You are on page 1of 9

Corrosion Science xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Degradation mechanisms of alumina-chromia refractories for secondary


copper smelter linings

Liugang Chena,b, , Annelies Malflieta, Jef Vleugelsa, Bart Blanpaina, Muxing Guoa
a
KU Leuven, Department of Materials Engineering, Arenberg Kasteelpark 44, Leuven, 3001, Belgium
b
Zhengzhou University, Henan Key Laboratory of High Temperature Functional Ceramics, Zhengzhou, 450001, China

A R T I C LE I N FO A B S T R A C T

Keywords: To assess the viability of an alumina-chromia refractory alternative for secondary copper smelters, the de-
Ceramic gradation behavior of alumina-chromia refractories in a secondary copper smelting slag and a Cu-CuxO mixture
High temperature corrosion was studied via lab scale refractory finger testing. Microstructural characterization of the worn refractory
EPMA samples resulted in a comprehensive understanding of the corrosion and penetration resistance and a compar-
SEM
ison with the degradation of magnesia-chromite refractories. Alumina grains in the refractory are preferentially
Copper
corroded by the slag. The alumina-chromia refractory exhibits superior infiltration resistance to the slag and the
Cu-CuxO mixture compared to magnesia-chromite refractories.

1. Introduction loads, the chemical degradation of the refractory lining caused by liquid
slags is typically the most severe and can lead to a loss of structural
Long-lasting refractory linings for furnaces in copper-making in- integrity of the brick [1,4], making the lining more susceptible to
dustries, where high cost magnesia-chromite bricks are the most ex- thermal shock, mechanical stresses and erosion. Therefore, under-
tensively used material, are desirable based on economic and en- standing the occurring degradation mechanisms in the refractory in
vironmental considerations. An extended lifespan of the refractory contact with slag is indispensable in assessing the viability of the alu-
lining can reduce production losses during downtime of the furnace [1]. mina-chromia refractory for copper-making furnaces.
Because of the possible leaching of Cr(VI), used Cr2O3-containing re- In some studies the corrosion and infiltration resistance of alumina-
fractories are treated as potentially hazardous waste [1,2]. A prolonged chromia and magnesia-chromite refractories in different copper pro-
lining lifetime could alleviate the generation of spent magnesia-chro- duction processes have been compared [8,10–12]. The alumina-
mite bricks. Therefore, continuous and consistent endeavors are being chromia brick exhibits a superior performance against slag attack,
made to improve the lifetime of refractory linings of various copper- thereby having a longer service life. Laboratory scale experiments
making furnaces via lining concept optimization [1], slag engineering comparing the wear behavior of alumina-chromia and magnesia-chro-
[3,4] ceramic/refractory engineering [5] and/or process engineering mite bricks against fayalite slag have been carried out by Zou et al. [11]
[1]. Among these approaches, the most clear-cut way to enhance the and Gregurek et al. [12] via rotary kiln tests. The alumina-chromia
performance of furnace linings is through the use of new refractory refractories were less worn than the magnesia-chromite refractories
types that better withstand the combination of mechanical, thermal and under the same testing conditions. Microstructural characterization and
chemical stresses imposed on the lining [1]. This has led to the con- thermodynamic analyses indicate that the newly-formed FeCr2O4 and
sideration of replacing the magnesia-chromite brick type by alternative FeAl2O4 spinel phases from the slag/refractory interactions generate a
refractory types [6–12]. protective layer and thus hinder the further corrosion of the refractory
Alumina-chromia, which was also written as Al2O3-Cr2O3 [6–8], [11,13]. Similarly, Rigby [10] figured out that a reaction layer was
refractories are characterized by a high corrosion resistance to acidic generated from the interactions between the Cr2O3-rich matrix of alu-
slags and excellent mechanical properties [6,7]. Therefore, this specific mina-chromia refractories and infiltrating CuxO, thereby preventing the
refractory type, which is commonly used to withstand rigorous en- further ingress of slags and the spalling of the refractory. As a result, for
vironments like waste incinerators [6], glass tanks [7] and coal gasifiers instance, the skimming lip of an anode vessel mouth lined with 30 wt%
[7,9], has been considered as an alternative material for the lining of Cr2O3 alumina-chromia refractories has shown a longer lifetime (over
copper-making furnaces. Among the mechanical, thermal and chemical nine months) than magnesia-chromite refractories (three months) [10].


Corresponding author at: KU Leuven, Department of Materials Engineering, Arenberg Kasteelpark 44, Leuven, 3001, Belgium.
E-mail address: liugang.chen@kuleuven.be (L. Chen).

https://doi.org/10.1016/j.corsci.2018.03.039
Received 6 December 2017; Received in revised form 19 March 2018; Accepted 21 March 2018
0010-938X/ © 2018 Elsevier Ltd. All rights reserved.

