Next Article in Journal
Dynamic Disorder of Fe3+ Ions in the Crystal Structure of Natural Barioferrite
Next Article in Special Issue
Origin of Platinum Group Minerals (PGM) Inclusions in Chromite Deposits of the Urals
Previous Article in Journal
Synthesis of 4A Zeolite from Kaolinite-Type Pyrite Flotation Tailings (KPFT)
Previous Article in Special Issue
Mineralogy of Platinum-Group Elements and Gold in the Ophiolite-Related Placer of the River Bolshoy Khailyk, Western Sayans, Russia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thalhammerite, Pd9Ag2Bi2S4, a New Mineral from the Talnakh and Oktyabrsk Deposits, Noril’sk Region, Russia

1
Czech Geological Survey, Geologická 6, 152 00 Prague 5, Czech Republic
2
Institute of Geology of Ore Deposits, Mineralogy, Petrography and Geochemistry RAS, Staromonetnyi per. 12, Moscow 119017, Russia
3
Oxford Instruments (Moscow Office), 26 Denisovskii Pereulok, Moscow 105005, Russia
4
Department of Earth Sciences, Natural History Museum, London SW7 5BD, UK
5
Institute of Physics, AS CR v.v.i. Na Slovance 2, 182 21 Prague 8, Czech Republic
6
Department of Applied Geosciences and Geophysics, University of Leoben, Peter Tunner Str. 5, A 8700 Leoben, Austria
*
Author to whom correspondence should be addressed.
Minerals 2018, 8(8), 339; https://doi.org/10.3390/min8080339
Submission received: 19 July 2018 / Revised: 2 August 2018 / Accepted: 3 August 2018 / Published: 8 August 2018

Abstract

:
Thalhammerite, Pd9Ag2Bi2S4, is a new sulphide discovered in galena-pyrite-chalcopyrite and millerite-bornite-chalcopyrite vein-disseminated ores from the Komsomolsky mine of the Talnakh and Oktyabrsk deposits, Noril’sk region, Russia. It forms tiny inclusions (from a few μm up to about 40–50 μm) intergrown in galena, chalcopyrite, and also in bornite. Thalhammerite is brittle and has a metallic lustre. In plane-polarized light, thalhammerite is light yellow with weak bireflectance, weak pleochroism, in shades of slightly yellowish brown and weak anisotropy; it exhibits no internal reflections. Reflectance values of thalhammerite in air (R1, R2 in %) are: 41.9/43.0 at 470 nm, 43.9/45.1 at 546 nm, 44.9/46.1 at 589 nm, and 46.3/47.5 at 650 nm. Three spot analyses of thalhammerite give an average composition: Pd 52.61, Bi 22.21, Pb 3.92, Ag 14.37, S 7.69, and Se 0.10, total 100.90 wt %, corresponding to the empirical formula Pd8.46Ag2.28(Bi1.82Pb0.32)Σ2.14(S4.10Se0.02)Σ4.12 based on 17 atoms; the average of five analyses on synthetic thalhammerite is: Pd 55.10, Bi 24.99, Ag 12.75, and S 7.46, total 100.30 wt %, corresponding to Pd8.91Ag2.03Bi2.06S4.00. The density, calculated on the basis of the empirical formula, is 9.72 g/cm3. The mineral is tetragonal, space group I4/mmm, with a 8.0266(2), c 9.1531(2) Å, V 589.70(2) Å3 and Z = 2. The crystal structure was solved and refined from the single-crystal X-ray-diffraction data of synthetic Pd9Ag2Bi2S4. Thalhammerite has no exact structural analogues known in the mineral system; chemically, it is close to coldwellite (Pd3Ag2S) and kravtsovite (PdAg2S). The strongest lines in the X-ray powder diffraction pattern of synthetic thalhammerite [d in Å (I) (hkl)] are: 3.3428(24)(211), 2.8393(46)(220), 2.5685(21)(301), 2.4122(100)(222), 2.3245(61)(123), 2.2873(48)(004), 2.2201(29)(132), 2.0072(40)(400), 1.7481(23)(332), and 1.5085(30)(404). The mineral honours Associate Professor Oskar Thalhammer of the University of Leoben, Austria.

1. Introduction

Thalhammerite, ideally Pd9Ag2Bi2S4, was observed in the same holotype specimen as kravtsovite, PdAg2S [1], and vymazalováite, Pd3Bi2S2 [2]. The type sample (polished section) comes from vein-disseminated pyrite-chalcopyrite-galena ore from the Komsomolsky mine in the Talnakh deposit of the Noril’sk district, Russia. The sample was found at coordinates: 69°30′20″ N and 88°27′17″ E. The mineralization is characterized by lack of Ni minerals and high galena content and Pt-Pd-Ag bearing minerals in an association of pyrite and chalcopyrite. The host rocks of pyrite-chalcopyrite-galena ore are diopside-hydrogrosssular-serpentine metasomatites developed in diopside-monticellite skarns below the lower exocontact of the Talnakh intrusion (the eastern part of the Komsomolsky mine). Thalhammerite, in pyrite-chalcopyrite-galena ores, occurs in association with cooperite, braggite, vysotskite, stibiopalladinite, telargpalite, sobolevskite, kotulskite, sopcheite, insizwaite, kravtsovite, vymazalováite, Au-Ag alloys, and Ag-bearing sulphides, selenides, sulphoselenides, and tellurosulphoselenides. The mineral was also observed in vein-disseminated millerite-bornite-chalcopyrite ore from the Talnakh and Oktyabrsk deposits of the Noril’sk region [3]. The host rocks of millerite-bornite-chalcopyrite ore are pyroxene-hornfels at the lower exocontact of the Kharaelakh intrusion (the western part of the Komsomolsky mine). In millerite-bornite-chalcopyrite ore, thalhammerite occurs in association with kotulskite, telargpalite, laflammeite, and Au-Ag alloys.
The mineral likely formed under the same conditions as kravtsovite and vymazalováite, with decreasing temperature [3], most likely below 400 °C. Thalhammerite was also observed, in intergrowths with sobolevskite, in PGE ores from the Fedorov-Pana Layered Intrusive Complex, Russia (V.V. Subbotin—per. communication). Furthermore, the occurrence of unknown phases corresponding to Pb- and Tl-analogues of thalhammerite from the Fedorov-Pana Layered Intrusive Complex has been reported [4].
Both the mineral and name were approved by the Commission on New Minerals, Nomenclature and Classification of the International Mineralogical Association (IMA No 2017-111). The mineral name is for Dr. Oskar Thalhammer (b. 1956) Associate Professor at the University of Leoben, Austria for his contributions to the ore mineralogy and mineral deposits of platinum group elements. The type specimen is deposited at the Department of Earth Sciences of the Natural History Museum, London, UK, catalogue no. BM 2016, 150.

2. Appearance, and Physical and Optical Properties

Thalhammerite forms very small inclusions (from a few μm up to about 40–50 μm) in galena, chalcopyrite (Figure 1), and also in bornite.
The mineral occurs in aggregates (100–200 μm in size) formed by intergrowths of telargpalite, braggite, vysotskite, sopcheite, stibiopalladinite, sobolevskite, moncheite, kotulskite, malyshevite, insizwaite, acanthite, aurian silver, kravtsovite, and vymazalováite in association with galena, chalcopyrite, bornite, millerite, and pyrite.
Thalhammerite is opaque with a metallic lustre. The mineral is brittle. The density calculated on the basis of the empirical formula is 9.72 g/cm3. In plane-polarized light, thalhammerite is light yellow with weak bireflectance, weak pleochroism, in shades of slightly yellowish brown and weak anisotropy. It exhibits no internal reflections.
Reflectance measurements were made in air relative to a WTiC standard on both natural and synthetic thalhammerite using a J and M TIDAS diode array spectrometer attached to a Zeiss Axiotron microscope. The results are tabulated (Table 1) and illustrated in Figure 2.

3. Chemical Composition

Electron probe micro-analyses (EPMA) on grains of thalhammerite were obtained using a WDA Inca Wave 500 (Oxford Instruments NanoAnalysis, High Wycombe, UK) installed on an SEM Lyra 3GM (Tescan), with analytical conditions of 20 kV, 10 nA, and counting times of 30 s (on peak positions)/2 × 15 s (background on the left and right positions). The spectra were collected on PbMα, BiMα, PdLα, AgLα, SKα, and SeLα lines with standards of pure Se, Pd, Ag, Bi, synthetic PbTe, and natural FeS2. Other elements were below the detection limit.
EPMA on synthetic thalhammerite were obtained using a CAMECA SX-100 electron probe microanalyzer in wavelength-dispersive mode with an electron beam focussed to 1–2 μm.
Pure elements and ZnS were used as standards and the radiations measured were BiMα PdLα, AgLα, and SKα, with an accelerating voltage of 15 kV, and a beam current of 10 nA measured on the Faraday cup.
EPMA compared with literature data are given in Table 2. The empirical formulae calculated on the basis of 17 apfu are Pd8.46Ag2.28(Bi1.82Pb0.32)Σ2.14(S4.10Se0.02)Σ4.12 for thalhammerite and Pd8.91Ag2.03Bi2.06S4.00 for its synthetic analogue, with the ideal formulae Pd9Ag2Bi2S4.

