Next Article in Journal
A Review on Direct Laser Deposition of Inconel 625 and Inconel 625-Based Composites—Challenges and Prospects
Next Article in Special Issue
The Synthesis of Lead-Bearing Jarosite and Its Occurrence Characteristic and Leaching Toxicity Evaluation
Previous Article in Journal
Evaluation of the Influence Exerted by Increased Silicon Contents on the Leaching Behavior of NMC-Based Black Mass
Previous Article in Special Issue
Preparation of Black Ceramic Tiles with Chromium Slag and Copper Smelting Waste Slag
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Relation between Viscosity and Conductivity of CaO-MgO-FeO-Al2O3-SiO2 System for Copper Smelting Slags

1
School of Metallurgy, Northeastern University, Shenyang 110819, China
2
Metallurgical Research and Design Institute, BGRIMM Technology Group, Beijing 100044, China
3
Zijin School of Geology and Mining, Fuzhou University, Fuzhou 350108, China
*
Author to whom correspondence should be addressed.
Metals 2023, 13(4), 786; https://doi.org/10.3390/met13040786
Submission received: 6 March 2023 / Revised: 7 April 2023 / Accepted: 15 April 2023 / Published: 17 April 2023

Abstract

:
The viscosity and conductivity of the smelting slag of copper oxide concentrate are important for reducing the operating temperature. In this study, seven slag samples were prepared by the reductive smelting of copper oxide concentrate with different ferrous oxide contents. The viscosity and conductivity data of these CaO-MgO-FeO-Al2O3-SiO2 samples were measured in the temperature range of 1290~1410 °C. Based on the structural features of aluminosilicate melts, the change and dependency relationships of their viscosity and conductivity were analyzed. The results show that there is a strong tendency to form orthosilicate even when the slag composition is acidic. The formation of fayalite would allow more Al3+ to form pyroxene with the six-coordinated structure. As a result, the polymerization degree and viscosity of the melt will be reduced. The [ AlO ] 4 5 as a network former will reduce the bonding strength of the structural units, thus reducing the slag viscosity at high temperature. In the experimental range, the logarithm of viscosity of each slag sample has a good linear relationship with its logarithm of conductivity. However, there is no uniform linear equation for these complex slags with wide composition variations. These results have potential guiding significance for the copper smelting process.

