Next Article in Journal
Process Mineralogy of Lithium and Rubidium in the Diantan Polymetallic Mining Area, Tengchong, Southwest China
Previous Article in Journal
Fracture Density Prediction of Basement Metamorphic Rocks Using Gene Expression Programming
Previous Article in Special Issue
Behavior of Tantalum in a Fe-Dominated Synthetic Fayalitic Slag System—Phase Analysis and Incorporation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stabilization of Mn4+ in Synthetic Slags and Identification of Important Slag Forming Phases

by
Alena Schnickmann
1,*,
Danilo Alencar De Abreu
2,
Olga Fabrichnaya
2 and
Thomas Schirmer
1
1
Department of Mineralogy, Geochemistry, Salt Deposits, Institute of Disposal Research, Clausthal University of Technology, Adolph-Roemer-Str. 2A, 38678 Clausthal-Zellerfeld, Germany
2
Institute of Materials Science, TU Bergakademie Freiberg, Gustav-Zeuner Str. 5, 09599 Freiberg, Germany
*
Author to whom correspondence should be addressed.
Minerals 2024, 14(4), 368; https://doi.org/10.3390/min14040368
Submission received: 29 February 2024 / Revised: 26 March 2024 / Accepted: 28 March 2024 / Published: 30 March 2024

Abstract

:
The expected shortage of Li due to the strong increase in electromobility is an important issue for the recovery of Li from spent Li-ion batteries. One approach is pyrometallurgical processing, during which ignoble elements such as Li, Al and Mn enter the slag system. The engineered artificial mineral (EnAM) strategy aims to efficiently recover critical elements. This study focuses on stabilizing Li-manganates in a synthetic slag and investigates the relationship between Mn4+ and Mg and Al in relation to phase formation. Therefore, three synthetic slags (Li, Mg, Al, Si, Ca, Mn, O) were synthesized. In addition to LiMn3+O2, Li2Mn4+O3 was also stabilized. Both phases crystallized in a Ca-silicate-rich matrix. In the structures of Li2MnO3 and LiMnO2, Li and Mn can substitute each other in certain proportions. As long as a mix of Mn2+ and Mn3+ is present in the slag, spinels form through the addition of Mg and/or Al.

1. Introduction

1.1. Recovery of Critical Elements from the Slag Phase

The usage of critically relevant elements, such as Li, will be increasingly in demand due to their use in various technological areas [1]. The EU has established a list of critical elements, which includes Li as well as Co [2]. In Li-ion batteries (LIBs), these two elements are used as cathode materials (LiCoO2). The geological supply of such elements is limited, thus the recovery of, e.g., Li from the industrial waste stream. such as from old LIBs, is indispensable [3,4,5]. Pyrometallurgical processes [6] are a promising method for the recovery. With this route, various elements (e.g., Co, Ni or Cu) can be recovered directly, while base elements (e.g., Li, Mg, Al or Mn) are either lost as dust or enter the slag phase and form complex compounds. For efficient recovery of critical elements such as Li, it is essential to concentrate the element of interest in a single phase within the slag [1,7]. In addition, this phase should have a high Li-content as well as good processing properties (e.g., habitus, crystal size, magnetic properties). The stabilization of Li in a single phase for an efficient recovery of critically relevant elements is the idea behind engineered artificial minerals (EnAMs) [8] (Figure 1).

1.2. Potential EnAMs for Efficient Recovery

A study by Elwert et al. [9] shows that the Li-aluminate LiAlO2 would be suitable as a potential EnAM, as the idiomorphic to hypidiomorphic crystals can be separated from the remaining slag via flotation [10]. In the presence of especially Mg, Al and Mn, spinel complexes are formed during solidification and solid solutions occurred between spinels (e.g., MgMn2O4, MgAl2O4, MnAl2O4 and Mn2+Mn3+2O4), which hampered the formation of LiAlO2 [11,12,13]. In addition, this Li aluminate can contain up to 3 wt.% Si, which complicates hydrometallurgical processing [10]. As the use of Mn in new Li batteries increases, Mn will become part of a complex slag system along with Li, Mg and Al. In addition, small amounts of Mg and Al in LiMnO2 and Li2MnO3 are expected to promote intralayer diffusion and Mg should also promote interlayer diffusion [14]. For this reason, current research is focusing on the stabilization of Li in Li-manganates as a potential EnAM [8]. The difficulties here are the redox-sensitive behavior of Mn and the Jahn–Teller effect on the Mn3+ at temperatures above 1445 K [12,13]. Schnickmann et al. [8] showed that it is possible to stabilize a Li-manganate in a complex synthetic oxide slag system (Li, Mg, Al, Si, Ca, Mn). Under a normal atmosphere, a mixture between Mn2+ and Mn3+ was present and pure LiMnO2 was formed. Nevertheless, the formation of spinels and spinel solid solutions could not be prevented, as well as the incorporation of Mn2+ into the Ca-silicate matrix. In a follow-up experiment, pure oxygen (100% O2) was used to stabilize a higher Mn oxidation stage in the slag. For this experiment, the same batch precursors were used as in Schnickmann et al. [8]. This experiment is designed to investigate whether it is possible to stabilize higher Mn oxidation states in synthetic slags and whether Mn4+ forms a compound with the other elements, especially with Mg and/or Al.