4. Synthetic Analogue

The small size of thalhammerite embedded in galena (bornite) prevented its extraction and isolation in an amount sufficient for the relevant crystallographic and structural investigations. Therefore, these investigations were performed on the synthetic Pd9Ag2Bi2S4.
The synthetic phase of Pd9Ag2Bi2S4 was prepared in an evacuated and sealed silica-glass tube in a horizontal furnace in the Laboratory of Experimental Mineralogy of the Czech Geological Survey in Prague. To prevent loss of material to the vapour phase during the experiment, the free space in the tube was reduced by placing a closely-fitting silica glass rod against the charge.
The temperature was measured with Pt-PtRh thermocouples and is accurate to within ±3 °C. A charge of about 300 mg was carefully weighed out from the native elements. We used, as starting chemicals, palladium (99.95%), silver (99.999%), bismuth (99.999%), and sulphur (99.999%). The starting mixture was sealed and annealed, quenched, and then ground in an agate mortar under acetone and reheated to 350 °C for 134 days. The sample was quenched by dropping the capsule in cold water.

5. X-ray Crystallography

5.1. Single-Crystal X-ray Diffraction

A small fragment of synthetic Pd9Ag2Bi2S4 was mounted on a glass fibre and examined using a Rigaku Super Nova single-crystal diffractometer with an Atlas S2 CCD detector utilizing MoKα radiation, provided by the microfocus X-ray tube and monochromatized by primary mirror optics. The ω rotational scans were used for collection of three-dimensional intensity data. From a total of 3659 reflections, 221 were classified as unique observed with I > 3ρ(I). Corrections for background, Lorentz effects and polarization were applied during data reduction with the CrysAlis software. Empirical absorption correction was performed using the same software yielding Rint = 0.034. The crystal structure was solved with a charge-flipping method using the program Superflip [5] and subsequently refined by the full-matrix least-squares algorithm of JANA2006 program [6]. Because of the similarity of atomic number of Pd and Ag (46 and 47, respectively), it is nearly impossible to distinguish between these atoms from single-crystal (MoKα radiation) diffraction data. The refinement indicated five metallic positions, which one of them was assigned as Bi. The remaining metallic sites show multiplicities 2:8:8:4. Considering the empirical chemical composition Pd8.91Ag2.03Bi2.06S4.00 (Z = 2) and coordination environment of the 4e site, which was very different from the others (see structure description), the 4e site was refined as Ag position. Next, refinement cycles included all anisotropic displacement parameters, which revealed too large a value for Pd(2) position (Ueq(Pd2) = 0.0146 Å2 cf. 0.0082 and 0.080 Å2 for Pd(1) and Pd(3), respectively). Refinement of occupancy factors yielded 0.88 occupancy for the Pd(2) position; other positions were found to be fully occupied. Final refinement in the I4/mmm space group for 21 parameters converged smoothly to the R = 0.0310 and wR = 0.0815 for 221 observed reflections. Details of data collection, crystallographic data, and refinement are given in Table 3.
Atom coordinates and displacement parameters are listed in Table 4. Table 5 shows selected bond lengths.
It should be noted that the refined tetragonal structure model of thalhammerite is only a substructure. As was revealed by subsequent Rietveld refinement (see below), the powder X-ray diffraction pattern of synthetic thalhammerite shows at medium and high diffraction angles a few very weak unindexed peaks and very subtle peak splitting, which cannot be fitted using the tetragonal model. Attempts to refine the structure from single-crystal data in rhombic subgroups of I4/mmm (i.e., Fmmm, Immm) led to negligible lowering of R-factors (e.g., from 0.0313 to 0.0293) with a rapid increase of the refined parameters and correlations between them. Refinements in monoclinic subgroups failed. Additionally, neither of these low-symmetry models describe all peak splitting observed in powder diffraction patterns of synthetic thalhammerite. Therefore, we proposed only the tetragonal average substructure of thalhammerite, leaving some aspects of the structure unclear.