Please cite this article as: Chen, L., Corrosion Science (2018), https://doi.org/10.1016/j.corsci.2018.03.039
L. Chen et al. Corrosion Science xxx (xxxx) xxx–xxx

The success of such initiatives consequently make alumina-chromia Table 2


refractories to be considered as replacement materials with a prolonged Composition and properties of the investigated alumina-chrome refractory.
lifetime in lining secondary copper smelters. Overall composition (in wt%) Bulk Apparent Cold
However, the degradation mechanisms and outstanding perfor- density porosity crushing
mance of alumina-chromia refractories in fayalitic slags may not be Al2O3 Cr2O3 MgO SiO2 Fe2O3 (g·cm−3) (%) strength
completely applicable for secondary copper smelting processes. (MPa)

Compared to the primary copper production, a broad stream of sec- 82.76 11.00 0.74 0.77 0.09 3.52 12.4 194
ondary raw materials containing different contaminations are fed in
secondary copper smelters [1,14]. This difference in feed will lead to a
difference in slag composition. ZnO, for instance, can be present in the Ar gas (passing the gas through silica gel and Mg turnings furnace
secondary copper smelting slag and changes both the physical and operating at 500 °C to remove traces of moisture and oxygen) was
chemical properties of the slag, thus affecting the corrosion resistance blown into the furnace at a flow rate of around 0.4 l/min to simulate the
of refractory linings. The presence of ZnO in the slag was found to in- reducing atmosphere used in the industrial process. A cylindrical re-
fluence the degradation behavior of magnesia-chromite refractories fractory finger sample was moved down and maintained at a distance of
[15–17]. The degradation behavior of alumina-chromia and magnesia- 10 mm above the Mo crucible for 10 min to preheat the sample. Then
chromite refractories in contact with calcium-ferrite type slags (57 wt% the refractory finger sample was immersed into the molten slag for 4 h
Fe2O3 and 19 wt% CaO) has also been tested by Gregurek et al. [12]. with a rotation speed of 12 rpm or 36 rpm. To investigate the infiltra-
Unlike in fayalitic slags, alumina–chromia bricks were more severely tion behavior of the Cu-CuxO mixture, a static refractory finger test was
corroded than magnesia-chromite refractories in calcium-ferrite slags. used. 400 g Cu-CuxO mixtures (1 wt% Al2O3, 4 wt% Cu2O and 95 wt%
Therefore, as a precondition for assessing the feasibility of the alumina- Cu) [17] were added in an alumina crucible thereby making sure that
chromia refractory alternative for secondary copper smelting, it is vital there was enough liquid to immerse the samples. The filled alumina
to identify and understand the occurring degradation mechanisms of crucible was put in the vertical tube furnace and heated to 1180 °C to
the refractory in slag systems that are relevant for secondary copper melt the mixtures. After melting the mixtures for 30 min, the finger
smelting. sample was inserted in the furnace and kept above the alumina crucible
In this work, the degradation mechanisms of alumina–chromia re- at a distance of 10 mm to preheat the sample for 10 min. Subsequently,
fractories in a Cu-CuxO mixture and a secondary copper smelting slag the refractory finger was immersed into the molten mixtures for 30 min.
were investigated. The influence of slag agitation, which was mimicked Ar gas, which generally contains around 2 ppm O2, was blown into the
via refractory finger rotation, and temperature on the corrosion beha- furnace tube with a flow rate of 0.6 l/min during the experiment. After
vior was also studied. The degradation mechanisms of alumina-chromia immersion for the required time, the refractory finger sample was re-
refractories are discussed and compared to those of the magnesia- moved from the molten slag or Cu-CuxO mixture and quenched in N2
chromite refractories reported in previous work [15–18] to evaluate the steam. The refractory samples were cut with a diamond saw at 1 cm
viability of using alumina-chromia refractory linings in secondary below the bath line for microstructural analysis. The slag was cooled
copper smelters. down to room temperature in the furnace, removed from the Mo cru-
cible and ground to a particle size smaller than 80 μm for compositional
2. Experimental analyses.

2.1. Materials
2.3. Sample analysis techniques
The composition of the secondary copper smelting slag, which is of
the fayalitic type, as supplied by a copper producer is listed in Table 1. The sliced refractory sections were embedded in a low viscosity
Minor elements, such as K, S, P, Cu, Mn, Pb, and Ni, are accounted as resin (Epofix) by vacuum impregnation, ground with silicon carbide
“others” in Table 1. The cylindrical refractory samples (dia- papers and polished with diamond paste. The polished specimens were
meter = about 15 mm; length = around 50 mm) were cut from a carbon coated for compositional and microstructural analysis.
commercially available alumina-chromia brick. The chemical compo- Compositional analyses were performed by full quantitative electron
sition and physical properties of the investigated alumina-chromia probe X-ray microanalysis coupled with wavelength dispersive spec-
brick are provided in Table 2. troscopy (EPMA-WDS, JEOL JXA-8530F). It was operated using an ac-
celeration voltage of 15 kV and a probe current of 15 nA. The oxygen
content was not measured directly. Instead the oxidation state of the
2.2. Experimental set-up and procedure
element was selected a priori. Periclase (MgO), hematite (Fe2O3),
willemite (Zn2SiO4), fluorspar (CaF2), chromium oxide (Cr2O3) and
The experimental set-up and the procedure for refractory finger
obsidian ((Na,K,Al,Fe) silicate glass) were used as standards.
tests are similar to the one used in our previous work in which the
Microstructural analyses were obtained using scanning electron mi-
degradation of magnesia-chromite refractory has been tested [15,17].
croscopy (SEM, FEI, XL-30). The ground slag before and after the cor-
The parameters for the refractory finger tests are listed in Table 3.
rosion tests was analysed with a wavelength-dispersive X-ray fluores-
Approximately 170 g of the as-delivered secondary copper smelting slag
cence spectrometer (WDXRF, Zetium PANalytical). Although both Fe2+
was placed in a Mo crucible (inside diameter = 40 mm;
and Fe3+ are possibly present in the slag and chromite spinel phases, all
height = 80 mm), heated in a vertical tube furnace (GERO HTRV
Fe is calculated in the form of “FeO” for the purpose of presentation.
100–250/18, with MoSi2 heating elements) to 1200 °C or 1180 °C, and
kept at that temperature for 50 min to ensure the molten state. Purified

