Next Article in Journal
Mine Backfilling in the Permafrost, Part I: Numerical Prediction of Thermal Curing Conditions within the Cemented Paste Backfill Matrix
Next Article in Special Issue
Formation Conditions and 40Ar/39Ar Age of the Gem-Bearing Boqueirão Granitic Pegmatite, Parelhas, Rio Grande do Norte, Brazil
Previous Article in Journal
Effects of Single and Mixed Energy Sources on Intracellular Nanoparticles Synthesized by Acidithiobacillus ferrooxidans
Previous Article in Special Issue
Emerald Deposits: A Review and Enhanced Classification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gem-Quality Green Cr-Bearing Andradite (var. Demantoid) from Dobšiná, Slovakia

1
Gemmological Institute, Faculty of Natural Sciences, Constantine the Philosopher University in Nitra, Nábrežie mládeže 91, 949 74 Nitra, Slovakia
2
Department of Mineralogy and Petrology, Faculty of Natural Sciences, Comenius University in Bratislava, Ilkovičova 6, Mlynská dolina, 842 15 Bratislava, Slovakia
3
Gemological Laboratory of e-gems.cz, 110 00 Prague, Czech Republic
4
Earth Science Institute of the Slovak Academy of Sciences, Ďumbierska 1, 974 01 Banská Bystrica, Slovakia
5
Department of Geological Sciences, Masaryk University, Kotlářská 2, 611 37 Brno, Czech Republic
6
Central European Institute of Technology, Masaryk University, Kamenice 5, 625 00 Brno, Czech Republic
7
Department of Environmental Ecology, Faculty of Natural Sciences, Comenius University in Bratislava, Ilkovičova 6, Mlynská dolina, 842 15 Bratislava, Slovakia
*
Author to whom correspondence should be addressed.
Minerals 2019, 9(3), 164; https://doi.org/10.3390/min9030164
Submission received: 11 February 2019 / Revised: 1 March 2019 / Accepted: 4 March 2019 / Published: 8 March 2019
(This article belongs to the Special Issue Mineralogy and Geochemistry of Gems)

Abstract

:
Andradite, variety demantoid, is a rare gem mineral. We describe gem-quality garnet crystals from serpentinized harzburgites from Dobšiná, Slovakia which were faceted. Both the andradite samples were transparent, with a vitreous luster and a vivid green color. They were isotropic with refractive indices >1.81. The measured density ranged from 3.82 to 3.84 g·cm−3. Andradite var. demantoid appeared red under Chelsea filter observation. Both samples contained fibrous crystalline inclusions with the typical “horsetail” arrangement. The studied garnet had a strong Fe3+ dominant andradite composition with 1.72–1.85 apfu Fe3+, Cr3+ up to 0.15 apfu, Al3+ 0.03 to 0.04 apfu, V3+ up to 0.006 apfu substituted for Fe3+, Mn2+ up to 0.002 apfu, and Mg up to 0.04 apfu substituted for Ca. Raman spectrum of garnet showed three spectral regions containing relatively strong bands: I—352–371 cm−1, II—816–874 cm−1, and III—493–516 cm−1. The optical absorption spectrum as characterized by an intense band at 438 nm and two broad bands at 587 and 623 nm and last one at 861 nm, which were assigned to Fe3+ and Cr3+. Transmission was observed in the ultraviolet spectral region (<390 nm), near the infrared region (700–800 nm), and around 530 nm in the green region of visible light, resulting in the garnet’s green color.

1. Introduction

Slovakia has only a few occurrences which produce gem-quality minerals. For a long time in history, precious opal from Červenica-Dubník was the only mineral considered as a gemstone [1,2,3,4,5,6,7]. Only a few Slovak minerals, including quartz var. smoky quartz [8,9], sphalerite [8,10], pyrite [8], garnet var. almandine [8], hematite [8], and corundum var. sapphire [11,12,13] are suitable for faceting.
Andradite garnet, Ca3Fe2Si3O12, is a rock-forming mineral typically found in metamorphic rocks such as skarns and rodingites. The andradite crystal structure consists of alternating SiO4 tetrahedra and FeO6 octahedra, and Ca with eight-fold coordination. Several studies have been carried out on the crystal chemistry and the thermodynamic properties of andradite garnets [14,15,16,17,18,19,20,21,22]. Green to yellow-green andradite garnets, demantoid variety, represent one of the most appreciated and precious gemstones among the garnet-group minerals, due to their color, brilliance, and rarity [23,24,25]. Gem-quality andradite-garnet sources occur in Russia [26,27,28,29], Italy [30], Iran [31,32,33], Pakistan [25,34,35,36,37], Namibia [38,39,40,41], Madagascar [42,43,44,45,46,47], Afganistan [24], Switzerland [24], Canada [48], Mexico [49], and China [50].
Our goal is a mineralogical and crystal-chemical investigation of andradite var. demantoid from a serpentinite body in Dobšiná, Slovakia using various analytical methods.

