Skip to main content
Erschienen in:
Buchtitelbild

Open Access 2024 | OriginalPaper | Buchkapitel

7. Interface Reactions Between the Metal Melt and the Filter Surface Activated by a Spark Plasma Sintering Process

verfasst von : Anton Salomon, Mykhaylo Motylenko, Martin Thümmler, David Rafaja

Erschienen in: Multifunctional Ceramic Filter Systems for Metal Melt Filtration

Verlag: Springer International Publishing

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

The reactions between newly developed filter materials and metal melts containing various inclusions were analyzed under laboratory conditions. For the melt production, a Spark Plasma Sintering (SPS) apparatus was utilized. The SPS process provides very high heating rates, which emulate the thermo-shock during a real filtration process, and variable reaction times, which allows to simulate both, short and long filtration processes. The short-time filtration processes are relevant for die-casting, where one ton of a steel passes the filter in approx. 15 s, the long-term ones for continuous casting taking several hours. In the SPS process, the convection of the steel melt is suspended, which makes the interpretation of the interfacial reactions and reaction kinetics more straightforward and trustworthy. Furthermore, the SPS process allows speeding up the reaction diffusion kinetics by using tiny diffusion couples having the form of powder mixtures that contain the metal or alloy and the functional filter material under study. In such samples, the equilibrium state that is suitable for a direct comparison with the results of thermodynamic simulations can be achieved very quickly.

7.1 Introduction

The understanding of the interface reactions between metal melts containing inclusions and the filter ceramic surfaces is essential for the development of functional materials for active and reactive filtration of metallic melts. On the laboratory scale, the filtration processes and the deposition of inclusions in ceramic filters are typically analyzed using impingement trials [13] or casting simulators [4], before the filters are subjected to prototype casting on industrial scale [5]. However, the results of such experiments are sometimes hardly to interpret. The main reason is that these experiments activate several simultaneous processes, which involve heterogeneous reactions between the metallic melt and the surface of the functional ceramics, production of inclusions in the melt and their deposition on the surface of the functional ceramics, and finally the solid state reactions and diffusion processes between the products of the heterogeneous reactions and the functionalized filter material. Furthermore, these processes are superimposed by the melt flow and/or by the macroscopic convection. Although the traditional laboratory experiments reveal important information about the thermal shock behavior [1, 6], melt flow rate [5, 7] and filtration efficiency [5, 7, 8] of the filters, which is usually complemented by the amount and chemical composition of the deposited non-metallic inclusions, they cannot substitute an in-depth investigation of the fundamental mechanisms of the interface and bulk reactions, which is required for a targeted development of the metal melt filters.
This chapter illustrates how the Spark Plasma Sintering (SPS) can be employed as a generally applicable but more controlled method to produce interface and reaction layers between metal melts and filter ceramics for advanced analysis of filtration processes. The SPS technology, which was originally developed for powder compaction and sintering [9], offers a high variability of process parameters, controlled heating and cooling rates, and widely adjustable dwell temperatures, holding times and atmospheres. In this work, SPS was utilized to melt the 42CrMo4 steel, aluminum and AlSi7Mg alloy that were in a direct contact with selected refractories like alumina, mullite, silicon oxide, titanium oxide, carbon-bonded alumina and carbon-bonded magnesia.
In contrast to other techniques, which bring molten metals in contact with the functionalized metal melt filters, the metal melting in a SPS apparatus minimizes the macroscopic melt flow, which is usually responsible for damage or even for removal of the newly formed reaction products and layers from the filter surface. Furthermore, the SPS allows working with extremely variable sample geometry including planar metal-ceramic interfaces in combination with bulk filter materials up to the powder mixtures of metal and ceramics with a huge contact area between the counterparts, with short reaction diffusion paths and with the finite geometry of the diffusion couples. The samples with planar geometry were utilized to analyze the sequence and morphology of new phases formed at the metal-ceramic interface. The experiments done on powder mixtures assisted in the identification and description of phases with narrow homogeneity ranges, and produced equilibrium-state samples for comparison with the thermodynamic calculations. Thus, the SPS-based melting technique provides complementary results with respect to the classical casting experiments or impingement tests, and helps to elucidate the reaction steps, which are frequently inaccessible, when the established methods are used.