Table 1
The chemical composition of the secondary copper smelting slag used in the experiments (by XRF).
A2O3 ZnO FeO SiO2 Na2O MgO CaO Cr2O3 PbO CuO Others

wt% 3.1 9.4 55.6 23.7 2.3 0.6 0.7 0.8 1.0 0.6 1.9
mol% 2.2 8.2 54.8 28.0 2.6 1.1 0.9 0.4 0.9 0.5 1.7

2
L. Chen et al. Corrosion Science xxx (xxxx) xxx–xxx

Table 3
Parameters of the refractory finger tests.
Exp. no. Rotation speed (rpm) Temperature (°C) Time (hours) pO2 (atm) Others

a
R12 12 1200 4 < 10−19 Sample size: D = 15 mm × L = 50 mm
R36 36 1180 Slag amount: ∼170 g
S0 0 1180 0.5 2 ppm O Mixture amount: 400 g

a
The oxygen partial pressure is measured in the outlet of the Ar gas (Rapidox 2100, Cambridge Sensotec Ltd.).

2.4. Methodology of thermodynamic calculation refractory matrix is rich in Cr2O3 and composed of alumina-chromia
solid solution grains, smaller alumina grains, zirconia and Al2O3·P2O5.
The solubility limit of Al2O3 was calculated according to the same The observed Al2O3·P2O5 originates from the phosphate binder for the
approach reported for MgO [16,18]. Thermodynamic calculations were brick preparation, and zirconia is commonly employed to improve the
performed with the FactSage software [19] (version 6.4). Equilibrium thermal shock resistance of this refractory type [6,8]. The CaO·6Al2O3
calculations were performed with the equilibrium module EQUILIB, and Al2O3·P2O5 impurity phases are dispersed in respectively the alu-
which is based on the minimization of the Gibbs free energy, and the FT mina grains and refractory matrix. Interconnected pores and cracks
oxide, FS stel and Fact PS databases. The following possible solution form a network in the as-delivered refractory brick.
phases are chosen in the calculations: (1) FT oxide-slag (molten oxide
phase); (2) FT oxide-monoxide (oxide solid solution); (3) FT oxide-
spinel (spinel solid solution); (4) FT oxide-olivine; (5) FT oxide-zincite; 3.2. Visual inspection and global slag analyses
(6) FT oxide-willemite, (7) FS stel-FCC (metallic iron) and (8) FACT PS
(gas phase). The dissolution of Al2O3 was modeled at 1200 °C by re- Fig. 2 shows the refractory fingers before and after corrosion testing.
lative addition of pure Al2O3 (x, in gram) into the slags ((100-x), in Before the rotating test, gray-appearing alumina grains with a max-
gram). ZnO, FeO, SiO2, Al2O3 and CaO are the slag components for the imum size of 5 mm and dark gray-appearing refractory matrix are ob-
calculation. FeO/SiO2, Al2O3/SiO2 and CaO/SiO2 mass ratios in the slag served on the surface of the refractory fingers (Fig. 2a and c). After
were kept at 1.4, 0.18 and 0.03, respectively. ZnO content in the slag testing for 4 h with a rotating speed of 12 rpm (Fig. 2b) or 36 rpm
varied from 0 to 20 wt%. The maximum Al2O3 dissolution amount into (Fig. 2d), large cavities appear at the locations of the alumina grains
the ZFS slag is determined when a solid phase precipitates from the although the surface is covered with a layer of frozen slag. These ob-
liquid slag. servations suggest that the alumina grains in the alumina-chromia re-
fractory are preferentially attacked by the secondary copper smelting
slag whereas the Cr2O3-rich refractory matrix is little corroded, similar
3. Result and discussion as direct-bonded magnesia-chromite refractories [15–17].
The preferential corrosion of alumina grains is also implied by the
3.1. As-delivered refractories combination of increased Al2O3 level with little increase in the Cr2O3
level in the tested slag (Table 5 and Fig. 3). The high solubility limit of
Fig. 1 shows backscattered electron (BSE) images of the as-delivered Al2O3 in the slag provides the thermodynamic driving force for the
alumina-chromia refractory. The chemical composition of the phases in alumina dissolution. The results also indicate that the decrease of op-
the refractory are presented in Table 4. The investigated alumina- erating temperature could hinder the corrosion of alumina grains. As
chromia refractory is a direct-bonded mixture of alumina (Al2O3, ‘Am’ shown in Table 5 and Fig. 3, the Al2O3 content increase in the slag in
in Fig. 1), alumina-chromia solid solution (Al2O3-Cr2O3, ‘AC’ in Fig. 1), test R36, where a stronger slag agitation and a lower testing tempera-
zirconia (ZrO2, ‘Z’ in Fig. 1), and CaO·6Al2O3 (CA6, ‘CA6’ in Fig. 1) and ture were employed, is 5.0 wt%, whereas that in the R12 test is 5.6 wt%.
Al2O3·P2O5 (‘AP’ in Fig. 1b) impurity phases. The alumina grains appear To compare the corrosion resistance of alumina-chromia and mag-
dark gray in BSE mode, with a maximum grain size of 3 mm (Fig. 1a) nesia-chromite refractories, the increase of Al2O3 and MgO in the tested
and a chromia-rich rim, directly binding with the refractory matrix slag and the predicted solubility limits of Al2O3 and MgO under the
(Fig. 1b). Between the alumina grains and the refractory matrix, mi- same experiment/calculation conditions are shown in Fig. 4. The so-
crocracks can be seen (Fig. 1a). Inside the alumina grains, micro-pores lubility limit of Al2O3 was calculated according to the same approach
and cracks, zirconia grains and CaO·6Al2O3 phases can be observed. The reported for MgO [16,18]. The MgO content increases with 0.8 wt%

Fig. 1. Backscattered electron (BSE) images of the as-delivered alumina-chromia refractory samples, (b) enlarged view of the selected area in (a); Am = alumina,
AC = alumina-chromia solid solution, Z = zirconia, CA6 = CaO·6Al2O3, AP = Al2O3·P2O5; P = pore.

3
L. Chen et al. Corrosion Science xxx (xxxx) xxx–xxx

Table 4
Chemical compositions of the phases in the as-delivered alumina-chrome refractory, as determined by EPMA-WDS.
Phases Chemical composition (wt%)

Al2O3 Cr2O3 ZrO2 CaO MgO SrO Ce2O3 P2O5

Alumina 93.6 ± 2.9 5.4 ± 2.7 – – – – – –


Alumina-chromia solid solution 47.8 ± 24.3 49.7 ± 23.2 – – – – – –
Zirconia – – 93.8 ± 1.0 – – – – –
CaO·6Al2O3 88.2 ± 0.6 – 5.2 ± 0.5 0.9 ± 0.2 1.6 ± 0.1 1.8 ± 0.7 –
Al2O3·P2O5 39.5 ± 1.3 – – – – – – 55.6 ± 2.0

Fig. 3. Comparison of the Al2O3 content in the secondary copper smelting slag
Fig. 2. Appearance of the refractory finger samples in R12 (a, b) and R36 (c, d)
before and after refractory finger tests.
tests, (a) and (c) before testing, and (b) and (d) after testing, showing the
corrosion of alumina grains.
lead to a comparable corrosion resistance of alumina-chromia and
after testing at 1200 °C with a rotation speed of 36 rpm (Fig. 4a) [16]. In magnesia-chromite refractories in the secondary copper smelting slag
comparison, the Al2O3 content increases with 2.5 wt % and 1.9 wt% in having 20 wt% ZnO. However, this merit needs to be further in-
the tests R12 (12 rpm at 1200 °C) and R36 (36 rpm at 1180 °C), re- vestigated.
spectively, for the same corrosion duration, although the slag before In summary, alumina grains in the alumina-chromia refractory are
testing has a higher Al2O3 level. The extent of dissolution of alumina preferentially corroded by the secondary copper smelting slag like
from alumina-chromia refractories is therefore more pronounced than periclase in magnesia-chromite refractories, and the extent of dissolu-
that of periclase from magnesia-chromite refractories in a slag with tion of alumina grains is more pronounced than that of periclase in
∼9 wt% ZnO. This would imply an inferior slag corrosion resistance of secondary copper smelting slag with 9 wt% ZnO, but would be com-
alumina-chromia bricks compared to magnesia-chromite refractories. parable in a slag with 20 wt% ZnO.
The different extents of dissolution of alumina and periclase should
be ascribed to the different solubility limits of Al2O3 and MgO. In the
slag containing 10 wt% ZnO the solubility limit of Al2O3 (4.4 wt%) is 3.3. Worn refractory microstructure analyses
higher than that of MgO (2.1 wt%) (Fig. 4b), thereby providing a
greater thermodynamic driving force for the Al2O3 dissolution. How- 3.3.1. Microstructure observations
ever, in the secondary copper smelting process, the ZnO level in the slag The worn microstructures from the tested refractory fingers are il-
can vary depending on the feed materials and/or operation conditions lustrated in Figs. 5–7. The chemical composition of the phases in the
[16], leading to different solubility limits of MgO and Al2O3 in the slags. worn refractory sample from the test R12 are presented in Table 6. The
The solubility limit of Al2O3 continuously decreases with increasing refractory surfaces are covered with frozen slag layers of 200 μm and
ZnO content (Fig. 4b), suggesting that the increase of ZnO concentra- 500 μm for respectively test R12 and R36 test (Fig. 5). The thicker slag
tion can suppress the Al2O3 dissolution. Once the ZnO content in the layer in test R36 is indicative for a higher slag viscosity, which is ex-
slag is raised to 20 wt%, the solubility limits of Al2O3 and MgO are pected from the lower testing temperature in R36. The concave shape of
comparable and ∼2.0 wt%. As a result, the extent of dissolution of the interface of the alumina grains with the slag indicates that the
alumina from alumina-chromia refractories is expected to be of the alumina grains were corroded by the slag. As a result, the Cr2O3-rich
same level as that of periclase in magnesia-chromite bricks. This will refractory matrix protrudes into the slag (Fig. 5b). These observations
further confirm that the alumina grains in the refractory are preferably

Table 5
Chemical compositions of the slags after rotating refractory finger tests as measured with XRF.
Exp. no. Chemical composition

Al2O3 ZnO FeO SiO2 Na2O MgO CaO Cr2O3 PbO CuO Others

R12 wt% 5.6 7.6 50.6 28.2 2.5 0.7 1.0 0.1 0.8 0.4 2.5
mol% 3.8 6.5 49.3 32.9 2.8 1.2 1.3 0.0 0.6 0.3 1.3
R36 wt% 5.0 8.6 51.3 26.6 2.5 0.6 0.9 0.7 0.9 0.4 2.5
mol% 3.5 7.5 50.3 31.3 2.8 1.1 1.1 0.3 0.6 0.4 1.1

4
L. Chen et al. Corrosion Science xxx (xxxx) xxx–xxx

Fig. 4. Comparison of (a) the increase in Al2O3 and MgO [16] concentration in secondary copper smelting slags and (b) the predicted solubility limits of Al2O3 and
MgO synthetic fayalite slags with a fixed FeO/SiO2 ratio of about 1.4.

corroded by the secondary copper smelting slag, whereas the Cr2O3-rich transformation of CaO·6Al2O3 phase into a liquid phase. As a result of
matrix is attacked less. At the refractory/slag interface (Figs. 5 and 6), a the CaO·6Al2O3 corrosion by the infiltrating slag, new paths are gen-
continuous, dense gray-appearing layer (‘H’ in Figs. 5 and 6) with a erated for the further slag infiltration.
thickness of 20 μm covers the fresh surface of the alumina grains. A number of large cracks with a maximum width of 15 μm are ob-
Chemical analyses by EPMA-WDS on this gray layer indicate that this served between the larger alumina grains and the refractory matrix in
would be a hercynite-based spinel consisting of 30.2 mol% FeO, the densified refractory zone (Figs. 5 and 7). These cracks are believed
20.5 mol% ZnO, 2.6 mol% MgO, 44.9 mol% Al2O3 and 1.2 mol% Cr2O3 to be generated during quenching or cutting of the refractory samples,
(Table 6). because these cracks are not infiltrated. Between the newly-formed
The infiltration behavior of secondary copper smelting slag in the spinel and alumina grains/alumina-chromia solid solution, no obvious
alumina-chromia refractory is shown in Fig. 7. A densified refractory cracks can be found (Figs. 5–7).
matrix is observed over a distance of 4.5 mm from the refractory surface
towards the center after 4 h testing. An elemental mapping of the
marked area in the densified zone indicated in Fig. 7 is shown in Fig. 8. 3.3.2. Spinel phase formation
Since Si, Fe and Ca are not detectable in the matrix of the as-delivered As discussed in Section 3.3.1, only spinel phases were generated
refractory (Table 4), the region containing Fe, Si and Ca is the in- from the slag/refractory interactions. The spinel phases are observed at
filtrated zone. Fig. 8 illustrates that the infiltrating slag is mainly found the refractory surface and in the densified refractory zone next to the
at the grain boundaries, and in the cracks and/or open pores in the infiltrating slag (Figs. 5–7).
refractory. Spinel phases, identified by the region composed of Al, Cr The composition analyses of the spinel phases at the refractory
and Fe in Fig. 8, are observed next to the infiltrating slag in the den- surface (Table 6) indicate that this is a (Zn,Fe,Mg)(Al,Cr)2O4 hercynite-
sified refractory zone. based spinel. It is believed that such hercynite-based spinel precipitates
Fig. 9 presents a detailed microstructure of the CaO·6Al2O3 phase in at the refractory/slag interface once the slag is locally saturated with
the bulk refractory sample from the R12 test, showing the corrosion Al2O3 according to the following reaction:
behavior of CaO·6Al2O3. Next to the CaO·6Al2O3, a Na, K, Ca, Al, Si and
FeOslag + ZnOslag + MgOslag + Al2O3,slag + Cr2O3,slag/refracory →
S containing phase is observed (point 001, Table 7), indicating the
(Zn,Fe,Mg)(Al,Cr)2O4,refractory (1)
diffusion of Na2O, K2O and SiO2 into the CaO·6Al2O3 phase. With the
continuous diffusion of Na2O, K2O and SiO2 from the slag, the melting This newly-formed continuous and dense hercynite-based spinel
point of CaO·6Al2O3 phase will decrease, thereby leading to the layer can act as a protective layer between the slag and the alumina

Fig. 5. Back scattering electron (BSE) images of the worn samples from R12 (a) and R36 (b) tests, showing the refractory surface zone: Am = alumina; AC = alumina-
chromia solid solution; Cra = crack; Z = zirconia; H = (Zn,Fe,Mg)(Al,Cr)2O4 hercynite-based spinel.

5
L. Chen et al. Corrosion Science xxx (xxxx) xxx–xxx

Fig. 6. BSE images of the worn samples from R12 (a and b) and R36 (c) tests, showing the slag/refractory interactions at the refractory surface: Am = alumina;
AC = alumina-chromia solid solution; P = pore; Z = zirconia; IS = infiltrating liquid slag; Fay = (Fex,Zn1−x)2SiO4 fayalite; M = (Fe,Zn)(Fe, Al)2O4 magnetite-based
spinel; H = (Zn,Fe,Mg)(Al,Cr)2O4 hercynite-based spinel; Sp = (Fe,Zn,Mg)(Cr,Al)2O4 chromia-rich spinel.