Geological Setting

The serpentinized harzburgites from Dobšiná are part of the Jurassic mélange complex (Meliata Unit). This inner Western-Carpathian unit forms small nappes over the Central Western Carpathians [51]. The Meliata Ocean formed after the continental break up in the Anisian [52], probably as a back-arc basin, due to northward subduction of the Paleotethys oceanic crust beneath the Eurasian continent [53]. The back-arc basin (BAB) basalts and mid-ocean ridge basalts (MORB) were dominant in the Meliatic oceanic rift [54,55]. The association of eupelagic and deepwater pelagic sediments, basalts, dolerite dikes, gabbros, and serpentinized mantle peridotites found in the Meliata Unit clearly indicates an oceanic suite [56,57]. The geochemistry of the basalts reveals a normal mid-ocean ridge basalts (N-MORB) [(La/Sm) n = 0.6–0.9] composition [54,55]. Traces of obducted oceanic crust (blocks of basalts alternating with radiolarite layers) were also identified from deepwater sedimentary breccias with a radiolaritic matrix [58]. The Neotethyan Meliata Oceanic back-arc basin closed in the Late Jurassic, according to 40Ar–39Ar ages of white mica from the blueschists and U/Pb LA-ICP-MS age of metamorphic-metasomatic perovskite [59,60]. The Meliatic subduction–accretionary wedge [61] was incorporated into the Cretaceous orogenic wedge of the Central Western Carpathians [62] between 130 and 50 Ma dated by 40Ar–39Ar method on white micas [63] in the form of mostly small nappe-sheeted bodies, thrust over the Gemeric and Veporic units of the Central Western Carpathians [64,65]. The zircon (U-Th)/He data recorded three evolutionary stages: (i) cooling through the ~180 °C isotherm at 130–120 Ma related to the startling collapse of the accretionary wedge, following exhumation of the high-pressure slices in the Meliatic accretionary wedge; (ii) postponed exhumation and cooling of some fragments through the ~180 °C isotherm from 115 to 95 Ma due to ongoing collapse of this wedge; and (iii) cooling from 80 to 65 Ma, postdating the thrusting (~100–80 Ma) during the Late Cretaceous compression related to formation of the Central Western Carpathians orogenic wedge [66].
The Dobšiná quarry (48°49.622′ N, 20°21.977′ E) (Figure 1) is in a tectonic fragment of the Meliata tectonic unit overlying the Gemeric tectonic unit. The locality is famous for the occurrence of yellow-green Cr-rich andradite (demantoid) in serpentinite veins [67,68,69]. Andradite-bearing serpentinites are part of the Late Jurassic (to Early Cretaceous) serpentinite–blueschist mélange complex. The mélange complex contains talc–phengite–glaucophane schists, marbles, metaconglomerates, blueschists of magmatic and sedimentary origin, and serpentinites. All rocks form decimeter- to 100 m-sized fragments within the Late Jurassic very low-temperature greenschist-facies metasediments [70].
Partly-preserved magmatic structures of harzburgite consist of olivine (~70 vol. %), orthopyroxene (~25 vol. %), accessory spinel (~4 vol. %), and rare clinopyroxene (~1 vol. %). Magmatically corroded and partly dissolved porphyric orthopyroxene (±clinopyroxene) is enclosed in olivine matrix with spinel. Both pyroxenes show exsolution lamellae, of clinopyroxene in orthopyroxene, or orthopyroxene in clinopyroxene, respectively. Orthopyroxene often forms sigmoidally rotated porphyroclasts, including clinopyroxene exsolution lamellae. The serpentinized harzburgite contains serpentinite group minerals (antigorite replaced by chrysotile and lizardite), Cr-Fe spinel, perovskite, pyrophanite, Ti andradite, magnetite, and hematite, and rarely carbonates with talc, quartz, and characteristic relics of olivine, orthopyroxene, clinopyroxene, and spinel [70,71].

2. Materials and Methods

2.1. Sample Description

Two pieces of transparent andradite var. demantoid were separated from the host rock and faceted into round brilliant cuts with dimensions of 2.80 × 2.80 × 1.64 mm (0.11 carat) and 1.71 × 1.74 × 1.14 mm (0.025 carat) dimensions (Figure 2).

2.2. Analytical Methods

The faceted samples were examined by standard gemmological methods at the Gemmological Institute, Faculty of Natural Sciences, Constantine the Philosopher University in Nitra, to determine their optical properties (optical character and refractive indices), hydrostatic specific gravity, ultraviolet fluorescence, and microscopic features. The refractive indices were measured with a refractometer Kruss and Presidium Refractive Index Meter II. A Kern ABT 120-5DM with extension KERN ABT A01 hydrostatic balance was used to determine the density of the gemstones. The ultraviolet fluorescence was investigated with a short (254 nm) and long (366 nm) wavelength ultraviolet lamp. We used a Chelsea filter.
The chemical composition of andradite garnet was measured using a Cameca SX100 electron microprobe operated in wavelength-dispersion mode at the Masaryk University, Brno, Czech Republic, under the following conditions: accelerating voltage 15 kV, beam current 20 nA, and beam diameter 5 μm. The samples were analyzed with the following standards: fluorapatite (PKα), sanidine (SiKα, AlKα, KKα), titanite (TiKα), chromite (CrKα), vanadinite (VKα), YAG (YLα), almandine (FeKα), spessartine (MnKα), forsterite (Mg Kα), gahnite (ZnKα), wollastonite (CaKα), albite (NaKα), and topaz (FKα). Lower detection limits of the measured elements varied between 200–400 ppm except for Na (530–570 ppm), Fe (670–730 ppm), Zn (880–900 ppm), and Y (470–500 ppm). The measured content of Y, Mn, Zn, Na, and K was always below the detection limit. Analytical times were 15 to 40 s during measurement depending on the expected concentration of the element in the mineral phase. Major elements were measured using shorter times, whereas longer times were applied for elements with a low concentration. The andradite chemical formula was calculated based on 12 anions assuming all Fe as ferric.
Instrumentation for laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) at the Department of Chemistry, Masaryk University, Brno, consists of a UP 213 laser ablation system (New Wave Research, Inc., Fremont, CA, USA) and Agilent 7500 CE ICP-MS spectrometer (Agilent Technologies, Santa Clara, CA, USA). A commercial Q-switched Nd-YAG laser ablation device worked at the 5th harmonic frequency corresponding to 213 nm wavelength. The ablation device was equipped with a programmable XYZ-stage to move the sample along a programmed trajectory during ablation. Target visual inspection and photographic documentation were achieved by the built-in microscope/CCD-camera system. A sample was enclosed in the SuperCell and was ablated by the laser beam, which was focused onto the sample surface through a quartz window. The ablation cell was flushed with a helium carrier gas which transported the laser-induced aerosol to the inductively coupled plasma (1 L·min−1). Sample argon gas flow was admixed with the helium carrier gas flow after the laser ablation cell to 1.6 L·min−1 total gas flow. NIST SRM 610 silicate glass reference material was used to optimize the gas flow rates, the sampling depth, and MS electrostatic lens voltage in LA-ICP-MS conditions. This provided maximum signal-to-noise ratio and minimum oxide formation (ThO+/Th+ count ratio 0.2%, U+/Th+ counts ratio 1.1%). Laser ablation required a 100 μm laser spot diameter, 8 J·cm−2 laser fluence, and a 20 Hz repetition rate. The fixed sample position during laser ablation enabled 60 s hole-drilling duration for each spot. All element measurements were normalized on 28Si in the investigated andradite. Analyses were gained from 10 spots in a line across the crystal.
Raman analyses were performed on the LabRAM-HR Evolution (Horiba Jobin-Yvon, Kyoto, Japan) spectrometer system with a Peltier cooled CCD detector and Olympus BX-41 microscope (Department of Geological Sciences, Masaryk University). Raman spectra excited by blue diode laser (473 nm) in the range of 50–4000 cm−1 were collected from each stone using 50× objectives. The acquisition time of 10 s per frame and 2 accumulations were used to improve the signal-to-noise ratio.
Optical absorption spectra of cut stones in the region (400–750 nm) were measured with the GL Gem SpectrometerTM at room temperature in the Gemmological Institute of Constantine the Philosopher University, Nitra. Both Raman and UV/Vis/NIR spectra were processed in Seasolve PeakFit 4.1.12 software. Raman and absorption bands were fitted by Lorentz function with the automatic background correction and Savitsky-Golay smoothing.