7.2 Adaptation of the Spark Plasma Sintering for Metal Melting

The Spark Plasma Sintering or Field Assisted Sintering Technology (SPS/FAST) is based on resistive Joule heating through the pulsed electric direct current (DC) that passes electrically conducting graphite tools and/or the sample, if the sample is electrically conducting as well [9]. The Joule heating by the graphite die or directly by the sample itself allows for high heating rates. The typical current used for SPS/ FAST is in the kA range. For sample having a diameter of about 2 cm, this current corresponds to the current density of approx. 1 kA/cm2. Although the voltage on the sample is relatively low (max. 10 V), the heating power is still about 30 kW. The sintering atmosphere is usually vacuum or some inert gas like nitrogen or argon [9]. As the SPS/FAST process was developed mainly for compacting and short time sintering of refractories, the standard SPS/FAST tools are not constructed for an extensive melting of one of the components. During the fast melting experiments, however, the metal has to be molten. Therefore, new sample environments must be developed for such experiments, because the molten metal would leak from the sample environment and react quickly with the carbon present in the graphite tools.
The first development is a corundum crucible, which is filled with the functionalized filter material and with the metal powder to be molten. The crucible itself is located in a tube made of high strength graphite, which serves as a heater (Fig. 7.1a). The outer diameter of the tube is about 100 mm, the thickness of the walls about 10 mm [10]. In this “high speed furnace”, the samples are heated indirectly by the heat irradiated from the heater and by the heat convection. Because of the indirect heating, the heating rate is below 200 K min−1, significantly faster than in a conventional furnace but slower than usual SPS. Still, this tool is useful for special experiments with large samples that should be kept at the dwell temperature for long holding times, for instance during the investigation of the metal infiltration into the real filter structures. A drawback of this tool is the expected inaccuracy of the temperature measurement by a pyrometer, which is focused to a graphite finger that is not in a direct contact with the sample under study.
The second sample environment is based on a “tool-in-tool” setup [11], in which the metal powder to be molten is placed within a crucible that is made from or coated by the functionalized filter ceramics. The ceramic parts were cut from suspension nozzles or produced by slip casting according to the procedures presented in references [1, 2, 7, 8, 1214]. The size of the ceramic crucible was usually 20 mm in diameter with an internal cavity of 5 to 10 mm. The whole inserts were located at the position, which corresponds to the sample position in a standard SPS experiment (Fig. 7.1b). Due to the small size of the inserts and the direct contact between the heater and the sample, very high heating rates up to 1500 K min−1 were achieved [3]. As this setup offered much higher heating rates and an easier handling than the previous one, it was used for the majority of the SPS experiments.

7.3 Methods of Structure and Microstructure Analysis

On the microscopic scale, the solidified products of the chemical reactions between the molten metals and the surface of the ceramic filter were analyzed using scanning electron microscopy with energy dispersive X-ray spectroscopy (SEM/EDX) and electron backscatter diffraction (SEM/EBSD). The elemental analysis using SEM/ EDX was complemented by the electron probe microanalysis with wavelength dispersive X-ray spectroscopy (EPMA/WDX). For SEM/EDX/EBSD, a high-resolution SEM LEO-1530 (Carl Zeiss AG, Germany) with field-emission cathode, an EDX detector (Bruker AXS) and a Nordlys II EBSD detector (HKL Technology) was used. The SEM imaging was performed using secondary electrons, back-scattered electrons or in a combined mode. The SEM/EDX/EBSD experiments were carried out at an acceleration voltage of 20 kV. The working distance for EBSD was 15 mm, the tilting angle of the sample 70° and the step size 0.3 μm. For identification of the Kikuchi patterns and for the evaluation of the measured data, the software package Channel 5 (HKL Technology) was used. For the EPMA/WDX measurements, an electron probe microanalyzer JXA-8230 SuperProbe (Jeol GmbH, Germany) with five-crystal spectrometers was used. The EPMA scans were performed with the step size of 0.5 µm.
The phase compositions of the solidified samples were analyzed using a Bragg–Brentano diffractometer URD 63 (Freiberger Praezisionsmechanik) that was equipped with a sealed X-ray tube with copper anode and with a curved graphite monochromator located in front of a scintillation detector. The X-ray diffraction (XRD) patterns patterns were collected between 2θ = 20° and 150° with the step size of ∆2θ = 0.04°, and with the counting time of 10 s per step. The phase composition of the samples was quantified by using the Rietveld method [15, 16] implemented in the computer program MAUD [17].
On the nanoscale, the samples were characterized using transmission electron microscopy (TEM), selected area electron diffraction (SAED) and energy dispersive X-ray spectroscopy (EDX/TEM). These analyses were done in a JEM 2200 FS transmission electron microscope (JEOL Ltd., Japan) at an acceleration voltage of 200 kV. The TEM samples were prepared by the focused ion beam method (FIB) with a Helios NanoLab 600i (FEI, USA) in form of thin slices.

7.3.1 Reactions Between Molten Steel and Corundum-Based Refractories with Different Carbon Contents

In analogy to the established entry nozzles [1821], the newly developed carbon-bonded alumina (Al2O3-C) filters [8, 13, 14] are expected to react with molten steels and to form new interface layers between the metallic melt and the filter surface. Already the formation of the interface layer should significantly contribute to the removal of unwanted oxygen or other contaminating elements [10]. The interface layer itself has to attract and to embed non-metallic inclusions. Generally, the chemical composition and the phase composition of the layer are considered as crucial factors influencing the agglomeration of non-metallic inclusions and their adherence to the filter surface [7, 22]. The carbon additions should improve the high-temperature mechanical properties of the alumina filters [2326], their thermal shock resistance (via higher thermal conductivity and low thermal expansion) [8, 25] and their resistance against crack initiation and propagation [26]. In this section, the formation of a secondary alumina layer at the interface between the Al2O3-C filters with different carbon contents and molten steel 42CrMo4 is described. The chemical composition of the steel is summarized in Table 7.1. The reaction experiments presented here were performed using the tool-in-tool setup with the respective carbon-containing alumina ceramic acting as the reaction vessel.
Table 7.1
Chemical composition of the utilized steel alloy 42CrMo4 as provided by the powder vender
Element
C
Cr
Mo
Si
Mn
Fe
Mass%
0.4
1.0
0.2
0.2
0.7
Balance

