Skip to main content
Erschienen in:
Buchtitelbild

Open Access 2024 | OriginalPaper | Buchkapitel

8. Dealing with Fe in Secondary Al-Si Alloys Including Metal Melt Filtration

verfasst von : Hanka Becker, Andreas Leineweber

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

Fe is a detrimental impurity element in secondary, i. e. recycled, Al–Si cast alloys (Zhang et al. in Miner. Process. Extr. Metall. Rev. 33:99, 2012;Raabe et al. in Prog. Mater. Sci. 128, 2022;). It leads to decrease of castability and promotes crack formation due to formation of primary, Fe-containing intermetallic particles, e.g. plate-shaped β-Al–Fe–Si, coarse αh-Al–Fe–Si or αc-Al–(Fe,Mn,Cr)–Si in presence of further transition metal elements e.g. Mn and Cr. Successfully, dealing with such secondary Al–Si cast alloys contributes to sustainability, circular economy and reduction of energy consumption (Raabe et al. in Prog. Mater. Sci. 128, 2022;Taylor in Mater. Sci. Forum 689:429, 2011;). In the present chapter, a systematic understanding is provided for dealing with Fe impurities in secondary Al–Si alloys by.
  • removal of Fe on the basis of melt conditioning and metal melt filtration and
  • modification of Fe-containing phases into harmless microstructural components.
In this context new insight is obtained into.
  • the crystal structures of some relevant intermetallic phases occurring in secondary Al–Si alloys,
  • thermodynamics and kinetics of phase formation during solidification and
  • the interaction of different filter materials with the transition metal containing Al–Si alloys.
The crystal structures of the β-Al–Fe–Si and δ-Al–Fe–Si phases and of their ordered variants were investigated. This allowed reliable distinction of occurring intermetallic phases, the αh-Al–Fe–Si, the αc-Al–(Fe,Mn,Cr)–Si, the β-Al–Fe–Si and the δ-Al–Fe–Si phase, especially by electron backscatter diffraction. While modification of the alloy composition by the Mn, Cr content and presence of other transition metal elements affect the thermodynamic properties of the phases, these elements also significantly affect the kinetics of phase formation during solidification at high cooling rates including the resulting phase morphology. The formation of primary, intermetallic phases during melt conditioning closely above the solidification temperature of the (Al)-solid solution can be utilized for the removal of Fe by separating the primary, Fe-containing, intermetallic particles from the Fe-depleted Al melt. Removal of such particles by application of filters to increase the Fe-removal efficiency extends the filters’ use beyond the removal of nonmetallic inclusions, contributing to production of high-quality, recycled Al–Si alloys. Evaluation of wettability, chemical reactions and microstructure in the interaction region between the filter material and Al–Si melts and the Fe-depleted melt reveals a beneficial effect of C-bonded Al2O3 filter material.

8.1 The Crystal Structures of the Relevant Intermetallic Phases

The intermetallic phases occurring during the various treatments of the secondary Al-Si alloys are introduced in the following. The focus is put on the phases αh-Al–Si–Fe, αc-Al–Si–(Fe,Mn,Cr) and β-Al–Si–Fe and δ-Al–Si–Fe which are most relevant in case of Al–Si–Fe alloys with Mn and Cr. The characteristics of these phases are summarized in Table 8.1. Note that in (additional) presence of other transition metal elements other or further Fe-containing intermetallic phases may occur.
Table 8.1
Summary of the intermetallic phases αh, αc, β and δ forming in hypoeutectic Al–Si alloys with Fe and Mn. The Greek letter that is commonly assigned to the phase and a frequently used formula is given next to space group and exemplary lattice parameters. Note that this table is meant to present an overview of the relevant phases and cannot cover variation of lattice parameters within the homogeneity ranges. Within the notes line further formulas for the phases and solubilities are given if available. The mentioned characteristic morphologies and kinetic preferences refer to solidification of hypoeutectic Fe-containing Al-Si casting alloys
Phase (Label)
Space group
Lattice parameters
αh-Al7.1Fe2Si
P63/mmc
a = 12.346 Å, c = 26.210 Å [4]
Used formulas: Al7.1Fe2Si [4], Al8Fe2Si [12, 13]
Homogeneity range: Narrow [4], Relative solubility of Mn: Fe0.97Mn0.03 [4]
Morphology: Large particles containing enclosures of Al [14]
Kinetic preference: Slow nucleation, mostly forming as primary phase, can be easily suppressed under high cooling rates [15]
αc-Al9Mn2Si1.8
Pm \(\overline{3 }\)
a = 12.68 Å [8]
Used formulas: Al9Mn2Si1.8 [8], Al20Fe4MnSi2 [9], Al15(Fe,M)3Si2 [10], Al34.1(Fe,M)9.5Si6 [6]
Crystal structure: Disordering for Mn/Fe ratios < 0.66 to Im \(\overline{3 }\) [9, 11],
Homogeneity range:
broad range of Si and of (Fe,Mn,Cr): Al: 68.8 [6] to 75 at% [10], Si: 7.4 [10] to 14 at% [8], Fe, Mn, Cr: 15 [10] to 19.2 at% [6]
Relative solubilities of Fe, Cr, Mn8.1: Fe0.97Mn0.03 to Fe0Mn1 [7], Fe0.88Cr0.12 [16] to Fe0.51Cr0.49 [17]
Morphology: Bulk facetted polyhedral, hopper, coarse and fine dendrites [16, 18]
Kinetic preference: Nucleates easily at all cooling rates during all stages of solidification (primary, co-dendritic, co-eutectic) in sufficient presence of e.g. Mn and Cr [15] and on various surfaces (if present) [19]
β-Al4.5FeSi2
A12/a1 [20, 21]
A21/e11 [22]
I41/acd [23]
a = 6.1676 Å, b = 6.1661 Å, c = 20.8093 Å, β = 91.0° [20]
Used formulas: Al4.5FeSi [10, 2022, 24], Al5Fe2Si [25]
Homogeneity range: Relative solubility of Mn: Fe0.87Mn0. 13 [26]
Morphology: Plate-shaped [15]
Kinetic preference: Unimpaired nucleation; can be suppressed under high cooling rates in sufficient presence of Mn and Cr in favour of αc or δ phase [15]
δ-Al3FeSi2
I4/mcm
a = 6.061 Å, c = 9.525 Å [27]
Used formulas: Al3FeSi2 [10, 26, 27], Al4FeSi2 [28]
Morphology: Plate-shaped [15, 26]
Kinetic preference: Forms in co-existence with β phase in the same particles with increasing δ phase fraction with increasing cooling rates [15]
1 Experiments of the kinetics of intermetallic phase formation on different substrates hints at the fact that under special conditions the αc phase forms as metastable phase in the ternary Al-Si-Fe system without the presence of further metallic element e.g. transition-metal elements
2 See text for explanation of the simultaneous presence of three structures