1.3. Important Slag Forming Phases

In a complex slag system consisting of Li2O, MgO, Al2O3, SiO2, CaO and MnO, mainly binary and ternary oxide compounds are formed, followed by a residual melt [8]. The compounds described below represent possible phases in a crystalline slag with a special focus on Li-rich compounds and the Mn oxides. Compounds between Mn4+ and Al/Mg have not yet been described in the literature. The ternary compounds are of minor importance for the results presented and are therefore not discussed in detail in this chapter.
The Li2O-MnOx system (1 ≤ x ≤ 2) must be considered for the investigation of possible Mn-containing EnAMs. Due to the redox sensitivity of Mn, the formation of various Li-manganates with different concentrations and Mn speciations are possible. The phase equilibria in the Li2O-MnOx system in air is shown in Figure 2. According to the experimental data obtained by Paulsen and Dahn [15], cubic spinel transforms to tetragonal spinel at high temperatures. A miscibility gap between hausmannite (Mn3O4) and t-spinel was proposed. LiMnO2 is stable at temperatures higher than 1223 K, while the Li2MnO3 phase is stable up to approximately 1240 K. In addition to Paulsen and Dahn [15], Longo et al. [16] also described such compounds in detail, as well as the influence of temperature, pressure and pH on the formation and stability of such Li-manganates. Below 400 °C, the cubic spinel phases with stoichiometry, LiMn1.75O4 and Li4Mn5O12, are found to be stable. They can also be part of the spinel solid solution, as shown in the calculations of [17], but to achieve equilibrium at these low temperatures would be problematic. The phase Li2Mn3O7 is calculated to be stable, but its stability has not yet been experimentally proven [15,16,18]. Between 400 and 800 °C, the Li-spinel (Li(1+x)Mn(2−x)O4) has a large stability field between LiMn2O4 (Mn3.5+) and Li4Mn5O12 (Mn4+) and can coexist with Li2MnO3 (Mn4+). Above 800 °C, Li2MnO3 and LiMn2O4 decompose into Mn3O4 and O2 alongside layered LiMnO2. From 1000 to 1060 °C, orthorhombic LiMnO2 can coexist with lithiated hausmannite (LixMn3−xO4) [15]. According to Mishra and Ceder [19], the Jahn–Teller distortion affects the tetragonal spinel LiMn2O4 as well as orthorhombic (e.g., LiMnO2) and monoclinic layered structures [19]. In addition, the investigations by Schnickmann et al. [8] showed that in the LiMnO2 system the Li and Mn contents can be exchanged and that this compound can incorporate up to 0.35 wt.% Al.
Due to their good flotation properties, Li-aluminates have been investigated as potential EnAMs for a long time. Recently, thermodynamic assessments of the Li2O-Al2O3 system using the CALPHAD approach have been published by Konar et al. [20] and De Abreu et al. [21]. Moreover, De Abreu et al. experimentally studied the phase equilibria in the Al2O3-rich part of the diagram and made calorimetric measurements of heat capacity for intermediate phases. The calculated phase diagram is shown in Figure 3a. Three stable intermediate phases were proposed: LiAl5O8, LiAlO2 and Li5AlO4. LiAlO2 has already been extensively analyzed and discussed for its potential as an EnAM [20,21]. Two polymorphs are reported for LiAlO2: the low-temperature α-phase (trigonal) and the high-temperature γ-phase (tetragonal). Two stable modifications of LiAl5O8 with spinel structure but with different space groups were found in the alumina-rich side of the phase diagram. High-temperature modification of the spinel is inversed and stable with some homogeneity range. No evidence of LiAl11O17 was found in the microstructures and thus LiAl11O17 is not treated as stable.
Another important area of Li-rich phases is the Li2O-SiO2 system, with the following stable phases: Li2Si2O5, Li2SiO3, Li4SiO4 and Li8SiO6 (ordered by decreasing Si concentration). According to the latest results, the Li6Si2O7 phase is treated as metastable. Li2SiO3 is formed from a Li2O-SiO2-rich melt.
Spinels are the most important Li-free compounds, whereas hausmannite (Mn2+Mn3+2O4) is the most important compound in the MnO-Mn2O3 system. Below 1172 °C, the Jahn–Teller effect has an influence on the Mn3+ position and causes deformation of the crystal parameters. The transformation from tetragonal to cubic is reversible and crystal lattice changes from cubic to tetragonal on cooling. Hausmannite can be regarded as a low-temperature phase as well as a deformed spinel under normal conditions (1 atm; 25 °C) and low oxygen partial pressure. Hausmannite forms from Mn2O3 at 1445 K and air oxygen partial pressure. It can form solid solutions with other spinel compounds. One of the most important spinels in the MnO-Al2O3 system is galaxite (MnAl2O4) [12,13,20]. The calculated phase diagrams with thermodynamic parameters optimized by [13] for the Al2O3-MnOx system in air and in the presence of metallic Mn are shown in Figure 3b,c, respectively. In the presence of metallic Mn, thus in varying oxygen partial pressure p(O2) with temperature, the homogeneity range of the spinel phase is narrow and its stoichiometry is close to MnAl2O4. According to crystallographic data, the spinel phase is normal at low temperature, with tetrahedral sites occupied by Mn2+ cations, but with temperature increases the degree of inversion (fraction of Al3+ in tetrahedral sites) increases. According to calculations, which agree with experimental results, the homogeneity range of spinel extends with an increase in oxygen partial pressure and in air conditions (p(O2) = 0.21 bar) the homogeneity range of the cubic spinel extends from Mn3O4 to MnAl2O4 (at high temperatures). It should be noted that the Mn3O4 spinel forming from Mn2O3 via oxygen release has a tetragonal structure. The tetrahedral sites are occupied by Mn2+ and the octahedral sites by Mn3+. Tetragonal spinel transforms to cubic types at 1443 K, probably due to the disproportionation reaction Mn3+ → Mn2+ + Mn4+ to reduce the Jahn–Teller distortion caused by Mn3+.
Phase diagrams of other important systems containing the spinel phase, like MgO-MnOx and MgO-Al2O3 systems are also shown in Figure 3d,e, respectively. The thermodynamic parameters obtained by [22] and [23] were applied to calculate the phase relationships for these binary systems. It should be noted that spinels with tetragonal and cubic structures were found in the MgO-MnOx system. Tetrahedral spinel has a homogeneity range from Mn3O4 to MgMn2O4, while cubic spinel extends from Mn3O4 to Mg2MnO4.

2. Materials and Methods

2.1. Chemicals and Preparation of Precursors and Reference Materials

For the melting experiment, the same batch of precursors was used as in the study by Schnickmann et al. [9], with the following chemical compositions: LiNO3, Mg(NO3)2 × 6 H2O, Al(NO3)3 × 9 H2O, Köstrosol® 0830 AS, Ca(NO3)2 × 4 H2O, Mn(NO3)2 × 4 H2O and C6H8O7.
The gel-combustion method was used to prepare the precursors. For this purpose, common procedures [24,25] were adapted to the experimental setup in the laboratory. The method has also been described by Schnickmann et al. [8]. The gel-combustion product was thermally treated with NH₄NO₃ in quartz crucibles up to 480 °C (10 °K/min; Nabertherm LE 1/11/R7, Nabertherm GmbH, Lilienthal, Germany) to remove the residual carbon. In Table 1, the elemental composition of the precursors can be found.

2.2. Melting Program

For the melting experiment, the thermally treated precursors were melted in Pt/Rh crucibles (filling capacity: 15 mL) under pure oxygen atmosphere at 200 L per hour in a chamber furnace (Nabertherm HT16/17, Nabertherm GmbH, Lilienthal, Germany).
The melt program of Schirmer et al. [10] and Schnickmann et al. [8] was adapted for this process (Figure 4). For each experiment, 1.20 g of sample material was used. The powder was not pressed or compacted in the crucible. All three samples were melted together in one run, in different crucibles, to achieve consistent conditions.
Due to the reaction-resistant material, no reaction took place between the slag and the crucible. However, not all of the slag material could be removed from the crucible, resulting in the loss of small amounts. For example, 1.03 g of S1, 1.01 g of S2 and 1.05 g of S3 were present after the melting experiment. The areas of slag near the surface took on a reddish color, while the rest of the slag was dark gray to black. EPMA was used to analyze the near-surface samples and the remaining slag, which is representative of the entire slag.