7.3.2 Time-Dependent Layer Growth and Reaction Scheme on the Microscopic Level

According to the thermodynamic model proposed by Zienert et al. [27], Al2O3 is partially decomposed by liquid iron. The decomposition of alumina is facilitated by the presence of carbon in the reaction zone, which also reacts with dissolved oxygen to CO and/or CO2 [27, 28]. The CO/CO2 gas leaves partially the reaction zone. The formation and evaporation of CO/CO2 decreases the carbon concentration in the reaction zone and decelerates the decomposition of Al2O3. Nevertheless, if sufficient amount of carbon is present in the filter material, alumina is permanently dissolved in the liquid iron [2729], which leads to an increase of the concentration of aluminum in the melt.
The microstructural consequences of a short-time reaction between the carbon binder and the dissolved oxygen stemming from the decomposed alumina are illustrated in Fig. 7.2. The contact region between the solidified steel and Al2O3(-C) becomes carbon deficient or even carbon-free. The thickness of the carbon-depleted zone does not exceed 40 – 50 µm and is almost independent of the dwell time. Liquid iron, which contributes substantially to the decomposition of alumina, can be observed in form of solidified droplets in the filter constituents containing simultaneously alumina and carbon. In the functional filter coating containing no carbon (Fig. 7.2a), the solidified droplets were observed only in the carbon-bonded alumina filter struts. In the functionalized filter coatings containing carbon (Fig. 7.2b and c), the solidified droplets were observed already in the functional coating. Increased concentrations of Al and O in the solidified steel melt were revealed by the SEM/EDS measurements [11, 29]. Figure 7.3a shows the grooves stemming from the escape of the CO/CO2 bubbles, which were formed during the reaction of free oxygen with the carbon binder. The escape of the CO/CO2 bubbles is the main reason for the decarburization of the C-bonded Al2O3 filter ceramic.
The results of these experiments confirmed that the local concentration of carbon is a crucial factor influencing the rate of the concurrent reactions, namely alumina dissolution and the formation of secondary corundum. Although these reactions occur in contact with liquid iron, the dissolution of Al2O3 is facilitated by the presence of carbon, which reacts with free oxygen, while the formation of secondary corundum is assisted by the absence of carbon. For this reason, secondary corundum forms preferentially at the filter surface, i.e., in the carbon-depleted zone, where a sufficiently high amount of dissolved aluminum and oxygen is present (Fig. 7.3b) [27, 29]. The formation of secondary corundum was observed mainly in samples that were kept at high temperature for a longer time (Fig. 7.4). Besides the dwell time, the rate of the CO/CO2 gas formation is another important factor controlling the secondary corundum formation, because it affects the local concentration of carbon in the filter ceramics.
Also the morphology of the Al2O3(-C) filters, their porosity, and the morphology of secondary corundum are additional important factors influencing both reaction processes [27, 29]. At a sufficiently high local concentration of carbon, the secondarily formed corundum can be reduced like the primary corundum in the filter, if it forms only small and separated crystallites that can be soaked by liquid iron containing carbon from the binder [27, 29]. On the other hand, compact and large grains of secondary corundum, which are in contact with carbon-depleted steel melt (Fig. 7.4a-c), stay stable. Hence, the formation of a dense, impenetrable layer of secondary corundum provides an efficient barrier for carbon diffusion from the Al2O3-C filter and iron penetration into the filter, and thus inhibits the decomposition of Al2O3 [29]. Still, the reaction layer serves as a docking site for non-metallic inclusions contained in the melt (Fig. 7.5).
The majority of the inclusions consisted of aluminum oxides. Other endogenous inclusions contained MnS (SG Fm\(\overline{3 }\)m) [11]. According to Sims and Dahle [30], the MnS precipitates can be classified as type II and III. They were found at the high-angle grain boundaries, free surface of the steel droplets and at the interfaces formed between the ceramics and the steel. In addition, MnS in form of a thin film covered partly the Al2O3 inclusions (Fig. 7.6). The formation of the MnS film was explained by the heterogeneous nucleation of the sulfide on oxide surfaces and by the supersaturation of Mn and S in the areas of final solidification [31]. Unfortunately, the described MnS formation on inclusion surfaces results in so-called duplex inclusions that are detrimental for the mechanical properties of any (cast) metallic part [32].