8.1.1 αh-Al–Si–Fe and αc-Al–Si–(Fe, Mn, Cr)

In pure Al–Si–Fe alloys, the hexagonal αh-Al–Si–Fe phase (space group P63/mmc) with a composition given by Al7.1Fe2Si develops [4]. However, in presence of as small amounts as > 0.03 wt% Cr [5] or > 0.3 wt% Mn [57] the cubic αc-Al–Si–(Fe,Mn,Cr) phase forms instead of αh with approximate compositions between Al9Mn2Si1.8 [8], Al20(Fe,Mn)5Si2 [9] and Al15(Fe,Mn)3Si2 [10]. It has space group \(Im\overline{3 }\) at low ratios of Mn or Cr to Fe content or \(Pm\overline{3 }\) at high ratios of Mn or Cr to Fe content [9, 11].
Both the structures of αh and αc can be derived from periodic arrangements of Mackay icosahedra (MI), supplemented by further atoms (Fig. 8.1a,b). Thereby, the ideal icosahedral symmetry of such clusters is, of course, broken (Fig. 8.1c). In αc the MI form a bcc arrangement. Additional sites of Al/Si atoms are located between the second-nearest neighbor clusters [9].
In αh the MI form a primitive hexagonal arrangement, where ordered occupation of 1/2 of the trigonal prismatic interstices by further clusters leads to a doubling of the c axis. There P63/mmc symmetry results by the latter clusters occupying the (1/3,2/3,1/4) and (2/3,1/3,3/4) positions forming an AB stacking and the MI end up on the octahedral sites. It is worth to stress that the αh and αc phase are clearly distinct phases with different properties and should be clearly distinguished during microstructure analysis.
The αh phase forms during primary crystallization from a Fe-containing Al–Si alloy melt under slow cooling rates in a shape of large polyhedra that can contain inclusions of Al [14, 15]. In co-dendritic and co-eutectic stages of solidification1 and at high cooling rates the evolution of the αh phase is suppressed [15]. The αc phase forms in Fe-containing Al–Si alloys containing further transition metal elements especially Mn and Cr. Furthermore, experiments on the kinetics of intermetallic phase formation from ternary Fe-containing Al–Si alloys on different substrates reveals that under special conditions the αc phase forms as metastable phase in the ternary Al–Si–Fe system. This might be attributed to the possible presence of icosahedral clusters already in the melt [31, 32]. Depending on the external conditions, e.g. elemental composition and cooling rate [18], the αc phase evolves into a broad variety of morphologies (Chinese script, bulk facetted polyhedra etc.) [16, 18]. The morphologies of the αc phase can be evaluated e. g. on the basis of EBSD data. Challenges arising from pseudosymmetry issues during indexing of EBSD patterns of the αc phase are scrutinized in [29].