2.3. Characterization Techniques

The mineralogical composition of the slags was quantitatively and qualitatively identified using powder X-ray diffraction (PXRD) and electron probe microanalysis (EPMA).
The crystalline phases in the slag sample were analyzed using PXRD (Panalytical X-Pert Pro Diffractometer, Malvern Panalytical GmbH, Kassel, Germany). For the measurement (2θ-angle range from 5° to 100°; step size 0.0066°; 150.45 s per step), a Co-X-ray tube (Malvern Panalytical GmbH, Kassel, Germany, λ = 1.7902 Å, 40 kV, 40 mA) was used. The individual crystalline phases were classified via the American Mineralogist Crystal Structure Database [26] and the pdf-2 ICCD XRD database [27].
The elemental composition of single grains/crystals in the prepared thin sections were determined with EPMA (Cameca SXFIVE FE Field Emission, CAMECA SAS, Gennevilliers Cedex, France) using the Kα lines (Mg, Al, Si, Ca, Mn). For the measurement (15 kV beam diameter 100–600 nm; Schottky type [28]), the device was calibrated beforehand with certified reference materials (CRM: P&H Developments Ltd.; Glossop, Derbyshire, UK and Astimex Standards Ltd.; Toronto, ON, Canada). The measured intensities of the emitted X-rays were evaluated using the X-PHI model [29]. With the chosen method, the Li content cannot be analyzed quantitatively with the required precision. Therefore, the element concentration of Li was determined using virtual compounds according to T. Schirmer [30].
In order to validate the measurement method and obtain consistent results, repeated measurements were carried out on different days using the international standard rhodonite (MnSiO3; Astimex). The low standard deviation (0.03) of Mn indicates that this element content can be analyzed very precisely and is suitable for calculating the Li content via virtual compounds (Table 2).

3. Results

3.1. PXRD

In all three samples, wollastonite (CaSiO3; ICDD PDF2: 01-084-0654), as well as the Li-manganates LiMnO2 (ICDD PDF2: 00-035-0749) and Li2MnO3 (ICDD PDF2: 01-081-1953) could be detected with PXRD. Rankinite (Ca3Si2O7; ICDD PDF2: 01-076-0623) was detected in slags 1 and 2 and larnite (Ca2SiO4; ICDD PDF2: 00-029-0369) in slag sample 3. The presence of hausmannite (Mn2+Mn3+2O4; ICDD PDF2 00-024-0734) and the Li-silicate Li2SiO3 (ICDD PDF2: 00-029-0828) is conceivable. However, the presence of these two phases cannot be clearly verified with this method due to line overlaps. Schnickmann et al. [8] have already discovered that within the Li-manganate LiMnO2, the elements Li and Mn can replace each other, which can result in shifting to smaller or larger lattice parameters. The same applies to the replacement of Mn2+ in Ca-silicate matrices. An exchange of elements in the crystal lattice is only possible as long as the exchangeable elements have approximately the same ionic radii (e.g., [31]). An overview of the recorded diffractograms of the three slags is given in Figure 5.

3.2. EPMA

With EPMA, the following phases were determined:
  • Li(1−x)Mn(1+0.33x)O2 / Li(1+x)Mn(1−0.33x)O2 (slags 1, 2, 3)
  • Li(2−x)Mn(1+0.33x)O3 / Li(2+x)Mn(1−0.33x)O3 (slag 2)
  • Li2SiO3 (slags 1, 2)
  • LixMn(3−x)O4 (slags 1, 2, 3)
  • CaSiO3 (slags 1, 2, 3)
  • Ca3Si2O7 (slags 1, 2)
  • Ca2SiO4 (slag 3)
  • Residual melt (slags 2, 3)
These phases were identified based on the elemental content of the individual grains (point measurements, line scans) and the crystal shape. The crystal shape provides additional information regarding the processes during solidification (Figure 6).

3.2.1. Li-Manganates

Two different Li-manganates (Li2Mn4+O3 and LiMn3+O2) were found in the slag samples. Li2MnO3 was only detected via EPMA in slag 2 in a prepared, near-surface region of the slag where 2 wt.% Mg was added. This phase theoretically contains 47.03 wt.% Mn and 11.88 wt.% Li. Within this phase, the average measured Mn content was 47.08 wt.% (min.: 46.88 wt.%; max.: 47.28 wt.%) (Figure 7). In addition, this phase forms idiomorphic crystals between 30 and 100 µm. Based on the crystal shape, it can be determined that Li2MnO3 forms early during solidification. Point measurements and recorded line scans within this phase show that no other elements (e.g., Mg, Ca, Si) have been incorporated. According to the line scan results, the Mn content within the crystal is constant (Figure 8).
The crystallized LiMnO2 formed idiomorphic to partially hypidiomorphic crystals with an average size of 50 µm. In addition, elongated LiMnO2 needles crystallized in the Ca-silicate matrix. In theory, this phase contains 58.52 wt.% Mn and 7.39 wt.% Li. Within the slag, the average Mn content in S1 was 58.52 wt.% (min.: 58.25 wt.%; max.: 58.76 wt.%), in S2 it was 58.67 wt.% (min.: 58.35 wt.%; max.: 58.88 wt.%) and in S3 it was 57.88 wt.% (min.: 57.02 wt.%; max.: 58.17 wt.%) (Figure 8). Slag 3 contains 0.83 wt.% Al, which results in a lower Mn concentration (Table 3). The results of the line scans show that the Mn content within the individual grains is almost homogeneous. In addition, no decrease in element concentration is observed towards the grain boundaries (Figure 9).

3.2.2. Hausmannite and Spinel

Lithium-rich hausmannite was found in all three slag samples. Hausmannite consists of a mixture of Mn2+ and Mn3+ occupying tetrahedral and octahedral sites, respectively. When Mn2+ is substituted by Li1+, Mn4+ can appear in the octahedral sublattice to balance the charge without a change in oxygen stoichiometry. However, there is not enough crystallographic data for the site occupancies of delithiated hausmannite. Additionally, tetragonal Mn3O4 and orthorhombic LiMnO2 are intergrown. With an average crystal size of 10–30 µm, these crystals are significantly smaller than LiMnO2. In all three slag samples, the incorporation of other elements into the crystal lattice was detected. The incorporation of Li was recorded in all slag samples. The largest amount of Li was incorporated into S1, with 1.59 wt.%, followed by S2 with 0.80 wt.% and S3 with 0.51 wt.%. In addition, an average incorporation of 0.84 wt.% Mg and 0.82 wt.% Al was observed in S3. For the calculation of the spinel solid solution in S3, the following virtual compounds were defined: Mn3O4, MnAl2O4, MgAl2O4 and Li2Mn2O4. This solid solution can be expressed by the following generalized stoichiometric formula: (Li(2x),Mg(x),Mn(1−2x))1+x(Al(2−z),Mn3+(z))2O4. For Li-rich hausmannite in S1 and S2, the following generalized stoichiometric formula was applied: (Li(2x)Mn(1−x))1+x(Mn3+)2O4 (Table 4).