7.3.3 Interface Reactions Preceding the Secondary Corundum Formation

High efficiency of the filtration processes requires a good adhesion of the non-metallic inclusions and/or reaction products to the functionalized filter surface in order to inhibit their spalling and the contamination of the metallic melt. It is assumed that materials having the same chemical composition or similar crystal structures possess better adhesion than fully incompatible compounds. As the secondary corundum grows on the filter surface that contains primary corundum, a local epitaxial growth was considered. However, the SEM/EBSD measurements, which were carried out on the metal melt filter covered by a functional Al2O3-C coating with 4 mass% C, revealed that the secondary corundum grows in form of interconnected, almost single-crystalline platelets, but without any pronounced orientation relationship to the corundum substrate (Fig. 7.7). This result motivated further studies, which should elucidate the early stages of the chemical reactions at the interface between the functionalized filter surface and the metallic melt that precede the growth of secondary corundum. These studies were carried out on functional Al2O3 coatings containing no carbon, 4 mass% C or 30 mass% C that were treated for a short time (1 – 2 min) at 1600 °C in the SPS apparatus or in a steel-casting simulator. The chemical composition of the interface layers was analyzed using EDX in TEM. The phase composition of the interlayers was concluded from the SAED patterns.
The chemical analysis of the interlayers using EDX confirmed the presence of iron, aluminum and oxygen, which are involved in the carbothermic reaction, as well as the presence of the alloying elements from the steel, mainly silicon [13, 29]. The SAED analysis revealed that the interlayers are almost amorphous (Figs. 7.8 and 7.9). Still, the contrasts observed in the TEM micrographs indicated fluctuations in the chemical and possibly in the phase composition, including the presence of nanocrystalline phases. Detailed analysis of the SAED patterns (Fig. 7.10) disclosed that the nanocrystalline phases contain wuestite (FeO, SG Fm\(\overline{3 }\)m), spinel-like phases with the chemical composition \({{\text{A}}}^{2+}{{\text{B}}}_{2}^{3+}{{\text{O}}}_{4}^{2-}\) (A and B being Fe, Al, Si and/or Mg contaminant) and the space group Fd\(\overline{3 }\)m and garnet-like structures, most probably Fe3Al2(SiO4)3 (SG Ia\(\overline{3 }\)d) [29]. After longer reaction times of a few minutes, the nanocrystalline oxide phases are replaced by corundum. A possible transient phase is metastable γ-Al2O3, which possesses a spinel-like crystal structure (SG Fd\(\overline{3 }\)m) containing highly mobile structural vacancies [33]. The structural vacancies facilitate the necessary fast exchange of metallic (cationic) species to remove iron and/or silicon from the spinel-like structure of \({{\text{A}}}^{2+}{{\text{B}}}_{2}^{3+}{{\text{O}}}_{4}^{2-}\).
When the seeds of secondary corundum in the interlayer are in a direct contact with the primary corundum from the functionalized coating, the secondary corundum grows epitaxially on the primary corundum (regions 1 and 2 in Fig. 7.11). Platelets of the secondary corundum, which grow from aluminum and oxygen dissolved in the steel melt on the amorphous interlayer, do not develop any pronounced orientation relationship to the filter surface (region 3 in Fig. 7.11), because the amorphous interlayer inhibits the epitaxial growth. Whereas the amorphous layers form on the surface of all Al2O3-C filters independently of their carbon content, the secondary corundum possesses different morphologies, which depend on the local carbon supply and thus on growth kinetics [29], as discussed in the previous section.

7.4 Reactions Between Molten Steel and Carbon-Bonded Corundum Coated with Carbon-Bonded Magnesia

Carbon-bonded magnesia (MgO-C) combines low thermal expansion and high thermal conductivity of the graphite binder with a low wettability of the composite against slags and metal melts [34]. In analogy to Al2O3-C (cf. previous sections), MgO in MgO-C is decomposed and CO/CO2 gas is formed, when the refractory is brought in contact with liquid steel [14, 27, 3437]. Dissolved magnesium and oxygen form secondary MgO, which deposits as a dense, thin layer at the steel/refractory interface [36, 38, 39], analogously to the formation of secondary Al2O3. After a longer reaction time, MgAl2O4 whisker-like fibers (SG Fd\(\overline{3 }\)m) formed at the interface between the MgO-C coating and the Al2O3-C filter substrate (Fig. 7.12). The fibers grow during a vapor–liquid-solid process [4, 37]. The reactants are Mg, Al and O that are dissolved in the liquid steel penetrating the porous MgO-C coating [40].
In addition to MgAl2O4 whiskers, a dense layer of MgAl2O4 formed at the former MgO-C/steel interface (Fig. 7.13a) [4, 37]. The Al enrichment on the edges of the pre-existing MgO grains of the MgO-C coating produced MgO/MgAl2O4 core/rim structures within the filter ceramic coating [37]. Vice versa, MgAl2O4 formed also on the edges or rims of exogenous Al2O3 inclusions that were present in the steel (Fig. 7.13b).
The SPS/FAST experiments helped to confirm the assumed “reactive” behavior of the carbon-bonded magnesia coatings deposited on the carbon-bonded alumina filter substrates. The MgAl2O4 formation on the filter surface and on exogenous inclusions was accompanied by a reduction of the oxygen concentration in the steel. The dense layer of MgAl2O4 is expected to represent a limiting factor for a further penetration of the molten steel into the interior of the filter material. Thus, a compact MgAl2O4 layer will slow down the reaction kinetics in analogy with a compact Al2O3 coating discussed above. In contrast, the MgAl2O4 whiskers act against the shrinkage of the filter during the filtration process [6] and could also be used to increase the strength of the functionalized filter ceramic [4, 37, 40].