3. Results

Both the andradite samples were transparent, with a vitreous luster and a yellow-green color, which defines the andradite variety. They were all isotropic with refractive indices >1.81. The measured density ranged from 3.82 to 3.84 g·cm−3, and the gemstones were inert in shortwave and longwave ultraviolet radiation. Andradite var. demantoid appeared red under Chelsea filter observation, which is typical for Cr-bearing gemstones. Both samples contained fibrous crystalline inclusions with the typical “horsetail” arrangement (Figure 3).
The studied garnet had a strongly Fe3+ dominant andradite composition with 1.72–1.85 apfu Fe3+ (Table 1 and Figure 4). Only Cr3+ (up to 0.15 apfu), Al3+ (0.03 to 0.04 apfu), and Mg (up to 0.04 apfu) substituted for Fe3+. From the trace elements, Cr was the most abundant, along with moderate to minor concentrations (in relative decreasing order of abundance) of V, Ti, and Mn. The other trace elements measured (Sc, Ge, As, Y, and Gd) were near or below their respective detection limit (Table 2). There was no specific chemical zoning of crystal observed in major or trace elements.
From the calculated crystal-chemical formulae, it is obvious that Z site occupancy by Si is slightly lower than three. It could be filled to three by Al or Fe3+, but their content is sufficient only to accommodate the Y site. Even after adding all possible octahedral cations, the Y-site sum is always slightly below two which, however, could result from the analytical error. However, the small deficiency at the Z site can be justified by the presence of F. The X site is almost uniformly occupied by Ca with a very small admixture of Mg.
Raman spectrum of garnet (Figure 5 and Table 3) showed three spectral regions containing relatively strong bands: I—352–371 cm−1, II—816–874 cm−1, III—493–516 cm−1. The other bands exhibited relatively lower intensities. Measured bands were assigned to various molecular motions (Table 3): translation of SiO4—T(SiO4), rotation of SiO4—R(SiO4), translation of YO6 (Y = Fe3+, Cr3+)—T(M), symmetric band ν2, asymmetric stretching frequency ν3, asymmetric bending frequency ν4 of SiO4 group. Our spectrum is consistent with published andradite spectra [72], except the 343 cm−1 band, which is not present in published data. Only bands with similar Raman shift are in almandine and pyrope spectra associated with rotation of SiO4 [72].
The optical absorption spectrum (Figure 6 and Table 4) is characterized by small absorption features below 370 nm, an intense band at 438 nm with shoulders at 470 and 488 nm, and two broad bands at 587 and 623 nm, and 861 nm with a shoulder at 964 nm. These spectral features were all assigned to Fe3+ and Cr3+. Transmission was observed in the near the infrared region (700–800 nm), and around 530 nm in the green region of visible light resulting in the garnet’s green color.

4. Discussion and Conclusions

Chromium-bearing garnets are relatively rare and are associated with specific geological environments and geochemical conditions. In the Western Carpathians, only a few occurrences have been described, including Pezinok–Rybníček [73], Čierna Lehota [74], and Dobšiná [67].
The garnet from Dobšiná occurs as the three following types [67]. The first type is present as relatively evenly dispersed minute grains in the rock. It originated preferentially at the expense of pseudomorphs after rhombic pyroxene. The garnet also forms lenticular or elongated layers 1 to 2 mm thick and several centimeters long. The second-type of garnet occurs as infilling of chrysotile veinlets. The veinlets’ thickness is up to 3 mm. The third type of garnet forms monomineral fills in 0.3 to 3 mm-thick fissures. The garnet crystals display usually euhedral habit. However, only the third type of garnets are emerald-green color. All three garnet types occur in pale-colored, pale-green varieties of serpentinite.
Andradite color differences are caused by the variable Cr content. The studied garnet belongs to the third type which is generally enriched in Cr compared to the other two types. In some crystals of andradite, the determined value exceeded 6 wt % Cr2O3 [67], but the studied crystals only contain up to 2.3 wt % Cr2O3. However, this content is still significantly higher than in other demantoid occurrences around the world [21,30,47]. Moreover, demantoid from Dobšiná is enriched in V (up to 0.09 wt % V2O3 from EMPA, 305–572 ppm from LA-ICP-MS) compared to other localities with V usually below the detection limit of EMPA and below 270 ppm determined by LA-ICP-MS [21,30,47]. The As content between 10–22 ppm (although very close to the detection limit, were below detection limit on other localities [21]) and absence of REE (except very low Y), which are usually between 1–90 ppm [21] could also be a fingerprint feature for Dobšiná demantoid.
Raman mode analysis predicts the existence of 25 active Raman modes in all garnet species 3A1g, 8Eg, and 14T2g. The vibrational spectra of garnet clearly possess two different groups of vibrational modes. The first group extends from 816 to 995 cm−1 in the Raman spectrum and possesses three strong bands and one weak one. In the garnet structure, these modes are commonly assigned to the ν1 Si–O stretching motions within SiO4 groups [72,75,76,77,78]. The second region extends from 174 to 576 cm−1 and possesses three strong and ten weak bands. The bands at 352, 452, 493, 516, 552, and 576 cm−1 are commonly assigned to the Si–O bending motions within SiO4 groups [72]. The bands at 343 and 371 cm−1 can be assigned to rotation of the [SiO4]4− tetrahedron R[SiO4]4− [72]. The bands at 174, 236, 265, 297, and 311 cm−1 can be assigned to translation modes of the tetrahedra and octahedra [72].
The optical absorption spectra of studied garnet revealed a few absorption and transmission regions. The absorption feature at 438 nm and its shoulders are attributed to octahedral Fe3+, although a further contribution from Cr3+ cannot be excluded [22,50,75,79,80,81,82], but this was not very significant due to the low Cr content. The weak absorption band at 488 nm is related to Fe3+ [30,47]. Two bands at 587 and 623 nm belong to octahedral Fe3+ and Cr3+, respectively [22,50,75,79,80,81,82]. Presumed contribution of Cr3+ to the band at 438 nm can be assigned, according to the Tanabe–Sugano diagram [83,84,85], to the dominant 4A2g4T2g transition in Cr3+. The second 4A2g4T1g transition in Cr3+ was at 623 nm but was overlapping with the Fe3+ absorption band at 587 nm. Moreover, a broad band at about 861 nm was due to octahedral Fe3+ [22,50,75,79,80,81,82].

