Publication Cover
Australian Journal of Earth Sciences
An International Geoscience Journal of the Geological Society of Australia
Latest Articles
289
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Systematic trends in the substitution mechanisms of minor elements in cassiterite

ORCID Icon, ORCID Icon, , ORCID Icon & ORCID Icon
Received 05 Jun 2023, Accepted 31 Jan 2024, Published online: 18 Apr 2024

Figures & data

Table 1. A selected summary of minimum Sn contents, and maximum concentrations of substituting elements measured by EPMA, as reported in the literature from different localities. The number of decimal places reflect the reported precision from the original source. All values converted to apfu*100, based on 2 atoms of oxygen per 1 cation. Not all values belong to the same analytical point, or even crystal, and totals will not equal 100. <dl = below the limit of detection, - = not analysed.

Figure 1. A schematic illustration of the physicochemical model of cassiterite mineral systems employed in this study. Silicate Liquid Systems, where cassiterite precipitates from a silicate liquid, include (a) pegmatite deposits, and (b) granite-hosted Disseminated Sn deposits. Hybrid Systems, where cassiterite may precipitate from both silicate liquids and aqueous fluids, are represented by (c) greisen deposits. Aqueous Fluid Systems, where cassiterite is precipitated from hydrothermal fluids only, are represented by (d) vein-hosted systems and (e) Sn-bearing skarn systems. In this model, the key distinction between vein-hosted and skarn systems is the degree of wall-rock alteration (metasomatism), with vein-hosted systems representing a low alteration endmember, and skarn systems representing a high alteration endmember. The classification of cassiterite crystals examined in this study is also listed in this figure.

Figure 1. A schematic illustration of the physicochemical model of cassiterite mineral systems employed in this study. Silicate Liquid Systems, where cassiterite precipitates from a silicate liquid, include (a) pegmatite deposits, and (b) granite-hosted Disseminated Sn deposits. Hybrid Systems, where cassiterite may precipitate from both silicate liquids and aqueous fluids, are represented by (c) greisen deposits. Aqueous Fluid Systems, where cassiterite is precipitated from hydrothermal fluids only, are represented by (d) vein-hosted systems and (e) Sn-bearing skarn systems. In this model, the key distinction between vein-hosted and skarn systems is the degree of wall-rock alteration (metasomatism), with vein-hosted systems representing a low alteration endmember, and skarn systems representing a high alteration endmember. The classification of cassiterite crystals examined in this study is also listed in this figure.

Table 2. Summary of the cassiterite samples and their localities used in this study, with approximate coordinates (accurate to within ±1 of the decimal place given). Each locality is classified into one of the simplified cassiterite mineral systems as outlined in . Any coexisting mineral phases noted by previous authors are shown along with any classifications of each deposit type, for comparison (Paragenesis), taken from the key references for each locality. The nature of the samples used in this study, and any other notable characteristics, are described in Sample Observations and Characteristics.

Table 3. EPMA acquisition routine for the cassiterite crystals analysed in this study.

Table 4. Oxygen fugacity reactions considered in this study. Buffer curves calculated from these equations are displayed in . The typical buffer series of Magnetite–Hematite (MH), Nickel–Nickel Oxide (NNO), Fayalite–Magnetite–Quartz (FMQ), Wüstite–Magnetite (WM), Iron–Wüstite (IW), and Quartz–Iron–Fayalite (QIF) were also calculated and displayed for comparison.

Figure 2. Intracrystalline and intercrystalline variability of substitutional elements in the cassiterite crystals analysed in this study, presented as violin plots in at% (bottom x-axis) and apfu*100 (top x-axis). See text for discussion. One-sigma uncertainty (σ) is numerically labelled on the bottom x-axis for each element, and the two-sigma uncertainty (2σ) is shown graphically as an error bar inset in each plot. (a) Variability in total Sn concentration. The total number of analyses (n), mean (orange circle) and median (vertical blue line) values are highlighted in the key. The dashed black line represents the stoichiometric ideal. (b) Variability in Ti. The lower limit of detection (LLD), the number of analyses below this threshold (n < LLD) and the number above (n > LLD) are highlighted in the key along with the mean and median and are used throughout the rest of the figure. (c) Variability in Zr. (d) Variability in W.