7.4.1 Reactions Between Molten Aluminumor AlSi7Mg0.6 Alloy and Selected Carbon-Free Oxide Coatings

Corundum foam filters are used as a standard tool for the aluminum melt filtration [2]. Still, alternative functional materials with specific wettability [41], selective reactivity and interactivity with respect to certain impurities and inclusions are sought. This contribution focusses on mullite (3Al2O3·2SiO2, SG Pbam), amorphous SiO2 and rutile (TiO2, SG P42/mnm), which are considered for production of functional coatings. The main phenomena under study were the reaction of the functional coatings with molten aluminum and molten AlSi7Mg0.6 alloy, formation of the reaction layers on the filter surface and the mechanisms of the adhesion between the reaction layers and the functionalized filter surface. Chemical composition of the AlSi7Mg0.6 alloy is presented in Table 7.2. The majority of the experiments discussed in this subchapter were performed in the SPS apparatus using the tool-in-tool setup. Additional experiments were done via impingement tests [13] and sessile drop experiments [41]. The temperature of the melt was always 750 °C.
Table 7.2
Chemical composition of the aluminum alloy powder used for the melt production as specified by the manufacturer
Alloy
Elemental composition [mass%]
Particle size
Source
Al
Si
Mg
Fe
Cu
Mn
Zn
AlSi7Mg0.6
Balance
7.07
0.61
0.09
0.09
0.02
0.28
45 – 100 µm
TLS Technik

7.4.2 Reaction Between AlSi7Mg0.6 and Corundum

For short dwell time (1 min), no reaction layers were detected on the surface of the corundum filters. The results of the SPS treatment and the impingement tests were identical, as for both methods the liquid AlSiMg0.6 did not penetrate into the almost dense ceramics. For the SPS-treated samples, slightly better wetting of the corundum filter by the molten aluminum alloy was observed. This phenomenon was attributed to the destruction of the thin oxide layer formed on the surface of original AlSi7Mg0.6 particles in the SPS melting process. The oxide layer is disrupted through the percolation of the electric current and removed by the reducing conditions in the SPS process, which are established by the application of the graphite tooling [3].
After a longer holding time (30 – 60 min), accumulation of Mg at the surface of the corundum filter was observed (Fig. 7.14). Although no distinct reaction layer was detected, the accumulation of Mg at the Al2O3/AlSi7Mg0.6 interface is regarded as a preliminary stage of the MgAl2O4 formation. Thermodynamic calculations supported this hypothesis and revealed that MgAl2O4 is the stable phase in the whole temperature range between 20 to 1000 °C [3]. According to the thermodynamic simulation, the formation of the MgAl2O4 spinel would, in the equilibrium state, fully consume the 0.6 mass% of Mg, which are available in AlSi7Mg0.6 [42, 43]. Consequently, no Mg-containing intermetallic phases, e.g., Mg2Si (SG Fm\(\overline{3 }\)m), should form. Silicon contained in the melt should precipitate upon cooling. These results of the thermodynamic simulation were proven experimentally. No Mg2Si was found, while Si formed needles in the solidified melt (Fig. 7.14).

7.4.3 Reaction of AlSi7Mg0.6 with AmorphousSiO2and Mullite

It is known from literature [4447] that aluminum silicates, e.g., mullite (3Al2O3 ⋅ 2SiO2), decompose in contact with molten aluminum, and form alumina and silica. Alumina usually crystallizes as corundum (α-Al2O3), while silica is reduced by liquid aluminum to silicon, which dissolves in the melt. This reaction results in additional formation of Al2O3 [44], which occurs in form of metastable alumina phases like η-, θ- and/or γ-Al2O3 [4850]. The presence of Mg in the aluminum alloy is expected to alter the interfacial reactions and to produce additional phases like MgAl2O4 spinel and/or MgSiO3 pyroxene.
Individual steps of the reaction between AlSi7Mg0.6 melt and SiO2 or mullite were visualized by model experiments that were performed with compacted powder samples containing 50 mol% of the aluminum alloy and 50 mol% of SiO2 or mullite. A short-time reaction between molten AlSi7Mg0.6 alloy and amorphous SiO2 at 750 °C led to the decomposition of SiO2, and to the formation of Si, γ/η-Al2O3 and MgAl2O4 in the solidified sample (Fig. 7.15a, 0 h). Note that Fig. 7.15 includes only crystalline phases, because the phase fractions were determined using XRD. With longer reaction time (24 h at 750 °C), the amount of Al2O3 increases, and a part of metastable alumina (γ/η-Al2O3) transforms to corundum (α-Al2O3).
Reaction experiments, which were carried out with an alumina plate coated with amorphous SiO2, revealed that Mg is attracted to the filter wall in the initial stages of the reaction process (Fig. 7.16a). However, as no magnesium silicate was found in the powder mixtures sintered for short time (Fig. 7.15a), it can be concluded that the functional SiO2 coating dissolves rather than a magnesium silicate, e.g., MgSiO3, forms. Free silicon forms Si precipitates, which were observed in the solidified AlSi7Mg0.6 melt (Fig. 7.16b). Oxygen reacts with Al and Mg to Al2O3 and to MgAl2O4 (Fig. 7.15a). In the planar sample, the coexistence of these compounds led to the formation of MgAl2O4 precipitates embedded in an Al2O3 layer (left part of Fig. 7.16b). The secondary Al2O3 layer is separated from the solidified melt by an Mg-rich and O-depleted interlayer (central part of Fig. 7.16b).
The reaction between molten AlSi7Mg0.6 and mullite leads to the formation of Si, Al2O3 and MgAl2O4 (Fig. 7.15b) as well. In this case, however, alumina exists in the thermodynamically stable form of corundum already in the initial stages of the reaction process. The formation of corundum instead of the metastable alumina phases is facilitated by the presence of α-Al2O3, which is a product of the mullite decomposition [4447]. Figure 7.17 depicts the Si needles in the solidified AlSi7Mg0.6 melt and illustrates the preferential diffusion of Mg into the mullite coating along the pore and grain boundaries.