Author Contributions

Concept of study, J.Š., R.H., J.F., and P.B.; methodology, J.Š., R.H., J.F., P.B, S.M., R.Š., T.V.; formal analysis, S.M., R.Š., T.V., and P.B.; formal investigation, J.Š., Ľ.I., J.F., and P.B.; compilation and writing of the manuscript, J.Š., J.F., and P.B.; review and editing of manuscript, J.Š., Ľ.I., J.F., S.Č., and P.B.

Funding

This present research was supported by the Ministry of Education of Slovak Republic grant agency under the contracts KEGA 062UKF-4/2016, VEGA-1/0079/15, VEGA-1/0499/16 and VEGA-1/0151/19.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kaličiak, M.; Ďud’a, R.; Burda, P.; Kaličiaková, E. Structural geological characteristics of the Dubník opal deposits. Zborník Vychodoslovenského Múzea v Košiciach 1976, 17, 7–22. [Google Scholar]
  2. Rondeau, B.; Fritsch, E.; Guiraud, M.; Renac, C. Opals from Slovakia (“Hungarian” opals): A re-assessment of the conditions of formation. Eur. J. Mineral. 2004, 16, 789–799. [Google Scholar] [CrossRef]
  3. Gaillou, E.; Fritsch, E.; Aguilar-Reyes, B.; Rondeau, B.; Post, J.; Barreau, A.; Ostroumov, M. Common gem opal: An investigation of micro- to nano-structure. Am. Mineral. 2008, 93, 1865–1873. [Google Scholar] [CrossRef]
  4. Gaillou, E.; Delaunay, A.; Rondeau, B.; Bouhnik-le-Coz, M.; Fritsch, E.; Cornen, G.; Monnier, C. The geochemistry of gem opals as evidence of their origin. Ore Geol. Rev. 2008, 34, 113–126. [Google Scholar] [CrossRef] [Green Version]
  5. Kiefert, L.; Karampelas, S. Use of the Raman spectrometer in gemmological laboratories: Review. Spectrochim. Acta-Part A Mol. Biomol. Spectrosc. 2011, 80, 119–124. [Google Scholar] [CrossRef] [PubMed]
  6. Caucia, F.; Marinoni, L.; Leone, A.; Adamo, I. Investigation on the gemological, physical and compositional properties of some opals from Slovakia (“Hungarian” opals). Periodico di Mineralogia 2013, 82, 251–261. [Google Scholar]
  7. Caucia, F.; Ghisoli, C.; Marinoni, L.; Bordoni, V. Opal, a beautiful gem between myth and reality. Neues Jahrbuch fur Mineralogie Abhandlungen 2013, 190, 1–9. [Google Scholar] [CrossRef]
  8. Ďuďa, R. Gemmological and genetic classification of precious and trim stones of Slovakia. Mineral. Slovaca 1987, 19, 353–362. (In Slovakian) [Google Scholar]
  9. Fridrichová, J.; Bačík, P.; Illášová, Ľ.; Kozáková, P.; Škoda, R.; Pulišová, Z.; Fiala, A. Raman and optical spectroscopic investigation of gem-quality smoky quartz crystals. Vib. Spectrosc. 2016, 85, 71–78. [Google Scholar] [CrossRef]
  10. Fridrichová, J.; Bačík, P.; Malíčková, I.; Illášová, Ľ.; Pulišová, Z. Optical-spectroscopic study of gem sphalerite from Banská Štiavnica and Hodruša-Hámre. Esemestník 2017, 6, 13–15. (In Slovakian) [Google Scholar]
  11. Uher, P.; Sabol, M.; Konečný, P.; Gregáňová, M.; Táborský, Z.; Puškelová, Ľ. Sapphire from Hajnáčka Cérová Highland, southern Slovakia. Slovak Geol. Mag. 1999, 5, 273–280. [Google Scholar]
  12. Uher, P.; Sabol, M.; Konečný, P.; Gregáňová, M.; Táborský, Z.; Puškelová, Ľ. Zafír vo výplni vrchnopliocénneho maaru pri Hajnáčke. Miner. Slovaca 2001, 33, 307–308. [Google Scholar]
  13. Uher, P.; Giuliani, G.; Szakáll, S.; Fallick, A.; Strunga, V.; Vaculovič, T.; Ozdín, D.; Gregáňová, M. Sapphires related to alkali basalts from the Cerová Highlands, Western Carpathians (southern Slovakia): Composition and origin. Geol. Carpath. 2012, 63, 71–82. [Google Scholar] [CrossRef]
  14. Quareni, S.; De Pieri, R. La struttura dell’andradite. Atti. e Memorie dell’. Accademia Patavina di Science Lettere ed Arti 1966, 78, 151–170. [Google Scholar]
  15. Novak, G.A.; Gibbs, G.V. The crystal chemistry of the silicate garnets. Am. Mineral. 1971, 56, 791–825. [Google Scholar]
  16. Hazen, R.M.; Finger, L.W. High-pressure crystal chemistry of andradite and pyrope: Revised procedure for high-pressure diffraction experiments. Am. Mineral. 1989, 74, 352–359. [Google Scholar]
  17. Lager, G.A.; Armbruster, T.; Rotella, F.J.; Rossman, G.R. OH substitution in garnets: X-ray and neutron diffraction, infrared, and geometric-modeling studies. Am. Mineral. 1989, 74, 840–851. [Google Scholar]
  18. Armbruster, T.; Geiger, C.A. Andradite crystal chemistry, dynamic X-site disorder and structural strain in silicate garnets. Eur. J. Mineral. 1993, 5, 59–71. [Google Scholar] [CrossRef]
  19. Armbruster, T.; Birrer, J.; Libowitzky, E.; Beran, A. Crystal chemistry of Ti-bearing andradites. Eur. J. Mineral. 1998, 10, 907–921. [Google Scholar] [CrossRef]