1. Introduction

The African Copperbelt is the second-largest and most important copper source in the world [1]. Its gangue component is mainly silica, in addition to a small amount of alumina, magnesium oxide, calcium oxide, and iron oxide. Blast furnace reduction smelting is very suitable for this high-grade copper oxide with Cu > 15% [2]. It has been used in the industrial practice of metal extraction from copper oxide concentrate in Africa. Enhancing the reductive atmosphere in the blast furnace is beneficial for improving production efficiency but will strengthen iron reduction [3]. Therefore, the slag type must be improved to reduce iron reduction.
The crude copper produced by the high-calcium slag has a high grade, but the slag has a high viscosity and requires a higher operating temperature. Although the appropriate addition of iron oxide will reduce the slag viscosity and operating temperature, the reduction degree of iron and the slag viscosity need to be controlled. The study of slag viscosity under different calcium oxide and ferrous oxide contents has important guiding significance for reducing operating temperature. In addition, when the slag viscosity is high, the product of blast furnace smelting needs to be pumped into the front bed to further separate copper from the slag. Generally, direct current power is used to maintain the temperature of the slag. Therefore, the conductivity of slag and its variation characteristics are also important for the separation of copper.
When the alumino-silicate melt is completely melted, it can be regarded as a Newtonian fluid, and the relation between the slag viscosity and temperature is in accordance with the Arrhenius relation (1) [4,5].
l n η = A η + E η / R T
The hole theory of melts considers that the liquid has solid-like structural units (SU), oscillating near the average position in their energy cells (potential wells) [6,7]. Oscillations of magnitude above the barrier cause structural units to move into adjacent empty units, or “holes”, provided the latter is empty. Thus, the viscosity of the liquid is determined by the ability of the structural units to jump over the barrier and the presence of “holes” in the liquid. E η and η 0 are related to the activation energy of structural units jumping into adjacent holes and the probability of hole formation, respectively.
The viscosities of alkaline earth silicate or aluminosilicate melts are determined by the polymerization degree of their networks [5]. It has been widely accepted to classify all cations into three behaviors, including glass former (Si4+, Ge4+, etc.), modifier (Na+, Mg2+, Ca2+, etc.), and amphoteric gelling agents (Al3+, Fe3+, etc.). However, the addition of basic oxides can transform the random network structure into a more depolymerized structure containing chains and rings [8,9,10]. Fincham and Richardson [11] suggested that there are three types of oxygen in silicate melts, including bridging oxygen (BO, O0), non-bridging oxygen (NBO, O), and free oxygen (FO, O2−). Schramm et al. [12,13] suggested that the Qn notation can be used to describe the different coordinated states of molten silicates and aluminates. Based on the Raman spectroscopy and deconvolution technique, Mysen et al. [14] evaluated the degree of polymerization of the melts by distinguishing Q0, Q1, Q2, and Q3 units and their relative abundance. Structural depolymerization occurs in three distinctive mechanisms, each of which dominates the structural depolymerization in each region of non-bridging oxygen number per tetrahedrally coordinated atom (NBO/T) [15]. In fact, both the n value in Qn and NBO/T reflect the number of shared vertices of [SiO4]4− basic structure units [16,17]. With the increase in polymerization degree of [SiO4]4−, the structural units gradually become larger and exist in the forms of rings, single chains, double chains, sheets, etc. Therefore, both the ability of structural units to jump potential barriers and the probability of creating “holes” in the liquid are decreased, thus leading to a high slag viscosity value.
Aluminosilicate melts consist of tetrahedral [SiO4]4− and [AlO4]5− units, while the [AlO4]5− tetrahedron requires charge compensation due to its valence electron difference from that of [SiO4]4−. Zhang et al. suggested [18] that the priority for charge compensation is Ca2+ > Mg2+. There are some obvious characteristics, such as the lubricant effect [19], the charge compensation effect for tetrahedrally coordinated aluminum [5], and the weak lubricant effect [20]. The local viscosity is maximum around the fayalite composition in the FeO-SiO2 system [21], which reflects the influence mechanism of composition and structure on the viscosity of aluminosilicate slags. The lubricant effect and charge compensation effect also have a significant influence on the viscosity by changing the depolymerization degree. For silicate slags, the depolymerization degree depends on the O/Si ratio. For aluminosilicate slags, Al2O3 will exacerbate the depolymerization of the network structure as an alkaline oxide, while the polymerization of the network structure will be improved by the charge compensation effect when it acts as an acidic oxide. However, few studies have analyzed the effect of alumina on the slag viscosity.
In this work, the viscosity and conductivity of the slags formed in the smelting process of copper oxide concentrate with different ferrous oxide contents were measured. The relation between the viscosity and conductivity of slags was analyzed. On the basis of the above theory, the reasons for the change in slag viscosity and conductivity with composition and temperature were discussed. These results will be instructive in reducing the slag viscosity and operating temperature of the copper smelting process.

2. Experiment

2.1. Preparation of Slag Samples

The copper oxide concentrate used in the experiment was from the Koluwezi mining area in the Democratic Republic of Congo. It contains 20.92 wt.% copper, and the main mineral phases are quartz, malachite, and clinochlorite. The coke ratio is 5 wt.% and the carbon content in coke is more than 83 wt.%. The calcium oxide and ferrous oxalate used for slag preparation are analytically pure. The CaO/SiO2 mass ratio was set to 0.4, and the FeO/SiO2 mass ratio varied between 0.41 and 1.25. The mixture of copper oxide concentrate with 8 wt.% water content, slagging agent, and coke powder was granulated under 40 MPa pressure and then dried at 300 °C for 1 h. The dry material was put into a corundum crucible to reduce and melt with a sufficient reductive atmosphere in a muffle furnace at 1400 °C for 2 h. After the furnace was cooled, the crucible was taken out and broken to separate slag and crude copper. The slag samples were weighted and used for subsequent experiments. The compositions of smelting slags were determined by inductively coupled plasma-atomic emission spectrometry (ThermoFisher, ICAP PRO XP, Waltham, MA, USA). The NBO/T parameter, which represents the polymerization degree of melt, was calculated by Equation (2). And the NBO/S (nonbridging oxygen per silicon) can be calculated by Equations (3) and (4) by considering the Al2O3 as a basic oxide or not, respectively [22].
NBO/T = 2(XCaO + XFeO + XMgO − XAl2O3)/(XSiO2 + 2XAl2O3)
NBO/S = (2XCaO + 2XFeO + 2XMgO + 6XAl2O3)/XSiO2
NBO/S = (2XCaO + 2XFeO + 2XMgO)/XSiO2
where the X is the mole fraction of each oxide.

2.2. Determination of Slag Viscosity

The slag viscosity was measured on a self-developed RTW-10 melt physical property tester (Northeast University, Boston, MA, USA) by a rotating cylinder method. The specifications of the apparatus and crucible were the same as the reference [23]. The pre-melted sample (150 g) was put inside a molybdenum crucible and heated in a rotatory viscometer. The highly purified Ar atmosphere (0.35 L/min) was continuously used to prevent the oxidation of the spindle and slag during the test process. After the desired temperature was reached and maintained for more than 30 min, the slag melt was homogenized. The molybdenum spindle was connected to the viscometer through an alumina shaft, then immersed into the liquid slag 10 mm above the crucible base and rotated at a fixed speed of 200 r/min. The viscosity was measured at different temperatures, with a temperature step of 30 °C during the temperature-dropping process.

2.3. Determination of Slag Conductivity

The electrical conductivities of slag melts were determined with the contact compensation method by measuring the alternating current resistance of a melt in an alundum crucible in an Ar atmosphere using an electrochemical comprehensive tester (Versa STAT 3F, AMETEK SI, Berwyn, PA, USA). The measuring cell consisted of two electrodes of a molybdenum wire enclosed in two channels of a single alundum rod. The distance between the measuring electrodes and their immersion depth in the melt was 20 mm. The melt temperature range was 1290–1410 °C, and measurements were performed upon steplike cooling of the melt at a step of 30 °C.

3. Results and Discussion

3.1. Chemical Composition of Slag Samples

The compositions of the obtained smelting slags under different FeO/SiO2 burden ratios are shown in Table 1. The reducing atmosphere in the slag production process basically maintains the iron in the slag in the Fe/FeO equilibrium state, so it can be considered that no Fe2O3 exists in the slag. The copper contents in the slags vary from 0.13 to 0.42 wt.%, and the equivalent molar fraction of Cu2O is between 0.001 and 0.002. Therefore, the influence of Fe2O3 and copper on slag viscosity and conductivity can be ignored due to its low content, indicating that the slag belongs to the typical CaO-FeO-MgO-Al2O3-SiO2 system. Table 2 shows the measurement results of slag viscosity and conductivity. It can be found that the viscosity and conductivity gradually decrease with the increase in the FeO/SiO2 ratio but increase sharply at 0.79 and 0.87, which may result from the structural change of the slags. In addition, according to the measured melting temperature of slags, except for the flow temperature of the Z6 slag, which is close to 1290 °C, the other melting temperatures are lower than the test temperature (1290~1410 °C). Therefore, the slags will not crystallize and are suitable for Newtonian fluid type.

3.2. Effect of NBO/T and Temperature on Viscosity

Based on Equations (2)–(4) and Table 1, the NBO/T and NBO/Si values can be obtained and shown in Table 3. In general, the increase in NBO/T implies depolymerization of the melt structural units, which will result in the decrease in slag viscosity. However, this feature cannot be found in Figure 1. It demonstrates a very complicated relationship between viscosity and NBO/T of slags. Moreover, the viscosity change of the same slag sample is more different at different temperatures. At 1290 °C, the viscosity of the slag increases first with the increase in the NBO/T value. When NBO/T is greater than 1.45, the viscosity value decreases rapidly and then stabilizes. The viscosity change of the slag follows a similar pattern, despite the increase in temperature. However, because high temperature is beneficial for reducing the slag viscosity, the fluctuation of the viscosity value with NBO/T is obviously weakened.
For simplicity, the viscosity changes of Z1, Z2, and Z3 samples were further analyzed based on slag basicity. Without considering Al2O3, the basicities of these three samples are 0.74, 0.81, and 0.90, respectively (Table 3), indicating all of them are acidic slags. If the Al2O3 in the slag is taken as basic oxide, the basicity of the Z1, Z2, and Z3 samples will increase to 1.01, 1.10, and 1.23, respectively. The corresponding NBO/Si values of the Z1, Z2, and Z3 slags also increase to 2.03, 2.20, and 2.46. Hence, their basicity becomes gradually stronger, and the added Al2O3 can act as a network-forming agent together with SiO2. Therefore, the polymerization degree of aluminosilicate melt increases gradually, which will result in an increase in viscosity. This variation feature can be verified by the viscosity change at 1290 °C. However, when Al2O3 acts as a network former, the polymerization strength of the structural units becomes weaker. With the increase in temperature, the polymerization degree of aluminosilicate will reduce, resulting in a decrease in slag viscosity.
Notably, the Z4 slag has a maximum viscosity value over the whole temperature range since more Al2O3 is available to act as a network-forming agent. The viscosity value becomes weaker with the increase in temperature due to the weak polymerization strength of [ AlO ] 4 5 . When Al2O3 is considered to act as a network former, the NBO/T value of the Z4 slag is 1.45. The structural units are still limited to the highly polymerized mica-like sheet and amphibole-like double chain. In addition, the NBO/T values of the Z1, Z2, Z3, and Z4 slags are in the range of 1.09~1.45. Hence, the number of shared vertices of SiO 4 4 and [ AlO ] 4 5 is between 2.91 and 2.51, and the structural units correspond to the transition region of sheet mica and amphibole. However, the Z5 slag has an NBO/T value of 1.62 (slightly higher than 1.5), indicating the formation of a structural unit intermediate between amphibole (double chain) and pyroxene (single chain). The sharp decrease in polymerization degree resulted in the lowest viscosity value for the Z5 slag at each temperature. The NBO/T value of the Z6 slag increases to 1.83, and its polymerization degree and viscosity should be lower than that of Z5. However, the Al2O3 mole fraction of slag Z6 reaches 0.069, which is higher than that of the Z5 slag, thus resulting in more [ AlO ] 4 5 to improve the polymerization of structural units. Therefore, the viscosity of Z6 slag slightly increases even at higher NBO/T values. Similarly, the Z7 slag has a molar fraction of 0.059 for Al2O3, resulting in an increase in viscosity.