8.1.2 β-Al–Si–Fe and δ-Al–Si–Fe

The structural characteristics of the β-Al–Si–Fe and δ-Al–Si–Fe phase in this section have been detailed in depth in [15, 22, 23]. The atomic structure of the β and δ phase consists of basic layers built from bicapped antiprisms with Al and Si atoms at the corners and with Fe atoms in the center (Fig. 8.2a). Within these layers, the bicapped square antiprisms share common edges, i.e. two (Al, Si) atoms. In the crystal structure of the β phase, pairs of these layers are condensed by sharing (Al, Si) atoms at the caps, thus, forming double layers (Fig. 8.2b,c). Stacking of these double layers by relative layer shifts of a/2 + cd or b/2 + cd allows four positions of the double layers, A, B, C, D, where cd is a vector perpendicular to unit cell basis vectors a and b. |cd|, thereby, is the distance between the double layers.
The δ phase is structurally closely related to the β phase, but thermodynamically β and δ are clearly distinct phases [23]. The atomic structure of the δ phase is composed of fully condensed basic layers by sharing all cap atoms. The δ phase is described in space group I4/mcm or space group Pbcn when ordering of Al and Si atoms is considered (Fig. 8.2c) [27]. The β phase has an ideal chemical composition of Al4.5FeSi [10, 20, 21, 23, 33, 34]. The δ phase is typically described with an ideal chemical composition Al3FeSi2 [27].
Several crystal structure variants for the β phase have been reported. All these structures can be described using near-tetragonal lattice parameters that can be grouped into monoclinic [10, 20, 33] e.g. with space group A12/a1 [20, 21], orthorhombic [24, 25] or tetragonal [24, 33] structures which provide lattice parameter c either around 20.8 Å [20, 21, 24, 25] or 41.6 Å [10, 24, 33]. Based on SEM, TEM including SAED and HAADF and DFT calculations, that series of structures was rationalized to a more complete image considering several hierarchical levels of the crystal structure [22].
  • Polytypes
From the stacking of the A, B, C and D double layers, ordered polytypes or disordered structures arise which were elaborated in [23]. The ordered polytypes that form from a periodic stacking of the layers by b/2 + cdb/2 + cd or b/2 + cda/2 + cdb/2 + cda/2 + cd result in an AB sequence or an ABCD sequence. The former ideally shows orthorhombic Aeam symmetry (with a pseudotetragonal unit cell shape) and the latter ideal tetragonal I41/acd symmetry (Fig. 8.2b). However, none of these ideal structures explains the clearly observed unit cell distortions away from a tetragonal or orthorhombic metric, nor some additional diffraction peaks hinting at further ordering effects. For the purpose of orientationally robust indexing e.g. of EBSD patterns of the β phase with locally varying stacking sequence including also disorderedly stacked regions, an artificial, partially disordered tetragonal crystal structure with I4/mmm symmetry has been devised and successfully applied in practice, e.g. upon indexing of EBSD data from δ + β containing microstructures [23, 35].
  • Ordering of the Al Versus Si atoms
Reconciling experimental evidence from literature and own studies in combination with new theoretical evidence from DFT calculations allowed to address these unsolved microstructure interpretations in [22]. Differently Al versus Si ordered model structures have been subjected to DFT calculations. Analysis of the relaxed lattice parameters imply that unit cell distortions away from an orthorhombic or tetragonal metric are generally enforced by the symmetry induced by the atomic arrangement of Al versus Si atoms. Two monoclinic structures differing solely by the orientation of the ordering in the double layers with respect to the AB stacking provide the lowest-energy. These two structures can be described in space groups.
  • A12/a1 with lattice parameters a = 6.1667 Å, b = 6.1683 Å, c = 20.7791 Å and β = 91.51° (in agreement with structures proposed in previous literature) and
  • A21/e11 (or P21/b11 with doubled set of atomic sites) with lattice parameters a = 6.1739 Å, b = 6.1602 Å, c = 20.7840 Å and α = 88.63°.
The monoclinic angle deviates by about 1.3–1.6° from 90°, which is compatible with results from powder-X-ray diffraction analysis. The ordering in these two lowest-energy variants can be transferred to the ABCD stacking of double layers, also leading to monoclinic symmetry. Thus, previously unexplained irreconcilable sets of additional diffraction peaks comply with the three ordered structures. Three different crystal structures coexisting simultaneously at constant composition of β-Al4.5FeSi are unexpected under thermodynamic equilibrium and can be non-equilibrium effects.
  • Defect structure
Various disordered stacking and Al versus Si ordering variants of the β can be considered as defect structure. In characteristically plate-shaped particles (Fig. 8.2d, e, f) neighboring extended phase regions of double layer stacked β phase and full condensed basic layer stacking of δ phase exist. Furthermore, next to these regions, short successive additionally condensed basic layer sequences are frequently present in the β phase. Such δ-like stacking sequences are typical planar defects of the β phase. Their presence can be related to compensation of excess Si, as implied by the higher Si content of the δ phase (Al3FeSi2) [22].
These structural details, on the one hand, can obscure the microstructure characterization when the full complexity is not considered during the microstructure analysis, e. g. when additional superstructure reflections would remain unidentified or misleadingly suggest the presence of further phases. On the other hand, some characterization methods might not resolve all details, so that it is recommended to use artificially higher symmetries in the microstructure analysis. The method and the structural model should address the same level of resolution.

8.2 Thermodynamic and Kinetic Considerations on the Phase Formation During Solidification

8.2.1 Melt Conditioning Under Influence of Mn, Cr and Mg

The results and conclusions highlighted in this section mainly reported in [14] outline the state of melt conditioning based on new results and literature. The current limitations and chances for dealing with Fe in Al–Si alloys by removal of Fe on the basis of melt conditioning and metal melt filtration are presented. Melt conditioning addresses the intended adjustment of the melt or semi-liquid state for a specific purpose. Here, the formation of primary, Fe-containing, intermetallic phases, the so-called sludge, in liquid aluminium is the precondition for removal of Fe by metal melt filtration. To maximally reduce the residual content of Fe in the Al melt by separating off those primary, Fe-containing, intermetallic particles, the melt has to be specifically conditioned for that purpose according to the parameters described in the following:
  • The modification of the alloy composition by addition of further elements like Mn and Cr can promote binding Fe in intermetallic phases by an increased tendency to form sludge [3640] and can reduce the solubility of Fe in the Al melt [4145]. Furthermore, these elements can change the type of primary phases constituting the sludge, i.e. αc, αh, β and/or other phases [14, 15, 44, 45] and, thus, influence the achievable Fe removal.
  • The amount of sludge increases with decreasing conditioning temperature. Thereby, the remaining amount of Fe and the added elements as Mn and Cr in the melt decreases [41, 46, 47]. In order to remove a high fraction of Fe, the conditioning temperature should be chosen closely above the solidification temperature of the (Al)-solid solution TAl-s.
  • The conditioning time should be sufficient to achieve a maximum (equilibrium) volume fraction of primary, intermetallic particles. The heat treatment time of 2 h was chosen according to [42, 4446] to achieve the equilibrium volume fraction of primary particles.
  • A coarse compact polyhedral morphology with a large particle size is preferred for the purpose of Fe removal by separation of primary, intermetallic particles from the Al melt from technological point of view. Such a coarse morphology is known to form as for the αc and also the αh phase under slow cooling or longer heat treatment [16, 18, 39, 44, 45, 48]. Note that various interfaces, especially naturally occurring oxide films are able to promote heterogeneous nucleation of the β, the αc and the αh phases [19, 49, 50].
The primary particles form between the liquidus temperature TL and the onset temperature of the (Al)-solid solution TAl-s. These temperatures as well as the type(s) of the primary phase(s) depend on the alloy composition. Information about these temperatures and phases is required to define the conditioning parameters for the melt conditioning. Three approaches can be evaluated to access these parameters: the sludge factor approach, thermodynamic calculations and thermal analysis.
Sludge factor: Conditions for the formation of primary, intermetallic particles, the sludge, can be estimated based on the sludge factor SF [3640]
$$ SF = w_{{{\text{Fe}}}} + { 2}w_{{{\text{Mn}}}} + { 3}w_{{{\text{Cr}}}} . $$
(8.1)
w are the mass percentages of the elements in the index. The sludge factor can be used to estimate the effect of the transition-metal elements on the sludge formation predominantly on the basis of the weighted sum of the mass percentages of the metal elements Fe, Mn and Cr. The effect of Mn is twice and that of Cr three times as large as that of Fe. The critical temperature below which sludge forms was proposed to depend solely on the sludge factor:
$$ T_{{{\text{sludge}}}} = \, \left( {{86}.{7}^\circ {\text{C }} \cdots SF/{\text{wt}}\% } \right) \, + { 5}0{6}^\circ {\text{C }}\left[ {{39}} \right]{\text{ based on }}\left[ {{37}} \right], $$
(8.2)
$$ T_{{{\text{sludge}}}} = \, \left( {{44}^\circ {\text{C }} \cdots SF/{\text{wt}}\% } \right) \, + { 552}^\circ {\text{C }}\left[ {{39}} \right]{\text{ based on }}\left[ {{36}} \right]. $$
(8.3)
Also another relationship for the critical temperature in dependence on the Fe content has been suggested:
$$ T_{{{\text{sludge}}}} = \, \left( {{34}.{2}^\circ {\text{C }} \cdots w_{{{\text{Fe}}}} /{\text{wt}}\% } \right)^{{2}} + { 645}.{7}^\circ {\text{C }}\left[ {{51}} \right]. $$
(8.4)
According to [36] such dependences represent a kind of liquidus temperature and in case of equilibrium conditions the sludge temperature corresponds to the liquidus temperature. Therefore, in the following, the sludge temperature Tsludge is referred to as liquidus temperature TL.
Thermodynamic calculations: A more advanced approach to access TL and TAl-s are thermodynamic calculations e.g. by using the CALPHAD method [52]. Additionally, these calculations provide information about the type(s) of primary phase(s) that form and provide information about the phase formation sequence during solidification. However, the current thermodynamic databases for thermodynamic calculations evidently require some fine-tuning to predict the effect of modifying elements as Mn and Cr correctly. That current limitation but also possible kinetic effects that cause a deviation from equilibrium conditions, make experimental investigations e.g. by thermal analysis necessary to achieve information on TL, TAl-s and evolving phases.
Thermal analysis: Thermal analysis data e.g. differential thermal analysis (DTA) curves allow experimentally evaluating reactions occurring during melting of alloys. Occurrence of such reactions is revealed by endothermic signals in DTA data. From data evaluation reaction temperatures e.g. TL and TAl-s can be extracted. The primary phases are identified based on their coarse morphology as compared to all other phases present in the microstructure. Thus, the formation of the primary phases can be assigned to the occurrence of the endothermic signal observed in the DTA signal corresponding to final melting of the alloy at TL. TAl-s can be derived from the high intensity signal in the DTA curve corresponding to melting of the main microstructural component Al in the alloy.
The three approaches have been applied to obtain the temperatures TL and TAl-s for the alloys of the compositions Al7.1Si(1.5-xM)Fe(xM)M with M = Mg, Mn, Cr and xM = 0, 0.3, 0.6, 0.9, 1.2 and 1.5 at%. The numbers preceding the elements indicate molar fractions. The resulting temperatures and calculated or observed primary phases are shown in Fig. 8.3.
  • Solidification of the (Al)-solid solution:
According to the results from the DTA experiments on the Al7.1Si(1.5-xM)Fe(xM)M alloys, solidification of the (Al)-solid solution starts closely below 620 °C. The thermodynamic calculations and DTA experiments show good agreement for TAl-s. Thus, the process window for the melt-conditioning temperature can be reasonably well predicted by the CALPHAD method [52] using the TCAL4 database [53]. Hence, holding the melts at 620 °C, as applied here for melt conditioning, is the reasonable lowest possible for potential development of primary, intermetallic phases without solidification of the (Al)-solid solution during melt conditioning.
  • Solidification of the primary, intermetallic particles
In contrast to TAl-s, the liquidus temperatures TL and primary phases obtained from DTA experiments, thermodynamic calculations or the sludge factor approach agree to only some limited extend.
In the alloys with Mg, the deviation is less than 8 K for all Mg-containing alloys including the alloys with xMg ≥ 0.6 at% and with xMg ≤ 0.3 at%. Note that, in the former, the (Al)-solid solution is the primary phase. In the latter, the β phase is the primary phase in the experiment while the αh phase results from thermodynamic calculations. The discrepancy of the formed primary intermetallic phase can be understood in relation to the retarded nucleation of the αh phase if the cooling rates are sufficiently high, as obviously the 0.5 K/s applied for cooling in the DTA experiments. Note that, after the actual melt-conditioning treatment of 2 h at 620 °C, the αh phase is observed as primary phase (Fig. 8.4).