grains. The infiltration of the slag into the alumina grains through the forming this Cr2O3-rich spinel phase can be written as:
grain boundaries can thus be physically inhibited. Like olivine on
FeOslag + ZnOslag + MgOslag + Al2O3,slag,refractory + Cr2O3,refractory →
periclase grains [15,16], the newly-formed spinel layer on alumina
(Fe,Mg,Zn)(Cr,Al)2O4,refractory (2)
grains would also protect alumina from the direct attack of the slag and
slow the alumina dissolution into the slag. Studies on the formation mechanisms of spinel layers in the FeO-
To identify the formation mechanism of the spinel next to the in- Cr2O3[20], MgO-Cr2O3 [20] and MgO-Al2O3 [21] etc. systems show
filtrating slag in the densified refractory zone, the spinel grains at a that in all systems the spinel phases form by the counter-diffusion of
distance of 2 mm from the refractory surface were selected for mi- cations, i.e. the outward diffusion of Al3+ and Cr3+ from the chromite
croprobe analysis. As shown in Table 6, this spinel contains a higher to the slag and the inward diffusion of Fe2+, Zn2+ and Mg2+ from the
level of Cr2O3 (22.5 mol%) compared to the hercynite-based spinel slag to the chromite, through the newly formed solid spinel layer.
(Cr2O3, 1.2 mol%) at the refractory surface. Because of the low solu- Therefore, this Cr2O3-rich spinel phase grew on both the Al2O3-Cr2O3/
bility of Cr2O3 in the secondary copper smelting slag [15], the high spinel and spinel/slag interfaces, due to the supply of fresh Al2O3 and
Cr2O3 content indicates that these spinel phases are most likely formed FeO from the bath through the liquid channels. The growth of this
by removing FeO, ZnO and MgO from the infiltrating slag. The reaction

Fig. 7. Overview of the worn refractory sample from the R12 test, showing the liquid slag infiltration in the refractory: Am = alumina; AC = alumina-chromia solid
solution; CA6 = CaO·6Al2O3; P = pore; Cra = crack; Z = zirconia; IS = infiltrating liquid slag; H = (Zn,Fe,Mg)(Al,Cr)2O4 hercynite-based based spinel;
Sp = (Fe,Zn,Mg)(Cr,Al)2O4 chromia-rich spinel.

6
L. Chen et al. Corrosion Science xxx (xxxx) xxx–xxx

Table 6
Chemical compositions of the phases in the alumina-chrome refractory from Test R12, as determined by EPMA-WDS (in mol%).
Phase Position Chemical composition

FeO ZnO Al2O3 Cr2O3 MgO SiO2 CaO Na2O P2O5 Cu2O

Spinel Surface 30.2 ± 0.6 20.5 ± 0.3 44.9 ± 0.9 1.2 ± 0.6 2.6 ± 0.2 – – – – –
2 mm from the surface 40.0 ± 1.0 5.3 ± 0.3 28.3 ± 6.3 22.5 ± 5.6 2.8 ± 0.4 – – – – –
Slag Surface 41.6 ± 2.0 4.7 ± 0.4 5.2 ± 0.2 – 2.0 ± 0.3 44.2 ± 1.8 1.4 ± 0.1 1.1 ± 0.1 – –
2 mm from the surface 16.5 ± 0.6 0.2 ± 0.1 16.3 ± 0.4 – 0.5 ± 0.1 56.5 ± 2.5 3.4 ± 0.5 2.6 ± 0.1 2.5 ± 1.0 –

newly-formed Cr2O3-rich spinel fills the open pores and cracks in the Al2O3·P2O5 phase from the refractory. Because of the depletion in net-
refractory, thereby densifying the refractory. work breaking oxides (FeO, ZnO and MgO) and the enrichment of
The microstructural observations could be summarized as follows: network forming oxides (SiO2 and P2O5), the viscosity of the infiltrating
(1) only spinel phases are generated from the slag/refractory interfacial slag increases. Additionally, as discussed in Section 3.3.2, the newly-
reactions; (2) slag infiltrated into the refractory along the open-pore formed Cr2O3-rich spinel could fill the pores and cracks in the refractory
network and the grain boundaries. Cr2O3-rich spinel was generated next matrix and thus block the channels for further slag infiltration into the
to the infiltrating slag. The newly-formed spinel continuously grew with refractory.
the supply of fresh Al2O3 and FeO from the infiltrating slag and thus Fig. 10 shows an overview microstructure of the alumina-chromia
filled the open pores and cracks in the refractory, leading to the den- refractory finger sample after testing for 30 min in molten Cu-CuxO.
sification of the refractory; (3) the corrosion of CaO·6Al2O3 phases oc- Limited Cu-CuxO penetrated into the alumina-chromia refractory. In-
curs in the slag infiltrated refractory part, which provides paths for the filtrated Cu-CuxO mixtures are mainly observed at the grain boundaries
further ingress of the slag. and/or at the interconnected open-pore network up to a distance of
400 μm from the refractory surface. Deeper in the refractory, the Cu-
3.4. Slag and Cu-CuxO infiltration behavior CuxO mixture is dispersed as micrometer sized particles. These ob-
servations indicate that the alumina-chromia refractory has a superior
The chemical composition of the infiltrating slag at a distance of infiltration resistance to molten Cu-CuxO.
2 mm from the refractory surface is shown in Table 6. Compared with Compared with magnesia-chromite refractories [15–17], alumina-
the frozen slag at the refractory surface, the infiltrating slag has a lower chromia refractories have a better infiltration resistance against both
FeO, ZnO and MgO content, but higher levels of SiO2, Na2O, CaO, Al2O3 the secondary copper smelting slag and the Cu-CuxO mixture. It is ob-
and P2O5. The depletion of FeO, ZnO and MgO is mainly due to the served in previous work [15–17] that both the slag and Cu-CuxO mix-
consumption of these oxides in the refractory/slag interactions forming ture infiltrated up to the center (7.5 mm) of magnesia-chromite re-
spinel phases, whereas the slag constituents which react little with the fractory finger samples after testing for 4 h and 30 min, respectively. In
refractory components, such SiO2 and Na2O, are enriched. The in- contract, the infiltration depths for the slag and Cu-CuxO mixture in the
creased Al2O3 and CaO content in the infiltrating slag is caused by the alumina-chromia refractory fingers are 4.5 mm and 400 μm after testing
combined effect of the depletion of slag constituents and the dissolution for 4 h and 30 min, respectively. The superior infiltration resistance to
of refractory components, for example CaO·6Al2O3. The presence of the slag and Cu-CuxO mixture of alumina-chromia refractory can thus
P2O5 in the infiltrating slag is mainly caused by the dissolution of hinder the corrosion in the infiltrated refractory part and severe