Figure 2. Intracrystalline and intercrystalline variability of substitutional elements in the cassiterite crystals analysed in this study, presented as violin plots in at% (bottom x-axis) and apfu*100 (top x-axis). See text for discussion. One-sigma uncertainty (σ) is numerically labelled on the bottom x-axis for each element, and the two-sigma uncertainty (2σ) is shown graphically as an error bar inset in each plot. (a) Variability in total Sn concentration. The total number of analyses (n), mean (orange circle) and median (vertical blue line) values are highlighted in the key. The dashed black line represents the stoichiometric ideal. (b) Variability in Ti. The lower limit of detection (LLD), the number of analyses below this threshold (n < LLD) and the number above (n > LLD) are highlighted in the key along with the mean and median and are used throughout the rest of the figure. (c) Variability in Zr. (d) Variability in W.

Figure 3. Intracrystalline and intercrystalline variability of substitutional elements in the cassiterite crystals analysed in this study, presented as violin plots in at% (bottom x-axis) and apfu*100 (top x-axis). Continued from . (a) Variability in Fe. (b) Variability in Nb. (c) Variability in Mn. (d) Variability in Ta.

Figure 3. Intracrystalline and intercrystalline variability of substitutional elements in the cassiterite crystals analysed in this study, presented as violin plots in at% (bottom x-axis) and apfu*100 (top x-axis). Continued from Figure 2. (a) Variability in Fe. (b) Variability in Nb. (c) Variability in Mn. (d) Variability in Ta.

Figure 4. Oxygen fugacity buffer curves for the standard set (QIF to MH, left) with the reactions of interest to this study (right). The grey area represents the expected fO2 of most cassiterite mineral systems (Bhalla et al., Citation2005; Ishihara, Citation1979; Linnen et al., Citation1995) and an expected upper T range of primary Sn-mineralising fluids in these systems (Heinrich, Citation1990).

Figure 4. Oxygen fugacity buffer curves for the standard set (QIF to MH, left) with the reactions of interest to this study (right). The grey area represents the expected fO2 of most cassiterite mineral systems (Bhalla et al., Citation2005; Ishihara, Citation1979; Linnen et al., Citation1995) and an expected upper T range of primary Sn-mineralising fluids in these systems (Heinrich, Citation1990).

Figure 5. Oxygen fugacity speciation curves for oxidation states of interest, calculated at 800 K. The overall geometry of these curves is similar at lower temperatures. The grey area shows the projected range in fO2 from . The mid-points (equal activities) of the QIF, WM, QFM and MH buffers are shown as dashed vertical lines, using the same colour scheme as . (a) The magenta line displays the total proportion of Sn4+ as a function of fO2 for the Sn:SnO2 reaction, the black to grey lines display the total proportion of Fe species. (b) The dark pink line displays the total proportion of Ta5+ for the Ta:Ta2O5 equation, the light pink line the proportion of Nb5+ for the NbO2:Nb2O5 equation. The blue line is the total proportion of W6+ for the WO2:WO3 equation, and black line is the total proportion of Ti4+ over the proportion of Ti3+. (c) The black to grey lines display the changing speciation of Mn species.

Figure 5. Oxygen fugacity speciation curves for oxidation states of interest, calculated at 800 K. The overall geometry of these curves is similar at lower temperatures. The grey area shows the projected range in fO2 from Figure 4. The mid-points (equal activities) of the QIF, WM, QFM and MH buffers are shown as dashed vertical lines, using the same colour scheme as Figure 4. (a) The magenta line displays the total proportion of Sn4+ as a function of fO2 for the Sn:SnO2 reaction, the black to grey lines display the total proportion of Fe species. (b) The dark pink line displays the total proportion of Ta5+ for the Ta:Ta2O5 equation, the light pink line the proportion of Nb5+ for the NbO2:Nb2O5 equation. The blue line is the total proportion of W6+ for the WO2:WO3 equation, and black line is the total proportion of Ti4+ over the proportion of Ti3+. (c) The black to grey lines display the changing speciation of Mn species.

