slag. Miner. Process. Extr. Metall. Rev. 2016, 37 (4), 255−263. (35) Li, J.; Wang, Q.; Liu, J. H.; Li, P. Synthesis process of forsterite refractory by iron ore tailings. J. Environ. Sci. 2009, 21, S92−S95. (36) Hossain, S. K. S.; Mathur, L.; Singh, P.; Majhi, M. R. Preparation of forsterite refractory using highly abundant amorphous rice husk silica for thermal insulation. J. Asian Ceram. Soc. 2017, 5 (2), 82. (37) Wojsa, J.; Podwórny, J.; Suwak, R. Thermal shock resistance of magnesia-chrome refractories experimental and criterial evaluation. Ceram. Int. 2013, 39, 1−12. (38) Gerlach, N.; Gehre, P.; Aneziris, C. G. Improvement of Magnesia Refractory Ceramics for Applications in Gasifiers. Chem. Ing. Tech. 2014, 86, 1761−1768. (39) Othman, A. G. M.; Khalil, N. M. Sintering of magnesia refractories through the formation of periclase-forsterite-spinel phases. Ceram. Int. 2005, 31 (8), 1117−1121. (40) Miretzky, P.; Cirelli, A. F. Cr (VI) and Cr (III) removal from aqueous solution by raw and modified lignocellulosic materials: A review. J. Hazard. Mater. 2010, 180, 1−19. (41) Liu, Q. S. A study on China’s magnesia export; Dongbei University of Finance and Economics: Dalian, China, 2012. (42) Quan, Y. Production and Application of Magnesia Refractories; Metallurgical Industry Press: Beijing, China, 2008. (43) Xue, Q. H.; Xu, W. Z. Refractory materials; Metallurgical Industry Press: Beijing, China, 2013. (44) Tavangarian, F.; Emadi, R.; Shafyei, A. Influence of mechanical activation and thermal treatment time on nanoparticle forsterite formation mechanism. Powder Technol. 2010, 198 (3), 412−416. (45) Lin, B. Y.; Hu, L. Raw materials of refractory; Metallurgical Industry Press: Beijing, China, 2015. (46) Mohapatra, D.; Sarkar, D. Preparation of MgO-MgAl2O4 composite for refractory application. J. Mater. Process. Technol. 2007, 189, 279−283. (47) Ganesh, I.; Bhattacharjee, B.; Saha, B. P.; Johnson, R.; Mahajan, Y. R. A new sintering aid for magnesium aluminate spinel. Ceram. Int. 2001, 27, 773−779. (48) Visloguzova, E. A.; Piontkevich, O. V.; Protasov, V. V. Experience in utilization of forsterite refractories. Metallurgy 2002, 2, 46−47. (49) Mustafa, E.; Khalil, N. M.; Gamal, A. G. Sintering and microstructure of spinel-forsterite bodies. Ceram. Int. 2002, 28, 663− 667.

J

DOI: 10.1021/acssuschemeng.7b04336 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX