Next Article in Journal
Mining Waste as a Potential Additional Source of HREE and U for the European Green Deal: A Case Study of Bangka Island (Indonesia)
Previous Article in Journal
A Population Balance Model for Shear-Induced Polymer-Bridging Flocculation of Total Tailings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Variations across Wolframite Solid Solutions, (Fe,Mn)WO4

Department of Geoscience, University of Calgary, Calgary, AB T2N 1N4, Canada
*
Author to whom correspondence should be addressed.
Minerals 2022, 12(1), 42; https://doi.org/10.3390/min12010042
Submission received: 30 November 2021 / Revised: 20 December 2021 / Accepted: 22 December 2021 / Published: 28 December 2021

Abstract

:
The crystal structure of four samples from natural wolframite solid solutions, (Fe,Mn)WO4, was obtained with synchrotron high-resolution powder X-ray diffraction (HRPXRD) data, Rietveld refinements, space group P2/c, and Z = 2. Wolframite solid solutions extend from ferberite (FeWO4) to hübnerite (MnWO4). The W and (Mn,Fe) cations are in six-fold coordination. This study shows that the unit-cell parameters, a, b, c, and β angle, vary linearly with the unit-cell volume, V, across the wolframite series. The average <Mn,Fe–O> distance increases linearly because of larger Mn2+ (0.83 Å) replacing smaller Fe2+ (0.78 Å) cations, whereas the average <W–O> distance increases slightly because of the higher effective charge of the smaller Fe2+ cation. The distortions of the two types of polyhedra across the series are discussed.

1. Introduction

Wolframite solid solutions extend from ferberite (FeWO4) to hübnerite (MnWO4). According to IMA CNMNC, wolframite with Fe > Mn is called ferberite and those with Mn > Fe is called hübnerite. Complete substitution occurs between Fe2+ and Mn2+ cations in wolframite, (Fe,Mn)WO4. Hereafter, we use the name wolframite to cover the entire series between the two end members.
Tungstates and molybdates form two isostructural groups with the general formula MTO4, where M is for divalent metal cations (M = Ca2+, Sr2+, Ba2+, Cd2+, and Pb2+) and T is for transition elements (T = W, Mo). The monoclinic wolframite-group minerals (space group P2/c) contain small divalent cations, such as Fe2+, Mn2+, Mg2+, Ni2+, and Co2+ in six-coordination with O atoms from six different octahedral WO6 groups (Figure 1). The wolframite structure is based on W and (Fe,Mn) distorted octahedra forming infinite zig zag chains along the c-axis (Figure 1c). Each chain contains just one type of cation, and each octahedron is joined to the next by a common edge. Each W octahedral chain is attached by common corners to (Fe,Mn) chains, and each (Fe,Mn) chain is also surrounded by four W chains. The same types of octahedra are joined by edges and different types of octahedra are connected by corners.
The tetragonal scheelite-group minerals (space group I41/a) contain large divalent cations, such as Ca2+ and Pb2+, in eight-coordination with O atoms from eight different tetrahedral (WO4)2− and (MoO4)2− groups. The structure of scheelite is close to that of anhydrite, CaSO4, and zircon, ZrSiO4, but differs from these in the manner of linking of the CaO8 polyhedra. The (WO4)2− and (MoO4)2− tetrahedra are flattened along the c-axis and join edges with CaO8 or PbO8 polyhedra in scheelite and wulfenite, respectively (these structures are not shown).
The problem of the O-atom coordinates in scheelite-type structure has been of interest in crystal chemistry for a long time [1]. These MTO4 compounds contain W and Pb atoms with high atomic numbers that dominate the X-ray scattering process, so the light O-atom positions were not determined to a high precision. In addition, Pb atoms are good absorbers of X-rays. Consequently, some authors choose neutrons instead of X-rays because the scattering amplitudes using neutrons are of the same order of magnitude for the atoms in these minerals.
The crystal structure of two similar ferberite samples from Liquinaste, Jujuy, Argentina is available [2,3] as well as another sample from Saxony, Germany [4]. The structure of a synthetic MnWO4 sample is also known as well as its high-pressure behavior [5,6]. However, a complete picture of the structural variations across wolframite solid solutions could not be deduced from the above studies, so the present study was carried out to fill this knowledge gap.
This study examines the structural variations across wolframite solid solutions based on the crystal structure of four natural samples that were obtained using synchrotron high-resolution powder X-ray diffraction (HRPXRD) data and Rietveld structure refinements. Some data from the literature are compared to the data from this study.