For Mn- and Cr-containing alloys, the observed type(s) of primary phase(s) after DTA experiments and melt conditioning agree, however, partly differ from CALPHAD based predictions. The appearance of the primary phases in the alloys quenched after melt conditioning is illustrated in Fig. 8.4.
In alloys with Mn, the difference of the thermodynamically predicted and observed liquidus temperatures is remarkable (>50 K) although the primary phase is the αc phase in both cases.
In alloys with Cr, the difference of the liquidus temperatures is less pronounced compared to Mn-containing alloys. However, the predicted phases for all compositions differ from the observed phases. The αc phase is observed as the only primary phase for xCr = 0.3, 0.6, 0.9 at%. For xCr = 1.2, 1.5 at%, the (Al,Si)2Cr, the Al13Cr4Si4 and the αc phase occur before the onset of solidification of (Al)-solid solution. The αc phase is not implemented in the description of the Al–Si–Fe–Cr system in the TCAL4 database [53] while the (Al,Si)2Cr is implemented but not predicted for the present alloy compositions. Furthermore, the observed (Al,Si)2Cr phase should generally dissolve during solidification and should not be present together with Al in the solid state [54]. Instead, experimentally not observed formation of Al5Cr and Al45Cr7 is predicted by thermodynamic calculations employing the TCAL4 database [53].
It can be concluded that the databases for prediction of the formation of primary intermetallic phases by CALPHAD generally need some further fine-tuning e.g. by implementation of the αc phase in the Al–Si–Fe–Cr system which is a true quaternary phase in this system not existing in any of the ternary subsystems.
In view of the sludge factor approach, the agreement of the liquidus temperatures TL is surprisingly good for the temperature estimation from Eq. 8.3 [36] with experimental values as long as an intermetallic phase is the primary phase (Fig. 8.3). The sludge factor approach was initially introduced for Cu-containing, secondary Al–Si alloys with similar Si content. The agreement to the temperature estimations from Eqs. 8.2 and 8.4 [37, 51] is less good which might be related to a remarkably higher Si content in the alloys in [51] compared to the present alloy compositions. Furthermore, it has been observed that Cu has a minor and Mg has no effect on the temperature of sludge formation [36] leading to the good agreement of the sludge factor approach in the present Cu-free Al–Si alloys. From this observation, it can be further assumed that the sludge factor approach can be applied to commercial Al–Si alloys in presence of alloying elements that do not take part in the sludge formation like Zn, Ni, Pb, Sn and Ti [36, 38] or when the amount of the element is negligible. However, in case of elements like Sc, B and V that promote the αc-phase formation [1], or if the effect of elements is unknown like Be, Sr, Co, Mo and S the sludge factor approach in form of Eqs. 8.18.4 should be used with care.
  • Fe and transition-metal reduction in the residual melt after its conditioning
The solubility of the elements Fe, Mn, Cr and Mg in the liquid Al at 620 °C determines the remaining minimum amount of these elements in the alloy after separation of the sludge from the liquid metal. These amounts, determined by EDS area measurements, are summarized in the quasi-ternary (Al,Si)–Fe–M projection of the Al-rich corner of the corresponding quaternary phase diagram (Fig. 8.5a) for the alloy series Al7.1Si(1.5-xM)Fe(xM)M with M = Mg, Mn, Cr and Al7.1Si1.5FexMM with M = Mn, Cr where xM = 0, 0.3, 0.6, 0.9, 1.2 and 1.5. at% in both alloy series.
For the evaluation of the success of Fe removal often relative Fe reductions are specified. However, those values depend on the initial Fe content. Consequently, it is recommended to use absolute elemental contents e. g. of the remaining melt in preference or in addition to the relative reductions. That facilitates the comparison of the success of melt conditioning for Fe removal with other studies.
In the Mg-containing alloys with xMg ≥ 0.6 at%, no solid phase was present at 620 °C and the chemical composition after melt conditioning equals the initial chemical alloy composition. In the other alloys, the formation of primary, Fe-containing, Mg-free intermetallic particles reduces the Fe content in the remaining material. In the ternary Al7.1Si1.5Fe alloy, 0.88 at% Fe remain dissolved in the liquid metal. The phase responsible for binding Fe in the alloys with xMg ≤ 0.3 at% is the αh phase. Note that Mg as typical additional alloying element of Al–Si cast alloys is not bound by sludge particles and remains in the melt.
In case of alloys with Mn and Cr, the Fe content and the total transition-metal content in the remaining melt decreases with increasing xMn/(xMn + xFe) and xCr/(xCr + xFe) ratios in the alloys (Fig. 8.5a). That is attributed to the decreased ability of the melt to dissolve Mn and Cr. In presence of Mn and Cr, the αc phase is responsible for binding Fe and also Mn and Cr in agreement with [14, 41, 4446]. Furthermore, a smaller amount of Cr than Mn for the same initial Mn and Cr content remains in the melt related to additionally binding Cr in the (Si,Al)2Cr and the Al13Cr4Si4 phase which, however, do not bind Fe in relevant amounts.
The optimum Mn and Cr content for Fe reduction sensitively depends on the actual alloy composition [4143]. Next to the experimentally determined residual melt composition, the chemical composition of the melt at 620 °C from thermodynamic calculations and from the sludge factor approach are illustrated in Fig. 8.5a. In the latter case, the minimum content of Fe in the melt is reached according to [36] when the sludge factor equals 1. In agreement with experimental observation, own and literature, reported in [14, 4145], the solubility of Fe decreases with increasing xMn/(xFe + xMn) ratio and xCr/(xFe + xCr) ratio. However, the optimum ratio with respect to the residual transition-metal content is controversial. While [43] achieved minimum xFe + xMn contents as low as 0.3 at% for an initial xMn:xFe ratio of 3:1, in the present investigation, a continuous decrease of the transition-metal content towards the ternary Al–Si–Mn and Al–Si–Cr alloy composition was observed. In addition to the chemical composition, the residual Fe, Mn and Cr content in the melt decreases with decreasing conditioning temperature [41, 46, 47] and approaches the lowest achievable content after a critical holding time. Then, the maximum amount of Fe and Mn are bound in primary, intermetallic particles [4244, 46] and their maximum volume fraction is reached. Consequently, a lower residual Fe and Mn content in the present study was achieved compared to similar alloys if these were conditioned at higher temperatures [41, 42, 46] or shorter times [42, 46].
Figure 8.5b depicts the intermetallic phases formed with the Fe, Mn and Cr which have remained in the melt despite optimum melt-conditioning at 620 °C and quenching. These intermetallic phases have formed during the co-dendritic and co-eutectic stage and are located in the Al–Si eutectic regions. Corresponding to the remaining contents of Fe, Mn and Cr, the volume fraction of intermetallic particles is lower in alloys with initially high xMn/(xMn + xFe) and xCr/(xCr + xFe) ratios. In case of the Mg-containing alloys, the reduced amount of the Fe-containing, intermetallic particles reflects the initially lower Fe content with increasing xMg/(xMg + xFe) ratio, although the Fe content is not (xMg ≥ 0.6 at%) or not remarkably (xMg = 0.3 at%) changed in presence of Mg. In Mn- and Cr-containing alloys, the reduced amount of the Fe-containing, intermetallic phases is attributed to the remarkably reduced Mn and Cr level, simultaneously to the reduced Fe level remaining in the melt after melt conditioning. Furthermore, although depending to some extent on the cooling rate, the morphology of the Mn- and Cr-containing, intermetallic particles is irregular and roundish compared to the plate-shaped particles in the eutectic region in alloys with low initial xMn/(xMn + xFe) and xCr/(xCr + xFe) ratios.