Fig. 8. (a) EPMA-WDS elemental mapping and (b) a combination of the elemental mappings of the selected zone ‘x’ in Fig. 7, showing the presence of infiltrating slag
in the refractory matrix and formation of (Fe,Zn,Mg)(Cr,Al)2O4 chromia-rich spinel: P = pore. The green area in (b) is because of the depth variation of the pore.

7
L. Chen et al. Corrosion Science xxx (xxxx) xxx–xxx

Fig. 9. (a) BSE image of CaO·6Al2O3 phase in an alumina grain, and (b) energy dispersive spectroscopy (EDS) spot measurements of selected point in the figure,
showing the corrosion of CA6 by the infiltrating slag: Am = alumina; CA6 = CaO·6Al2O3; IS = infiltrating slag.

Table 7 Additionally, the alumina-chromia refractory has a superior infiltration


The chemical composition of the selected point in Fig. 9a, as determined EDS resistance against both the slag and Cu-CuxO mixture compared to
(atom%). magnesia-chromite refractories. These results can lead to a conclusion
O Ce Na Mg Al Si S K Ca Ti that the alumina-chromia refractory alterative may have a longer life-
time compared to magnesia-chromite refractories for secondary copper
52.61 0.03 0.15 1.53 15.67 17.47 1.23 1.72 9.55 0.04 smelters operating with ZnO-rich fayalitic slags. However, since the
wear of refractory linings is caused by a combination of chemical,
thermal and mechanical loads, industrial trials are deemed necessary to
spalling [10]. The lower apparent porosity of 13.6% for the alumina-
validate our results.
chromia refractory compared with 15–16.5% for the magnesia-chro-
mite refractories [15,17], is considered to be the cause for this. Besides,
4. Conclusion
the differences in alumina-alumina and periclase-periclase interfaces,
and alumina-chromia and periclase-chromite interfaces, can affect the
Refractory finger corrosion tests were performed to investigate the
wettability, and thereby the infiltration, of liquid slag and the Cu-CuxO
degradation mechanism of alumina-chromia refractories in the sec-
mixture. For instance, the Cu–CuxO preferentially wets the Fe-enriched
ondary copper smelting slag with a ZnO content of ∼9 wt% as well as
surface layer of chromite grains in fused-grain magnesia-chromite re-
Cu-CuxO mixture. From the macroscopical and microstructural ob-
fractories instead of the chromite grains in direct-bonded refractories
servations, the following conclusions are drawn:
[17]. In the alumina-chromia refractory, the Cr2O3-rich matrix does not
contain Fe2O3 (Table 4), and thus would be less wetted by the Cu–CuxO
(1) It is preferentially that the alumina in the alumina-chromia re-
mixture. However, supporting literature data on the wetting behavior
fractory is corroded by the slag. The Cr2O3-rich phase in the re-
and contact angles have not been found. Additionally, the blockage
fractory is little attacked. The high solubility limit of Al2O3 in the
effect of the newly-formed spinel phase would be another reason to
slag is the thermodynamic driving force for the faster dissolution of
prevent the further ingress of slag and Cu-CuxO mixture.
the alumina grains.
(2) The slag infiltrated into the alumina-chromia refractory through the
3.5. Industrial relevance open-pore network and the grain boundaries, and by the corrosion
of the CaO·6Al2O3 impurity phases. The infiltration of Cu-CuxO
Due to the higher solubility limit of Al2O3, the extent of dissolution mixture into the refractory via the open-pore network and the grain
of alumina grains is pronounced compared to that of periclase in the boundaries was limited.
slag having 9 wt% ZnO. When the ZnO level reaches to 20 wt%, the (3) (Fe,Mg,Zn)(Al,Cr)2O4 spinel phases are generated from the slag/
solubility limits of Al2O3 and MgO are comparable. We expect that this refractory interactions at the refractory surface and in the in-
will lead to a similar extent of dissolution of alumina and periclase, and filtrated refractory. The spinel was formed next to the infiltrating
thus a comparable corrosion resistance of alumina-chromia and mag- slag and filled the open pores and cracks in the refractory, leading
nesia-chromite refractories in the slag containing 20 wt % ZnO. to a densified refractory zone.

Fig. 10. Light optical microscopy images of the infiltration of Cu-CuxO mixture in the alumina-chromia refractory for 30 min at 1180 °C; Am = alumina, Cu = Cu-
CuxO mixture.