5.2. Powder X-ray Ddiffraction

The powder XRD pattern of synthetic thalhammerite was collected in the Bragg-Brentano geometry on a Bruker D8 Advance diffractometer equipped with the LynxEye XE detector and CuKα radiation. The data were collected in the range from 10° to 100° 2θ with a step size of 0.005° 2θ and 2 s counting time per step. The structure model obtained from a single-crystal XRD study of synthetic thalhammerite was used as a starting structural model in the subsequent Rietveld refinement. The FullProf program [7] was used and the pseudo-Voigt function was used to generate the shape of the diffraction peaks. The refined parameters include those describing peak shape and width, peak asymmetry, unit-cell parameters, the occupancy parameter of the Pd(2) position, and six isotropic displacement parameters.
In total, 17 parameters were refined. No fractional coordinates were refined. The final cycles of Rietveld refinement converged to the agreement factors Rp = 0.077 and Rwp = 0.115. The refinement indicated 7 wt % Pd3Bi2S2 (I213) impurity in the investigated sample.
Figure 3 depicts two details of final Rietveld plot showing weak, however discernible, peak splitting at middle and high diffraction angles of 2θ (i.e., above 50°). Attempts to index all observed diffractions in the powder pattern in the large and/or lower symmetry unit-cell remained unsuccessful and, therefore, the structure refinement was limited to the tetragonal substructure. Table 6 presents powder diffraction data for thalhammerite.

6. Structure Description

The tetragonal substructure of thalhammerite contains three Pd, one Ag, Bi, and S sites, respectively. All sites, except the Pd(2) position, were found to be fully occupied. Its crystal structure is shown in Figure 4.

6.1. Coordination of Cations

The Pd(1) position is in the centre of regular square of S atoms with Pd(1)-S distances of 2.362(3) Å. The coordination is perfectly planar. Similar coordination was observed in vysotskite, PdS [8], which shows similar Pd–S separation of 2.34 Å. Such coordination geometry is typical for low-spin 4d8 Pd2+ cation in normal sulfides with M:S ratio equal to or smaller to one [9]. The Pd(1) coordination is further completed by two Ag atoms at 2.919(2) Å lying perpendicular to the [M(1)S4] squares.
The Pd(2) (refined to 0.88 occupancy of Pd) and Pd(3) sites form complex polyhedron. Both Pd positions are coordinated by two S atoms at distances 2.3372(7) and 2.343(3) Å, a value very close to the Pd–S distance of 2.334(4) Å observed for the zig-zag chains in the structure of kravtsovite, PdAg2S [1]. Whereas the S–Pd(2)–S group is perfectly linear, the S–Pd(3)–S shows a bonding angle of 177.9(1)°. Pd(2) is further coordinated by two Bi (2.8378(1) Å), two Ag (2.9073(4) Å), and two Pd(3) (3.0670(2) Å) atoms. Pd(3) also shows two Bi (2.8080(8) Å), two Ag (2.905(2) Å), and four Pd(3) (3.0670(2) Å) short contacts.
Ag site is surrounded by nine Pd atoms (Figure 5) forming a mono-capped tetragonal antiprismatic coordination. The Ag–Pd distances are in the range of 2.905(1) Å to 2.919(2) Å, comparable to those observed in lukkulaisvaaraite (Pd–Ag: 2.891(4)–3.037(4) Å; [10], where Ag atoms display tetragonal antiprismatic coordination.
As is shown in Figure 5, the Bi atom is coordinated by eight Pd atoms to form a bi-capped trigonal prism with Bi–Pd bond distances ranging from 2.808(1) to 2.8378(1) Å, values slightly shorter than those observed in structure of monoclinic PdBi (2.84–2.95) Å; [11]. There are no short (<3.5 Å) Bi–S contacts in the thalhammerite crystal structure. This contrasts with the environment of Bi in structure of chemically-related vymazalováite, Pd3Bi2S2 [2,12], where Bi atoms show one additional S contact at 3.22(3) Å.

6.2. Modular Description

The thalhamerite crystal structure forms a three-dimensional framework. It contains features typical for intermetallic compounds (e.g., complex crystallochemical environment of metals) and, therefore, cannot be presented using a traditional cation-based coordination polyhedra approach.
Alternatively, the structure of thalhammerite can be conveniently described as an arrangement of two types of building blocks (cuboids) having common S atoms at the corners (Figure 6).
The first block (green in Figure 6) contains the [PdS4] squares forming one face of the block and Ag atoms in its centre. Pd atoms are approximately located to the midpoints of the longer S-S edges. The second block (orange in Figure 6) contains Bi atoms in its centre. By analogy with the first block, the Pd atoms are located to the midpoints of the longer S–S edges. In the thalhammerite structure, two types of block alternate in a chess-boar fashion within the (001) plane and form chains along the c axis (Figure 6). It should be mentioned that, neglecting the Ag and Bi atoms, the packing of the green blocks automatically generates their duals, and the orange block, vice versa.