2. Experimental Methods

2.1. Electron-Probe Microanalysis (EPMA)

The wolframite samples were analyzed using a JEOL JXA-8200 WD-ED electron-probe microanalysis (EPMA). The samples were mounted in epoxy and polished for EPMA. The JEOL operating program on a Solaris platform was used for ZAF (atomic number, Z; absorption, A; fluorescence, F) correction and data reduction. The wavelength-dispersive (WD) analyses were conducted quantitatively using an accelerated voltage of 15 kV, a beam current of 20 nA, and a beam diameter of 5 μm. Relative analytical errors were 1% for major elements and 5% for minor elements. Various standards were used: chromite (Cr), pyromorphite (Pb), scheelite (W), CaMoO4 (Ca, Mo), NiO (Ni), hornblende (Mg), ZnO (Zn), V2O5 (V), ilmenite (Fe, Ti), barite (S), and UC09059 hübnerite (Mn). Our four samples appeared to be homogeneous based on optical observations and EPMA analyses. Table 1 contains the average chemical composition from 10 spots for each sample. The localities of our four samples are given in Table 2. Based on EPMA and HRPXRD data, sample 1 contained two different phases with different compositions. The composition of two data points that may represent the two different phases are given in Table 1.

2.2. Synchrotron High-Resolution Powder X-ray Diffraction (HRPXRD)

Single crystals of our wolframite samples, of about 0.2 mm in diameter, were hand-picked under a binocular microscope and finely ground in a corundum mortar and pestle for synchrotron high-resolution powder X-ray diffraction (HRPXRD) experiments that were performed at beamline 11-BM, Advanced Photon Source, Argonne National Laboratory. Each sample was loaded into a Kapton capillary (0.8 mm internal diameter) and rotated during the experiment at a rate of 90 rotations per second. The data were collected to a maximum 2θ of about 50° with a step size of 0.001° and a step time of 0.1 s per step. The HRPXRD traces were collected with a unique multi-analyzer detection assembly consisting of 12 independent silicon (111) crystal analyzers and LaCl3 scintillation detectors that reduce the angular range to be scanned and allow for the rapid acquisition of data. A silicon (NIST 640c) and alumina (NIST 676a) standard (ratio of ⅓ Si:⅔ Al2O3 by weight) was used to calibrate the instrument and to refine the monochromatic wavelength (λ = 0.41279(2) Å) used in the experiment (Table 2). Additional details of the experimental set-up are given elsewhere [8,9,10]. The experimental techniques used in this study are well established [11,12,13,14,15,16].

2.3. Rietveld Structural Refinement

The crystal structure was modeled with the Rietveld method [17] that is incorporated in the GSAS program [18] and using the EXPGUI interface [19]. The starting atom coordinates, unit-cell parameters, and space group P2/c, Z = 2 were taken from Cid-Dresdner and Escobar [2]. The background was modeled with a Chebyschev polynomial (8 terms). The reflection-peak profiles were fitted using type-3 profile (pseudo-Voigt; [20,21]). The structure refinements were carried out by varying parameters in the following sequence: scale factor, background, unit-cell parameters, zero shift, profile, atom positions, and isotropic displacement parameters, U. The site-occupancy factor (sofs) for the Fe/Mn site is fixed based on EPMA values because these elements differ by just one electron. Finally, all variables were refined simultaneously. The U(O1) was set equal to U(O2). The sofs for the W and O sites were set equal to 1.
Table 2 contains unit-cell parameters and other information regarding data collection and refinement. Table 3 contains the atom coordinates and isotropic displacement parameters. Table 4 contains the bond distances. CIF files for the four samples are available as supplementary materials.

3. Results and Discussion

3.1. Chemical Composition of Natural Wolframite Samples

The compositions of the wolframite samples are given in Table 1. The atom per formula unit (apfu) is based on four oxygen atoms. The W site is filled exclusively with W atoms. The divalent cation site is essentially filled with Mn and Fe atoms. All of our samples are Mn-rich with the % hübnerite [=(Mn/(Mn + Fe)) × 100] ranging from 64 to 100%.
Two possible compositions for phase-1a and phase-1b in sample 1 are given in Table 1. The X-ray diffraction data below show that sample 1 contains two separate phases that are slightly different from each other. These two phases differ in their Mn and Fe content. The Fe-rich phase-1a has a smaller unit-cell volume than the Mn-rich phase-1b, as expected, whereas sample 2 has values intermediate between phase-1a and phase-1b.