8.2.2 Effect of Different Cooling Rates and Mn on the Phase Formation

In this section results and conclusions from [15] about the microstructures that have formed under different cooling rates and xMn/(xMn + xFe) ratios in Al7.1Si(1.5-xMn)Fe(xMn)Mn with xMn = 0, 0.3, 0.375, 0.6 and 0.75 at% and a focus on the Fe-containing, intermetallic phases is presented. Mn is the most frequently used element for modification of the microstructure. Depending on the casting technique, relevant cooling rates range from 1–10 K/s in permanent mold casting to 10–100 K/s in high pressure die casting [55]. Modification of intermetallic particles into harmless microstructural components in order to deal with Fe impurities in secondary Al–Si casting alloys requires detailed understanding of the solidification path in presence of modifying elements and non-equilibrium solidification. The present data is further analyzed to provide a systematic understanding on the solidification path for a broad range of Fe- and Mn-containing Al–Si casting alloys. Thus, it is shown how seemingly controversial results of intermetallic phases in the microstructures of related alloys observed in other studies well integrate to give a consistent picture when considering the effect of cooling rate and presence of further elements especially Mn. Solid-state phase transformations that could occur during subsequent heat treatments are beyond the scope of the presented investigations. No signs of such transformation have been observed in [15], except of marginal transformation of the αh phase into the β phase.
The microstructures of the Al7.1Si(1.5-xMn)Fe(xMn)Mn alloys solidified with cooling rates of 0.05, 1.4, 11.4 and 200 K/s are shown in Fig. 8.6. The type of phases corresponding to the intermetallic particles with different morphology and location within the microstructure and within the sample are indicated. The volume fractions of the intermetallic particles in the solidified microstructures are presented in Fig. 8.7. The intermetallic particles have been quantitatively analyzed of in view of specific morphologies and type of the phases.
  • Mn-free Al7.1Si1.5Fe alloy
In the Mn-free alloy, at the edge of the slowest cooled samples, where the alloy is in contact with the surrounding oxide film, coarse polyhedral particles of the αh phase are present. Plate-shaped particles appear associated with the Al–Si eutectic, i.e. interdendritic, region, but partly reach into the Al-dendrites. All other higher cooling rates have led to microstructures without any αh-phase particles which is attributable to retarded nucleation of the αh phase. Only plate-shaped particles occur. The plate-shaped particles in samples solidified with intermediate cooling rates appear within the Al-dendrites. In samples solidified with the highest cooling rate, the plate-shaped particles are located within the Al–Si eutectic region. These plate-shaped particles consist mainly of the β phase after solidification with low cooling rates and of the δ phase after solidification with high cooling rates. The observed suppression of the αh phase and the formation of the δ phase under high cooling rates contradict thermodynamically calculated solidification paths for equilibrium and Scheil solidification conditions which predict the αh and β phase or solely β phase.
  • Mn-containing Al7.1Si(1.5-xMn)Fe(xMn)Mn alloys
In Mn-containing alloys, at the edge of the samples cooled up to intermediate cooling rates, coarse polyhedral and coarse dendritic particles of the αc phase form during solidification. After solidification with low cooling rates plate-shaped particles accompany the coarse particles located in association with the Al–Si eutectic region. Under the highest cooling rates, plate-shaped particles also form, however, with smaller size and within the Al–Si eutectic. At intermediate cooling rates Chinese-script particles have formed being located within the Al-dendrites. The αc-phase particles form as the only intermetallic microstructural component above a Mn-content dependent cooling rate. The relevant range of cooling rates for the solely αc-phase formation increases with increasing Mn content. In presence of Mn, the discrepancy to the thermodynamically calculated solidification paths is less pronounced. The equilibrium calculation by trend agrees with the solidification at low cooling rates. Scheil solidification calculations are in accordance with the solidification at low to intermediate cooling rates. However, the formation of the δ phase is not predicted by the thermodynamic calculations similar to the Mn-free Al–Si alloys.
These observed deviations from the calculated equilibrium and Scheil solidification path substantiate the reported importance of kinetic effects during solidification of secondary Al–Si alloys [19, 5658]. These kinetic effects can include:
  • suppression of phases which are expected to occur according to equilibrium and Scheil solidification conditions e.g. due to retarded nucleation,
  • occurrence of phases which are not expected to occur according to equilibrium and Scheil solidification conditions and
  • formation of metastable phases which are not present in the equilibrium phase diagrams (not observed in the investigation in [15]).
The influence of such kinetic effects during non-equilibrium solidification cannot be fully accounted for by Scheil solidification conditions or alternatively by setting a certain phase or several phases dormant. Therefore, it is suggested to construct so-called apparent liquidus projections (Fig. 8.8) which provide a tool to estimate the phase formation during solidification. Their construction is based on the experimental microstructures, which are analyzed according to the present phases and their relative location in the microstructures and in the sample.
As obvious from the apparent liquidus projections, the kinetic effects on the microstructure formation in secondary Al–Si alloys are pronounced. It can be assumed that two aims are followed by addition of further elements and Mn and solidification with a high cooling rate: avoiding sludge formation and suppressing the formation of plate-shaped particles. However, some trade-offs must be considered when the parameters are optimized.
The general alloy composition including the Si content and the total transition metal content, as well as the xMn/(xMn + xFe) ratio, determine the potentially forming phases. Aiming at an optimized microstructure with a low intermetallic phase fraction, without plate-shaped particles and without sludge particles, Mn should be present, but the xMn/(xMn + xFe) ratio and total transition metal content should be as low as required and the minimum cooling rate should be chosen according to the chemical composition of the alloy.

8.3 Interaction of Different Filter Materials with the Transition Metal Containing Al–Si Alloys

The following sections elucidate the effect of oxide- and carbon-containing filter materials in contact with Al7.1Si1.5Fe alloy on the Fe-removal efficiency in order to produce high-quality, secondary Al–Si alloys. Details exploring the utilization of filter materials to support the Fe removal can be found in [59]. The basic interaction effects investigations were investigated on flat substrates of Al2O3, mullite (3Al2O3 · 2SiO2), spinel (MgAl2O4), silicon carbide (SiC) and carbon bonded alumina (Al2O3-C).
Sessile-drop experiments have been carried out at 950 °C with the alloy in fully liquid state and at 625 °C with the alloy in the melt conditioned state (see Sect. 8.2.1) to assess the interaction behavior. For the interaction with Al7.1Si and Al7.1Si0.75Fe0.75Mn alloys as well as interaction experiments in small scale crucibles see [59].
Evaluation of the effectiveness of filter materials for Fe-removal requires information about:
(1)
the contact of substrate and alloy melt in view of wetting, infiltration, reactive chemical interaction and formation of new layers at the interface,
 
(2)
the effect of the filter material on the alloy melt composition concerning the desired removal of impurity Fe and possibly further transition metal elements, but also changes of the chemical concentration of intended alloying elements and contamination by external elements from the filter material,
 
(3)
the effect on the formation of primary Fe-containing intermetallic particles regarding the type of phase and its amount which might include nucleation impacts and
 
(4)
the position of primary Fe-containing intermetallic particles at the interface to the substrate or within the droplet volume.
 