  20. Pavese, A.; Diella, V.; Pischedda, V.; Merli, M.; Bocchio, R.; Mezouar, M. Pressure-temperature equation of state of andradite and grossular, by high-pressure and -temperature powder diffraction. Phys. Chem. Miner. 2001, 28, 242–248. [Google Scholar] [CrossRef]
  21. Bocchio, R.; Adamo, I.; Diella, V. Trace-element profiles in gem-quality green andradite from classic localities. Can. Mineral. 2010, 48, 1205–1216. [Google Scholar] [CrossRef]
  22. Adamo, I.; Gatta, G.D.; Rotiroti, N.; Diella, V.; Pavese, A. Green andradite stones: Gemmological and mineralogical characterisation. Eur. J. Mineral. 2011, 23, 91–100. [Google Scholar] [CrossRef]
  23. Balčiūnaitė, I.; Kleišmantas, A.; Norkus, E. Chemical composition of rare garnets, their colours and gemmological characteristics. Chemija 2015, 26, 18–24. [Google Scholar]
  24. O’Donoghue, M. Gems, 6th ed.; Butterworth-Heinemann: Oxford, UK, 2006; p. 936. [Google Scholar]
  25. Adamo, I.; Bocchio, R.; Diella, V.; Caucia, F.; Schmetzer, K. Demantoid from Balochistan, Pakistan: Gemmological and Mineralogical Characterization. J. Gemmol. 2015, 34, 428–433. [Google Scholar] [CrossRef]
  26. Phillips, W.R.; Talantsev, A.S. Russian demantoid, czar of the garnet family. Gems Gemol. 1996, 32, 100–111. [Google Scholar] [CrossRef]
  27. Krzemnicki, M.S. Diopside needles as inclusions in demantoid garnet from Russia: A Raman microspectrometric study. Gems Gemol. 1999, 35, 192–195. [Google Scholar] [CrossRef]
  28. Laurs, B.M. A new find of demantoid at a historic side in Kladovka, Russia. Gems Gemol. 2003, 39, 54–55. [Google Scholar]
  29. Mayerson, W.M. Lab Notes: Large demantoid of exceptional color. Gems Gemol. 2006, 42, 261–262. [Google Scholar]
  30. Adamo, I.; Bocchio, R.; Diella, V.; Pavese, A.; Vignola, P.; Prosperi, L.; Palanza, V. Demantoid from Val Malenco, Italy: Review and update. Gems Gemol. 2009, 45, 280–287. [Google Scholar] [CrossRef]
  31. Douman, M.; Dirlam, D. Update on demantoid and cat´s-eye from Iran. Gems Gemol. 2004, 40, 67–68. [Google Scholar]
  32. Du Toi, G.; Mayerson, W.; van Der Bogert, C.; Douman, M.; Befi, R.; Koivula, J.I.; Kiefert, L. Demantoid from Iran. Gems Gemol. 2006, 42, 131. [Google Scholar]
  33. Karampelas, S.; Gaillou, E.; Fritsch, E.; Douman, M. Les grenats andradites-démantoïde d’Iran: Zonage de couleur et inclusions. Rev. Gemmol. 2007, 160, 14–20. [Google Scholar]
  34. Fritz, E.A.; Laurs, B.M. Gem News International: Andradite from Balochistan, Pakistan. Gems Gemol. 2007, 43, 373. [Google Scholar]
  35. Milisenda, C.C.; Henn, U.; Henn, J. Demantoide aus Pakistan. Gemmol. Z. Dtsch. Gemmol. Ges. 2001, 50, 51–56. [Google Scholar]
  36. Quinn, E.P.; Laurs, B.M. Demantoid from northern Pakistan. Gems Gemol. 2005, 41, 176–177. [Google Scholar]
  37. Palke, A.C.; Pardieu, V. Gem News International: Demantoid from Baluchistan Province in Pakistan. Gems Gemol. 2014, 50, 302–303. [Google Scholar]
  38. Lind, T.; Henn, U.; Bank, H. New occurrence of demantoid in Namibia. Aust. Gemmol. 1998, 20, 75–79. [Google Scholar]
  39. Cairncross, B.; Bahmann, U. The Erongo Mountains, Namibia. Mineral. Rec. 2006, 37, 361–470. [Google Scholar]
  40. Koller, F.; Niedermayar, G.; Pintér, Z.; Syabó, C. The demantoid garnets of the Green Dragon Mine (Tubussi, Erongo region, Namibia). Acta Mineral.-Petrogr. 2012, 7, 72. [Google Scholar]
  41. Kaiser, H.; Enzersdorf, M. Demantoid aus der Green Dragon Mine, Erongo Region, Namibia. Gemmo News 2012, 32, 5. [Google Scholar]
  42. Danet, F. Gem News International: New discovery of demantoid from Ambanja, Madagascar. Gems Gemol. 2009, 45, 218–219. [Google Scholar]
  43. Rondeau, B.; Mocquet, B.; Lulzac, Y.; Fritsch, E. Les nouveaux grenats démantoïde d’Ambanja, Province d’Antsiranana, Madagascar. Le Règne Mineral 2009, 90, 41–45. [Google Scholar]
  44. Rondeau, B.; Fritsch, E.; Mocquet, B.; Lulyac, Y. Ambanja (Madagascar)—New source of gem demantoid garnet. InColor 2009, 11, 16–20. [Google Scholar]
  45. Praszkier, T.; Gajowniczek, J. Demantoide aus Antetezambato auf Madagascar. Miner. Welt 2010, 21, 32–41. [Google Scholar]
  46. Pezzotta, F. Andradite from Antetezambato, North Madagascar. Mineral. Rec. 2010, 41, 209–229. [Google Scholar]
  47. Pezzotta, F.; Adamo, I.; Diella, V. Demantoid and topazolite from Antetezambato, northern Madagascar: Review and new data. Gems Gemol. 2011, 47, 2–14. [Google Scholar] [CrossRef]