3.2. High-Resolution Powder X-ray Diffraction (HRPXRD) Traces for Natural Wolframite Samples

Except for sample 1, the traces for the other three samples contain sharp and symmetrical peaks with no impurity. An example for such a trace is shown for sample 2 that contains a single phase (Figure 2c,d). However, sample 1 contains two different phases as indicated by the split reflections (Figure 2a,b) compared to the un-split reflections in Figure 2c,d. These features are shown clearly in the expanded traces (Figure 2b,d). The crystal structure of both phases in sample 1 were obtained as well as their compositions (Table 1 and Table 4). Splitting of reflections were observed in several other minerals [22,23,24,25,26,27,28].

3.3. Variations of the Unit-Cell Parameters across Wolframite, (Fe,Mn)WO4, Solid Solutions

Across natural wolframite solid solutions, the unit-cell parameters a, b, c, and β with volume, V, vary linearly (Table 2, Figure 3). The a-axis (Figure 3a), b axis (Figure 3b), c-axis (Figure 3c), and β angle (Figure 3d) increase linearly with the increase in V. For comparison to our data, those from other studies are shown [3,4,5]. Based on the chemical analyses, the substitution of larger Mn2+ (0.83 Å) occurs for smaller Fe2+ (0.78 Å) cations and causes the increase in unit-cell parameters with V.

3.4. Variations of Structural Parameters across Wolframite, (Fe,Mn)WO4, Solid Solutions

The essential features of the wolframite structure obtained in this study are the same as those observed in previous studies. The wolframite structure consists of two main structural units: distorted (Fe,Mn)O6 and WO6 octahedra (Figure 1). In our refinements of the wolframite structure, the site occupancy factors (sofs) for the Fe/Mn site were fixed by the EPMA data because these two elements differ by only one electron (Table 1). In fact, the Mn content varies linearly with the increase in unit-cell volume, V, as shown in Figure 4. It appears that a nearly pure natural FeWO4 end member was studied by Ülkü [4] and a pure synthetic MnWO4 end member was studied by Macavei and Schulz [5]. Based on our extrapolated data, it appears that the pure FeWO4 may have a unit-cell volume of about 132 Å3. As such, the volume range were extended from 132 to about 139.4 Å3 (Figure 3, Figure 4 and Figure 5).
The average <(Fe,Mn)–O> distance increases linearly with V and has values from about 2.154 (sample 1a) to 2.184 (sample 4), a difference of 0.030Å (Table 4; Figure 5a). The average <W–O> distance increases linearly and has values from about 1.943 (sample 1a) to 1.948 (sample 3), a difference of 0.005Å, which is about 6 times less than the difference for the average <(Fe,Mn)–O> distance (Figure 5b). The average <(Fe,Mn)–O> distance increases linearly because of the substitution of larger Mn2+ (0.83 Å) for smaller Fe2+ (0.78 Å) cations. End-member data from the literature [4,5] match the data from this study quite well (Figure 5a,b). However, the two similar data points from Escobar et al. [3] and Cid-Dresdner and Escobar [2] appear unreliable (see Figure 5).
Shannon [7] gives the following radii for atoms in six coordination in wolframite: Fe2+ = 0.78, Mn2+ = 0.83, W6+ = 0.60 Å, and three-coordination O2− = 1.36 Å, so the radii sum gives Fe–O = 2.14 Å, which is close to the projected average <Fe–O> = 2.095 Å (Figure 5a). The radii sum gives Mn–O = 2.19 Å, which is close to the average <Mn–O> = 2.184 Å in sample 4. In addition, the radii sum gives W–O = 1.96 Å, which is close to the average <W–O> = 1.948 Å in sample 3 (Table 4). In danalite, Fe8[Be6Si6024]S2, the average <Fe–O> <3> distance is 2.033(3) Å [29], whereas in magnesioferrite spinel, MgFe2O4, the average <Fe–O> <6> distance is 2.061(1) [30]. The values are similar to those for the Fe end-member of wolframite.
The distortion indices of the W and (Fe,Mn) polyhedra are given in Table 4. Among our samples, sample 1b has an intermediate composition and contains the shortest Fe–O1 and W–O2 distances. However, the average <(Fe,Mn)–O> distance is comparable to those in the other samples as indicated in Figure 5a. The distortion indices across the series represent the data quite well (Figure 6), especially those in Figure 6a,d. Because sample 1b is about 44 wt. %, the refinement results may not be as accurate as for the other samples, including sample 1a. However, the possibility of some Fe3+ and vacancies replacing Fe2+ cannot be ruled out.