6.3. Relation to Other Minerals

The thalhammerite structure represents a unique structure type, and no exact structural analogue is hitherto known. It is worth noting that its structure merges structure motives typical for polar chalcogenides and intermetallic compounds. The [Pd(1)S4] square-planar coordination is a hallmark of Pd-bearing sulfides with an M:S ratio equal to, or slightly smaller than, one. Contrary to that, (almost) linear coordination of Pd by two S atoms and number of further metal-metal contacts resulting in complex coordination geometry, can be observed in sulphides with intermetallic behaviour (e.g., kravtsovite PdAg2S, Vymazalová et al., 2017 [1]).
Another chemically-related mineral, coldwellite, Pd3Ag2S (McDonald et al., 2015 [13]), adopts a cubic β-Mn-like structure and, hence, differs substantially from that of thalhammerite.

7. Proof of Identity of Natural and Synthetic Thalhammerite

The structural identity between the synthetic Pd9Ag2Bi2S4 and the natural material was confirmed by electron back-scattering diffraction (EBSD) and Raman spectroscopy.

7.1. Electron Back-Scattering Diffraction

The structural identity between the natural material and the synthetic Pd9Ag2Bi2S4 was confirmed by EBSD. A TESCAN Lyra 3GM field emission scanning electron microscope combined with EBSD system (Oxford Instruments AztecHKL system with NordlysNano EBSD camera) was used for the measurements. The surface of natural sample was prepared for investigation by broad beam argon ion milling using Gatan PECS II system operated at 1 kV. The solid angles calculated from the patterns were compared with our structural model for Pd9Ag2Bi2S4 synthetic phase match containing 12 reflectors to index the patters. The EBSD patterns (also known as Kikuchi patterns) obtained from the natural material (>50 measurements on different spots on natural thalhammerite grains) were found to match the patterns generated from our structural model for Pd9Ag2Bi2S4 synthetic phase, Figure 7.
The values of the mean angular deviation (MAD, i.e., goodness of fit of the solution) between the calculated and measured Kikuchi bands range between 0.22° and 0.48°. These values reveal a very good match; as long as values of mean angular deviation are less than 1°, they are considered as indicators of an acceptable fit (HKL Technology, 2004).

7.2. Raman Spectroscopy

The Raman spectroscopy technique was applied to verify the structural identity between the synthetic Pd9Ag2Bi2S4 and the natural material (Figure 8).
Raman spectra were obtained using a LABRAM (ISA Jobin Yvon) instrument installed at the University of Leoben, Austria. A frequency-doubled 100 mW Nd:YAG laser with an excitation of a wavelength of λ = 532.6 nm was used. The obtained Raman spectra of natural and synthetic Pd9Ag2Bi2S4 show four discernible absorption bands at the following values: 122, 309, 362, and 483 cm−1 (see Figure 8).
The EBSD study, Raman spectra, chemical identity and optical properties confirmed the identity of the natural and synthetic materials and thereby legitimise the use of the synthetic phase for the complete characterization of thalhammerite.

Author Contributions

All the authors (A.V., F.L., S.F.S., V.V.K., C.J.S., J.P., F.Z., G.G. and R.B.) discussed the obtained results, evaluated the data and wrote the article together. A.V. designed the article and conceived experiments; S.F.S. provided the samples and geological background; F.L., V.V.K., J.P. obtained the crystallographic data; C.J.S. provided optical properties; F.Z., G.G. and R.B. studied thalhammerite by Raman and evaluated the chemical data. All the authors revised and edited the manuscript.

Funding

The work was supported by the Grant Agency of the Czech Republic (project no. 18-15390S to A.V.), through an internal project 331400 from the Czech Geological Survey, the Russian Foundation for Basic Research (project RFBR 18-05-70073), and C.J.S. acknowledges Natural Environment Research Council, grant NE/M010848/1, Tellurium and Selenium Cycling and Supply.