7.4.4 Reaction Between Al and AlSi7Mg0.6 Melts and TiO2

In contact with molten Al or aluminum alloys, TiO2 is reduced in analogy to SiO2. Typical reaction products are Ti3O5, Ti2O3, TiO and Ti [12, 5157]. Metallic Ti is dissolved in the melt. In our SPS experiments that were performed with the powder mixtures of Al and TiO2 (rutile, SG\(P{4}_{2}/mnm\)), α-Al2O3 and Ti2O3 with the corundum crystal structures (SG\(R\overline{3 }c\)) and Al3Ti (SG I4/mmm) were found after 24 h at 750 °C (Fig. 7.18). In literature, also the formation of intermetallic phases like (Al,Si)3Ti (SG I4/mmm), Al60Si12Ti28 or Ti7Al5Si12 [12, 52, 55, 58, 59] was reported. Ti2O3 present in our samples is a product of the TiO2 reduction. α-Al2O3 is a product of the reaction of Al with released oxygen. The formation of corundum (α-Al2O3) is possibly facilitated by the presence of Ti2O3 having the same crystal structure like α-Al2O3. Al3Ti formed, because the Ti concentration in the melt exceeded locally the solubility limit for Ti in Al (< 1 at% at 750 °C).
The experiments carried out with planar samples disclosed that α-Al2O3 grows in form of a compact layer on the surface of the functional TiO2 coating (Fig. 7.19). The analysis of the Al2O3/TiO2 interface using SEM/EBSD revealed that Al2O3 grows on TiO2 frequently with the orientation relationship \({\left(001\right)}_{{{\text{TiO}}}_{2}}\parallel {\left(100\right)}_{{{\text{Al}}}_{2}{{\text{O}}}_{3}}\) and \({\left[010\right]}_{{{\text{TiO}}}_{2}}\parallel {\left[001\right]}_{{{\text{Al}}}_{2}{{\text{O}}}_{3}}\) (Fig. 7.20). The round brackets stand for parallel lattice planes, the square brackets for parallel crystal axes. The slight misorientations (Fig. 7.20b) compensate the differences in the interatomic distances. The pronounced orientation relationship between α-Al2O3 and TiO2 is a consequence of the similarity of their crystal structures (Fig. 7.20c, orientation relationships were plotted using VESTA 3 [60]) and of the possible presence of Ti2O3 as a thin interlayer. Such heteroepitaxial growth improves the adhesion of Al2O3 to TiO2 significantly. Moreover, a compact corundum layer impedes the direct contact between molten aluminum and the TiO2 coating, which inhibits a further reduction of rutile (and other titanium oxides) and the production of free titanium. The lack of reduced titanium retards or even hinders the formation of Al3Ti.
In contact with molten AlSi7Mg0.6 alloy, TiO2 forms quite quickly MgTiO3 (SG R\(\overline{3 }\)), cf. Fig. 7.21a. Additional oxygen, which is required for this reaction, is supplied by the melt. Silicon present in the AlSi7Mg0.6 melt forms Si precipitates. The highly variable thickness of the MgTiO3 layer (Fig. 7.21b) suggests that the diffusion of Mg and O into TiO2 is accelerated, when the TiO2 coating contains cracks, voids or grain boundaries. The analysis of the MgTiO3/TiO2 interface using SEM/EBSD revealed that the grains of magnesium titanate and rutile possess frequently an orientation relationship \({\left(100\right)}_{{{\text{MgTiO}}}_{3}}\parallel {\left(001\right)}_{{{\text{TiO}}}_{2}}\) and \({\left[001\right]}_{{{\text{MgTiO}}}_{3}}\parallel {\left[010\right]}_{{{\text{TiO}}}_{2}}\) (Fig. 7.22). Round brackets denote the lattice planes, while square brackets denote the crystal axes. This orientation relationship was confirmed by a local analysis of the MgTiO3/TiO2 interface using SAED in TEM (Fig. 7.23). The ab initio simulations using the density functional theory [52] revealed that the above orientation relationship reduces the total energy of the MgTiO3/TiO2 interface in comparison with the total energy of the individual bulk components.