4. Concluding Remarks

Understanding (Fe,Mn)WO4 solid solutions requires good structural data across the series. Although a few single-crystal structure refinements are available for a few members across the series, it was not possible to examine the structural variations across the series using those studies. This study shows that good structural information can be obtained with synchrotron HRPXRD data and the Rietveld refinement method to show expected linear systematic variations across wolframite solid solutions. The unit-cell parameters, a, b, c, and β angle vary linearly with the unit-cell volume, V. The average <W–O> distance increases slightly with the increase in V from ferberite to hübnerite because of the higher effective charge of the smaller Fe2+ cation. The average <Mn,Fe–O> distance increases linearly with V because of the larger Mn2+ (0.83 Å) replacing the smaller Fe2+ (0.78 Å) cations.

Supplementary Materials

CIF files for the four samples are available online at https://www.mdpi.com/article/10.3390/min12010042/s1.

Author Contributions

The academic editor and three anonymous reviewers are thanked for comments that helped to improve this manuscript. D.A.U. performed the EPMA experiment and analyzed the chemical composition data. S.M.A. conceived the project and analyzed the HRPXRD data. Both authors contributed to writing and editing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a NSERC Discovery Grant to SMA, grant number 10013896.

Acknowledgments

Robert Marr is thanked for his help with the EPMA analysis. The HRPXRD data were collected at the X-ray Operations and Research beamline 11-BM, Advanced Photon Source (APS), Argonne National Laboratory (ANL). Use of the APS was supported by the U.S. Dept. of Energy, Office of Science, Office of Basic Energy Sciences, under Contract No. DE-AC02-06CH11357.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vegard, L. Results of crystal analysis. In Space Lattices and Atomic Dimensions; I Kommission hos J. Dybwad: Oslo, Norway, 1925. [Google Scholar]
  2. Cid-Dresdner, H.; Escobar, C. The crystal structure of ferberite, FeWO4. Z. Krist. 1968, 127, 61–72. [Google Scholar] [CrossRef]
  3. Escobar, C.; Cid-Dresdner, H.; Kittl, P.; Dümler, I. The relation between “light wolframite” and common wolframite. Am. Mineral. 1971, 56, 489–498. [Google Scholar]
  4. Ülkü, D. Untersuchungen zur Kristallstruktur und magnetischen Struktur des Ferberits FeWO4. Z. Krist. 1967, 124, 192–219. [Google Scholar] [CrossRef]
  5. Macavei, J.; Schulz, H. The crystal structure of wolframite type tungstates at high pressure. Z. Krist. 1993, 207, 193–208. [Google Scholar]
  6. Errandonea, D.; Ruiz-Fuertes, J. A brief review of the effects of pressure on wolframite-type oxides. Crystals 2018, 8, 71. [Google Scholar] [CrossRef] [Green Version]
  7. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  8. Lee, P.L.; Shu, D.; Ramanathan, M.; Preissner, C.; Wang, J.; Beno, M.A.; Von Dreele, R.B.; Ribaud, L.; Kurtz, C.; Antao, S.M.; et al. A twelve-analyzer detector system for high-resolution powder diffraction. J. Synchrotron Radiat. 2008, 15, 427–432. [Google Scholar] [CrossRef]
  9. Wang, J.; Toby, B.H.; Lee, P.L.; Ribaud, L.; Antao, S.M.; Kurtz, C.; Ramanathan, M.; Von Dreele, R.B.; Beno, M.A. A dedicated powder diffraction beamline at the advanced photon source: Commissioning and early operational results. Rev. Sci. Instrum. 2008, 79, 85105. [Google Scholar] [CrossRef]
  10. Antao, S.M.; Hassan, I.; Wang, J.; Lee, P.L.; Toby, B.H. State-of-the-art high-resolution powder X-ray diffraction (HRPXRD) illustrated with Rietveld structure refinement of quartz, sodalite, tremolite, and meionite. Can. Mineral. 2008, 46, 1501–1509. [Google Scholar] [CrossRef]