  48. Wilson, B.S. Coloured gemstones from Canada. Rocks Miner. 2009, 85, 24–43. [Google Scholar] [CrossRef]
  49. Ostrooumov, M. Mexican demantoid from new deposits. Gems Gemol. 2015, 51, 450–452. [Google Scholar]
  50. Guobin, L.; Xu, K.; Lin, Z. On the Genesis of Demantoid from Xinjiang, China. Chin. J. Geochem. 1986, 5, 381–390. [Google Scholar] [CrossRef]
  51. Mock, R.; Sýkora, M.; Aubrecht, R.; Ožvoldová, L.; Kronome, B.; Reichwalder, P.; Jablonský, J. Petrology and stratigraphy of the Meliaticum near the Meliata and Jaklovce Villages, Slovakia. Slovak Geol. Mag. 1998, 4, 223–260. [Google Scholar]
  52. Kozur, H. The Evolution of the Meliata-Hallstatt ocean and its significance for the early evolution of the Eastern Alps and Western Carpathians. Palaeogeogr. Palaeoclimatol. Palaeoecol. 1991, 87, 109–135. [Google Scholar] [CrossRef]
  53. Stampfli, G.M. The Intra-Alpine terrain: A Paleotethyan remnant in the Alpine Variscides. Eclogae Geol. Helv. 1996, 89, 13–42. [Google Scholar]
  54. Ivan, P. Relics of the Meliata Ocean crust: Geodynamic implications of mineralogical, petrological and geochemical proxies. Geol. Carpath. 2002, 53, 245–256. [Google Scholar]
  55. Faryad, S.W.; Spišiak, J.; Horváth, P.; Hovorka, D.; Dianiška, I.; Józsa, S. Petrological and geochemical features of the Meliata mafic rocks from the sutured Triassic Oceanic Basin, Western Carpathians. Ofioliti 2005, 30, 27–35. [Google Scholar]
  56. Kozur, H.; Mock, R. New paleogeographic and tectonic interpretations in the Slovakian Carpathians and their implications for correlations with the Eastern Alps and other parts of the Western Tethys, Part II: Inner Western Carpathians. Miner. Slovaca 1997, 29, 164–209. [Google Scholar]
  57. Kozur, H.; Mock, R. Erster Nachweis von Jura der Meliata-Einheit der südlichen Westkarpaten. Geologisch-Paläontologische Mitteilungen Innsbruck 1985, 13, 223–238. [Google Scholar]
  58. Putiš, M.; Radvanec, M.; Hain, M.; Koller, F.; Koppa, M.; Snárska, B. 3-D analysis of perovskite in serpentinite (Dobšiná quarry) by X-ray micro-tomography. In Proceedings of the Petros Symposium; Ondrejka, M., Šarinová, K., Eds.; Univerzita Komenského: Bratislava, Slovakia, 2011; pp. 33–37. [Google Scholar]
  59. Dallmeyer, R.D.; Neubauer, F.; Handler, R.; Fritz, H.; Müller, W.; Pana, D.; Putiš, M. Tectonothermal evolution of the internal Alps and Carpathians: 40Ar/39Ar mineral and whole rock data. Eclogae Geol. Helv. 1996, 89, 203–277. [Google Scholar]
  60. Putiš, M.; Yang, Y.-H.; Koppa, M.; Dyda, M.; Šmál, P. U/Pb LA-ICP-MS age of metamorphic-metasomatic perovskite from serpentinized harzburgite in the Meliata Unit at Dobšiná, Slovakia: Time constraint of fluid-rock interaction in an accretionary wedge. Acta Geol. Slovaca 2015, 7, 63–71. [Google Scholar]
  61. Faryad, S.W.; Henjes-Kunst, F. Petrological and K-Ar and 40Ar/39Ar age constraints for the tectonothermal evolution of the high-pressure Meliata unit, Western Carpathians (Slovakia). Tectonophysics 1997, 280, 141–156. [Google Scholar] [CrossRef]
  62. Plašienka, D. Cretaceous tectonochronology of the central western Carpathians, Slovakia. Geol. Carpath. 1997, 48, 99–111. [Google Scholar]
  63. Putiš, M.; Frank, W.; Plašienka, D.; Siman, P.; Sulák, M.; Biroň, A. Progradation of the Alpidic Central Western Carpathians orogenic wedge related to two subductions: Constrained by 40Ar/39Ar ages of white micas. Geodin. Acta 2009, 22, 31–56. [Google Scholar] [CrossRef]
  64. Dal Piaz, G.; Martin, S.; Villa, I.M.; Gosso, G.; Marschalko, R. Late Jurassic blueschist facies pebbles from the Western Carpathians orogenic wedge and paleostructural implications for Western Tethys evolution. Tectonics 1995, 14, 874–885. [Google Scholar] [CrossRef]
  65. Putiš, M.; Gawlick, H.J.; Frisch, W.; Sulák, M. Cretaceous transformation from passive to active continental margin in the Western Carpathians as indicated by the sedimentary record in the Infratatric Unit. Int. J. Earth Sci. 2008, 97, 799–819. [Google Scholar] [CrossRef]
  66. Putiš, M.; Danišík, M.; Ružička, P.; Schmiedt, I. Constraining exhumation pathway in an accretionary wedge by (U-Th)/He thermochronology—Case study on Meliatic nappes in the Western Carpathians. J. Geodyn. 2014, 81, 80–90. [Google Scholar] [CrossRef]
  67. Fediuková, E.; Hovorka, D.; Greguš, J. Compositional zoning of andradite from serpentinite at Dobšiná (West Carpathias). Věst. Ústr. Úst. Geol. 1976, 51, 339–345. [Google Scholar]