3.2.3. Li-Silicate

The Li-silicate Li2SiO3 was only found in S1 and S2. This phase forms 20 to 50 µm, rounded, elongated- to needle-shaped crystals. The hypidiomorphic to xenomorphic crystal shape indicates crystallization towards the end of solidification. The ideal crystals of this phase contain an average Si content of 31.22 wt.% and an average Li content of 15.43 wt.%. In S1, the average Si content is 31.66 wt.% and in S2 it is 31.68 wt.%. A higher Si content indicates a lower Li content.

3.2.4. Matrix-Forming Phases

The main matrix-forming mineral is wollastonite (CaSiO3), followed by rankinite (Ca3Si2O7) in S1 and S2 and larnite (Ca2SiO4) in S3. All three phases form xenomorphic crystals, which indicates crystallization at the end of solidification. Due to the approximately equal ionic radii of Mg2+, Ca2+ and Mn2+, Ca can be replaced by Mg and Mn in the crystal lattice [31]. In S1, 0.85 wt.% Mn was incorporated in wollastonite and 1.28 wt.% was incorporated in rankinite; in S2, there was 0.59 wt.% Mn in wollastonite and 2.22 wt.% in rankinite; in S3, there was 0.75 wt.% Mn in wollastonite and 6.97 wt.% Mn and 0.76 wt.% Mg in larnite (Table 5).

4. Discussion

The aim of this study was to investigate the stabilization of Mn4+ in the slag and to also verify which Li-rich phases are formed. Moreover, it should be investigated whether Mn4+ forms a compound with Mg or Al. Based on the phase diagrams, the experiment provides an overview of which phases crystallize in a synthetic slag consisting of Li2O, MgO, Al2O3, SiO2, CaO and MnOx.

4.1. Influence of Mg and Al on Mn4+

As can be seen from Figure 3e, the spinel phase in the MgO-Al2O3 system is the stable phase from low temperatures and has an extension towards both Al2O3 and MgO concentrations at high temperatures. The addition of MgO together with Al2O3 will increase the stability range of cubic spinel in the Li2O-MnOx-MgO-Al2O3 system.
The tetragonal-to-cubic transformations for the spinel phase in the MgO-MnOx and Al2O3-MnOx systems are driven by the Jahn–Teller distortion of octahedral sites, primarily occupied by Mn3+ ions. In the cubic phase, assuming a disproportionation of Mn3+ into Mn4+ and Mn2+ and the occupation the octahedral sites by Mn4+ and distribution of Mn2+ between tetrahedral and octahedral sites, a general formula for cubic spinel (Al3+, Mn2+)1(Al3+, Mn3+, Mn2+, Mn4+, Va)2(Mn2+, Va)2(O2−)4 is employed in the Al2O3-MnOx system and (Mg2+, Mn2+)1(Mg2+, Mn3+, Mn2+, Mn4+,Va)2(Mg2+, Mn2+, Va)2(O2−)4 in the MgO-MnOx system. This formulation accommodates the variation in oxidation states of Mn within the crystal lattice. The cubic spinel phase is further characterized by the inclusion of vacancies (Va) on the octahedral sites and Mn2+/Mg2+ in the interstitial site. The vacancies are introduced into the spinel model to extend the homogeneity ranges towards the Al2O3 side and interstitial site to model extension towards the MnOx/MgO side. The applied models account for deviations from a perfectly ordered structure to describe the disordering in cationic sites and to describe the homogeneity ranges of spinel, thus enhancing the model applicability in a large range of conditions (compositions, temperature, oxygen partial pressure). The modeling of tetragonal spinel differs from that for cubic spinel by introducing Mn3+ into the tetrahedral site and via the absence of Mn4+ in the octahedral site in both systems. The tetragonal phase is stable at lower temperatures and in a more narrow composition range than cubic spinel.
The octahedral sites are substantially occupied by Mn4+ in cubic spinel in the MgO-MnOx system in the whole possible composition range. A low concentration of Mn4+ is found in the spinel phase, with compositions close to stoichiometric MnAl2O4 in the Al2O3-MnOx system even at high oxygen partial pressures. Thus, the introduction of Mg into the spinel phase not only enhances the stability but also effectively increases the Mn4+ concentration. This suggests a more pronounced role for Mg in maintaining the desired Mn oxidation state within the spinel structure.
The Li addition introduces a notable impact on the Mn4+ ion concentration on the octahedral sites. Substitution of Mn2+ for Li1+ in the tetrahedral sites reduces the charge of cations in spinel. In order to preserve the electroneutrality of this structure, the fraction of Mn4+ in the octahedral sites occupied by Mn3+ in tetragonal spinel should be increased. The increase in Mn4+ in octahedral sites should also reduce the Jahn–Teller effect caused by the electronic structure of the Mn3+ cation. It should be noted that in the stoichiometric spinel LiMn2O4, the cation Li1+ completely occupies the tetrahedral site and the ratio of Mn4+/Mn3+ in octahedral site is equal to one. In cases where the Li content is higher than in stoichiometric spinel, Li1+ partially occupies octahedral site and ratio Mn4+/Mn3+ becomes larger than one. With a temperature increase, the spinel becomes enriched by Mn, with a shift of composition to the stoichiometric spinel LiMn2O4 and an even higher content of Mn. Therefore, spinel with a high Li content is stable at lower temperatures and with a temperature increase it will decompose to spinel with a lower Li content and a Li2MnO3 phase (Figure 2). At a temperature of 1230 K, LiMnO2 forms from spinel and Li2MnO3 and, at temperatures of 1250 K, Li2MnO3 decomposes into LiMnO2 and Li2O. Tetragonal spinel with a composition close to the stoichiometric one is stable in equilibrium with hausmannite with a much lower Li content from one side and from other side it is in equilibrium with LiMnO2. The high-temperature stability limit of tetragonal spinel is defined by decomposition to hausmannite and LiMnO2 at 1270 K.
According to the phase diagram shown in Figure 3b, no reactions should take place between MnO2 and Al2O3 up to 688 K. The experimental results indicate that Mg and Al react to form spinel (Mg2AlO4) and therefore small amounts of these elements are not a problem for Li-manganate (LiMnO2 or Li2MnO3) forming processes. As long as Mn2+ and Mn3+ are present in the slag, spinels with Mg, Al and Mn are formed, which can interfere with the formation of EnAMs. In order to produce large and pure EnAM crystals that can also be efficiently separated from the rest of the slag, it must be ensured that LiMnO2 or Li2MnO3 crystallize out of the melt first.