Figure 6. Lattice strain model for the possible substitution mechanisms of minor elements in cassiterite. (a) Predicted partition coefficients for ions only, including strain due to charge mismatch. (b) Partition coefficients using stoichiometrically averaged ionic radii (i.e. the ionic radius for the substitution of 2Nb5+Fe2+ is modelled as [2rNb5++rFe2+]/3). Charge is balanced in this instance, so strain only results from the combined radii mismatch. Note that the combination of larger and smaller cations results in an overall effective ionic radius similar to Sn4+.

Figure 6. Lattice strain model for the possible substitution mechanisms of minor elements in cassiterite. (a) Predicted partition coefficients for ions only, including strain due to charge mismatch. (b) Partition coefficients using stoichiometrically averaged ionic radii (i.e. the ionic radius for the substitution of 2Nb5+Fe2+ is modelled as [2rNb5++rFe2+]/3). Charge is balanced in this instance, so strain only results from the combined radii mismatch. Note that the combination of larger and smaller cations results in an overall effective ionic radius similar to Sn4+.

Table 5. Possible substitution mechanisms for minor element incorporation into cassiterite. The different mechanisms are grouped into classes defined by the highest valency in the substitution reaction. The stoichiometric coupling reflects the ratio between cations in coupled substitutional pairs. The equations show the full substitution replacement reactions with Sn4+ into the cassiterite lattice. Each reaction has an assigned number, which will be used throughout the text. The ionic radius in six-fold (octahedral) coordination (Shannon & Prewitt, Citation1969) is shown for each cation (in Angstroms, Å), along with the radius for Sn4+ for comparison. Where the reaction contains multiple cations, the ionic radius shown is the stoichiometrically averaged radius for that reaction. An asterisk (*) denotes substitution reactions disregarded or not considered in previous literature.

Table 6. Existing evidence for the substitution mechanisms outlined in . Each cell contains a tick (✓) if the evidence is in favour of the substitution mechanism, a cross (×) if the evidence does not support this mechanism, and a question mark (?) if the evidence is ambiguous or uncertain. The analytical methods are also described in each cell, as one of stoichiometric analysis via electron probe microanalysis (EPMA), direct measurement of oxidation states via electron paramagnetic resonance (EPR) or Mössbauer Spectroscopy (Möss), or consideration of the local electronic environment via cathodoluminescence theory (CL). Direct measurement of OH content by Fourier Transform Infrared spectroscopy (FTIR) is also included.

Table 7. Possible end-member phases for (limited) solid solution series with cassiterite, corresponding to the substitution mechanisms in (listed under Eq.#). The class (Tet., Tetragonal; Orth., Orthorhombic; Mon., Monoclinic; Trig., Trigonal), symmetry (space group) and unit cell parameters (lengths in Å, angles in degrees) for each phase are also listed. Data sourced from mindat.org (accessed 5th April 2020), except where referenced. Phases synthesised in a laboratory, and not known in nature, are labelled as synthetic.

Figure 7. Binary plot of the substitutional relationship between Fe + Mn and W. The grey trend lines represent stoichiometries of 1:2 for Equations 22 and 23, 1:1 for Equations 20 and 21, and 2:1 for Equations 18 and 19. Points plotting in regions defined between these lines require a component of each bounding substitutional stoichiometry. Samples are grouped according to (a) pegmatitic and granite-hosted systems, (b) greisen systems, (c) vein-hosted systems, and (d) skarn systems. Note that the horizontal trend along the x-axis observed for Sifleetes Reward (in A) is due to the substitution of Fe coupled with Nb and Ta.