  68. Hovorka, D.; Jaroš, J.; Kratochvíl, M.; Mock, R. The Mesozoic ophiolites of the Western Carpathians. Krystalinikum 1984, 17, 143–157. [Google Scholar]
  69. Hovorka, D.; Ivan, P.; Jaroš, J.; Kratochvíl, M.; Reichwalder, P.; Rojkovič, I.; Spišiak, J.; Turanová, L. Ultramafic rocks of the Western Carpathians, Czechoslovakia; Geological Institute of Dionýz Štúr Publishers: Bratislava, Czechoslovakia, 1985; p. 258. [Google Scholar]
  70. Putiš, M.; Koppa, M.; Snárska, B.; Koller, F.; Uher, P. The blueschist-associated perovskite-andradite-bearing serpentinized harzburgite from Dobšiná (the Meliata Unit), Slovakia. J. Geosci. 2012, 57, 221–240. [Google Scholar] [CrossRef]
  71. Putiš, M.; Yang, Y.-H.; Vaculovič, T.; Koppa, M.; Li, X.-H.; Uher, P. Perovskite, reaction product of a harzburgite with Jurassic-Cretaceous accretionary wedge fluids (Western Carpathians, Slovakia): Evidence from the whole-rock and mineral trace element data. Geol. Carpath. 2016, 67, 133–146. [Google Scholar] [CrossRef]
  72. Hofmeister, A.M.; Chopelas, A. Vibrational spectroscopy of end-member silicate garnets. Phys. Chem. Miner. 1991, 17, 503–526. [Google Scholar] [CrossRef]
  73. Uher, P.; Kováčik, M.; Kubiš, M.; Shtukenberg, A.; Ozdín, D. Metamorphic vanadian-chromian silicate mineralization in carbon-rich amphibole schists from the Malé Karpaty Mountains, Western Carpathians, Slovakia. Am. Mineral. 2008, 93, 63–73. [Google Scholar] [CrossRef]
  74. Bačík, P.; Uher, P.; Kozakova, P.; Števko, M.; Ozdín, D.; Vaculovič, T. Vanadian and chromian garnet- and epidote-supergroup minerals in metamorphosed Paleozoic black shales from Cierna Lehota, Strazovske vrchy Mountains, Slovakia: Crystal chemistry and evolution. Mineral. Mag. 2018, 82, 889–911. [Google Scholar] [CrossRef]
  75. Peters, T. A water-bearing andradite from the Totalp serpentine (Davos, Switzerland). Am. Mineral. 1965, 50, 1482–1486. [Google Scholar]
  76. Passaglia, E.; Rinaldi, R. Katoite, a new member of the Ca3Al2(SiO4)3-Ca3Al2(OH)12 series and a new nomenclature for the hydrogrossular group of minerals. Bulletin de la Société Française de Minéralogie et de Cristallographie 1984, 107, 605–618. [Google Scholar] [CrossRef]
  77. Galuskina, I.O.; Galuskin, E.V. Garnets of the hydrogrossular-‘‘hydroandradite”-‘‘hydroschorlomite” series. Spec. Pap. Mineral. Soc. Pol. 2003, 22, 54–57. [Google Scholar]
  78. Galuskin, E. Minerals of the Vesuvianite Group from the Achtarandite Rocks (Wiluy River, Yakutia); University of Silesia Publishing House: Katowice, Poland, 2005; p. 192. [Google Scholar]
  79. Moore, R.K.; White, W.B. Electronic spectra of transition metal ions in silicate garnets. Can. Mineral. 1972, 11, 791–811. [Google Scholar]
  80. Manning, P.G. Optical absorption spectra of Fe3+ in octahedral and tetrahedral sites in natural garnets. Can. Mineral. 1972, 11, 826–839. [Google Scholar]
  81. Amthauer, G. Crystal chemistry and colour of chromium bearing garnets. Neues Jahrbuch für Mineralogie Abhandlungen 1976, 126, 158–186. [Google Scholar]
  82. Stockton, C.M.; Manson, D.V. Gem andradite garnets. Gems Gemol. 1983, 19, 202–208. [Google Scholar] [CrossRef]
  83. Tanabe, Y.; Sugano, S. On the absorption spectra of complex ions II. J. Phys. Soc. Jpn. 1954, 9, 766–779. [Google Scholar] [CrossRef]
  84. Tanabe, Y.; Sugano, S. On the absorption spectra of complex ions I. J. Phys. Soc. Jpn. 1954, 5, 753–766. [Google Scholar] [CrossRef]
  85. Tanabe, Y.; Sugano, S. On the absorption spectra of complex ions III. J. Phys. Soc. Jpn. 1956, 11, 864–877. [Google Scholar] [CrossRef]
Figure 1. Locations of andradite var. demantoid source (small point—capital town of region) (CZ—Czech Republic, AT—Austria, HU—Hungary, UA—Ukraine, PL—Poland).
Figure 1. Locations of andradite var. demantoid source (small point—capital town of region) (CZ—Czech Republic, AT—Austria, HU—Hungary, UA—Ukraine, PL—Poland).
Minerals 09 00164 g001
Figure 2. Andradite var. demantoid from Dobšiná, Slovakia (0.11 and 0.025 carat). Photo by J. Štubňa.
Figure 2. Andradite var. demantoid from Dobšiná, Slovakia (0.11 and 0.025 carat). Photo by J. Štubňa.
Minerals 09 00164 g002
Figure 3. Fibrous crystalline inclusions of serpentine group minerals in andradite var. demantoid from Dobšiná, Slovakia. Photomicrograph by J. Štubňa; magnified 50×.