4.2. Li-Rich Phases as a New Potential EnAM

A total of four Li-containing phases crystallized out in the slag. These phases include the two Li-manganates LiMnO2 and Li2MnO3, as well as the Li-silicate Li2SiO3 and the Li-containing hausmannite. With an average of 0.51 wt.% (S3) to 1.59 wt.% Li (S1), the Li content in hausmannite is too low to be a potential EnAM. The Li1+ is incorporated at the tetrahedral site and replaces Mn2+ in the crystal lattice. The incorporation of Li1+ into cubic LiMn2O4 can occur by balancing Mn4+ on the octahedral site, but it can also occur by deviation from the stoichiometry of oxygen by introducing vacancies [32,33,34]. There is not enough crystallographic data for hausmannite showing how substitution of Mn2+ by Li1+ is compensated. Since it is not possible to detect small amounts of Mn4+ with EPMA, more advance techniques should be used to further investigate whether Mn4+ can be incorporated into the crystal lattice of hausmannite for charge compensation. Hausmannite forms in addition to relatively tiny crystals and only occurs in small amounts. Due to the partially inhomogeneous mixture of Mn2+ and Mn3+ in this phase, further processing would be necessary after successful separation from the slag. With an average of 15.43 wt.% Li, Li2SiO3 would be more suitable as a new EnAM. However, this phase crystallizes towards the end of solidification and thus forms relatively small and needle-shaped crystals that are inhibitors for a successful separation of Li. Furthermore, Li and Si must be separated from each other after the recycling step. Therefore, Li-manganates like LiMnO2 or Li2MnO3 would be more suitable as new potential EnAMs. The Li-manganate Li2MnO3, with 11.88 wt.% Li, could only be found in slag 2 where 2 wt.% Mg was added. No incorporation of other elements could be observed within this phase. As shown in Figure 2, the Li2MnO3 phase is stoichiometric, which corroborates with the absence of the solubility of other elements. Furthermore, the Li2MnO3 phase has some advantages as a new EnAM. Firstly, it is an early crystallite with large, idiomorphic crystals that contains more Li than LiAlO2 (10.35 wt.% Li). The stabilization of only Mn4+ in the slag has the additional advantage that spinel formation with Mn could be suppressed and that the Jahn–Teller effect does not cause deformation. In addition, LiMnO2 (7.39 wt.% Li) can also be considered as a potential EnAM. The low incorporation of Al (0.83 wt.%) does not seem to have any effect on the crystal lattice. According to Kong et al. [14], small amounts of Al in the Li-rich cathode material should promote intralayer diffusion.
Nevertheless, it remains to be determined which of the two Li manganates, LiMnO2 or Li2MnO3, formed first. It is possible that LiMnO2 forms first as a high-temperature phase and remains stable during cooling. At low temperatures, according to phase diagram, the stable phases are spinel and Li2MnO3. The alternative would be that Li2MnO3 forms first and decomposes to LiMnO2 and Li2O, which reacts with SiO2 forming Li2SiO3.

5. Conclusions and Outlook

The presented results indicate that it was possible to stabilize the Mn4+-containing Li-manganate Li2MnO3 in a synthetic slag, next to LiMnO2. Both Li-manganates, Li2MnO3 and LiMnO2, have proven to be extremely pure phases. Using EPMA for phase identification, LiMnO2 was found in all areas of the prepared slag, while Li2MnO3 was found in only one area, the near-surface area of the slag that was affected by oxygen. Accordingly, based on the literature data for other slag systems, it is assumed that the added oxygen only affects the near-surface region due to the low viscosity of the melt [35,36]. In addition, it needs to be clarified which elements can be added to the system to reduce the viscosity. For this purpose, future experiments should investigate whether, as described in the literature (e.g., [37]), the viscosity can be reduced by adding FeO and Fe2O3 so that the oxygen can affect the entire melt, with the aim of stabilizing Mn4+ throughout the slag. Stabilization of Mn4+ could provide further important experimental data on whether spinel formation can be prevented or whether Mn4+ from compounds with elements other than Li.
However, the conditions under which Li2MnO3 was formed and how Mn4+ was stabilized throughout the slag need to be clarified. A promising approach would be to investigate phase relations (experimentally and theoretically) and to develop thermodynamic databases for a complex system in order to understand the processes taking place during solidification. Further investigations should indicate which slag modification should be performed to maximize EnAM yield.

Author Contributions

Conceptualization, A.S. and D.A.D.A.; methodology, A.S. and D.A.D.A.; validation, A.S. and D.A.D.A.; formal analysis, A.S. and D.A.D.A.; investigation, A.S., D.A.D.A., O.F. and T.S.; data curation, A.S. and D.A.D.A.; writing—original draft preparation, A.S., D.A.D.A., O.F. and T.S.; writing—review and editing, A.S., D.A.D.A., O.F. and T.S.; visualization, A.S. and D.A.D.A.; supervision, O.F. and T.S.; project administration, O.F. and T.S.; funding acquisition, O.F. and T.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Open Access Publishing Fund of Clausthal University of Technology and the German Research Foundation as part of the priority program: Engineered Artificial Minerals (EnAM)—a geometallurgical tool for the recovery of critical elements from waste streams (SPP 2315, ProjNo. 470309740 and 470392360).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We acknowledge the support by Open Access Publishing Fund of Clausthal University of Technology.