3.3. Effect of NBO/T and Temperature on Conductivity

Figure 2 shows that the conductivity values of the Z1, Z2, and Z3 slags drop gradually with the increase in NBO/T. These results imply that the increase in network former and the polymerization degree of structural units leads to a gradual increase in the binding of conductive cations, which in turn results in a decrease in conductivity. The Z4 slag contains enough Al2O3 to act as a network former. However, the intervention of excess [ AlO ] 4 5 will lead to a reduction in the bonding strength of the structural units, thus reducing its binding ability to the cations. Meanwhile, the increase in NBO/T will also improve the polymerization of structural units, which will make cation migration more difficult. These two effects contribute to the slight increase in conductivity.
As for the Z5 slag, the structural units have passed through the transition region of infinite sheet clays and into the amphibole and pyroxene chains because of its NBO/T = 1.62. As a result, the binding between the slag structural units and conductive cations would be significantly reduced, and its conductivity would increase rapidly. However, the Z6 slag has a significantly higher Al2O3 content than the Z5 slag and needs more calcium ions for charge compensation. The increase in NBO/T comes mainly from the increase in FeO content, but both FeO and SiO2 have a strong tendency to form fayalite, resulting in a reduction in free cations that can migrate and conduct electricity, thus reducing the conductivity.

3.4. Relationship between Conductivity and Viscosity

Several studies illustrated the relationship between conductivity and viscosity of various electrolytes [24], based on Walden’s rule [25]. This empirical rule indicated that the product of viscosity and equivalent ionic conductance at infinite dilution in electrolyte solution is constant. In oxide melts, however, there is little evidence that Walden’s rule is valid. The prediction of conductivity based on melt viscosity cannot be reliably done, but a strong correlation between these two properties is evident in some melts but not in all [24]. The general rule is that there is a linear relationship between the logarithm of conductivity and viscosity, but the slope and its corresponding influence degree are different for diverse systems. Even for the same CaO-MgO-Al2O3-SiO2 system, the deviation reaches 30.7% for the data of Winterhager [26] and 6.3% for the data of Sarkar [27], respectively. For the CaO-MgO-FeO-Al2O3-SiO2 system, Wanli Li et al. [28] obtained the relation l n η   = 1.89 − 1.14 l n κ with a mean deviation of 14.92%.
Figure 3a shows the good linear fitting result between the logarithm of conductivity and viscosity for each slag. The results in Table 4 show that both viscosity and conductivity can be well described by the Arrhenius law (Equations (5) and (6)). Meanwhile, as temperature increases, the slag viscosity decreases while its conductivity increases. Therefore, the temperature dependence of the conductivity is opposite to that of the viscosity.
ln κ = A κ E κ / R T
l n η + nln κ = C + Δ E / R T
where C = A η + n A κ , Δ E = ( E η − n E κ ). For a specific system, if there exists a constant ‘‘n’’ that always fulfills the condition of E η = n E κ regardless of composition, the logarithm of the viscosity will be the linear function of the logarithm of the conductivity. Therefore, it is theoretically possible to build a linear function between them. The linear relationship between the logarithm of viscosity and conductivity can be described with Equation (7).
l n η   =   C n ln κ
The values of E η and A η can be calculated based on Equation (1) and the data in Table 2. Therefore, the E κ and A κ   values can be obtained and summarized in Table 4. It can be seen from Figure 3b that the linear fitting slope of CaO-MgO-FeO-Al2O3-SiO2 systems is highly dependent on the NBO/T value of slag. The above data show that the viscosity and conductivity of slag depend on the depolymerization and ion transport behaviors of the aggregated structural units, respectively. The gradually increasing fitting slope indicates that at NBO/T less than 1.29, the structural transition of the slag leads to a more pronounced negative correlation between viscosity and conductivity. This phenomenon has been weakened when NBO/T further increases, suggesting that increasing the FeO component enhances the depolymerization of slag and thus decreases its viscosity.