Figure 7. Binary plot of the substitutional relationship between Fe + Mn and W. The grey trend lines represent stoichiometries of 1:2 for Equations 22 and 23, 1:1 for Equations 20 and 21, and 2:1 for Equations 18 and 19. Points plotting in regions defined between these lines require a component of each bounding substitutional stoichiometry. Samples are grouped according to (a) pegmatitic and granite-hosted systems, (b) greisen systems, (c) vein-hosted systems, and (d) skarn systems. Note that the horizontal trend along the x-axis observed for Sifleetes Reward (in A) is due to the substitution of Fe coupled with Nb and Ta.

Figure 8. Monte Carlo simulation of a pure cassiterite EDS spectrum (black line), compared with simulated spectra of cassiterite doped with 1 apfu*100 of either Si (green line), Ta (red line) or W (blue line). The location of the Si Kα peak is indistinguishable from the combined X-ray emissions of the Ta Mα and Mβ lines. Overlap from the W Mαβ lines may also cause analytical difficulties in this region.

Figure 8. Monte Carlo simulation of a pure cassiterite EDS spectrum (black line), compared with simulated spectra of cassiterite doped with 1 apfu*100 of either Si (green line), Ta (red line) or W (blue line). The location of the Si Kα peak is indistinguishable from the combined X-ray emissions of the Ta Mα and Mβ lines. Overlap from the W Mαβ lines may also cause analytical difficulties in this region.

Figure 9. Net charge versus W for a cassiterite crystal from Saltwater Creek. The net charge is calculated as the sum of each component (in atomic percent) multiplied by their assumed charges (Nb5+, Ta5+, Sn4+, Ti4+, Zr4+, Fe2+, Mn2+), and assuming a stoichiometric concentration of 66.6̅ at% O. A linear correlation between charge and W concentration is observed. Charge is overbalanced (excess positive charge) if all W were assumed to be W6+, and underbalanced (excess negative charge) if all W were assumed to be W4+. Charge is only balanced if 25% of total W is W4+.

Figure 9. Net charge versus W for a cassiterite crystal from Saltwater Creek. The net charge is calculated as the sum of each component (in atomic percent) multiplied by their assumed charges (Nb5+, Ta5+, Sn4+, Ti4+, Zr4+, Fe2+, Mn2+), and assuming a stoichiometric concentration of 66.6̅ at% O. A linear correlation between charge and W concentration is observed. Charge is overbalanced (excess positive charge) if all W were assumed to be W6+, and underbalanced (excess negative charge) if all W were assumed to be W4+. Charge is only balanced if 25% of total W is W4+.

Figure 10. Binary plots of the substitutional relationships between Fe + Mn and Nb + Ta. The grey trend lines represent slopes of 1:1 for the stoichiometric substitutions of Equations 10–13 and 2:1 for Equations 14–17. Substitution via Equations 1 and 2 will plot along the x-axis, likewise, substitution via Equations 7 and 8 will plot along the y-axis. Points plotting in regions defined between these lines require a component of each bounding substitutional stoichiometry. Samples are grouped according to (a) pegmatitic and granite-hosted systems, (b) greisen systems, (c) vein-hosted systems, and (d) skarn systems.

Figure 10. Binary plots of the substitutional relationships between Fe + Mn and Nb + Ta. The grey trend lines represent slopes of 1:1 for the stoichiometric substitutions of Equations 10–13 and 2:1 for Equations 14–17. Substitution via Equations 1 and 2 will plot along the x-axis, likewise, substitution via Equations 7 and 8 will plot along the y-axis. Points plotting in regions defined between these lines require a component of each bounding substitutional stoichiometry. Samples are grouped according to (a) pegmatitic and granite-hosted systems, (b) greisen systems, (c) vein-hosted systems, and (d) skarn systems.

Figure 11. Projection of cassiterite samples onto the columbite–tantalite quadrilateral, showing Mn# vs Ta#. The tapiolite–tantalite miscibility gap of Černý et al. (Citation1992) is included for reference, with the dotted field representing the range of compositions between coexisting pairs of tantalite and tapiolite, and the grey dashed lines representing the compositional bounds for single phases. (a) Cassiterite from silicate liquid systems. (b) Cassiterite from hybrid and aqueous fluid systems.