Diffraction contrasts visible in the TEM micrograph (Fig. 7.23a) explained the slight misorientation between MgTiO3 and TiO2, which was first concluded from the results of the SEM/EBSD analysis (Fig. 7.22b). The diffraction contrasts stem from geometrically necessary misfit dislocations, which are distributed almost equidistantly along the MgTiO3/TiO2 interface and which compensate the lattice misfit between MgTiO3 and TiO2. These dislocations produce small angle grain boundaries that are visible by SEM/EBSD as slight departures from the ideal orientation relationship.
Furthermore, the local analysis of the reaction zone between TiO2 and the solidified AlSi7Mg0.6 melt using TEM and SAED contributed essentially to the understanding of the reaction kinetics in this system. Within the original TiO2 coating, bands of Ti2O3 (SG \(R\overline{3 }c\), corundum type) and MgTiO3 (SG \(R\overline{3 }\)) having a pronounced heteroepitaxial orientation relationship \({\left(001\right)}_{{{\text{Ti}}}_{2}{{\text{O}}}_{3}}\parallel {\left(001\right)}_{{{\text{MgTiO}}}_{3}}\) and \({\left[100\right]}_{{{\text{Ti}}}_{2}{{\text{O}}}_{3}}\parallel {\left[100\right]}_{{{\text{MgTiO}}}_{3}}\) at their interfaces were detected (Fig. 7.24). The thickness of the individual stripes was about 200 nm. The formation of such structures is facilitated by a local oxygen deficiency. When a sufficient amount of magnesium but a low amount of oxygen diffuse into the TiO2 coating, the oxygen that is required for the formation of MgTiO3 is produced by the reduction of TiO2, which leads to the formation of Ti2O3. Note that Ti2O3 and MgTiO3 are miscible in a broad range of the Ti and Mg concentrations [61] and that MgTiO3 belongs to the group of ilmenites, which can accommodate various metallic species in their crystal structure, e.g., Fe and Mn in addition to or instead of Mg and Ti [62]. Thus, the MgTiO3/Ti2O3 composite layer can incorporate a variety of alloying or foreign elements.
When TiO2, Mg and O come into contact with alumina, e.g., near the interface between the Al2O3 filter wall and the functional rutile coating, MgAl2O4 spinel forms as an additional phase (Fig. 7.25). Whereas the titanium magnesium oxides exist in the form of ilmenite (MgTiO3, SG\(R\overline{3 }\)) and spinel (Mg(Ti1-xMgx)O4, SG\(Fd\overline{3 }m\)) [63], MgAl2O4 exists only in the spinel form. Furthermore, Mg(Ti1-xMgx)O4 and MgAl2O4 are not miscible. Therefore, a complex microstructure consisting of the TiO2, MgTiO3 and MgAl2O4 phases with segregated Ti and Al atoms forms in the reaction zone containing Ti, Mg, O and Al (Fig. 7.25). Nevertheless, the SAED patterns (Fig. 7.25) indicated that these phases possess a pronounced orientation relationship, which was described as \({\left(100\right)}_{{{\text{TiO}}}_{2}}\parallel {\left(11\overline{1 }\right)}_{{{\text{MgAl}}}_{2}{{\text{O}}}_{4}}\parallel {\left(001\right)}_{{{\text{MgTiO}}}_{3}}\) and \({\left[001\right]}_{{{\text{TiO}}}_{2}}\parallel {\left[011\right]}_{{{\text{MgAl}}}_{2}{{\text{O}}}_{4}}\parallel {\left[120\right]}_{{{\text{MgTiO}}}_{3}}\) and which is substantiated by the crystal structure models depicted in Fig. 7.26. Also in this case, the heteroepitaxy of the neighboring phases is believed to improve the adhesion of the reaction layers formed during the filtration process.
The layer formation in each case provided evidence for the “reactive” interaction between metal melt and the functional filter ceramic coating, i.e. rutile, consuming magnesium and dissolved detrimental oxygen.
The binding mechanism between the newly formed magnesium titanate layer and rutile was also investigated systematically to make sure that no spalling or peel-off of MgTiO3 could contaminate the filtered alloy melt. The MgTiO3 layer growth at the interface between the AlSi7Mg0.6 melt and rutile depends strongly on the local orientation of the TiO2 grains.