3.5. SEM Analysis of Z3 Slag

Because the Z3 slag has low viscosity at high temperatures and does not need to add too much iron ore ingredient, it is a suitable slag type for site application. Therefore, the Z3 slag sample was further analyzed by SEM analysis, as shown in Figure 4. The element content at point 1 (Figure 4a,b) shows the presence of the pyroxene phase with a single-chain structure, although the pyroxene phase may form when NBO/T is between 1.5 and 2.0. However, the NBO/T of the Z3 slag is only 1.29, which means that the structural unit will be between the sheet (mica) and the double chain (amphibole). Moreover, the NBO/Si of Z3 will reach up to 2.46 when Al2O3 is considered a basic oxide. Therefore, a large fraction of Al2O3 must be used as a basic oxide to satisfy the requirement for the formation of pyroxene. The general chemical formula of pyroxene group minerals is W1−p(X, Y)1+pZ2O6, where W represents Ca2+ and Na+, X represents Mg2+ and Fe2+, Y represents smaller Fe3+, Al3+ and Ti3+, and Z represents Si4+ and Al3+ [27]. It can be inferred that a large part of Al3+ will enter the pyroxene phase as a six-coordinated form represented by Y, while other Al3+ will become the network former of the pyroxene chain as a four-coordinated form represented by Z.
The elemental composition in point 2 (Figure 4a,c) corresponds to the fayalite phase. The results show that a small amount of Ca2+ and Mg2+ ions are mixed in the fayalite, which is an orthosilicate formed at NBO/T = 4. Therefore, the presence of fayalite in the Z3 slag indicates that ferrous oxide and silicon dioxide have a strong tendency to form orthosilicate. The basicity of Z3 was only 0.9 when Al2O3 was ignored, hence the formation of fayalite, which will allow more Al3+ to form pyroxene with a six-coordinated structure. As a result, the polymerization degree and viscosity of the melt will be reduced. It is in accordance with the viscosity change feature in Figure 1. The composition of point 3 (Figure 4a,d) is a copper-bearing Fe alloy with a particle size of 25 to 30 μm, which belongs to metal droplets that can be separated from the slag. It can be seen that the slag is mainly composed of pyroxene and fayalite phases (Figure 4e). There are three main reasons for the certain deviation from the designed fayalite slag type. First, in order to reduce the added dosage for the desired forsterite slag type, the original basic components of the concentrate, alumina, and silica, were designed as metasilicate components. At the same time, the added amount of iron oxide was reduced. The formation of fayalite is incomplete because the slag composition could not achieve the alkalinity required for complete protosilicate. Moreover, the partial reduction of the added ferrous oxide reduces the basicity of the slag and makes the pyroxene mineral the main phase of the slag.

4. Conclusions

Seven slag samples were prepared by reductive smelting of copper oxide concentrate under the different ferrous oxide contents. The viscosity and conductivity data of these slag samples were measured and discussed based on the change of structural units of the aluminosilicate melts. The results indicate that even if the slag composition is acidic, there is a strong tendency to form protosilicates. The formation of fayalite would allow more Al3+ to form pyroxene with a six-coordinated structure. [ AlO ] 4 5 as a network former will reduce the bonding strength of the structural units, thus reducing its viscosity at high temperatures. In the experimental range, the viscosity of each slag sample has a good linear logarithm relationship with its conductivity. However, there is no uniform linear equation for the CaO-MgO-FeO-Al2O3-SiO2 system with a wide range of composition variation.