Figure 11. Projection of cassiterite samples onto the columbite–tantalite quadrilateral, showing Mn# vs Ta#. The tapiolite–tantalite miscibility gap of Černý et al. (Citation1992) is included for reference, with the dotted field representing the range of compositions between coexisting pairs of tantalite and tapiolite, and the grey dashed lines representing the compositional bounds for single phases. (a) Cassiterite from silicate liquid systems. (b) Cassiterite from hybrid and aqueous fluid systems.

Figure 12. Ternary plots showing stoichiometric relationships between W, Fe + Mn, and Nb + Ta. These ternaries allow for comparison of W–Fe relationships in Nb + Ta bearing samples. The trend lines shown in project along the Fe + Mn and Nb + Ta join, and those shown in project along the Fe + Mn, and W join. Potential tie lines joining these substitutional stoichiometries are shown in grey. Transparency of the points decreases away from W concentrations at 2σ of the LLD, so that samples with the highest W contents are darker (more solid). Samples are grouped according to (a) pegmatitic and granite-hosted systems, (b) greisen systems, (c) vein-hosted systems, and (d) skarn systems.

Figure 12. Ternary plots showing stoichiometric relationships between W, Fe + Mn, and Nb + Ta. These ternaries allow for comparison of W–Fe relationships in Nb + Ta bearing samples. The trend lines shown in Figure 10 project along the Fe + Mn and Nb + Ta join, and those shown in Figure 7 project along the Fe + Mn, and W join. Potential tie lines joining these substitutional stoichiometries are shown in grey. Transparency of the points decreases away from W concentrations at 2σ of the LLD, so that samples with the highest W contents are darker (more solid). Samples are grouped according to (a) pegmatitic and granite-hosted systems, (b) greisen systems, (c) vein-hosted systems, and (d) skarn systems.

Figure 13. Ternary plots of the substitutional relationships between Ti, Fe + Mn, and Nb + Ta. The trend lines representing the 1:1 and 2:1 stoichiometric substitution reactions from project onto the Fe + Mn and Nb + Ta join, and tie lines joining these to the Ti apex are shown in grey. Colour represents the total SnO2 content of each analysis, and the colour scheme was chosen to visually down-weight values close to stoichiometric SnO2 (i.e. those with little to no substituting elements present). Samples are grouped according to (a) pegmatitic and granite-hosted systems, (b) greisen systems, (c) vein-hosted systems, and (d) skarn systems.

Figure 13. Ternary plots of the substitutional relationships between Ti, Fe + Mn, and Nb + Ta. The trend lines representing the 1:1 and 2:1 stoichiometric substitution reactions from Figure 10 project onto the Fe + Mn and Nb + Ta join, and tie lines joining these to the Ti apex are shown in grey. Colour represents the total SnO2 content of each analysis, and the colour scheme was chosen to visually down-weight values close to stoichiometric SnO2 (i.e. those with little to no substituting elements present). Samples are grouped according to (a) pegmatitic and granite-hosted systems, (b) greisen systems, (c) vein-hosted systems, and (d) skarn systems.

Figure 14. The Ti–(Fe + Mn)–(Nb + Ta) ternary diagram from , with the data from each sub-figure () contoured as probability density distributions, coloured according to (a, b) pegmatite and granite-hosted disseminated Sn systems, (c) greisen systems, (d) vein systems, and (e) skarn systems. Analyses from the cassiterite crystal from Beechworth, which has an unknown paragenesis, are plotted as points with the same colour scheme as .

Figure 14. The Ti–(Fe + Mn)–(Nb + Ta) ternary diagram from Figure 13, with the data from each sub-figure (Figure 13a–d) contoured as probability density distributions, coloured according to (a, b) pegmatite and granite-hosted disseminated Sn systems, (c) greisen systems, (d) vein systems, and (e) skarn systems. Analyses from the cassiterite crystal from Beechworth, which has an unknown paragenesis, are plotted as points with the same colour scheme as Figure 13.
Supplemental material

Supplemental Material

Download PDF (2.4 MB)