  11. Antao, S.M. Crystal-structure analysis of four mineral samples of anhydrite, CaSO4, using synchrotron high-resolution powder X-ray diffraction data. Powder Diffr. 2011, 26, 326–330. [Google Scholar] [CrossRef]
  12. Parise, J.B.; Antao, S.M.; Michel, F.M.; Martin, C.D.; Chupas, P.J.; Shastri, S.; Lee, P.L. Quantitative high-pressure pair distribution function analysis. J. Synchrotron Radiat. 2005, 12, 554–559. [Google Scholar] [CrossRef]
  13. Hassan, I.; Antao, S.M.; Parise, J.B. Haüyne: Phase transition and high-temperature structures obtained from synchrotron radiation and Rietveld refinements. Mineral. Mag. 2004, 68, 499–513. [Google Scholar] [CrossRef]
  14. Hassan, I.; Antao, S.M.; Hersi, A.A. Single-crystal XRD, TEM, and thermal studies of the satellite reflections in nepheline. Can. Mineral. 2003, 41, 759–783. [Google Scholar] [CrossRef]
  15. Antao, S.M.; Hassan, I.; Mulder, W.H.; Lee, P.L. The R-3cR-3m transition in nitratine, NaNO3, and implications for calcite, CaCO3. Phys. Chem. Miner. 2008, 35, 545–557. [Google Scholar] [CrossRef]
  16. Zaman, M.; Schubert, M.; Antao, S. Elevated radionuclide concentrations in heavy mineral-rich beach sands in the Cox’s Bazar region, Bangladesh and related possible radiological effects. Isot. Environ. Health Stud. 2012, 48, 512–525. [Google Scholar] [CrossRef]
  17. Rietveld, H.M. A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr. 1969, 2, 65–71. [Google Scholar] [CrossRef]
  18. Larson, A.C.; Von Dreele, R.B. General Structure Analysis System (GSAS); Los Alamos National Laboratory Report, LAUR 86-748; Los Alamos National Laboratory: Los Alamos, NM, USA, 2000. [Google Scholar]
  19. Toby, B.H. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef] [Green Version]
  20. Cagliotti, G.; Paoletti, A.; Ricci, F.P. Choice of collimators for a crystal spectrometer for neutron diffraction. Nucl. Instrum. 1958, 3, 223–228. [Google Scholar] [CrossRef]
  21. Thompson, P.; Cox, D.E.; Hastings, J.B. Rietveld refinement of Debye-Scherrer synchrotron X-ray data from alumina. J. Appl. Crystallogr. 1987, 20, 79–83. [Google Scholar] [CrossRef] [Green Version]
  22. Antao, S.M. Is near-endmember birefringent grossular non-cubic? New evidence from synchrotron diffraction. Can. Mineral. 2013, 51, 771–784. [Google Scholar] [CrossRef]
  23. Antao, S.M. The mystery of birefringent garnet: Is the symmetry lower than cubic? Powder Diffr. 2013, 28, 281–288. [Google Scholar] [CrossRef]
  24. Antao, S.M.; Hassan, I. A two-phase intergrowth of genthelvite from Mont Saint-Hilaire, Quebec. Can. Mineral. 2010, 48, 1217–1223. [Google Scholar] [CrossRef]
  25. Antao, S.M.; Klincker, A.M. Crystal structure of a birefringent andradite-grossular from Crowsnest Pass, Alberta, Canada. Powder Diffr. 2014, 29, 20–27. [Google Scholar] [CrossRef]
  26. Antao, S.M.; Mohib, S.; Zaman, M.; Marr, R.A. Ti-rich andradites: Chemistry, structure, multi-phases, optical anisotropy, and oscillatory zoning. Can. Mineral. 2015, 53, 133–158. [Google Scholar] [CrossRef]
  27. Antao, S.M. Crystal structure of morimotoite from Ice River, Canada. Powder Diffr. 2014, 29, 325–330. [Google Scholar] [CrossRef] [Green Version]
  28. Antao, S.M.; Round, S.A. Crystal chemistry of birefringent spessartine. Powder Diffr. 2014, 29, 233–240. [Google Scholar] [CrossRef] [Green Version]
  29. Antao, S.M.; Hassan, I.; Parise, J.B. The structure of danalite at high temperature obtained from synchrotron radiation and Rietveld refinements. Can. Mineral. 2003, 41, 1413–1422. [Google Scholar] [CrossRef]