Author Contributions

Conceptualization, F.X. and D.L.; methodology, K.J.; validation, F.X. and K.J.; formal analysis, L.Z.; investigation, L.Z.; resources, D.L.; writing—original draft preparation, L.Z.; writing—review and editing, D.L.; visualization, L.Z.; supervision, F.X.; funding acquisition, D.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Key R&D Program of China [2019YFC1907304]. And The APC was funded by [2019YFC1907304].

Data Availability Statement

All data are available from the authors as requested.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mambwe, P.; Swennen, R.; Cailteux, J.; Mumba, C.; Dewaele, S.; Muchez, P. Review of the origin of breccias and their resource potential in the Central Africa Copperbelt. Ore Geol. Rev. 2023, 156, 105389. [Google Scholar] [CrossRef]
  2. Zhou, W.; Liu, X.; Lyu, X.; Gao, W.; Su, H.; Li, C. Extraction and separation of copper and iron from copper smelting slag: A review. J. Clean. Prod. 2022, 368, 133095. [Google Scholar] [CrossRef]
  3. Souza, R.; Queiroz, C.; Brant, J.; Brocchi, E. Pyrometallurgical processing of a low copper content concentrate based on a thermodynamic assessment. Miner. Eng. 2019, 130, 156–164. [Google Scholar] [CrossRef]
  4. Tang, K.; van der Eijk, C.; Gouttebroze, S.; Du, Q.; Safarian, J.; Tranell, G. Rheological properties of Al2O3–CaO–SiO2 slags. Calphad 2022, 77, 102421. [Google Scholar] [CrossRef]
  5. Kim, T.S.; Park, J.H. Viscosity-Structure Relationship of CaO–Al2O3–FetO–SiO2–MgO Ruhrstahl-Heraeus (RH) Refining Slags. ISIJ Int. 2021, 61, 724–733. [Google Scholar] [CrossRef]
  6. Weymann, H.D. On the hole theory of viscosity, compressibility, and expansivity of liquids. Kolloid-Z. Und Z. Für Polym. 1962, 181, 131–137. [Google Scholar] [CrossRef]
  7. Suzuki, M.; Jak, E. Development of a Quasi-chemical Viscosity Model for Fully Liquid Slags in the Al2O3–CaO–‘FeO’–MgO–SiO2 System: The Revised Model to Incorporate Ferric Oxide. ISIJ Int. 2014, 54, 2134–2143. [Google Scholar] [CrossRef]
  8. Urbain, G. Viscosity estimation of slags. Steel Res. 1987, 58, 111–116. [Google Scholar] [CrossRef]
  9. Dai, B.; Wu, X.; Zhang, L. Establishing a novel and yet simple methodology based on the use of modified inclined plane (M-IP) for high-temperature slag viscosity measurement. Fuel 2018, 233, 299–308. [Google Scholar] [CrossRef]
  10. Zhang, G.-H.; Chou, K.-C.; Xue, Q.-G.; Mills, K.C. Modeling Viscosities of CaO-MgO-FeO-MnO-SiO2 Molten Slags. Metall. Mater. Trans. B 2012, 43, 64–72. [Google Scholar] [CrossRef]
  11. Fincham, C.J.B.; Richardson, F.D. The behavior of sulphur in silicate and aluminate melts. Philos. Trans. R. Soc. A 1954, 223, 40–62. [Google Scholar]
  12. Jong, B.D.; Schramm, C.M.; Parziale, V.E. Silicon-29 magic angle spinning NMR study on local silicon environments in amorphous and crystalline lithium silicates. J. Am. Chem. Soc. 1984, 106, 4396–4402. [Google Scholar] [CrossRef]
  13. Schramm, S.; Oldfield, E. High-resolution oxygen-17 NMR of solids. J. Am. Chem. Soc. 1984, 106, 2502–2506. [Google Scholar] [CrossRef]
  14. Mysen, B.O.; Virgo, D.; Scarfe, C.M. Relations between the anionic structure and viscosity of silicate melts—A Raman spectroscopicstudy. Am. Mineral. 1980, 65, 290–301. [Google Scholar]
  15. Min, D.J.; Tsukihashi, F. Recent Advances in Understanding Physical Properties of Metallurgical Slags. Met. Mater. Int. 2017, 23, 1. [Google Scholar] [CrossRef]