8
L. Chen et al. Corrosion Science xxx (xxxx) xxx–xxx

Acknowledgements [9] J.P. Bennett, K. Kwong, Refractory liner materials used in slagging gasifiers,
Refract. Appl. New 9 (2004) 20–25.
[10] A.J. Rigby, The use of alumina-chrome refractories in the high wear areas of anode
The authors are grateful to the support from the Hercules refining vessels, Proceeding Copper, Santiago, Chile, 2013, pp. 497–505.
Foundation (project ZW09-09) in the use of the FEG-EPMA system and [11] M. Zou, M. Jiang, Y. Qian, X. Shi, Simulation study on corrosion resistance of Al2O3-
the close co-operation with engineers from Metallo-Chimique. L. Chen Cr2O3 and MgO-Cr2O3 bricks to ISA furnace slag, Refract. Naihuo Cailiao 41 (2007)
180–182 (in Chinese).
would like to thank the Research Fund of KU Leuven for his post-doc- [12] D. Gregurek, A. Ressler, V. Reiter, A. Franzkowiak, A. Spanring, T. Prietl, Refractory
toral fellowship (PDM/16/140). wear mechanisms in the nonferrous metal industry: testing and modeling results,
JOM 65 (2013) 1622–1630.
[13] Z. Chen, On refractory material and formation of protection layer for Ausmelt
References copper smelting furnace, China Nonferrous Metall. 2 (2005) 27–30.
[14] J. Wood, S. Creedy, R. Matusewicz, M. Reuter, Secondary copper processing using
[1] A. Malfliet, S. Lotfian, L. Scheunis, V. Petkov, L. Pandelaers, P.T. Jones, B. Blanpain, Outotec Ausmelt TSL technology, Proceeding of Met Plant, (2010), pp. 460–467.
Degradation mechanisms and use of refractory linings in copper production pro- [15] L. Chen, M. Guo, H. Shi, L. Scheunis, P.T. Jones, B. Blanpain, A. Malfliet, The in-
cesses: a critical review, J. Eur. Ceram. Soc. 34 (2014) 849–876. fluence of ZnO in fayalite slag on the degradation of magnesia-chromite refractories
[2] V. Petkov, P.T. Jones, E. Boydens, B. Blanpain, P. Wollants, Chemical corrosion during secondary Cu smelting, J. Eur. Ceram. Soc. 35 (2015) 2641–2650.
mechanisms of magnesia-chromite and chrome-free refractory bricks by copper [16] L. Chen, M. Guo, H. Shi, S. Huang, P.T. Jones, B. Blanpain, A. Malfliet, Effect of ZnO
metal and anode slag, J. Eur. Ceram. Soc. 27 (2007) 2433–2444. level in secondary copper smelting slags on slag/magnesia-chromite refractory in-
[3] K.H.T. Matsui, T. Ikemoto, K. Sawano, A consideration on corrosion of magnesia teractions, J. Eur. Ceram. Soc. 36 (2016) 1821–1828.
refractory brick by silicate slag, J. Tech. Assoc. Refract. 22 (2002) 302–309. [17] L. Chen, S. Li, P.T. Jones, M. Guo, B. Blanpain, A. Malfliet, Identification of mag-
[4] L. Scheunis, A.F. Mehrjardi, M. Campforts, P.T. Jones, B. Blanpain, E. Jak, The nesia-chromite refractory degradation mechanisms of secondary copper smelter
effect of phase formation during use on the chemical corrosion of magnesia – linings, J. Eur. Ceram. Soc. 36 (2016) 2119–2132.
chromite refractories in contact with a non-ferrous PbO–SiO2 based slag, J. Eur. [18] L. Chen, M. Guo, H. Shi, P.T. Jones, B. Blanpain, A. Malfliet, Influence of FeO/SiO2
Ceram. Soc. 34 (2014) 1599–1610. and CaO/SiO2ratios in iron-saturated ZnO-rich fayalite slags on the corrosion of
[5] Y.Y. Deng, H.Z. Wang, H.Z. Zhao, Influence of chrome-bearing sols vacuum im- MgO, J. Am. Ceram. Soc. 99 (2016) 3754–3760.
pregnation on the properties of magnesia-chrome refractory, Ceram. Int. 34 (2008) [19] C.W. Bale, E. Bélisle, P. Chartrand, S.A. Decterov, G. Eriksson, A.E. Gheribi, K. Hack,
573–580. I.H. Jung, Y.B. Kang, J. Melançon, A.D. Pelton, S. Petersen, C. Robelin, J. Sangster,
[6] D. Chen, A. Huang, H. Gu, M. Zhang, Z. Shao, Corrosion of Al2O3–Cr2O3 refractory P. Spencer, M.A. Van Ende, FactSage thermochemical software and databases,
lining for high-temperature solid waste incinerator, Ceram. Int. 41 (2015) 2010–2016, Calphad 54 (2016) 35–53.
14748–14753. [20] K. Nagata, R. Nishiwaki, Y. Nakamura, T. Maruyama, Kinetic mechanisms of the
[7] M. Nath, A. Ghosh, H.S. Tripathi, Hot corrosion behavior of Al2O3-Cr2O3refractory formations of MgCr2O4 and FeCr2O4 spinels from their metal oxides, Solid State
by molten glass at 1200 °C under static condition, Corros. Sci. 102 (2016) 153–160. Ionics 49 (1991) 161–166.
[8] S. Breyner, K. Santowski, T. Prietl, A. Spanring, A. Franzkowiak, Thermal shock [21] I. Ganesh, A review on magnesium aluminate (MgAl2O4) spinel: synthesis,proces-
resistant alumina-chromia products for the copper industry, Proceeding Copper, sing and applications, Int. J. Mater. Res. 58 (2016) 63–112.
Santiago, Chile, 2013, pp. 451–460.

You might also like