  30. Antao, S.M.; Hassan, I.; Parise, J.B. Cation ordering in magnesioferrite, MgFe2O4, to 982 °C using in situ synchrotron X-ray powder diffraction. Am. Mineral. 2005, 90, 219–228. [Google Scholar] [CrossRef]
Figure 1. Structure of wolframite, (Fe,Mn)WO4, (space group P2/c) projected (a) down [001] (b-axis is vertical and the a-axis is horizontal), (b) down [010] (c-axis is vertical and the a-axis is horizontal), and (c) down [100] (c-axis is vertical and the b-axis is horizontal). The O (grey spheres) oxygen atoms are indicated. The monoclinic unit-cell edges are outlined (black frame). The structure contains distorted WO6 octahedra (yellow) and (Fe,Mn)O6 octahedra (blue) forming infinite zig-zag chains along the c-axis, as shown in (c).
Figure 1. Structure of wolframite, (Fe,Mn)WO4, (space group P2/c) projected (a) down [001] (b-axis is vertical and the a-axis is horizontal), (b) down [010] (c-axis is vertical and the a-axis is horizontal), and (c) down [100] (c-axis is vertical and the b-axis is horizontal). The O (grey spheres) oxygen atoms are indicated. The monoclinic unit-cell edges are outlined (black frame). The structure contains distorted WO6 octahedra (yellow) and (Fe,Mn)O6 octahedra (blue) forming infinite zig-zag chains along the c-axis, as shown in (c).
Minerals 12 00042 g001
Figure 2. Comparison of HRPXRD traces for sample 1 (a,b) and sample 2 (c,d) together with the calculated (continuous line) and observed (crosses) profiles. The difference curve (IobsIcalc) with the same intensity scale is shown at the bottom of each trace. Short vertical lines indicate allowed reflection positions. The intensities for the trace and difference curve that are above 20° in 2θ are multiplied by 10 in both (a,c). Expanded traces are displayed in (b,d). The traces in (c,d) contain sharp and symmetrical peaks with no impurity peaks. Sample 1 contains two similar phases as indicated by the split reflections in (a,b) compared to the un-split reflections in (c,d).
Figure 2. Comparison of HRPXRD traces for sample 1 (a,b) and sample 2 (c,d) together with the calculated (continuous line) and observed (crosses) profiles. The difference curve (IobsIcalc) with the same intensity scale is shown at the bottom of each trace. Short vertical lines indicate allowed reflection positions. The intensities for the trace and difference curve that are above 20° in 2θ are multiplied by 10 in both (a,c). Expanded traces are displayed in (b,d). The traces in (c,d) contain sharp and symmetrical peaks with no impurity peaks. Sample 1 contains two similar phases as indicated by the split reflections in (a,b) compared to the un-split reflections in (c,d).
Minerals 12 00042 g002
Figure 3. Variations of the unit-cell parameters (a) a, (b) b, (c) c, and (d) β with volume, V, across the wolframite series, (Fe,Mn)WO4. The literature data are from Macavei and Schulz [5] for a pure synthetic end member, MnWO4, Escobar et al. [3], and Ülkü [4]. The linear trend lines and equations are based only on data from this study. The unit-cell parameters increase with V across the series.
Figure 3. Variations of the unit-cell parameters (a) a, (b) b, (c) c, and (d) β with volume, V, across the wolframite series, (Fe,Mn)WO4. The literature data are from Macavei and Schulz [5] for a pure synthetic end member, MnWO4, Escobar et al. [3], and Ülkü [4]. The linear trend lines and equations are based only on data from this study. The unit-cell parameters increase with V across the series.
Minerals 12 00042 g003
Figure 4. The Mn apfu varies linearly with the unit-cell volume, V. The inserted data points are from Macavei and Schulz [5] for a pure synthetic end member, MnWO4, Escobar et al. [3], and Ülkü [4].