  16. Engelhardt, L.G.; Zeigan, D.; Jancke, H.; Wieker, W.; Hoebbel, D. 29Si-NMR-Spektroskopie an Silicatlösungen. II. Zur Abhängigkeit der Struktur der Silicatanionen in wäßrigen Natriumsilicatlösungen vom Na: Si-Verhältnis. Z. Für Anorg. Und Allg. Chem. 1975, 418, 17–28. [Google Scholar] [CrossRef]
  17. De Jong, B.; Keefer, K.D.; Brown, G.E., Jr.; Taylor, C.M. Polymerization of silicate and aluminate tetrahedra in glasses, melts, and aqueous solutions—III. Local silicon environments and internal nucleation in silicate glasses. Geochim. Cosmochim. Acta 1981, 45, 1291–1308. [Google Scholar] [CrossRef]
  18. Zhang, G.H.; Chou, K.C.; Mills, K. A Structurally Based Viscosity Model for Oxide Melts. Metall. Mater. Trans. B 2014, 45, 698–706. [Google Scholar] [CrossRef]
  19. Avramov, I.; CRüssel Keding, R. Effect of chemical composition on viscosity of oxide glasses. J. Non-Cryst. Solids 2003, 324, 29–35. [Google Scholar] [CrossRef]
  20. Wu, G.; Yazhenskikh, E.; Hack, K.; Wosch, E.; Müller, M. Viscosity model for oxide melts relevant to fuel slags. Part 1: Pure oxides and binary systems in the system SiO2–Al2O3–CaO–MgO–Na2O–K2O. Fuel Process. Technol. 2015, 137, 93–103. [Google Scholar] [CrossRef]
  21. Wu, G.; Seebold, S.; Yazhenskikh, E.; Hack, K.; Müller, M. Viscosity model for oxide melts relevant to fuel slags. Part 3: The iron oxide containing low order systems in the system SiO2–Al2O3–CaO–MgO–Na2O–K2O–FeO–Fe2O3. Fuel Process. Technol. 2018, 171, 339–349. [Google Scholar] [CrossRef]
  22. Mills, K.; Yuan, L.; Li, Z.; Zhang, G.; Chou, K. A review of the factors affecting the thermophysical properties of silicate slags. High Temp. Mater. Process. 2012, 31, 301–321. [Google Scholar] [CrossRef]
  23. Shen, X.; Chen, M.; Wang, N.; Wang, D. Viscosity Property and Melt Structure of CaO–MgO–SiO2–Al2O3–FeO Slag System. ISIJ Int. 2019, 59, 9–15. [Google Scholar] [CrossRef]
  24. Zhang, G.H.; Chou, K.C. Correlation Between Viscosity and Electrical Conductivity of Aluminosilicate Melts. Metall. Mater. Trans. B 2012, 43, 849–855. [Google Scholar] [CrossRef]
  25. Wang, W.; Dai, S.; Zhou, L.; Zhang, J.; Tian, W.; Xu, J. Viscosity and structure of MgO–SiO2-based slag melt with varying B2O3 content. Ceram. Int. 2020, 46, 3631–3636. [Google Scholar] [CrossRef]
  26. Winterhager, H.; Greiner, L.; Kammel, R. Investigation of the Density and the Electrical Conductivity of Metals of the Systems CaO-Al2O3-SiO2 and CaO-MgO-Al2O3-SiO2; Westdeutscher Verlag: Cologne, Germany, 1966. [Google Scholar]
  27. Talapaneni, T.; Yedla, N.; Sarkar, S.; Pal, S. Effect of Basicity. Al2O3 and MgO content on the softening and melting properties of the CaO-MgO-SiO2-Al2O3 high alumina quaternary slag system. Metall. Res. Technol. 2016, 113, 501. [Google Scholar] [CrossRef]
  28. Li, W.; Cao, X.; Jiang, T.; Yang, H.; Xue, X. Relation between Electrical Conductivity and Viscosity of CaO–SiO2–Al2O3–MgO System. ISIJ Int. 2016, 56, 205–209. [Google Scholar] [CrossRef]
Figure 1. Effect of NBO/T and temperature on slag viscosity.
Figure 1. Effect of NBO/T and temperature on slag viscosity.
Metals 13 00786 g001
Figure 2. Effect of NBO/T and temperature on slag conductivity.
Figure 2. Effect of NBO/T and temperature on slag conductivity.
Metals 13 00786 g002