7.5 Summary and Conclusions

In this contribution, the reaction processes between functionalized metal melt filters and molten 42CrMo4 steel, pure aluminum melt and molten AlSi7Mg0.6 alloy were studied. The samples were produced in a Spark Plasma Sintering apparatus with modified tooling. This experimental setup allowed very fast heating of the samples, controlled melting of the metals and the investigation of reaction processes at constant reaction temperatures and without convection of the melt.
The initial products of the reaction between molten 42CrMo4 steel and carbon-bonded corundum (Al2O3-C) filters were dissolved aluminum and oxygen, CO/CO2 gas and an amorphous reaction layer, which contained Fe, Al, O as well as the alloying elements and impurities from the steel. After a longer reaction time, nanocrystalline wuestite, garnets, spinel-like phases including metastable alumina, and secondary corundum formed. The production of CO/CO2, the formation of reaction layers and the growth of secondary corundum contributed significantly to the reduction of the oxygen content in the steel melt. When the secondary corundum grows directly on the primary corundum from the metal melt filter, the growth is epitaxial. The habitus of the secondary corundum (attached platelets or compact layers) depends on the local availability of carbon in the reaction zone, on the carbon concentration in the Al2O3-C filter and on the filter morphology.
When the carbon-bonded corundum is coated with carbon-bonded magnesium oxide, MgAl2O4 spinel forms in a carbothermic reaction between corundum from the Al2O3-C substrate and MgO. This reaction is facilitated by the presence of iron from the molten 42CrMo4 steel. Iron that penetrates into the porous MgO-C coating promote the formation of MgAl2O4 whiskers at the interface between the MgO-C coating and the Al2O3-C substrate. These whiskers are supposed to strengthen the filter ceramics, which is an additional benefit to the removal of oxygen from the melt, to the increased flotation of oxygen and inclusions present in the melt by the CO/CO2 bubbles and to growth of secondary MgAl2O4 as docking sites for further deposition of endogenous inclusions.
In contact with molten aluminum, functional coatings made from corundum stay intact. Still, they serve as substrate for secondary corundum, which forms from aluminum and oxygen dissolved in the melt. In contrast to Al2O3, SiO2 dissolves completely. Reduced silicon is solved in the melt. Free oxygen reacts with Al to metastable alumina (γ/η-Al2O3), which transforms later to corundum (α-Al2O3). Mullite (3Al2O3 ·2SiO2) decomposes to α-Al2O3 and SiO2. SiO2 dissolves quickly; α-Al2O3 serves as substrate for secondary corundum. Consequently, no metastable alumina phases are formed, when mullite is brought in contact with molten Al. In functional coatings made from rutile, TiO2 is reduced by molten Al to Ti2O3. Free oxygen (including the original oxygen present in the melt) reacts with Al to corundum.
The reactions between molten AlSi7Mg0.6 alloy and the functional coatings made from corundum, mullite and SiO2 are accompanied by the formation of MgAl2O4, which is a product of the reaction of Mg and O with Al2O3. The other reactions and processes are the same like for the Al melt. Mullite decomposes into corundum and SiO2. SiO2 dissolves. Reduced Si is solved in the melt. Oxygen reacts with Al and Mg to Al2O3 and MgAl2O4. On the surface of the corundum and mullite filters, where corundum is present (in mullite as a product of the 3Al2O3 ·2SiO2 decomposition), Al2O3 crystallizes as corundum. If Al2O3 cannot grow directly on corundum substrate, e.g., on the surface of a SiO2 coating, metastable alumina phases are formed first. The TiO2 coatings react with Mg and O from the melt to MgTiO3. If the local concentration of Mg is higher than the local concentration of O, TiO2 is reduced to Ti2O3. When MgTiO3 comes in contact with Al2O3 that stems, e.g., from the skeleton of the filter, MgAl2O4 and TiO2 or Ti2O3 are formed.
Our study illustrated the importance of the spinel-like structures, which were found in all filter/melt combinations under consideration. In all cases, the spinel-like phases (in particular MgAl2O4) accommodated oxygen and impurities from the melts. Furthermore, the spinel-like phases are expected to dock the inclusions having the spinel structure, e.g., the metastable alumina phases. For these reasons, the spinel-like phases are considered as possible candidates for production of the functional coatings covering the metal melt filters.
Another finding of this study concerns the role of the epitaxial or heteroepitaxial growth of secondary coatings during the filtration process. Although many of the (primary) functional coatings under study decomposed in contact with the metal melt, the solved elements were captured in oxides, which formed in a secondary “deposition” process. It was shown that almost all oxides involved in this process are able to grow by mutual heteroepitaxy. This was proven for Al2O3, Ti2O3 (both having the corundum structure), TiO2 (rutile), MgAl2O4 (spinel) and MgTiO3 (ilmenite).

Acknowledgements

The former CRC 920 project collaborators are highly acknowledged for their contributions: Milan Dopita, Tilo Zienert, Lilit Amirkhanyan, Marcus Emmel, Claudia Voigt, Eva Jäckel and Anne Schmidt. Furthermore, we thank Christiane Ullrich, Steffen Dudczig, Jana Hubalkova, Brigitte Bleiber, Dietrich Heger, Beate Kutzner, Christian Schimpf, Diane Hübgen, Katrin Becker, Astrid Leuteritz, Karin Müller and Galina Savinykh for their support with production, preparation and analysis of the samples. The electron probe microanalyzer was funded by the German Research Foundation in the frame of the Major Research Instrumentation Program under the project number 395240765. The German Research Foundation (DFG) is acknowledged for funding the Collaborative Research Centre 920—Project-ID 169148856, subproject A06.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the chapter's Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
Literatur
10.
Zurück zum Zitat H.U. Kessel, J. Hennicke, Interceram 56, 164 (2007) H.U. Kessel, J. Hennicke, Interceram 56, 164 (2007)
30.
Zurück zum Zitat C.E. Sims, F.B. Dahle, Trans. Am. Foundrymen’s Assoc. 46, 65 (1938) C.E. Sims, F.B. Dahle, Trans. Am. Foundrymen’s Assoc. 46, 65 (1938)
50.
Zurück zum Zitat P. Mossino, D. Vallauri, F.A. Deorsola, L. Pederiva, R. Dal Maschio, G. Scavino, I. Amato, Metallurgia Italiana 97, 25 (2005) P. Mossino, D. Vallauri, F.A. Deorsola, L. Pederiva, R. Dal Maschio, G. Scavino, I. Amato, Metallurgia Italiana 97, 25 (2005)
62.
Zurück zum Zitat C. Klein, C.S. Hurlbut, Jr., Manual of mineralogy (after James D. Dana), 21st ed., (New York: Wiley, 1993), pp. 380–381, ISBN 047157452X C. Klein, C.S. Hurlbut, Jr., Manual of mineralogy (after James D. Dana), 21st ed., (New York: Wiley, 1993), pp. 380–381, ISBN 047157452X
Metadaten
Titel
Interface Reactions Between the Metal Melt and the Filter Surface Activated by a Spark Plasma Sintering Process
verfasst von
Anton Salomon
Mykhaylo Motylenko
Martin Thümmler
David Rafaja
Copyright-Jahr
2024
DOI
https://doi.org/10.1007/978-3-031-40930-1_7

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.