Figure 4. The Mn apfu varies linearly with the unit-cell volume, V. The inserted data points are from Macavei and Schulz [5] for a pure synthetic end member, MnWO4, Escobar et al. [3], and Ülkü [4].
Minerals 12 00042 g004
Figure 5. Structural variations across the wolframite series with increasing V: (a) the average <(Fe,Mn)–O> distance increases linearly because of Mn2+ (0.83 Å) replacing Fe2+ (0.78 Å) cations. (b) The average <W–O> distance increases slightly with V because Mn2+ has a less effective polarizing power to the O atoms than the smaller Fe2+ cation. The data from Escobar et al. [3] appear to be unreliable because they are far away from the red trend lines.
Figure 5. Structural variations across the wolframite series with increasing V: (a) the average <(Fe,Mn)–O> distance increases linearly because of Mn2+ (0.83 Å) replacing Fe2+ (0.78 Å) cations. (b) The average <W–O> distance increases slightly with V because Mn2+ has a less effective polarizing power to the O atoms than the smaller Fe2+ cation. The data from Escobar et al. [3] appear to be unreliable because they are far away from the red trend lines.
Minerals 12 00042 g005
Figure 6. (Fe,Mn)O6 polyhedral distortion indices across the wolframite series with increasing V: All the trend lines appear to represent variations across the series quite well. However, those in (a) and (d) have the highest correlation factors, but the values are lower in (b,c).
Figure 6. (Fe,Mn)O6 polyhedral distortion indices across the wolframite series with increasing V: All the trend lines appear to represent variations across the series quite well. However, those in (a) and (d) have the highest correlation factors, but the values are lower in (b,c).
Minerals 12 00042 g006
Table 1. Chemical compositions for four wolframite (Fe,Mn)WO4 samples.
Table 1. Chemical compositions for four wolframite (Fe,Mn)WO4 samples.
OxidesRadii (Å) *1a1b234
wt. %
MnO 15.1420.0717.66(29)21.89(46)23.41(12)
FeO 8.713.425.89(26)1.51(56)0.07(2)
ZnO 0.020.040.02(3)0.01(1)0.04(2)
MgO 0.020.000.00(2)0.00(1)0.00(1)
NiO 0.020.040.00(2)0.03(2)0.02(2)
CaO 0.000.000.00(1)0.00(0)0.00(0)
PbO 0.000.000.00(0)0.00(0)0.00(0)
WO3 76.1176.0076.4776.29(16)76.05(39)
MoO3 0.000.000.00(0)0.00(0)0.00(0)
SO3 0.060.210.26(8)0.05(9)0.08(11)
TiO2 0.020.000.00(5)0.00(0)0.01(1)
V2O3 0.000.000.00(0)0.00(0)0.00(0)
Cr2O3 0.010.000.00(1)0.01(1)0.00(1)
Total 100.1199.78100.3099.7999.68
Cations-apfu
Mn2+ 0.830.6450.8560.748(12)0.935(19)1.000(5)
Fe2+0.780.3670.1440.247(11)0.064(23)0.003(1)
Zn2+0.740.0010.0010.001(1)0.000(0)0.002(1)
Mg2+0.720.0010.0000.000(1)0.000(1)0.000(1)
Ni2+0.690.0010.0010.000(1)0.001(1)0.001(1)
Ca2+1.000.0000.0000.000(0)0.000(0)0.000(0)
Pb2+1.190.0000.0000.000(0)0.000(0)0.000(0)
ΣA 1.0151.0020.9961.0001.006
W6+0.600.9920.9910.992(2)0.998(3)0.994(3)
Mo6+0.590.0000.0000.000(0)0.000(0)0.000(0)
S6+0.290.0020.0080.010(3)0.002(3)0.003(4)
Ti4+0.610.0010.0000.000(2)0.000(0)0.000(0)
V3+0.640.0000.0000.000(0)0.000(0)0.000(0)
Cr3+0.620.0000.0000.000(0)0.000(0)0.000(0)
ΣB 0.9950.9991.0021.0000.997
Σ(A + B) 2.0102.0011.9982.0002.003
% Hü 63.7485.6075.1893.5999.70
* Ionic radii are from Shannon [7]. The apfu are based on 4 oxygen atoms. % Hü = % hübnerite = [Mn/(Mn + Fe)] × 100. Sample 1 contains two similar phases, so 2 possible compositions are given from 2 data points, whereas the average composition from 10 data points is given for the other samples.
Table 2. Wolframite samples: locality, unit-cell parameters, and Rietveld refinement statistical indicators, R (F2), χ2, etc.
Table 2. Wolframite samples: locality, unit-cell parameters, and Rietveld refinement statistical indicators, R (F2), χ2, etc.