Figure 3. Fibrous crystalline inclusions of serpentine group minerals in andradite var. demantoid from Dobšiná, Slovakia. Photomicrograph by J. Štubňa; magnified 50×.
Minerals 09 00164 g003
Figure 4. Ternary classification diagram of garnets.
Figure 4. Ternary classification diagram of garnets.
Minerals 09 00164 g004
Figure 5. Raman spectrum of the andradite var. demantoid from Dobšiná, Slovakia.
Figure 5. Raman spectrum of the andradite var. demantoid from Dobšiná, Slovakia.
Minerals 09 00164 g005
Figure 6. Optical absorption spectrum of the andradite var. demantoid from Dobšiná, Slovakia.
Figure 6. Optical absorption spectrum of the andradite var. demantoid from Dobšiná, Slovakia.
Minerals 09 00164 g006
Table 1. Electron microprobe analyses of demantoids, first provided in wt % of elements and then recalculated on the basis of 12 anions.
Table 1. Electron microprobe analyses of demantoids, first provided in wt % of elements and then recalculated on the basis of 12 anions.
12345678
P2O50.020.020.020.020.040.000.020.02
SiO235.2135.4135.3835.3335.4735.4335.7435.72
TiO20.050.050.040.030.000.030.080.04
Al2O30.330.340.310.320.400.340.370.32
V2O30.090.070.070.080.070.090.060.05
Cr2O31.241.502.312.141.541.340.970.63
Fe2O329.1529.1527.6928.3428.6428.8229.1629.65
MgO0.200.230.200.170.340.230.200.21
CaO33.8833.8033.7433.7233.8533.8533.7433.65
F0.080.110.090.110.090.080.090.10
O=F−0.04−0.05−0.04−0.06−0.04−0.04−0.05−0.05
Total100.19100.6399.80100.20100.40100.16100.38100.34
P5+0.0020.0010.0010.0010.0030.0000.0010.001
Si4+2.9752.9772.9932.9812.9842.9903.0033.004
F0.0210.0280.0230.0300.0230.0220.0240.026
Σ2.9973.0063.0183.0123.0103.0123.0293.031
Ti4+0.0030.0030.0030.0020.0000.0020.0050.003
Al3+0.0330.0340.0310.0320.0400.0340.0370.032
V3+0.0060.0050.0050.0060.0050.0060.0040.003
Cr3+0.0830.1000.1550.1430.1020.0890.0640.042
Fe3+1.8451.8371.7571.7931.8061.8231.8401.873
Σ1.9691.9791.9501.9751.9531.9541.9501.952
Mg2+0.0250.0290.0260.0220.0430.0280.0250.026
Ca2+3.0673.0443.0593.0493.0523.0603.0373.032
Σ3.0923.0733.0843.0713.0953.0893.0633.058
Table 2. The trace-elements contents (in ppm) in demantoid from Dobšiná measured by laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS). (LOD—limit of detection).
Table 2. The trace-elements contents (in ppm) in demantoid from Dobšiná measured by laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS). (LOD—limit of detection).
KScTiV51Cr53MnGe72As75Y89Mo95Gd157
Min392161033053814103411200
Max78740389572133312241322512
Average523252444449199143916311
LOD48.72.111.40.79.27.83.89.10.10.10.1
Table 3. Raman shifts (cm−1) of Raman-active bands in the studied sample. Assignment of band modes was made according to Reference [72].
Table 3. Raman shifts (cm−1) of Raman-active bands in the studied sample. Assignment of band modes was made according to Reference [72].
SymmetryAssignmentDemantoid DobšináAndradite [72]
T2g + T1uT(M)174173
T2g + T1uT(M)236235
T2gT(SiO4)mix265264
EgT(SiO4)297296
T2g + T1uT(SiO4)311311
??343-
Egν2352352
A1gR(SiO4)371370
T2g + T1uν2452452
Egν2493494
A1gν2516516
T2g + T1uν4552553
Egν4576576
T2g + T1uν3816816
T2g + T1uν3843842
Egν3874874
T2g + T1uν3995995
Table 4. The optical spectral bands (nm) in the studied sample. Assignment of bands was made according to Reference [22] and references therein.
Table 4. The optical spectral bands (nm) in the studied sample. Assignment of bands was made according to Reference [22] and references therein.
ChromophoresDemantoid Dobšiná
Fe3+438
Fe3+470
Cr3+488
Fe3+587
Cr3+623
Fe3+861
Fe3+964

Share and Cite

MDPI and ACS Style

Štubňa, J.; Bačík, P.; Fridrichová, J.; Hanus, R.; Illášová, Ľ.; Milovská, S.; Škoda, R.; Vaculovič, T.; Čerňanský, S. Gem-Quality Green Cr-Bearing Andradite (var. Demantoid) from Dobšiná, Slovakia. Minerals 2019, 9, 164. https://doi.org/10.3390/min9030164

AMA Style

Štubňa J, Bačík P, Fridrichová J, Hanus R, Illášová Ľ, Milovská S, Škoda R, Vaculovič T, Čerňanský S. Gem-Quality Green Cr-Bearing Andradite (var. Demantoid) from Dobšiná, Slovakia. Minerals. 2019; 9(3):164. https://doi.org/10.3390/min9030164

Chicago/Turabian Style

Štubňa, Ján, Peter Bačík, Jana Fridrichová, Radek Hanus, Ľudmila Illášová, Stanislava Milovská, Radek Škoda, Tomáš Vaculovič, and Slavomír Čerňanský. 2019. "Gem-Quality Green Cr-Bearing Andradite (var. Demantoid) from Dobšiná, Slovakia" Minerals 9, no. 3: 164. https://doi.org/10.3390/min9030164

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