The interaction characteristics and the consequences for the alloy are illustrated in Fig. 8.9 and summarized in Table 8.2.
Table 8.2
Summary of the interaction characteristics of the substrate material with the Fe-containing Al-Si alloys. The table was reproduced from [59], Copyright Wiley
 
Al2O3
Mullite
Spinel
SiC
Al2O3-C
Reaction
Non-reactive
3 SiO2(s)
 + 4 [Al](l)
→ 3 [Si](l)
 + 2 Al2O3(s)
3 MgAl2O4(s)
 + 2 [Al](l)
→ 3 [Mg](l)
 + 4 Al2O3
3 SiC(s)
 + 4 [Al](l)
→ 3 [Si](l)
 + AlxCy(s)
3 C(s)
 + 4 [Al](l)
→ Al4C3(s)
Contact angles (60 min, 625 °C, Al7.1Si1.5Fe)
115°
104°
119°
72°
84°
Phase in contact
to the alloy melt
Al2O3
Al2O3
(main
component)
Al2O3
(main
component)
Al4C3
Al4C3
Alloy
composition
As-initial
Si
enrichment8.1
Mg
enrichment8.1
Fe, Cr, V, Ni, Al and Ti
0.9 at% Fe
Fe-containing
primary
intermetallic
αh
β, δ
αh
αc, αh
αh
Location of
primary
intermetallic
Non-
specific
At substrate
At substrate
Non-
specific
At substrate
1 While the Si enrichment was clearly measurable, the Mg enrichment was marginal
(1)
Contact region of substrate and alloy melt
 
The investigated alloy-substrate combinations can be classified into two main groups regarding the wetting behavior in view of the final reaction product in contact region with the alloy melt. The reaction equations with the individual substrates are listed in Table 8.2. One group contains the substrate materials Al2O3, mullite and spinel. Characteristic for these substrate materials is that the final reaction product in contact with the alloy melt is Al2O3. The final contact angles are in the range of 104 to 119° after melt conditioning at 625 °C in agreement with [60]. The other group contains SiC and Al2O3-C. In these systems, the final reaction product Al4C3 is in contact with the alloy melt. Consequently, final contact angles are in the range of 71 to 79°.
Hence, the final wettability mainly depends on the wettability of the liquid melt on the new reaction product [60, 61]. Despite the same reaction product, the spreading kinetics are governed by the reactive interaction at the triple line of the droplet to the substrate [6163]. That temperature–time dependence might, as one aspect, causes the slightly different contact angles.
(2)
Consequences for the alloy melt composition
 
The Al−Si melt in contact with Al2O3 represents a non-reactive, low-wetting system acting as a reference for the reactive systems with the other substrates. Consequently, no contamination of the melt after the interaction has been recognized. After melt conditioning the composition of the remaining melt, without primary Fe-containing αh phase, is in agreement with the Fe-content after melt conditioning in Al2O3 crucibles (Sect. 8.2.1).
In the reactive systems with SiO2 in 3Al2O3 · 2SiO2 mullite and with the MgAl2O4 spinel, the Al alloy melt reduces the oxides forming Al2O3. Thereby, Si enriches in significant large amounts in the melts while Mg enriches in the melt only to a minor extent.
While the interaction of the melt with SiC and Al2O3-C leads, in both cases to a dense layer of Al4C3 carbide, the consequences for the alloy were different. The SiC substrate material is an example for the effect of impurities in the substrate material. Solution of transition-metal elements from spots that are rich in transition metal elements Fe, Cr, V, Ni, and Ti lead to a detectable modification of the solidified microstructure. A detectable enrichment of Si or C in the melt has not been observed. In the case of the Al2O3-C substrate, no contamination of the alloy has been observed. Note that contamination of aluminum alloys with C is undesirable. It could lead to the formation of Al4C3 particles in the bulk microstructure which are known to deteriorate the mechanical properties [64, 65]. However, the solubility of C at 960 °C in molar fraction is approximately 6.4 · 10–4 [66] and thus, might not be relevant for the significant formation of Al4C3 during solidification. However, spalling off of Al4C3 particles from the reaction layer into the melt must be avoided. Furthermore, it has been suggested by [64] that the presence of Si in the Al-Si melt prevents the formation of the carbide layer. This was not observed in the present study although the resulting carbide layer is thinner on SiC than on Al2O3-C after sessile drop experiments.
(3)
Primary Fe-containing intermetallic particles
 
The type of primary Fe-containing intermetallic particles is directly related to the alloy composition after the interaction experiment. As expected for the inert system with Al2O3, the primary phase is the αh phase. The αh phase is also the primary phase in contact to the MgAl2O4 spinel and Al2O3-C. In these systems the introduced impurity content is very low and the elements do not affect the type of primary phase. In case of 3Al2O3 · 2SiO2 mullite, the remarkable increase of the Si content in the alloy, leads to the primary formation of β and δ phase. Although the introduced impurity content of Fe, Cr, V, Ni, and Ti from the SiC substrate is small, their effect is clearly detectable in the microstructure. The transition elements, best known for the modifying effect of small amounts of Cr [5], change the type of intermetallic phase in dependence on the total amount of the element and the cooling rate from αh and β to αc [15]. Although such a microstructural modification towards the presence of αc is often desirable [1], it should be applied as intended and controlled action.
(4)
Position of primary Fe-containing intermetallic particles
 