No.LocalitiesabcβV3* R(F2)χ2NobsNptsVar.
1aOregon Mine, Boulder, Colorado (VC758)4.78219(4)5.73389(4)4.98336(3)90.5631(6)136.640(2)0.02081.699213443,06743
1bVC7584.81212(3)5.74991(4)4.99266(3)90.9304(6)138.125(2)
2Tae Wha mine, Korea (VC315)4.79440(2)5.74073(2)4.98734(2)90.7109(3)137.258(1)0.01781.623106543,06728
3Silverton, Colorado (VC692)4.82435(2)5.75627(2)4.99611(1)91.0773(2)138.7188(8)0.02713.66498341,44428
4Pasto Bueno, Peru (UC09059)4.82991(2)5.75974(2)4.99795(2)91.1475(2)139.0101(9)0.02382.791103243,06728
All samples were refined in space group P2/c with λ = 0.41279(2) Å. * R(F2) = R-structure factor based on observed and calculated structure amplitudes = [∑(Fo2Fc2)/∑(Fo2)]1/2. The 2θ range for each sample is from 2° to 45°. The values for Nobs (number of observed reflections), Npts (number of data points), and Var. (number of variables) are given. Sample 1 contains two different phases. The weight % for phase-1a is 55.84(6) and phase-1b is 44.16(7).
Table 3. Wolframite samples: atom coordinates and isotropic displacement parameters (U × 102 Å2).
Table 3. Wolframite samples: atom coordinates and isotropic displacement parameters (U × 102 Å2).
Atom-Site1a1b234
Mn/Fex½½½½½
y0.3202(2)0.3149(3)0.3181(1)0.3139(1)0.3131(1)
2fz¾¾¾¾¾
U0.29(1)0.29(1)0.31(1)0.54(2)0.31(1)
sof0.65 Mn
0.35 Fe
0.86 Mn
0.24 Fe
0.75 Mn
0.25 Fe
0.94 Mn
0.06 Fe
1 Mn
0 Fe
Wx00000
y0.17974(6)0.18025(8)0.17999(3)0.17988(4)0.17975(4)
2ez¼¼¼¼¼
U0.109(4)0.109(4)0.109(4)0.323(4)0.273(4)
O1x0.2149(6)0.2084(7)0.2139(3)0.2137(4)0.2136(4)
y−0.0980(5)−0.1063(6)−0.1032(3)−0.0992(3)−0.1035(3)
4gz0.4433(7)0.4351(8)0.4396(4)0.4392(4)0.4458(4)
U0.07(3)0.07(3)0.05(3)0.47(3)0.25(5)
O2x0.2628(6)0.2479(7)0.2544(3)0.2539(3)0.2523(3)
y0.3655(5)0.3766(6)0.3702(3)0.3730(3)0.3680(3)
4gz0.3972(7)0.3880(8)0.3945(4)0.3893(4)0.3952(4)
Space group P2/c, choice 1. The sofs for W, O1, and O2 were fixed at 1.0; and U(O1) = U(O2).
Table 4. Bond distances (Å) and polyhedral distortion indices for four wolframite samples.
Table 4. Bond distances (Å) and polyhedral distortion indices for four wolframite samples.
Bond1a1b234
(Fe,Mn)O6 octahedron
Fe–O1 × 22.106(3)2.074(3)2.081(2) 2.093(2)2.094(2)
Fe–O2 × 22.099(4)2.189(4)2.137(2)2.166(2)2.143(2)
Fe–O2 × 22.256(3)2.263(3) 2.265(2)2.274(2)2.316(2)
<(Fe,Mn)–O> <6>2.1542.1762.1612.1782.184
Average edge length3.02393.04613.03063.05083.0590
Average angle104.570104.048104.378104.125104.058
Polyhedral volume12.952613.250313.062213.319413.4182
Octahedral angle variance65.424383.626070.042978.633482.4303
Mean octahedral quadratic elongation1.01991.02521.02121.02361.0255
WO6 octahedron
W–O1 × 22.121(3)2.132(4)2.136(2)2.122(2)2.155(2)
W–O1 × 21.909(3)1.926(4)1.919(2)1.936(2)1.906(2)
W–O2 × 21.798(3)1.773(4)1.782(2)1.786(2)1.776(2)
<W–O> <6>1.9431.9441.9461.9481.945
Average edge length2.71452.72152.72222.72582.7218
Average angle103.519104.010103.821103.840103.808
Polyhedral volume9.27549.38369.39259.41199.3755
Octahedral angle variance118.679297.5426102.6663104.9690103.9077
Mean octahedral quadratic elongation1.04071.03471.03631.03621.0377
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Umbsaar, D.A.; Antao, S.M. Structural Variations across Wolframite Solid Solutions, (Fe,Mn)WO4. Minerals 2022, 12, 42. https://doi.org/10.3390/min12010042

AMA Style

Umbsaar DA, Antao SM. Structural Variations across Wolframite Solid Solutions, (Fe,Mn)WO4. Minerals. 2022; 12(1):42. https://doi.org/10.3390/min12010042

Chicago/Turabian Style

Umbsaar, Darren A., and Sytle M. Antao. 2022. "Structural Variations across Wolframite Solid Solutions, (Fe,Mn)WO4" Minerals 12, no. 1: 42. https://doi.org/10.3390/min12010042

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