Acknowledgments

The authors acknowledge Ulf Hålenius, Chairman of the CNMNC and its members for helpful comments on the submitted data. The authors are grateful to Zuzana Korbelová (Institute of Geology AS CR, v.v.i.) for carrying out the electron microprobe analyses. We thank two anonymous reviewers and the Editorial Board members, for their comments and improvements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vymazalová, A.; Laufek, F.; Sluzhenikin, S.F.; Stanley, C.J.; Kozlov, V.V.; Chareev, D.A.; Lukashova, M.L. Kravtsovite, PdAg2S, a new mineral from Norilsk-Talnakh deposit, Russia. Eur. J. Mineral. 2017, 29, 597–602. [Google Scholar] [CrossRef]
  2. Sluzhenikin, S.F.; Kozlov, V.V.; Stanley, C.J.; Lukashova, M.I.; Dicks, K. Vymazalováite, Pd3Bi2S2, a new mineral from Norilsk -Talnakh deposit, Krasnoyarskiy region, Russia. Mineral. Mag. 2018, 82, 367–373. [Google Scholar] [CrossRef]
  3. Sluzhenikin, S.F.; Mokhov, A.V. Gold and silver in PGE-Cu-Ni and PGE ores of the Norilsk deposit, Russia. Mineral. Depos. 2015, 50, 465–492. [Google Scholar] [CrossRef]
  4. Subbotin, V.V.; Gabov, D.A.; Korchagin, A.U.; Savchenko, E.E. Gold and Silver in the Composition of PGE Ores of the Fedorov-Pana Layered Intrusive Complex. Her. Kola Sci. Cent. RAS 2017, 1, 53–65. (In Russian) [Google Scholar]
  5. Palatinus, L.; Chapuis, G. SUPERFLIP—A computer program for the solution of crystal structures by charge flipping in arbitrary dimensions. J. Appl. Crystallogr. 2007, 41, 786–790. [Google Scholar] [CrossRef]
  6. Petříček, V.; Dušek, M.; Palatinus, L. Crystallographic Computing System JANA2006: General features. Z. Kristallogr. 2014, 229, 345–352. [Google Scholar] [CrossRef]
  7. Rodríguez-Carvajal, J. Full Prof. 2k Rietveld Profile Matching & Integrated Intensities Refinement of X-ray and/or Neutron Data (Powder and/or Single-Crystal); Laboratoire Léon Brillouin, Centre d’Etudes de Saclay: Gif-sur-Yvette, France, 2006. [Google Scholar]
  8. Brese, N.E.; Squattrito, P.J.; Ibers, J.A. Reinvestigation of the structure PdS. Acta Crystallogr. Sect. C 1985, C41, 1829–1830. [Google Scholar] [CrossRef]
  9. Dubost, V.; Balić-Žunić, T.; Makovicky, E. The crystal structure of Ni9.54Pd7.46S15. Can. Mineral. 2007, 45, 847–855. [Google Scholar] [CrossRef]
  10. Vymazalová, A.; Grokhovskaya, T.L.; Laufek, F.; Rassulov, V.A. Lukkulaisvaaraite, Pd14Ag2Te9, a new mineral from Lukkulaisvaara intrusion, northern Russian Karelia, Russia. Mineral. Mag. 2014, 78, 1743–1754. [Google Scholar] [CrossRef]
  11. Bhatt, Y.C.; Schubert, K. Kristallstruktur von PdBi.r. J. Less Common Met. 1979, 64, 17–24. [Google Scholar] [CrossRef]
  12. Weihrich, R.; Matar, S.F.; Eyert, V.; Rau, F.; Zabel, M.; Andratschke, M.; Anusca, I.; Bernert, T. Structure, ordering and bonding of half antiperovskites PbNi3/2S and BiNi3/2S. Prog. Solid State Chem. 2007, 35, 309–327. [Google Scholar] [CrossRef]
  13. McDonald, A.M.; Cabri, L.J.; Stanley, C.J.; Good, D.J.; Redpath, J.; Spratt, J. Coldwellite, Pd3Ag2S, a new mineral species from the Marathon deposit, Coldwell Complex, Ontario, Canada. Can. Mineral. 2015, 53, 845–857. [Google Scholar] [CrossRef]
Figure 1. Digital image in reflected plane polarized light showing inclusions of thalhammerite in galena (gn) in association with (a) chalcopyrite (ccp) and (b) vymazalováite (vym).
Figure 1. Digital image in reflected plane polarized light showing inclusions of thalhammerite in galena (gn) in association with (a) chalcopyrite (ccp) and (b) vymazalováite (vym).
Minerals 08 00339 g001aMinerals 08 00339 g001b
Figure 2. Reflectance data for thalhammerite compared to synthetic analogue, in air. The reflectance values (R%) are plotted versus the wavelength λ in nm.
Figure 2. Reflectance data for thalhammerite compared to synthetic analogue, in air. The reflectance values (R%) are plotted versus the wavelength λ in nm.
Minerals 08 00339 g002
Figure 3. Details of the Rietveld profiles of synthetic thalhammerite showing the weak peak splitting, which cannot be fitted using the tetragonal cell. The observed (circles), calculated (solid), and difference profiles are shown. The vertical bars correspond to Bragg reflections.
Figure 3. Details of the Rietveld profiles of synthetic thalhammerite showing the weak peak splitting, which cannot be fitted using the tetragonal cell. The observed (circles), calculated (solid), and difference profiles are shown. The vertical bars correspond to Bragg reflections.
Minerals 08 00339 g003
Figure 4. Crystal structure of thalhammerite showing the [PdS4] squares and Pd–S bonds. Unit-cell edges are highlighted. Details show the Rietveld profiles of synthetic thalhammerite showing the weak peak splitting, which cannot be fitted using the tetragonal cell.
Figure 4. Crystal structure of thalhammerite showing the [PdS4] squares and Pd–S bonds. Unit-cell edges are highlighted. Details show the Rietveld profiles of synthetic thalhammerite showing the weak peak splitting, which cannot be fitted using the tetragonal cell.
Minerals 08 00339 g004
Figure 5. Coordination polyhedra of Ag (mono-capped tetragonal antiprism) and Bi (bi-capped trigonal prism) in the thalhammerite structure.
Figure 5. Coordination polyhedra of Ag (mono-capped tetragonal antiprism) and Bi (bi-capped trigonal prism) in the thalhammerite structure.