For effective filtration, the specific attachment of primary particles on the substrates is beneficial. The observed location of αh particles in the alloy on Al2O3 filter material and αhc particles in the alloy on SiC filter material is non-specific. In the alloys on Al2O3-C few large primary αh particles are attached to the interface with the non-reactive Al2O3 filter material. In the case of 3Al2O3 · 2SiO2 mullite and the MgAl2O4 spinel and especially Al2O3-C primary αh particles are specifically attached to the substrate. Thereby, this attachment of the αh particles to Al2O3–C with the carbide layer is associated with oriented growth of the carbide layer.
In conclusion, under the tested filter materials, Al2O3-C is the most promising candidate as a filter material. Although in case of Al2O3-C, the Fe content in the remaining melt of approx. 0.9 at% Fe is the same as in the remaining melt with contact to inert Al2O3 substrate (as expected from thermodynamic calculations). The reactive interaction of the intermetallic towards the Al2O3-C and the resulting attachment to this substrate could possibly be beneficially utilized to increase the Fe-reduction efficiency for secondary Al-Si alloys. Furthermore, the interaction forming a dense thin layer Al4C3 does not contaminate the Al–Si alloy.

Acknowledgements

This study was financially supported by the German Research Foundation within the Collaborative Research Centre 920 “Multi-Functional Filters for Metal Melt Filtration—A Contribution towards Zero Defect Materials” (Project-ID 169148856) in the subproject A07. The authors are grateful for the financial support for the transmission electron microscope investigations related to activities within the centre for research-based innovation SFI Manufacturing in Norway, partial funding by the Research Council of Norway under contract number 237900 and financial support by the Research Council of Norway to the NORTEM project (197405). One of the authors, H. Becker likes to thank for the financial support by the Department of Materials Science and Engineering at NTNU during her visit in Trondheim (Norway) and by the Department of Civil and Mechanical Engineering at DTU during her visit in Lyngby (Denmark).
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.
Fußnoten
1
For explanation of stages of solidification of Fe-containing Al–Si cast alloys see [67].
 
Literatur
2.
Zurück zum Zitat D. Raabe, D. Ponge, P.J. Uggowitzer, M. Roscher, M. Paolantonio, C. Liu, H. Antrekowitsch, E. Kozeschnik, D. Seidmann, B. Gault, F. De Geuser, A. Deschamps, C. Hutchinson, C. Liu, Z. Li, P. Prangnell, J. Robson, P. Shanthraj, S. Vakili, C. Sinclair, L. Bourgeois, S. Pogatscher, Prog. Mater. Sci. 128, 100947 (2022). https://doi.org/10.1016/j.pmatsci.2022.100947CrossRef D. Raabe, D. Ponge, P.J. Uggowitzer, M. Roscher, M. Paolantonio, C. Liu, H. Antrekowitsch, E. Kozeschnik, D. Seidmann, B. Gault, F. De Geuser, A. Deschamps, C. Hutchinson, C. Liu, Z. Li, P. Prangnell, J. Robson, P. Shanthraj, S. Vakili, C. Sinclair, L. Bourgeois, S. Pogatscher, Prog. Mater. Sci. 128, 100947 (2022). https://​doi.​org/​10.​1016/​j.​pmatsci.​2022.​100947CrossRef
5.
Zurück zum Zitat D. Munson, J. Inst. Metals 95, 217 (1967) D. Munson, J. Inst. Metals 95, 217 (1967)
36.
Zurück zum Zitat J. Gobrecht, Schwereseigerungen von Eisen, Mangan und Chrom in Aluminium-Silicium-Gußlegierungen (Teil 1). Giesserei 62, 263 (1975) J. Gobrecht, Schwereseigerungen von Eisen, Mangan und Chrom in Aluminium-Silicium-Gußlegierungen (Teil 1). Giesserei 62, 263 (1975)
37.
Zurück zum Zitat J.L. Jorstad, Understanding “Sludge.” Die Cast Eng. 30, 30 (1986) J.L. Jorstad, Understanding “Sludge.” Die Cast Eng. 30, 30 (1986)
38.
Zurück zum Zitat R. Dunn, Aluminum melting problems and their influence on furnace selection. Die Cast Eng. B 9, 8 (1965) R. Dunn, Aluminum melting problems and their influence on furnace selection. Die Cast Eng. B 9, 8 (1965)
40.
Zurück zum Zitat M.M. Makhlouf, D. Apelian, Casting characteristics of aluminum die cast alloys, work performed under contract No. DEFC07–99ID13716 prepared for US department of energy office of industrial technologies prepared by the advanced casting research center Worcester Polytechnic Institute, pp. 1–46 (2002). https://doi.org/10.2172/792701 M.M. Makhlouf, D. Apelian, Casting characteristics of aluminum die cast alloys, work performed under contract No. DEFC07–99ID13716 prepared for US department of energy office of industrial technologies prepared by the advanced casting research center Worcester Polytechnic Institute, pp. 1–46 (2002). https://​doi.​org/​10.​2172/​792701
57.
Zurück zum Zitat L. Bäckerud, G. Chai, J. Tamminen, Aluminum-Silicon alloys, Chapter 5. In: Solidification Characteristics of Aluminum Alloys Vol.2:, Foundry Alloys, AFS and Skanaluminium, Oslo, Norway (1990) L. Bäckerud, G. Chai, J. Tamminen, Aluminum-Silicon alloys, Chapter 5. In: Solidification Characteristics of Aluminum Alloys Vol.2:, Foundry Alloys, AFS and Skanaluminium, Oslo, Norway (1990)
63.
Zurück zum Zitat V. Laurent, C. Rado, N. Eustathopoulos, Wetting kinetics and bonding of A1 and A1 alloys on u-SiC. Mater. Sci. Eng. A 205, 1 (1996)CrossRef V. Laurent, C. Rado, N. Eustathopoulos, Wetting kinetics and bonding of A1 and A1 alloys on u-SiC. Mater. Sci. Eng. A 205, 1 (1996)CrossRef
Metadaten
Titel
Dealing with Fe in Secondary Al-Si Alloys Including Metal Melt Filtration
verfasst von
Hanka Becker
Andreas Leineweber
Copyright-Jahr
2024
DOI
https://doi.org/10.1007/978-3-031-40930-1_8

    Marktübersichten

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