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Brückner, L.; Frank, J.; Elwert, T. Industrial Recycling of Lithium-Ion Batteries—A Critical Review of Metallurgical Process Routes. Metals 2020, 10, 1107. [Google Scholar] [CrossRef]
  2. European Commission. Directorate General for Internal Market, Industry, Entrepreneurship and SMEs. In Study on the EU’s List of Critical Raw Materials (2020): Final Report; Publications Office, Luxembourg 2020. Available online: https://data.europa.eu/doi/10.2873/11619 (accessed on 20 December 2023).
  3. Bae, H.; Kim, Y. Technologies of lithium recycling from waste lithium ion batteries: A review. Mater. Adv. 2021, 2, 3234–3250. [Google Scholar] [CrossRef]
  4. Ma, X.; Azhari, L.; Wang, Y. Li-ion battery recycling challenges. Chem 2021, 7, 2843–2847. [Google Scholar] [CrossRef]
  5. U.S. Geological Survey. Data Release for Mineral Commodity Summaries 2023; U.S. Geological Survey: Reston, VA, USA, 2023.
  6. Baum, Z.J.; Bird, R.E.; Yu, X.; Ma, J. Lithium-Ion Battery Recycling—Overview of Techniques and Trends. ACS Energy Lett. 2022, 7, 712–719. [Google Scholar] [CrossRef]
  7. Sommerfeld, M.; Vonderstein, C.; Dertmann, C.; Klimko, J.; Oráč, D.; Miškufová, A.; Havlík, T.; Friedrich, A. Combined Pyro- and Hydrometallurgical Approach to Recycle Pyrolyzed Lithium-Ion Battery Black Mass Part 1: Production of Lithium Concentrates in an Electric Arc Furnace. Metals 2020, 10, 1069. [Google Scholar] [CrossRef]
  8. Schnickmann, A.; Hampel, S.; Schirmer, T.; Fittschen, U.E.A. Formation of Lithium-Manganates in a Complex Slag System Consisting of Li2O-MgO-Al2O3-SiO2-CaO-MnO—A First Survey. Metals 2023, 13, 2006. [Google Scholar] [CrossRef]
  9. Elwert, T.; Goldmann, D.; Schirmer, T.; Strauß, K. Recycling von Li-Ionen-Traktionsbatterien–Das Projekt LiBRi. Recycl. Rohst. 2012, 5, 679e690. [Google Scholar]
  10. Schirmer, T.; Qiu, H.; Li, H.; Goldmann, D.; Fischlschweiger, M. Li-Distribution in Compounds of the Li2O-MgO-Al2O3-SiO2-CaO System—A First Survey. Metals 2020, 10, 1633. [Google Scholar] [CrossRef]
  11. Pal, S.; Bandyopadhyay, A.K.; Mukherjee, S.; Samaddar, B.N.; Pal, P.G. MgAl2O4-γ-Al2O3 solid solution interaction: Mathematical framework and phase separation of α-Al2O3 at high temperature. Bull. Mater. Sci. 2011, 34, 859–864. [Google Scholar] [CrossRef]
  12. Chatterjee, S.; Jung, I.-H. Critical evaluation and thermodynamic modeling of the Al–Mn–O (Al2O3–MnO–Mn2O3) system. J. Eur. Ceram. Soc. 2014, 34, 1611–1621. [Google Scholar] [CrossRef]
  13. Ilatovskaia, M.; Fabrichnaya, O. Critical assessment and thermodynamic modeling of the Al-Mn-O system. J. Alloys Compd. 2021, 884, 161153. [Google Scholar] [CrossRef]
  14. Kong, F.; Longo, R.C.; Park, M.-S.; Yoon, J.; Yeon, D.-H.; Park, J.-H.; Wang, W.-H.; Kc, S.; Doo, S.-G.; Cho, K. Ab initio study of doping effects on LiMnO2 and Li2MnO3 cathode materials for Li-ion batteries. J. Mater. Chem. A 2015, 3, 8489–8500. [Google Scholar] [CrossRef]
  15. Paulsen, J.M.; Dahn, J.R. Phase Diagram of Li−Mn−O Spinel in Air. Chem. Mater. 1999, 11, 3065–3079. [Google Scholar] [CrossRef]
  16. Longo, R.C.; Kong, F.T.; Kc, S.; Park, M.S.; Yoon, J.; Yeon, D.-H.; Park, J.-H.; Doo, S.-G.; Cho, K. Phase stability of Li–Mn–O oxides as cathode materials for Li-ion batteries: Insights from ab initio calculations. Phys. Chem. Chem. Phys. 2014, 16, 11233–11242. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, W.; Cupid, D.M.; Gotcu, P.; Chang, K.; Li, D.; Du, Y.; Seifert, H.J. High-Throughput Description of Infinite Composition–Structure–Property–Performance Relationships of Lithium–Manganese Oxide Spinel Cathodes. Chem. Mater. 2018, 30, 2287–2298. [Google Scholar] [CrossRef]
  18. Müller, H.A.; Joshi, Y.; Hadjixenophontos, E.; Peter, C.; Csiszár, G.; Richter, G.; Schmitz, G. High capacity rock salt type Li2MnO3−δ thin film battery electrodes. RSC Adv. 2020, 10, 3636–3645. [Google Scholar] [CrossRef] [PubMed]
  19. Mishra, G.; Jha, R.; Meshram, A.; Singh, K.K. A Review on Recycling of Lithium-Ion Batteries to Recover Critical Metals. J. Environ. Chem. Eng. 2022, 10, 108534. [Google Scholar] [CrossRef]
  20. Konar, B.; Van Ende, M.-A.; Jung, I.-H. Critical Evaluation and Thermodynamic Optimization of the Li2O-Al2O3 and Li2O-MgO-Al2O3 Systems. Metall. Mater. Trans. B 2018, 49, 2917–2944. [Google Scholar] [CrossRef]
  21. De Abreu, D.A.; Löffler, M.; Kriegel, M.J.; Fabrichnaya, O. Experimental investigation and thermodynamic modeling of the Li2O-Al2O3 system. J. Phase Equilibria Diffus. 2024, 45, 36–55. [Google Scholar] [CrossRef]
  22. Dilner, D.; Pavlyuchkov, D.; Zienert, T.; Kjellqvist, L.; Fabrichnaya, O. Thermodynamics of the Mg-Mn-O system—Modeling and heat capacity measurements. J. Am. Ceram. Soc. 2017, 100, 1661–1672. [Google Scholar] [CrossRef]
  23. Zienert, T.; Fabrichnaya, O. Thermodynamic assessment and experiments in the system MgO–Al2O3. Calphad 2013, 40, 1–9. [Google Scholar] [CrossRef]
  24. Blank, D.H.A.; Kruidhof, H.; Flokstra, J. Preparation of YBa2Cu3O7−δ by citrate synthesis and pyrolysis. J. Phys. D Appl. Phys. 1988, 21, 226–227. [Google Scholar] [CrossRef]
  25. Ehi-Eromosele, C.O.; Ajayi, S.O.; Onwucha, C.N. Influence of Fuels in the Sol-Gel Combustion Synthesis of Li2MnO3 Positive Electrode Material for Li-Ion Battery. Mater. Chem. Phys. 2021, 259, 124055. [Google Scholar] [CrossRef]
  26. Downs, R.T.; Hall-Wallace, M. The American Mineralogist crystal structure database. Am. Mineral. 2003, 88, 247–250. [Google Scholar]
  27. Gates-Rector, S.; Blanton, T. The Powder Diffraction File: A quality materials characterization database. Powder Diffr. 2019, 34, 352–360. [Google Scholar] [CrossRef]
  28. Jercinovic, M.J.; Williams, M.L.; Allaz, J.; Donovan, J.J. Trace analysis in EPMA. IOP Conf. Ser. Mater. Sci. Eng. 2012, 32, 012012. [Google Scholar] [CrossRef]