Figure 3. (a) Logarithm relation between viscosity and conductivity and (b) the linear fitting slope of CaO-MgO-FeO-Al2O3-SiO2 systems at different NBO/T values.
Figure 3. (a) Logarithm relation between viscosity and conductivity and (b) the linear fitting slope of CaO-MgO-FeO-Al2O3-SiO2 systems at different NBO/T values.
Metals 13 00786 g003
Figure 4. (a) SEM of Z3 slag, (bd) EDS patterns at three points, and (e) the corresponding elemental composition.
Figure 4. (a) SEM of Z3 slag, (bd) EDS patterns at three points, and (e) the corresponding elemental composition.
Metals 13 00786 g004
Table 1. The composition of slag samples at different FeO/SiO2 burden ratios.
Table 1. The composition of slag samples at different FeO/SiO2 burden ratios.
SlagFeO/SiO2Xi (Mole Fraction)Cu
(wt.%)
CaOMgOAl2O3FeOSiO2
Z10.410.220.130.0500.0500.550.23
Z20.560.230.130.0500.0680.520.39
Z30.710.220.130.0550.0970.500.27
Z40.790.210.120.0660.160.450.13
Z50.870.190.110.0500.200.450.24
Z61.100.160.0880.0690.300.380.23
Z71.170.150.0850.0590.320.390.42
Table 2. Measured viscosity (Pa·s) and conductivity (S/m) of slags at different temperatures (°C).
Table 2. Measured viscosity (Pa·s) and conductivity (S/m) of slags at different temperatures (°C).
Slag1290 °C1320 °C1350 °C1380 °C1410 °C
κ η κ η κ η κ η κ η
Z111.963.2215.092.4519.481.8123.791.3728.671.15
Z27.943.4910.352.4112.991.7216.611.2919.761.02
Z36.724.038.632.6410.661.7313.041.0416.430.71
Z49.654.4912.433.0415.522.3518.871.4922.921.04
Z512.942.5117.021.7322.571.2228.480.9335.550.67
Z65.333.287.422.019.141.5612.541.2216.080.69
Z75.892.687.812.177.791.678.531.5711.291.28
Table 3. Calculated O/Si, NBO/Si, and NBO/T values of different slags (mol).
Table 3. Calculated O/Si, NBO/Si, and NBO/T values of different slags (mol).
SlagBasicityBasicityNBO/SiNBO/SiNBO/T
(Al2O3 Ignored)(Al2O3 as Base)(Al2O3 as Base)(Al2O3 Ignored)
Z10.741.012.031.481.09
Z20.811.102.201.631.21
Z30.901.232.461.801.29
Z41.091.533.062.181.45
Z51.101.432.862.201.62
Z61.421.963.922.841.83
Z71.411.873.742.831.94
Table 4. Logarithm relation between viscosity and conductivity and the related parameters.
Table 4. Logarithm relation between viscosity and conductivity and the related parameters.
Slag Equation R2 E η E κ     A η A κ  
Z1 l n η = 4.133 − 1.195 l n κ 0.997186.35156.23−13.6014.84
Z2 l n η = 4.071 − 1.366 l n κ 0.998227.18166.4116.22−8.89
Z3 l n η = 5.284 − 2.018 l n κ 0.994320.22158.94−23.2114.12
Z4 l n η = 5.372 − 1.688 l n κ 0.985261.97155.32−18.5514.17
Z5 l n η = 4.168 − 1.274 l n κ 0.997234.03184.18−17.1516.73
Z6 l n η = 3.371 − 1.313 l n κ 0.966261.38198.91−18.9617.01
Z7 l n η = 3.095 − 1.184 l n κ 0.892130.14110.36−9.0810.29
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

Zhang, L.; Jiang, K.; Xie, F.; Lu, D. Relation between Viscosity and Conductivity of CaO-MgO-FeO-Al2O3-SiO2 System for Copper Smelting Slags. Metals 2023, 13, 786. https://doi.org/10.3390/met13040786

AMA Style

Zhang L, Jiang K, Xie F, Lu D. Relation between Viscosity and Conductivity of CaO-MgO-FeO-Al2O3-SiO2 System for Copper Smelting Slags. Metals. 2023; 13(4):786. https://doi.org/10.3390/met13040786

Chicago/Turabian Style

Zhang, Lei, Kaixi Jiang, Feng Xie, and Diankun Lu. 2023. "Relation between Viscosity and Conductivity of CaO-MgO-FeO-Al2O3-SiO2 System for Copper Smelting Slags" Metals 13, no. 4: 786. https://doi.org/10.3390/met13040786

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