Minerals 08 00339 g005
Figure 6. (a) Arrangement of two types of building blocks in the thalhammerite structure. (b) Detailed view showing the block containing Ag (green) and Bi (orange) atoms.
Figure 6. (a) Arrangement of two types of building blocks in the thalhammerite structure. (b) Detailed view showing the block containing Ag (green) and Bi (orange) atoms.
Minerals 08 00339 g006
Figure 7. EBSD image of natural thalhammerite; in the right pane, the Kikuchi bands are indexed.
Figure 7. EBSD image of natural thalhammerite; in the right pane, the Kikuchi bands are indexed.
Minerals 08 00339 g007
Figure 8. Comparison of Raman spectra in the synthetic Pd9Ag2Bi2S4 and in the natural material.
Figure 8. Comparison of Raman spectra in the synthetic Pd9Ag2Bi2S4 and in the natural material.
Minerals 08 00339 g008
Table 1. Reflectance data for natural and synthetic thalhammerite.
Table 1. Reflectance data for natural and synthetic thalhammerite.
NaturalSynthetic
λ (nm)R1 (%)R2 (%)R1 (%)R2 (%)
40040.041.240.642.1
42040.641.841.342.6
44041.142.342.043.2
46041.742.842.643.9
47041.943.042.944.3
48042.243.343.144.6
50042.743.943.845.3
52043.244.444.646.0
54043.744.945.346.6
54643.945.145.646.9
56044.245.445.947.2
58044.745.946.447.7
58944.946.146.747.9
60045.246.346.948.1
62045.646.847.348.5
64046.147.347.748.9
65046.347.547.949.1
66046.547.848.049.2
68047.048.348.349.5
70047.448.948.649.8
Note. The values required by the Commission on Ore Mineralogy are given in bold.
Table 2. Electron-microprobe analyses of natural and synthetic thalhammerite.
Table 2. Electron-microprobe analyses of natural and synthetic thalhammerite.
wt %PdAgPbBiSSeTotal
Thalhammerite
52.8014.572.6022.567.750.07100.35
53.4014.293.0522.097.620.03100.47
51.6414.256.1221.987.700.19101.87
average52.6114.373.9222.217.690.10100.90
13/B-92 *53.8512.51 24.847.90 99.1
52.7712.271.7724.297.450.5799.12
1/K-92 *53.8512.8 24.247.89 99.01
52.7711.83 25.737.99 99.53
54.0812.21 25.347.95 101.21
Synthetic Sample
Exp3754.1813.69 25.027.59 100.48
54.7412.91 25.047.50 100.19
54.6612.46 25.907.39 100.42
56.1412.01 24.707.36 100.21
55.7812.67 24.277.44 100.16
55.1012.75 24.997.46 100.29
average55.1012.75 24.997.46 100.30
* Sluzhenikin and Mohkov [2].
Table 3. Crystallographic data for the selected crystal of synthetic thalhammerite, Pd9Ag2Bi2S4.
Table 3. Crystallographic data for the selected crystal of synthetic thalhammerite, Pd9Ag2Bi2S4.
Crystal Data
Chemical formula (idealized)Pd9Ag2Bi2S4
Space groupI4/mmm (No. 139)
a [Å]8.0266(2)
c [Å]9.1531(2)
V3]589.70(2)
Z2
Crystal size (mm)0.034 × 0.027 × 0.013
Data Collection
DiffractometerSuperNova
Temperature (K)293
RadiationMoKα (0.7107 Å)
Theta range (°)5.08–27.62
Reflections collected3659
Independent reflections 226
Unique observed reflections [I > 3(σ)]221
Index ranges−10 < h < 10
−10 < k < 10
−11 < l < 11
Absorption correction methodEmpirical
Structure Refinement
Refinement methodFull matrix least-squares on F2
Parameters/restrains/constrains21/0/0
R, wR (obs)0.0310/0.0815
R, wR (all)0.0318/0.0817
Largest diff. peak and hole (e3)1.20/−5.20
Table 4. Fractional coordinates and anisotropic displacement parameters (Å2) for synthetic thalhammerite.
Table 4. Fractional coordinates and anisotropic displacement parameters (Å2) for synthetic thalhammerite.
AtomPd(1)Pd(2) *Pd(3)AgBiS
Wyckoff Position2a8f8j4e4d8h
x1/21/41/21/21/20.2081(4)
y1/21/40.2027(2)1/200.2081(4)
z1/21/400.1810(2)1/40
U110.0069(8)0.0104(8)0.0077(7)0.0098(6)0.0090(4)0.0071(12)
U220.0069(8)0.0104(8)0.0097(7)0.0098(6)0.0090(4)0.0071(12)
U330.0109(13)0.0051(10)0.0077(7)0.0082(9)0.0079(6)0.012(2)
U1200.0017(6)000−0.0018(15)
U1300.0011(4)0000
U2300.0011(4)0000
Ueq0.0083(6)0.0086(50.0084(4)0.0106(4)0.0086(3)0.0087(9)
* Refined with 0.88 occupancy.
Table 5. Selected bond distances (Å) in the thalhammerite crystal structure.
Table 5. Selected bond distances (Å) in the thalhammerite crystal structure.
Pd(1)4 × S2.362(3)Ag4 × Pd(3)2.905(1)
2 × Ag2.919(2) 4 × Pd(2)2.9073(4)
Pd(2)2 × S2.3372(7)
2 × Bi2.8378(1)Bi14 × Pd(3)2.808(1)
2 × Ag2.9073(4) 4 × Pd(2)2.8378(1)
4 × Pd(3)3.0670(2)
Pd(3)2 × S2.343(3)
2 × Bi2.8080(8)
2 × Ag2.905(2)
4 × Pd(2)3.0670(2)
Table 6. X-ray powder diffraction data of thalhammerite (CuKα radiation, Bruker D8 Advance, Bragg-Brentano geometry). Only reflections with I(obs) ≥ 1 are listed.
Table 6. X-ray powder diffraction data of thalhammerite (CuKα radiation, Bruker D8 Advance, Bragg-Brentano geometry). Only reflections with I(obs) ≥ 1 are listed.
I(obs)hkld(meas)d(calc)
111016.03646.0338
111105.67905.6767
130024.57524.5736
82004.01554.0140
181123.56203.5615
242113.34283.3420
22023.01813.0169
91032.85102.8504
462202.83932.8383
213012.56852.5684
1002222.41222.4117
611232.32452.3241
480042.28732.2868
291322.22012.2197
22312.16372.1634
171142.12132.1212
404002.00722.0070
33301.89231.8922
84021.83771.8378
152331.79811.7982
182241.78051.7807
233321.74811.7485
41341.69911.6991
24221.67111.6710
51431.64131.6410
12151.62991.6300
44311.58141.5814
15101.57431.5744
80351.51021.5103
304041.50851.5085
91161.47231.4724
134401.41931.4192
74421.35541.3554
122261.34311.3431
92531.33951.3393
13521.31851.3184
193161.30701.3070
71541.29691.2968
96201.26941.2693
31271.22791.2279
186221.22311.2231
11631.21131.2112
104441.20591.2058
113361.18721.1871

Share and Cite

MDPI and ACS Style

Vymazalová, A.; Laufek, F.; Sluzhenikin, S.F.; Kozlov, V.V.; Stanley, C.J.; Plášil, J.; Zaccarini, F.; Garuti, G.; Bakker, R. Thalhammerite, Pd9Ag2Bi2S4, a New Mineral from the Talnakh and Oktyabrsk Deposits, Noril’sk Region, Russia. Minerals 2018, 8, 339. https://doi.org/10.3390/min8080339

AMA Style

Vymazalová A, Laufek F, Sluzhenikin SF, Kozlov VV, Stanley CJ, Plášil J, Zaccarini F, Garuti G, Bakker R. Thalhammerite, Pd9Ag2Bi2S4, a New Mineral from the Talnakh and Oktyabrsk Deposits, Noril’sk Region, Russia. Minerals. 2018; 8(8):339. https://doi.org/10.3390/min8080339

Chicago/Turabian Style

Vymazalová, Anna, František Laufek, Sergey F. Sluzhenikin, Vladimir V. Kozlov, Chris J. Stanley, Jakub Plášil, Federica Zaccarini, Giorgio Garuti, and Ronald Bakker. 2018. "Thalhammerite, Pd9Ag2Bi2S4, a New Mineral from the Talnakh and Oktyabrsk Deposits, Noril’sk Region, Russia" Minerals 8, no. 8: 339. https://doi.org/10.3390/min8080339

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