  29. Merlet, C. Quantitative Electron Probe Microanalysis: New Accurate Φ(pz) Description. In Electron Microbeam Analysis; Springer: Berlin/Heidelberg, Germany, 1992; pp. 107–115. [Google Scholar]
  30. Schirmer, T. Stoichiometric Calculation of Lithium-Containing Phases Based on Spatially Resolved X-ray Analysis and Virtual Compounds. Adv. X-ray Anal. 2022, 65. [Google Scholar]
  31. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  32. Tarascon, J.M.; McKinnon, W.R.; Coowar, F.; Bowmer, T.N.; Amatucci, G.; Guyomard, D. Synthesis Conditions and Oxygen Stoichiometry Effects on Li Insertion into the Spinel LiMn2O4. J. Electrochem. Soc. 1994, 141, 1421–1431. [Google Scholar] [CrossRef]
  33. Yu, F.-D.; Wang, Z.-B.; Chen, F.; Wu, J.; Zhang, X.-G.; Gu, D.-M. Crystal structure and multicomponent effects in Li1+xMn2−x−yAlyO4 cathode materials for Li-ion batteries. J. Power Sources 2014, 262, 104–111. [Google Scholar] [CrossRef]
  34. Luo, C.; Martin, M. Stability and defect structure of spinels Li1+xMn2−xO4−δ: I. In situ investigations on the stability field of the spinel phase. J. Mater. Sci. 2007, 42, 1955–1964. [Google Scholar] [CrossRef]
  35. Ilyushechkin, A.; Kondratiev, A. Viscosity of Slags with Solids: The Effect of Solids Morphology and Concentration. J. Rheol. 2019, 63, 719–733. [Google Scholar] [CrossRef]
  36. Park, J.H.; Todoroki, H. Control of MgO·Al2O3 Spinel Inclusions in Stainless Steels. ISIJ Int. 2010, 50, 1333–1346. [Google Scholar] [CrossRef]
  37. Nagata, K.; Sasabe, M. Permeability Coefficient of Oxygen in Molten Slags and Its Mechanisms. ISIJ Int. 2020, 60, 1872–1877. [Google Scholar] [CrossRef]
Figure 1. Sketch of the EnAM strategy for efficient Li recovery. The aim is to form a Li-manganate as an early crystallizate from the liquid melt. The other elements (Mg, Al, Si, Ca) should form matrix-forming phases. This early crystallizate (Li-Manganate) should be efficiently separated from the matrix.
Figure 1. Sketch of the EnAM strategy for efficient Li recovery. The aim is to form a Li-manganate as an early crystallizate from the liquid melt. The other elements (Mg, Al, Si, Ca) should form matrix-forming phases. This early crystallizate (Li-Manganate) should be efficiently separated from the matrix.
Minerals 14 00368 g001
Figure 2. Adapted phase diagram of the Li2O-MnOx system in air based on experimental data obtained from Paulsen and Dahn [15].
Figure 2. Adapted phase diagram of the Li2O-MnOx system in air based on experimental data obtained from Paulsen and Dahn [15].
Minerals 14 00368 g002
Figure 3. Calculated phase diagrams of the: (a) Li2O-Al2O3 system, (b) Al2O3-MnOx system in air; (c) Al2O3-MnOx system in the presence of metallic Mn; (d) MgO-MnOx system in air and (e) Al2O3-MgO system with the thermodynamic parameters optimized from [13,21,22,23].
Figure 3. Calculated phase diagrams of the: (a) Li2O-Al2O3 system, (b) Al2O3-MnOx system in air; (c) Al2O3-MnOx system in the presence of metallic Mn; (d) MgO-MnOx system in air and (e) Al2O3-MgO system with the thermodynamic parameters optimized from [13,21,22,23].
Minerals 14 00368 g003
Figure 4. The heating program used for the experiments. Slow heating of the precursor (2.90 °C/min; 1.49 °C/min) should minimize the loss of Li. By cooling down more slowly, the phases should have enough time to form crystals that are as large as possible. During the entire melting experiment, the oxygen supply was 200 L/h.
Figure 4. The heating program used for the experiments. Slow heating of the precursor (2.90 °C/min; 1.49 °C/min) should minimize the loss of Li. By cooling down more slowly, the phases should have enough time to form crystals that are as large as possible. During the entire melting experiment, the oxygen supply was 200 L/h.
Minerals 14 00368 g004
Figure 5. (a) Recorded PXRD pattern of the three slag samples. The identified main phases wollastonite (W), rankinite (R), larnite (L), LiMnO2 (*) and Li2MnO3 (**) were marked. (b) Identified LiMnO2 reflex in the slag samples (S1–S3). (c) Identified Li2MnO3 reflex in the slag samples (S1–S3).
Figure 5. (a) Recorded PXRD pattern of the three slag samples. The identified main phases wollastonite (W), rankinite (R), larnite (L), LiMnO2 (*) and Li2MnO3 (**) were marked. (b) Identified LiMnO2 reflex in the slag samples (S1–S3). (c) Identified Li2MnO3 reflex in the slag samples (S1–S3).
Minerals 14 00368 g005
Figure 6. Recorded BSE(Z) micrographs of the three slags. In light gray: hausmannite (Mn3O4); medium gray: Li-manganate (LiMnO2); dark gray: wollastonite (CaSiO3) and rankinite (Ca3Si2O7)/larnite (Ca2SiO4). (a) S1; (b) S2; (c) S3.
Figure 6. Recorded BSE(Z) micrographs of the three slags. In light gray: hausmannite (Mn3O4); medium gray: Li-manganate (LiMnO2); dark gray: wollastonite (CaSiO3) and rankinite (Ca3Si2O7)/larnite (Ca2SiO4). (a) S1; (b) S2; (c) S3.
Minerals 14 00368 g006
Figure 7. Overview of the measured Mn concentrations within the Li-manganates in the three slags. The dotted lines indicate the Mn concentrations of the pure stoichiometric compounds. In all three slag samples, the phase LiMnO2 was found; Li2MnO3 was only found in slag 2 (2.1). The results shown are based on point measurements in different crystals of the respective phase. Measurements on the reference material rhodonite demonstrate that the Mn content can be measured with an accuracy of ±0.03%. X marks the mean value.
Figure 7. Overview of the measured Mn concentrations within the Li-manganates in the three slags. The dotted lines indicate the Mn concentrations of the pure stoichiometric compounds. In all three slag samples, the phase LiMnO2 was found; Li2MnO3 was only found in slag 2 (2.1). The results shown are based on point measurements in different crystals of the respective phase. Measurements on the reference material rhodonite demonstrate that the Mn content can be measured with an accuracy of ±0.03%. X marks the mean value.
Minerals 14 00368 g007
Figure 8. Elemental composition of the matrix (CaSiO3) and the two Li-manganates (Li2MnO3) and (LiMnO2) in slag 2. (a) The line scan results show that minimal amounts of Mn2+ were incorporated into the matrix. Furthermore, there is no decrease or increase in the element distribution towards the grain boundaries. (b) BSE(Z) image showing the path of the line scan (A: start, B: end, step size: 5.3 µm).
Figure 8. Elemental composition of the matrix (CaSiO3) and the two Li-manganates (Li2MnO3) and (LiMnO2) in slag 2. (a) The line scan results show that minimal amounts of Mn2+ were incorporated into the matrix. Furthermore, there is no decrease or increase in the element distribution towards the grain boundaries. (b) BSE(Z) image showing the path of the line scan (A: start, B: end, step size: 5.3 µm).
Minerals 14 00368 g008
Figure 9. Elemental composition of the matrix (CaSiO3) and Li-manganate (III) (LiMnO2) in slag 2. (a) The line scan results show that minimal amounts of Mn2+ were incorporated into the matrix. Furthermore, there is no decrease or increase in the element distribution towards the grain boundaries. (b) BSE(Z) image showing the path of the line scan (A: start, B: end, step size: 4.4 µm).
Figure 9. Elemental composition of the matrix (CaSiO3) and Li-manganate (III) (LiMnO2) in slag 2. (a) The line scan results show that minimal amounts of Mn2+ were incorporated into the matrix. Furthermore, there is no decrease or increase in the element distribution towards the grain boundaries. (b) BSE(Z) image showing the path of the line scan (A: start, B: end, step size: 4.4 µm).
Minerals 14 00368 g009
Table 1. Elemental composition of the precursors in wt.% and at.%. The precursors have the same composition as [8].
Table 1. Elemental composition of the precursors in wt.% and at.%. The precursors have the same composition as [8].
Precursor 1Precursor 2Precursor 3
Elementwt.%at.%wt.%at.%wt.%at.%
Li1.96.31.96.31.85.8
Mg0.00.01.21.11.21.1
Al0.00.00.00.01.10.9
Si19.315.919.115.618.915.4
Ca29.917.229.416.929.016.7
Mn9.84.19.23.98.63.7
O39.156.439.256.239.456.4
Total100.0100.0100.0100.0100.0100.0
Table 2. The CRM rhodonite was used to verify the measurement accuracy of Mn. The Mn content can be determined with an inaccuracy of ±0.03%. Indication in wt.%. * The iron content was not measured.
Table 2. The CRM rhodonite was used to verify the measurement accuracy of Mn. The Mn content can be determined with an inaccuracy of ±0.03%. Indication in wt.%. * The iron content was not measured.
Wt.%Average Rhodonite%StDev. RhodoniteRef. Rhodonite
MgO1.920.012.0
Al2O30.0110.002n.a.
SiO246.760.0646.8
CaO4.740.044.6
MnO42.300.0342.3
FeO *4.34n.a.4.3
Table 3. For the two Li-manganates LiMnO2 and Li2MnO3, the average structural formula was calculated from all measuring points. To clarify the possible exchange of Li and Mn ions in the crystal lattice, a structural formula was also calculated for the lowest and highest Mn concentration according to: Li(1−x)Mn(1+0.33x)O2/Li(1+x)Mn(1−0.33x)O2 and Li(2−x)Mn(1+0.33x)O3 / Li(2+x)Mn(1−0.33x)O3.
Table 3. For the two Li-manganates LiMnO2 and Li2MnO3, the average structural formula was calculated from all measuring points. To clarify the possible exchange of Li and Mn ions in the crystal lattice, a structural formula was also calculated for the lowest and highest Mn concentration according to: Li(1−x)Mn(1+0.33x)O2/Li(1+x)Mn(1−0.33x)O2 and Li(2−x)Mn(1+0.33x)O3 / Li(2+x)Mn(1−0.33x)O3.
SampleLiMnO2 (Average)LiMnO2 (min.)LiMnO2 (max.)
S1 (n = 25)Li1.01+Mn3+1.0O2Li1+1.01Mn3+0.99O2Li1+0.98Mn3+1.01O2
S2 (n = 25)Li1+1.0Mn3+1.0O2Li1+1.01Mn3+1.00O2Li1+0.97Mn3+1.01O2
S3 (n = 25)Li1+1.0(Mn3+0.97Al3+0.03)O2Li1+1.02(Mn3+0.96Al3+0.03)O2Li1+0.96(Mn3+0.99Al3+0.02)O2
SampleLi2MnO3 (average)Li2MnO3 (min.)Li2MnO3 (max.)
S2 (n = 25)Li1+2.0Mn3+1.00O3Li1+2.02Mn3+0.99O3Li1+1.97Mn3+1.01O2
Table 4. Determined structural formula of lithiated hausmannite in slags 1,2 and 3.
Table 4. Determined structural formula of lithiated hausmannite in slags 1,2 and 3.
SampleMn2+Mn3+2O4
S1 (n = 12)(Li1+0.50Mn2+0.75)Mn3+2.0O4
S2 (n = 16)(Li1+0.26Mn2+0.87)Mn3+2.0O4
S3 (n = 16)(Li1+0.16Mn2+0.84Mg2+0.08)(Mn3+1.93Al3+0.07)O4
Table 5. Calculated structural formulas for the matrix-forming phases wollastonite (CaSiO3) and rankinite (Ca3Si2O7) in S1 and S2 and of larnite (Ca2SiO4) in S3.
Table 5. Calculated structural formulas for the matrix-forming phases wollastonite (CaSiO3) and rankinite (Ca3Si2O7) in S1 and S2 and of larnite (Ca2SiO4) in S3.
SampleCaSiO3
S1 (n = 40)(Ca2+0.99Mn2+0.02)Si4+0.99O3
S2 (n = 35)(Ca2+0.99Mn2+0.01)Si4+1.0O3
S3 (n = 30)(Ca2+0.99Mn2+0.02)Si4+1.0O3
SampleCa3Si2O7
S1 (n = 40)(Ca2+2.96Mn2+0.07)Si4+1.99O7
S2 (n = 40)(Ca2+2.86Mn2+0.12Mg2+0.02)Si4+2.0O7
SampleCa2SiO4
S3 (n = 28)(Li1+0.33Ca2+1.63Mn2+0.14Mg2+0.05)Si4+0.99O4
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Schnickmann, A.; De Abreu, D.A.; Fabrichnaya, O.; Schirmer, T. Stabilization of Mn4+ in Synthetic Slags and Identification of Important Slag Forming Phases. Minerals 2024, 14, 368. https://doi.org/10.3390/min14040368

AMA Style

Schnickmann A, De Abreu DA, Fabrichnaya O, Schirmer T. Stabilization of Mn4+ in Synthetic Slags and Identification of Important Slag Forming Phases. Minerals. 2024; 14(4):368. https://doi.org/10.3390/min14040368

Chicago/Turabian Style

Schnickmann, Alena, Danilo Alencar De Abreu, Olga Fabrichnaya, and Thomas Schirmer. 2024. "Stabilization of Mn4+ in Synthetic Slags and Identification of Important Slag Forming Phases" Minerals 14, no. 4: 368. https://doi.org/10.3390/min14040368

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop