Skip to main content

ORIGINAL RESEARCH article

Front. Earth Sci., 27 March 2023
Sec. Economic Geology

Gemological features of diaspore in sodra—Milas (mugla) region

  • 1Department of Jewelry and Jewelry Design, Batman University TBMYO, Batman, Turkey
  • 2Department of Geological Engineering, Konya Technical University, Konya, Turkey

Diaspore, is related to bauxite in terms of its formation mechanism. Diaspore is one of the gemstone quality stones used in jewelry and jewelry making. It is possible to come across gemstones in various regions of the world. In our country, there are gemstone deposits in all regions. In these deposits, diaspore crystals are considered very valuable. Within the scope of this study, it was aimed to determine the gemological features of the diaspore formations in Sodra (Milas-Mugla) and its surroundings, as well as to determine other minerals and their formation conditions. Main oxide, trace, and rare earth element analyses were performed on seven collected samples. The classification was made using the Al2O3-SiO2-Fe2O3 triangular variation and it was determined that bauxite samples were in the “ferric bauxite” and “bauxite” areas. SEM imaging and EDS analyzes were performed on the samples and according to this result, it was determined that the samples mainly contained O, Al, and Si. The data obtained as a result of Raman Spectroscopy were matched with the device’s own data library and it was determined that the samples taken from the Sodra region were diaspore. In addition, according to the results of the FTIR analysis performed on the faceted sample taken from the Sodra region, it was seen that the samples were compatible with the reference standards and it was scientifically proven to be a diaspore crystal with gemstone quality. During the field studies carried out in the study area, diaspore crystals were encountered in the Milas-Sodra region, together with chloritoid and muscovite, within the metabauxite formations and in the altered zones close to the metabauxite limestone contact. Rock were classified as “muscovitequartzschist,” “quartzschist” with grano-lepidoblastic texture, “micaquartzschist” with lepidoblastic texture, “phyllite” with lepidoblastic texture, and “chlorite-epidote fels” with nematoblastic texture as a result of petrographic analyzes performed on the country rock samples. It is predicted that these gemstone-quality diaspore crystals can be evaluated economically.

1 Introduction

Diaspore is a bauxite mineral, with the chemical formula [Al(OH)O], it has the feature of being a single mineral, and about 40% of its crystals have a transparent-translucent structure. Diaspore, which has the ability to change color like Alexandrite, is a rare stone due to this feature. While the diaspore is green-yellow in white light, it turns pink-red in daylight. Since it is not too hard, it can break while cutting. Diaspore, which has a hardness level of 6.5–7 according to the Mohs Hardness Scale, is used in jewelry and jewelry making. This is the main reason why the Turkish diasporic stone is so popular and in demand. Since diaspore is a bauxite mineral, its formation mechanism is related to bauxite formation. Bauxite rock contains different minerals in varying amounts, and it has more than 45% Al2O3 in its composition. In addition to aluminum hydroxide minerals, this rock also contains iron oxide, hydroxide minerals, clay minerals, secondary amounts of silica, titanium, and other minerals. Iron compounds in bauxites are amorphous iron hydroxide, goethite, hematite, lepidocrosite, magnetite, maghemite, pyrite, marcasite, siderite, and leukoxene. The main clay minerals are kaolinite, halloysite, chlorite and illite. Silica is found in the form of quartz and chalcedony. Apart from these, bauxite may contain up to 4% rutile and anatase (Temur, 2001).

In order to avoid confusion/difficulties in the classification of bauxite deposits, various researchers have proposed a general classification model based on previous studies. Accordingly, bauxite deposits can be genetically divided into three main groups: lateritic, sedimentary (tikhvin), and karstic species (Bogatyrev et al., 2009). Lateritic-type bauxites are residual deposits derived from underlying aluminosilicate rocks whose preserved texture and composition can be directly associated with the underlying source rocks. Sedimentary (tikhvin) type bauxites are clastic bauxite deposits that cover the eroded surface of aluminosilicate rocks, and karstic type deposits overlie the karstified surfaces of carbonate rocks (Bárdossy and Aleva, 1990). Karstic type bauxites are subdivided into autochthonous, para-autochthonous, allochthonous, and para-alloctonous subtypes (Bárdossy, 1982). In Turkey, there are mostly karstic type and less laterite type bauxite deposits among the main three deposit types (Robertson and Ustaömer, 2009).

There are important and current studies on bauxite in the world and our country (Karadag et al., 2003; Hanilci, 2013; Yalcin and Ilhan, 2013; Nyamsari and Yalcin, 2017; Sidibe and Yalcin, 2019; Nyamsaria et al., 2020; Ozer and Yalcin, 2020; Ozturk et al., 2021; Kinaci and Ozturk, 2022). Karadag et al. (2003), determined that the study area did not differ much from other deposits in the Seydişehir region, and they distinguished five types of ores in the deposit: massive, oolitic, pisolithic, clayey, ferrous, and brecciated. They stated that the prevalence of oolitic-type ore in the deposit indicates the presence of a warm, shallow and turbulent environment during the ore formation. Hanilci (2013) stated that bauxitization processes started with the formation of lateritic soil rich in kaolinite after the Upper Triassic and pre-Aalenian. In particular, it was stated that Rb, K, Ba, Ce, Si, Y, Sr, P, Ca, Mg, and U elements were lost by weathering/abrasion during bauxitization, while Fe, V, Pb, Ni, and Cr elements were gained in the bauxite area (Hanilci, 2013). Yalcin and Ilhan (2013) found the average modulus of bauxites to be 18.04 according to the Al2O3/SiO2 ratio, and it was determined that Küçükkoraş bauxite was the best quality bauxite. Geochemical data showed that the limestone was formed as a result of the transformation of limestone into bauxite first and then into terra rossa. Nyamsari and Yalcin (2017), origin rock research showed that the source of bauxite is mafic, basaltic andesite igneous rocks with medium pH (basic-acidic characteristic). Sidibe and Yalcin (2019) determined that the mineralogy of both bauxite types (sedimentary and lateritic) of the Balaya plateau is almost the same. Nyamsari et al. (2020) stated that the rocks in the Haléo-Danielle Plateau bauxite deposit were subjected to intense weathering, which led to the formation of industrial bauxite deposits. Examination of the bedrocks revealed that bauxite was formed due to weathering of magnesium-iron andesite-basalts under conditions with near-neutral redox potential. Ozturk et al. (2021) stated that bauxites in the Payas region occur in two types, cover and pocket, are chemically and texturally homogeneous, and have a thrust structure with ophiolitic melange formations. They stated that bauxite is mainly composed of diaspore, hematite, rutile, anatase, rare kaolinite, boehmite, and pyrite minerals, and prismatic and stave-shaped euhedral rutile in bauxites, and bauxites are formed in situ from Ti-rich basaltic pyroclastics.

Diaspore is an extremely rare, gemstone-quality form of aluminum oxide hydroxide, one of the three main mineral components of bauxite, an important aluminum ore widely used for many industrial purposes. It has the unique feature of changing color under different lighting conditions. The color-changing diaspore is easily distinguishable from other similar gemstones by the unique presence of pleochroism and its excellent color-changing ability. Choo and Kim (2003) determined that tourmaline found in the Milyang clay deposit, diaspore nodules from the Milyang clay deposit in southeast Korea form as clusters or diffused crystals in the clay matrix, and diaspore nodules occur next to or near fracture zones that develop in the clay bed. Hatipoglu et al. (2010) revealed that the diaspore [AlO(OH)] is polycrystalline in their study in the Ilbirdagı diasporic metaboxite (diasporite) deposit in Milas (Mugla) region of Turkey. The reason why the diaspore crystals in this deposit are more unique than those in other deposits is that the diaspore is optically uniform and pure. Hatipoglu (2011) detected diaspore, gibbsite, corundum, specular hematite, ilmenite, chloritoid, goethite, ferroalluaudite and Fe-Mg-rich muscovite in his microscopic studies in the diaspore mineralizations in the Upper Cretaceous Early Paleogene aged marble succession in the Menderes Massif. It was also concluded that the composition of the primary fine-grained mineralization in the unaltered (original) bauxite ore is sufficient to be considered a source for much later mineralization in the fracture zones.

The increase in diaspore prices in recent years (96-carat diaspore jewelry designed in London is valued at 1.5 million dollars) has increased the attention and importance of this mineral. Within the scope of this study, it was aimed to determine the diaspore and other accompanying minerals in the metabauxite ores found in Sodra and its surroundings (Supplementary Figure S1), and to determine their formation conditions and gemological properties. It is thought that diaspore is very rare in the world and it will make significant contributions to the semi-precious stone industry as a result of bringing the minerals discovered for the first time in the Milas—Sodra region with this study to the literature.

2 Materials and methods

Detailed geological and structural investigations were made at the contact of core and cover units, which are important in terms of diaspore mineralization in field studies; samples from the Sodra (Milas) region were collected in numbers and qualities that could represent the region, and a total of 16 samples were sent for analysis. In the laboratory studies, the preparation of the samples and petrographic analyzes were performed in the Geological Engineering Department of Konya Technical University (Turkey), and the chemical analyzes and gemological analyzes were performed in domestic (BİTAM) and foreign accredited laboratories (ACME).

Petrographic thin sections were prepared from the samples. Petrographic studies in mineral determinations were completed by using a scanning electron microscope (SEM-EDS) for local element determination and imaging purposes. Qualitative element analysis was performed with the Hitachi-SU 1510 device at 10 and 100 µ zooms with point and mapping methods. SEM-EDS imaging procedures were performed in the Necmettin Erbakan University BİTAM laboratory (Turkey).

Whole rock (main oxide, trace element, and rare earth element) and mineral chemistry analyzes samples were chemically analyzed to make geochemical and geostatistical evaluations on the rock and mineral samples taken from the study area. The samples collected from the field for geochemical investigations were reduced to less than 0.5 cm by hammering or through jaw crushers. Then, rock samples weighing 20–50 g were ground in automatic mortars or agate mortars so that the grain size was below 70 µ. Analyzes were made at Bureau Veritas (Canada) ACME Laboratory.

Since diaspore samples are few and economically important, non-destructive analysis methods are preferred in gemology. For this purpose, scans with FTIR (Fourier Transform Infrared Spectroscopy/Fourier transform infrared spectroscopy) device equipped with MAGILABS GemmoFtir™ spectrometer low noise DLATGS- detector obtained at a resolution of 4 cm-1 for a measurement time of 20 s. With these methods, peaks were formed by detecting the characteristic fingerprint of each sample and these peaks were used for chemical identification.

Measurements were made with Renishaw in via Reflex Raman, which is used for another non-destructive identification purpose. A silicon wafer was used to calibrate the device before measurement. Raman Spectrometer analysis was performed in Necmettin Erbakan University BİTAM laboratories to verify the FTIR analysis.

3 Research results and findings

3.1 Regional geology

Studies conducted in the last two decades have largely revealed the relationship between the Menderes Massif and the Lycian nappes in the area between Bafa Lake and Mugla, and the stratigraphy of these units (Konak et al., 1987). In the north of Kale-Tavas regions, the extent to which the Menderes Massif extends to the east and its relationship with the Lycian nappes is little known. The region between Bafa Lake and Mugla, where the geology of the Menderes Massif is best known, should be taken as a standard region. The main points defining the Menderes Massif in this region are given below (Okay, 1985).

From bottom to top, the Menderes Massif sequence consists of Precambrian gneisses, Lower Paleozoic mica schists, Permo-Carboniferous metaquartzite, black phyllite, and black recrystallized limestones, Mesozoic age, bauxite-level, thick-bedded, recrystallized, neritic limestones and recrystallized limestones up to the Lower Eocene flysch. Lycian nappes, which were emplaced in the Middle Eocene with a tectonic contact, overlie the Eocene flysch, the youngest unit of the massif. In the Menderes Massif, regional metamorphism of Eocene aged Barrovian type, gradually decreasing from bottom to top, is observed. This regional metamorphism (Sengor et al., 1984b), which developed due to the emplacement of the Lycian nappes on the Menderes Massif, also affected the lower parts of the Lycian nappes. In terms of metamorphism, there is no discontinuity between the Menderes Massif and the nappes overlying it. Apart from the Eocene metamorphism in the Menderes Massif, traces of old metamorphism of the Pan-African age were found in Precambrian gneisses. The Menderes Massif constitutes the lowest tectonic unit between Lake Bafa and Mugla and is in a relatively autochthonous position. The massif has a simple south-dipping structure, which developed due to regional uplift and erosion during the Oligocene. Large-scale thrusts and isoclinal folds are not observed in the massif (Satir and Friedrichsen, 1986).

Metabauxite formations were observed as a karstic fill at the border of white-colored recrystallized limestone and dark-colored recrystallized limestone (recrystallized dolomitic limestone). Diaspore formations, on the other hand, were observed in the carbonate-altered zone at the contact of metabauxite and recrystallized dolomitic limestones and in the cracks of metabauxites, generally together with mica (muscovite)-chloritoid formations (Supplementary Figure S2).

3.2 Geochemistry

A triangular variation of Al2O3-SiO2-Fe2O3 (Meshram and Randive, 2011; Gu et al., 2013; Zamanian et al., 2016) was used. With the bauxite samples taken according to the classification, Milas-Yatagan Area (ORT_1), Seydisehir-Akseki Area (ORT_2), Isparta-Sarkikaraagac area (ORT_3), Kokaksu area (ORT_4), Bolkardagi area (ORT_5) and Payas-Islahiye area (ORT_6) in Turkey was evaluated together with the bauxite formations (Hanilci, 2019) and it was determined that the bauxite formations in the study area had “ferric bauxite” and “bauxite” characteristics (Supplementary Figure S3). In addition, it is seen that the bauxite formations have undergone a high degree of chemical change/alteration according to the Chemical Index of Alteration (CIA) degree of alteration.

The Si ratios in the metabauxite and diaspore samples taken to represent the general study area vary between 0.88% - 15.52% and 0.18%–0.32%, respectively. The arithmetic value of the population (Samples) is 4.25% (metabauxite) and 0.25% (diaspore). These values obtained are the Al value is between 16.27% and 28.80% (metabauxite) and 38.80%–40.15% (diaspore) and their average values are 24.59% and 39.31%, respectively. The Al value in the earth’s crust is 8.13% (Krauskopf, 1979). The average values obtained are approximately 3 and 5 times higher, respectively than the value in the earth’s crust. The Fe average was 13.90% (metabauxite) and 0.56% (diaspore). The lowest values among the samples taken were 9.76% (metabauxite) and 0.59% (diaspore); the highest values are 16.79% and 0.64% (diaspore). These values are approximately 2.5 (metabauxite) times higher than the average Fe in the earth’s crust, which has a value of 5% (Krauskopf, 1979) (Supplementary Tables S1–S3).

Diaspore samples collected from the study area were normalized according to chondrite (Taylor and McLennan, 1981; Boynton, 1984) (Supplementary Table S4). Their normalization graph was prepared according to LREE (light rare earth elements), MREE (medium rare earth elements) and HREE (heavy rare earth elements) (Supplementary Figure S4). According to this chart; a clear agreement is seen in the normalization graph of S-1, S-2, S-3, and S-4 diaspore samples according to chondrite.

In order to determine the color changes of the diaspore samples in the study area, the main and trace elements that can cause color changes were made logarithmically according to the chondrites (Pearce et al., 1984) and the analysis values were compared (Supplementary Figure S5).

It is seen that the lightly rare earth element enrichment (LREE) in the samples is higher than the medium (MREE) and heavy rare earth element (HREE) enrichment. This may be due to material coming from the upper continental crust to the source magma. In addition, a negative Eu anomaly is observed. The presence of negative Eu anomaly indicates that plagioclase differentiation (Kalkan and Ozpinar, 2018) plays an important role in the development of these rocks.

3.3 SEM-EDS studies

According to the results of the point and mapping method made for the determination of the elements that can be found in the crystal lattice structure of three diaspore samples collected from the region, oxygen was determined as 55% and aluminum as 42%. As a result of SEM imaging and EDS analyses, it was determined that the samples mainly contain O, Al, and Si, and the presence of K, Na, Fe, Ca, and Mg in addition to these elements in close-ups (Supplementary Figures S6–S9).

3.4 Raman spectroscopy studies

Due to its non-destructive nature and fast measurement times, low material and sample preparation requirements, Raman spectroscopy is routinely used for gemstone identification, examination of inclusions, and possible value enhancement (Bersani and Lottici, 2010). As a result of the shooting, in the light of the match values taken from the microscope’s own database, it was determined that the sample was diaspore, and Brucite [Mg(OH)2)] was found at 79 cm−1 (Supplementary Figure S10).

Raman spectra of the diaspore crystal are shown in Supplementary Figure S10. Raman bands showing diaspore in the spectra of the samples were observed at peaks at 867, 675, 627, and 571 cm−1.

3.5 Studies for the determination of gemological characteristics

In mineralogical research, FTIR spectroscopy is widely used to characterize molecular water and hydroxyl groups in various minerals such as silicates, clays, aluminum oxides, and iron oxides. In gemology, FTIR spectra can also be applied to determine the molecular and/or structural vibration of–OH. This vibration is the result of reflections of OH-containing mineral inclusions such as boehmite, diaspore, goethite, and mica (Phlayrahan et al., 2018).

FTIR analysis was performed on the faceted sample taken from the study area to determine the origin, and according to the results of this analysis, the strong peaks in the range of (600–640) cm−1, (680–720) cm−1, 1,000 cm−1, 1,040 cm−1, 1,140 cm−1, weak peaks in the 1,200 range (in red) were found to be consistent with reference to the standard wavelength (colored black) (Supplementary Figure S10).

It has been scientifically proven that the facetted samples collected from the region (Supplementary Figure S11) are gemstone-quality diaspore crystals with various color variations.

The sample was cut with a standard diamond cut, with 45o, 90o, and 43o angles on the cone part, and then on the stone part with 47o, 42o, 27o, and 0o angles, 57 facets on the stone and cone part, and 16 facets on the belt part, a total of 73 facet surfaces were processed (Supplementary Figure S12).

It is a cutting technique applied to faceted gemstones for the highest shine and best appearance and is usually applied for transparent, precious gemstones. The reflection or brilliance depends on the optical, physical, and facets of the gemstone. The facet-cutting technique can turn a gemstone containing inclusions into a valuable piece, or it can turn a high-end gemstone into a rather worthless piece.

4 Discussion

The Sodra bauxite deposit is located very close to the Milas settlement and 23 km from the port of Güllük. Etibank (1975) extracted about 4,000 tons of ore from this deposit. The ore zone is located within the Menderes Massif marble. The bauxite zone is E-W to N80° W and slopes 40° Northeast. It is also 150 m long and about 5.5 m thick. Etibank (1975) estimated the ore reserves to be 0.23 Mt with an average content of 56% Al2O3, 5% SiO2, and 2.8% TiO2 (Hanilci, 2019).

Inert element compositions (Cr, Zr, Ga, V, Ni, Nb, etc.) and various diagrams were used to elucidate the possible source of bauxite and diaspore formations in the region (Supplementary Figure S13).

According to the Cr-Ni distributions of the samples collected from the study area, bauxite occurrences are in the karst-bauxite region, and diaspore occurrences are in the high-iron bauxite and nepheline syenite bauxite regions.

V/(V+Ni) and Ni/Co diagrams of bauxite and diaspores were drawn to reveal oxic, semi-oxic, anoxic, and redox conditions in the study area (Supplementary Figure S14A). Ni/Co ratios of <5 in rocks reflect oxygenated conditions, >5 reflects suboxic and anoxide conditions (without oxygen) in the formation of clays in the basin, and V/(V+Ni) ratios (0.4–0.60) reflect dioxide conditions (Rimmer, 2004; Akinyemi et al., 2012). In light of these data, bauxite formations in the region are concentrated mainly in the semi-oxic region, and diaspores are concentrated in the anoxic region according to their V/(V+Ni) ratios.

The pH value in the environment during bauxite formation is expressed as the dominant control parameter for the mobility of REEs and it is stated that REEs cause separation/decomposition in acidic conditions and accumulation in alkaline conditions. Among the trace elements, REEs are generally accepted as the most suitable geochemical tools for evaluating the source of sedimentary rocks in general and a bauxite deposit in particular (Calagari and Abedini, 2007; Mameli et al., 2007; Mongelli et al., 2016; Sinisi, 2018). REEs are thought to be relatively inert elements during decomposition processes. In conclusion, the enrichment of REEs in bauxite indicates that alkaline conditions were dominant during its evolution (Abedini and Calagari, 2014). Measuring changes in the La/Y ratio is a useful geochemical tool to determine changes in pH during bauxitization (Salamab Ellahi et al., 2016). La/Y <1 values indicate the prevalence of acidic conditions and La/Y >1 values indicate alkaline conditions/basic conditions. It was determined that the La/Y ratio of bauxite samples collected from the study area had values above 1. In diaspore samples, on the other hand, the La/Y ratio was found below 1. According to these data, we can state that diaspore formations in the Sodra region occur under acidic conditions.

However, Cerium (Ce) is degradable relative to other rare earth elements because its geochemical behavior during weathering can be strongly influenced by changes in the redox state. As a result, the altered/decomposed material may be enriched or depleted in cerium and show positive or negative Ce anomalies accordingly. Although the stable valence is trivalent for most REEs, Ce has been shown to be redox-sensitive under weathering and oxidized to the tetravalent state (Ma et al., 2007). It has been shown that positive Ce anomalies in lateritic profiles and bauxites are mineralogically related to cerianite (CeO2) precipitation due to Ce 3+ to Ce 4+ redox change (Mongelli et al., 2016). In addition, Ce 4+ is prone to hydrolysis. It is stated that it is separated from other REEs by precipitating after adsorption by Fe and Mn oxides, and this effect leads to a Ce deficit. It is stated that the traces of goethite and hematite in the examinations indicate that the region is oxic and therefore Ce is probably in the tetravalent state and therefore precipitated as CeO2 (Yusoff et al., 2013). It is stated that such a process causes the formation of a positive Ce anomaly (Sinisi, 2018). In addition, studies generally accept 0.78 (Ce) as a reference value to indicate a redox environment, and Ce > 0.78 is expressed as a positive anomaly and Ce < 0.78 as a negative anomaly (Ji et al., 2000; Zhang et al., 2020). Negative Ce values indicate that the ore formation process is likely under suboxic water body/deoxidized conditions. According to the Ceanom–Nd diagram, Ce/Ce* (Ceanom-NASC). If > −0.1, the water in the environment where the mineralizations occur or are stored in the sedimentary storage environment is insufficient in terms of oxygen, Ceanom <-0.1 indicates that the sedimentary water in the environment is rich in oxygen (Ozyurt et al., 2020). Ce/Ce* of bauxite mineralization values in the study area (NASC) were obtained as 0.86 on average, and 0.90 of diaspore formations. With these data, it was determined that these formations occur in an oxygen-poor environment (anoxic) and bauxite formation processes occur slowly, while diaspore formation processes occur rapidly (Supplementary Figure S14B).

In order to determine the origin of bauxite and diaspore formations in the study area, the relations between Eu anomaly (Eu/Eu*) and Sm/Nd and TiO2/Al2O3 ratios were also used. Similar to Eu anomalies, the Sm/Nd ratio is also used as an index of chemical differentiation during intense tropical weather conditions (Viers and Wasserburg, 2004) (Supplementary Figure S15).

Eu anomaly is an index of chemical differentiation and represents the proportion of crust exposed in the source area(s) (Condie et al., 2001; Sinisi, 2018) during bauxidization (Mameli et al., 2007; Mondillo et al., 2011). In addition, Eu anomaly values can be of great help in understanding the physicochemical conditions of various geochemical systems, such as the sedimentation environment of limestones. Positive Eu anomalies (normalized to PAAS) are generally observed in limestones affected by hydrothermal processes (Siby et al., 2008; Madhavaraju and Lee, 2009). Negative Eu/Eu* anomalies are a geochemical feature of post-Archean cratonic deposits and are generally associated with mafic source rocks (Sinisi, 2018). In addition, intense diagenesis, and differences in the amount of feldspar and wind input are other factors that can have a positive effect (Madhavaraju and Lee, 2009). Also, based on the SHAB (Hard/Soft of Acids and Bases) theory proposed by Dai (1987), some reduced species such as CO, CH4, S2O3 2−, S2- and HS- increase in the aqueous system under low oxygen fugacity. In such an environment, Eu2+ prevails widely and remains in solution as a stable Eu 2+- complex and is therefore suitable for inclusion in Eu precipitates (positive anomaly). In contrast, under higher oxygen fugacity, Eu3+ dominates in the aqueous system and requires hard bases such as OH, SO4 2− and CO3 2− (especially OH) to precipitate, as a result, negative Eu anomaly can be considered. In addition, Abedini et al. (2019) suggested that the strong negative Eu anomaly may be associated with higher O2 and lower pH of the solutions. Post-Archean Australian Shale (PAAS) =0.61 in the evaluation of Eu anomaly (Eu/Eu*); Post/Archean upper continental crust (UCC)=0.66 is accepted as the post-Archean upper continental crust (Sinisi, 2018). The Eu/Eu* values (PASS) of the bauxite mineralizations in the study area were found to be 1.05 on average, and 1.33 (SO) in the diaspore formations. The bauxite formations in the study area are concentrated in the region between cratonic sandstones-PAAS-bedrock limestone and we can consider these formations as the dominant clastic material product, probably derived from a distant source, carried by wind/water (?).

It has been stated that the bauxites in the region have undergone metamorphism under the conditions of approximately 10–25 km depth, 5–10 kb pressure and 500°C–620 °C temperature conditions due to burial due to tectonic movements. In addition, as a result of metamorphism, gibbsite and boehmite formations were transformed into diaspore and corundum, while ferrous compounds transform into magnetite, and clays were thought to transform into minerals such as muscovite and chloritoid (Temur et al., 2003).

In recent years, stones of dubious origin have been encountered quite frequently in the precious stone market, and it is becoming more and more difficult to distinguish synthetic and/or fake gemstones produced with developing technologies from the originals. Raman spectroscopy becomes an ideal method in terms of the fact that gemstone identification is non-destructive and gives fast results. The disappearance of the human factor in Raman shots seems to prevent possible concerns (Jenkins and Larsen, 2004). With the Raman spectrometry method used in this study, diaspore peak values were determined to be 867, 675, 627, and 571 cm−1 depending on the wavelengths. With increasing pressure, the Raman peaks move towards the high wave number direction, the intensity of the Raman peaks increases, and the vibrational mode of the diaspore changes linearly (Huang et al., 2020). Juan-Farfan et al. (2011) obtained 22 Raman peaks in the range of 150–1,300 cm−1 by theoretical calculation. However, only 13 Raman peaks were observed in the same wavenumber range in the experiments because the other peaks were determined to be too weak to be observed in the experiments (Ruan et al., 2001; Juan-Farfán et al., 2011). Raman spectrometer values of diaspores in the Sodra region are in agreement with literature values. This shows that diaspora has an ornamental quality.

5 Results

The properties and distributions of major and trace elements were investigated to reveal bauxite formation processes. It was thought that the clayey/clastic metamorphic-igneous rocks with high Al content in the study area were the source of bauxite formation and the enrichment was repeated more than once. In addition, it has been determined that bauxite formations are related to a bedrock with intermediate magmatic or clayey composition, while diaspore formations provide data close to ultrabasic-basic properties.

It has been observed that the bauxites in the region have undergone metamorphism at a depth of approximately 10–25 km, a pressure of 5–10 kb, and a temperature of 500°C–620 °C due to burial due to tectonic movements. With this metamorphism, it was thought that while gibbsite and boehmite were transformed into diaspore and chorus, ferrous compounds turned into magnetite, and clays into minerals such as muscovite and chloritoid. As a result of SEM imaging and EDS analyses, it was observed that the samples in the study area mainly contained O, Al, and Si, and the presence of K, Na, Fe, Ca, and Mg in addition to these elements were detected in close-ups. Chemical and EDS analysis results support each other, and no different results were found during sample scanning. It was observed that the diaspores collected from the Sodra field had gemstone qualities in terms of physical and chemical properties. When the crystals suitable for shaping collected in these samples were subjected to facet processing and cut at appropriate angles, they were well polished and attained a glassy shine. As a result of the Raman shots, it was determined that the sample was diaspore in the light of the match values obtained from the microscope’s own database, and Brucite [Mg(OH)2)] was determined to be at 79 cm−1. Raman spectra of the diaspore crystal are shown in Supplementary Figure S10. Raman bands showing diaspore in the ranges of the samples were observed at peaks at 867, 675, 627, and 571 cm−1. Although the presence of diaspore crystals within the borders of Mugla province in the Southwest Anatolia region has been determined by previous studies, the discovery of diaspore crystals was revealed for the first time in the Sodra region, where this study was conducted. It is predicted that these gemstone-quality diaspore crystals can be evaluated economically.

Data availability statement

The datasets presented in this study can be found in online repositories. The names of the repository/repositories and accession number(s) can be found in the article/Supplementary Material.

Author contributions

The manuscript was prepared from EH PhD’s thesis. Supervisor AO determined the area where the study was conducted, and after determining the sample locations together, EH made the samples. Samples were sent for analysis with the approval of supervisor AO. EH evaluated the results of chemical analysis, and the manuscript was prepared with the recommendations and approval of supervisor AO. AO made a great contribution in the preparation of this publication. All authors have read and approved this article.

Funding

This study was supported by Konya Technical University Scientific Research Project Coordinator (BAP Project No: 201107007).

Acknowledgments

The authors thank the Konya Technical University Scientific Research Project Coordinator for their financial support in the conduct of this study.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/feart.2023.1145674/full#supplementary-material

SUPPLEMENTARY FIGURE S1 | Location map of the study area ((A): Turkey map, (B): Map of the study area, (C): Geological map of the study area) (modified from MTA, 2002).

SUPPLEMENTARY FIGURE S2 | Diaspore formations in the study area (A) Diaspore crystal (B) Diaspore containing vein (C) formation containing diaspore crystals in recrystallized limestone.

SUPPLEMENTARY FIGURE S3 | SiO2- Fe2O3- Al2O3 diagram (Meshram and Randive, 2011; Gu et al., 2013; Zamanian et al., 2016) and Chemical Change Index (CIA) with alteration (Nesbitt and Young 1982; Shao et al., 2012) degree (Ozturk et al., 2021).

SUPPLEMENTARY FIGURE S4 | Normalized logarithmic (y-axis) distribution plot of diaspore samples in the study region according to chondrite.

SUPPLEMENTARY FIGURE S5 | Logarithmic (y-axis) distribution graphs of some elemental major and trace elements of diaspore formations.

SUPPLEMENTARY FIGURE S6 | 10μ magnification SEM-EDS dot and mapping images of the S-1 diaspore sample in the Sodra region.

SUPPLEMENTARY FIGURE S7 | SEM-EDS dot and mapping images with 10 μ magnification of the S-2 diaspore sample in the Sodra region.

SUPPLEMENTARY FIGURE S8 | SEM-EDS dot and mapping images with 10 μ magnification of the diaspore sample S-3 in the Sodra region.

SUPPLEMENTARY FIGURE S9 | 10 μ magnification SEM-EDS dot and mapping images of the S-4 diaspore sample in the Sodra region.

SUPPLEMENTARY FIGURE S10 | Mineral contents determined by Raman spectroscopy (S-1, S-2, S-3 and S-4) in diaspore samples from the Sodra region. FTIR plots of diaspore samples taken in the Sodra region (blue lines are FTIR standards, red lines are diaspore samples).

SUPPLEMENTARY FIGURE S11 | The image of the diaspore samples collected from the field (A) Round-Brilliant Cut (GIA, 2016). (B,C) The image of the sample cut from the samples under yellow light and white light.

SUPPLEMENTARY FIGURE S12 | Schematic representation of the technique used on diaspore samples collected from the field (Gemmology Project, 2022).

SUPPLEMENTARY FIGURE S13 | Cr-Ni diagram showing the distribution of Bauxite-Diaspore formations in the study area (Sinisi, 2018).

SUPPLEMENTARY FIGURE S14 | (A): V/(V + Ni) (Hatch and Leventhal, 1992). and Ni/Co; (Jones and Manning, 1994) diagram (Rimmer, 2004; Aydogan, 2021). (B): Diagram showing oxic, semi-oxic, anoxic, and redox conditions (Wright et al., 1987; Hatch and Leventhal, 1992; Jones and Manning, 1994; Aydogan, 2021).

SUPPLEMENTARY FIGURE S15 | Diagram of Sm/Nd and Eu/Eu* (A) TiO2/Al2O3 and Eu/Eu* (B) outlining the comparison between mean standard lithologies (Viers and Wasserburg, 2004).

References

Abedini, A., and Calagari, A. A. (2014). REE geochemical characteristics of titanium-rich bauxites: The permian kanigorgeh horizon, NW Iran. Turkish J. Earth Sci. 23, 513–532. doi:10.3906/yer-1404-11

CrossRef Full Text | Google Scholar

Abedini, A., Mehr, M. H., Khosravi, M., and Calagari, A. A. (2019). Geochemical characteristics of the karst-type bauxites: An example from the kanirash deposit, NW Iran. Arabian J. Geosciences 12, 475. doi:10.1007/s12517-019-4601-z

CrossRef Full Text | Google Scholar

Akinyemi, S. A., Gitari, W. M., Akinlua, A., and Petrik, L. F. (2012). Mineralogy and geochemistry of sub-bituminous coal and its combustion products from mpumalanga province. South Africa: Open Access.

Google Scholar

Aydogan, S. (2021). The Genesis and geochemical characteristics of manganese mineralization in the southern part of the Cihanpasa (Yozgat) area. Turk. Geol. Bull. 65 (1), 1–34. doi:10.25288/tjb.975220

CrossRef Full Text | Google Scholar

Bárdossy, G., and Aleva, J. Y. Y. (1990). Lateritic bauxites. Developments in economic geology, 27. Amsterdam: Elsevier.

Google Scholar

Bárdossy, G. (1982). Karst bauxites, bauxite deposits on carbonate rocks. Amsterdam: Elsevier Science and Technology Books, 1–441.

Google Scholar

Bersani, D., and Lottici, P. P. (2010). Applications of Raman spectroscopy to gemology. Anal. Bioanal. Chem. 397, 2631–2646. doi:10.1007/s00216-010-3700-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Bogatyrev, B. A., Zhukov, V. V., and Tsekhovsky, Yu. G. (2009). Formation conditions and regularities of the distribution of large and superlarge bauxite deposits. Lithology Mineral Resour. 44 (2), 135–151. doi:10.1134/s0024490209020035

CrossRef Full Text | Google Scholar

Boynton, W. V. (1984). “Cosmochemistry of the rare Earth elements: Meteoric studies,” in Rare Earth element geochemistry (Amsterdam, Netherlands: Elsevier), 63–114.

CrossRef Full Text | Google Scholar

Calagari, A. A., and Abedini, A. (2007). Geochemical investigations on permo–triassic bauxite horizon at kanisheeteh, east of bukan, west–azarbaidjan, Iran. J. Geochem. Explor. 94, 1–18. doi:10.1016/j.gexplo.2007.04.003

CrossRef Full Text | Google Scholar

Choo, C. O., and Kim, Y. (2003). Textural and spectroscopic studies on hydrothermal dumortierite from an Al-rich clay deposit, southeastern Korea. Mineral. Mag. 67, 799–806. doi:10.1180/0026461036740136

CrossRef Full Text | Google Scholar

Condie, K. C., Marais, D. J., and Abbott, D. H. (2001). Precambrian superplumes and supercontinents: A record in black shales, carbon isotopes, and paleoclimates? Precambrian Res. 106, 239–260. doi:10.1016/s0301-9268(00)00097-8

CrossRef Full Text | Google Scholar

Gemmology Project (2022). Standard Round brillant. Available at: https://www.gemologyproject.com/wiki/images/4/40/SRB800px2.JPG (Accessed September 05, 2022).

Google Scholar

GIA (2016). Round brilliant cut diamond i?llustration. Available at: https://4cs.gia.edu/wp-content/uploads/2016/07/155880-960x960-round-brilliant-cut-diamond-illustration.jpg (Accessed July 18, 2021).

Google Scholar

Gu, J., Huang, Z., Fan, H., Jin, Z., Yan, Z., and Zhang, J. (2013). Mineralogy, geochemistry, and Genesis of lateritic bauxite deposits in the wuchuan–zheng'an–daozhen area, northern guizhou province, China. J. Geochem. Explor. 130, 44–59. doi:10.1016/j.gexplo.2013.03.003

CrossRef Full Text | Google Scholar

Hanilci, N. (2019). Bauxite deposits of Turkey. Modern Approaches in solid earth sciences. MASE 16, 681–730. doi:10.1007/978-3-030-02950-0_15

CrossRef Full Text | Google Scholar

Hanilci, N. (2013). Geological and geochemical evolution of the Bolkardagi bauxite deposits, Karaman, Turkey: Transformation from shale to bauxite. J. Geochem. Explor. 133, 118–137. doi:10.1016/j.gexplo.2013.04.004

CrossRef Full Text | Google Scholar

Hatch, J. R., and Leventhal, J. S. (1992). Relationship between inferred redox potential of the depositional environment and geochemistry of the Upper Pennsylvanian (Missourian) Stark Shale Member of the Dennis Limestone, Wabaunsee County, Kansas, U.S.A. Chem. Geol. 99, 65–82. doi:10.1016/0009-2541(92)90031-y

CrossRef Full Text | Google Scholar

Hatipoglu, M. (2011). Al(Fe, Ti, Si)-mobility and secondary mineralization implications: A case study of the karst unconformity diasporite-type bauxite horizons in Milas (Mugla), Turkey. J. Afr. Earth Sci. 60, 175–195. doi:10.1016/j.jafrearsci.2011.02.009

CrossRef Full Text | Google Scholar

Hatipoglu, M., Helvaci, C., Chamberlain, S. C., and Babalik, H. (2010). Mineralogical characteristics of unusual “Anatolian” diaspore (zultanite) crystals from the Ilbirdagı diasporite deposit, Turkey. J. Afr. Earth Sci. 57, 525–541. doi:10.1016/j.jafrearsci.2010.01.002

CrossRef Full Text | Google Scholar

Huang, S., Xu, J., Chen, C., Li, B., Ye, Z., Chen, W., et al. (2020). Topaz, A Potential Volatile-Carrier in Cold Subduction Zone: Constraint from Synchrotron X-Ray Diffraction and Raman Spectroscopy at High Temperature and High Pressure. Minerals 10 (9), 780. doi:10.3390/min10090780

CrossRef Full Text | Google Scholar

Jenkins, A. L., and Larsen, R. A. (2004). Gemstone Identification Using Raman Spectroscopy. Spectroscopy 19 (4), 20–25.

Google Scholar

Ji, H., Ouyang, Z., Wang, S., and Zhou, D. (2000). Element geochemistry of weathering profile of dolomitite and its implications for the average chemical composition of the upper-continental crust. Sci. China Ser. D. 43, 23–35. doi:10.1007/bf02877828

CrossRef Full Text | Google Scholar

Jones, B., and Manning, D. A. C. (1994). Comparison of Geochemical Indices Used for the Interpretation of Palaeoredox Conditions in Ancient Mudstones. Chem. Geol. 111, 111–129. doi:10.1016/0009-2541(94)90085-x

CrossRef Full Text | Google Scholar

Juan-Farfán, R. E., Bayarjargal, L., Winkler, B., Haussühl, E., Avalos-Borja, M., Refson, K., et al. (2011). Pressure Dependence of the Lattice Dynamics of Diaspore, α-AlO(OH), from Raman Spectroscopy and Density Functional Perturbation Theory. Phys. Chem. Miner. 38 (9), 693–700. doi:10.1007/s00269-011-0442-3

CrossRef Full Text | Google Scholar

Kalkan, M., and Ozpinar, Y. (2018). Geological, Petrographic and Geochemical Characteristics of Pyroclastic Rocks of the Civanadag Region (Gugu-Balıkesir), NW-Turkey. Aksaray Univ. J. Sci. Eng. 2 (2), 95–115. doi:10.29002/asujse.331754

CrossRef Full Text | Google Scholar

Karadag, M. M., Temur, S., Arik, F., and Ozturk, A. (2003). Geological and petrographic characteristics of the Catmakaya (Seydisehir-Konya) bauxite deposit. Geosciences 24 (28), 109–122.

Google Scholar

Kinaci, E. H., and Ozturk, A. (2022). Investigation of Geological, Geochemical and Gemological Characteristics of Diaspore Formations in the Southwest Wing of the Menderes Massif (Bafa Lake - Milas). Turk. Geol. Bull. 65 (3), 309–330. doi:10.25288/tjb.1084167

CrossRef Full Text | Google Scholar

Konak, N., Akdeniz, N., and Ozturk, E. M. (1987). “Geology of the south of Menderes Massif,” in Guide book for the field excursion along Western Anatolia, for the IGCP Project No. 5: Correlation of variscan and pre-variscan events of the alpine-mediterranean mountain belt (Ankara: Mineral Research and Exploration Institute), 42–53.

Google Scholar

Krauskopf, K. B. (1979). Current perspectives in geologic disposal of radioactive wastes. Proc. Symposium Waste Manag., 99–107.

Google Scholar

Ma, J., Wei, G., Xu, Y., Long, W., and Sun, W. (2007). Mobilization and redistribution of major and trace elements during extreme weathering of basalt in Hainan Island, South China. Geochim. Cosmochim. Ac 71, 3223–3237. doi:10.1016/j.gca.2007.03.035

CrossRef Full Text | Google Scholar

Madhavaraju, J., and Lee, I. I. (2009). Geochemistry of the Dalmiapuram Formation of the Uttatur Group (Early Cretaceous), Cauvery Basin, Southeastern India: Implications on Provenance and Paleo-Redox Conditions. Rev. Mex. Ciencias 26, 380–394.

Google Scholar

Mameli, P., Mongelli, G., Oggiano, G., and Dinelli, E. (2007). Geological, geochemical and mineralogical features of some bauxite deposits from Nurra (Western Sardinia, Italy): insights on conditions of formation and parental affinity. Int. J. Earth Sci. 96 (5), 887–902. doi:10.1007/s00531-006-0142-2

CrossRef Full Text | Google Scholar

Meshram, R., and Randive, K. (2011). Geochemical study of laterites of the Jamnagar district, Gujarat, India: Implications on parent rock, mineralogy and tectonics. J. Asian Earth Sci. 42, 1271–1287. doi:10.1016/j.jseaes.2011.07.014

CrossRef Full Text | Google Scholar

Mondillo, N., Balassone, G., Boni, M., and Rollinson, G. K. (2011). Karst bauxites in the Campania Apennines (southern Italy): a new approach. Period. Di Mineral. 80, 1. doi:10.2451/2011PM0028

CrossRef Full Text | Google Scholar

Mongelli, G., Buccione, R., Gueguen, E., Langone, A., and Sinisi, R. (2016). Geochemistry of the apulian allochthonous karst bauxite, Southern Italy: Distribution of critical elements and constraints on Late Cretaceous Peri-Tethyan palaeogeography. Ore Geol. Rev. 77, 246–259. doi:10.1016/j.oregeorev.2016.03.002

CrossRef Full Text | Google Scholar

Mta, (2002). 1:250.000 scaled active fault map of Turkey. Ankara: Denizli Map.

Google Scholar

Nesbitt, H. W., and Young, G. M. (1982). Early Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature 299, 715–717. doi:10.1038/299715a0

CrossRef Full Text | Google Scholar

Nyamsari, D. G., and Yalcin, M. G. (2017). Statistical analysis and source rock of the Minim-Martap plateau bauxite, Cameroon. Arabian J. Geosciences 10, 415–416. doi:10.1007/s12517-017-3172-0

CrossRef Full Text | Google Scholar

Nyamsari, D. G., Yalcin, M. G., and Wolfson, I. (2020). Alteration, Chemical Processes, and Parent Rocks of Haléo-Danielle Plateau Bauxite, Adamawa–Cameroon. Lithology Mineral Resour. 55, 231–243. doi:10.1134/s0024490220030049

CrossRef Full Text | Google Scholar

Okay, A. I., (1985). Relationship and metamorphism of the Menderes massif and allochthonous units between Lake bafa-mugla-usak: Tpao, Exploration Group. Rep., Los Angeles.

Google Scholar

Ozer, O., and Yalcin, M. G. (2020). Correlation of chemical contents of Sutlegen (Antalya) bauxites and regression analysis. AIP Conf. Proc. 2293, 180008. doi:10.1063/5.0026731

CrossRef Full Text | Google Scholar

Ozturk, H. S., Hanilci, N., Cansu, Z., and Kasapci, C. (2021). Formation of Ti-rich bauxite from alkali basalt in continental margin carbonates, Payas region, SE Turkey: implications for sea level change in the Upper Cretaceous. Turkish J. Earth Sci. 30, 116–141. doi:10.3906/yer-2006-6

CrossRef Full Text | Google Scholar

Ozyurt, M., Kirmaci, M., Al-Aasm, I. S., Hollis, C., Taslı, K., and Kandemir, R. (2020). REE Characteristics of Lower Cretaceous Limestone Succession in Gümüshane, NE Turkey: Implications for Ocean Paleoredox Conditions and Diagenetic Alteration. Minerals 10 (8), 683. doi:10.3390/min10080683

CrossRef Full Text | Google Scholar

Pearce, J. A., Harris, N. B. W., and Tindle, A. G. (1984). Trace element discrimination diagrams for the tectonic interpretation of granitic rocks. J. Pet 25, 956–983. doi:10.1093/petrology/25.4.956

CrossRef Full Text | Google Scholar

Phlayrahan, A., Monarumit, N., Satitkune, S., and Wathanakul, P. (2018). Role of Ti Content on the Occurrence of the 3309-cm–1 Peak in FTIR Absorption Spectra of Ruby Samples. J. Appl. Spectrosc. 85, 385–390. doi:10.1007/s10812-018-0662-0

CrossRef Full Text | Google Scholar

Rimmer, S. M. (2004). Geochemical Paleoredox Indicators in Devonian-Mississippian Black Shales, Central Appalachian Basin (USA). Chem. Geol. 206, 373–391. doi:10.1016/j.chemgeo.2003.12.029

CrossRef Full Text | Google Scholar

Robertson, A. H. F., and Ustaömer, T. (2009). Upper Palaeozoic subduction/accretion processes in the closure of Palaeotethys: evidence from the Chios Melange (E Greece), the Karaburun Melange (W Turkey) and the Teke Dere Unit (Sw Turkey). Sediment. Geol. 201, 29–59. doi:10.1016/j.sedgeo.2009.06.005

CrossRef Full Text | Google Scholar

Ruan, H. D., Frost, R. L., and Kloprogge, J. T. (2001). Comparison of Raman Spectra in Characterizing Gibbsite, Bayerite, Diaspore, and Boehmite. J. Raman Spectrosc. 32 (9), 745–750. doi:10.1002/jrs.736

CrossRef Full Text | Google Scholar

Salamab Ellahi, S., Taghipour, B., Zarasvandi, A., Bird, M. I., and Somarin, A. K. (2016). Mineralogy, geochemistry, and stable isotope studies of the Dopolan Bauxite Deposit, Zagros Mountain, Iran. Minerals 6:11, doi:10.3390/min6010011

CrossRef Full Text | Google Scholar

Satir, M., and Friedrichsen, H. (1986). The origin and evolution of the Menderes Massif, W-Turkey: A rubidium/strontium and oxygen isotope study. Geol. Rundsch. 75, 703–714. doi:10.1007/bf01820642

CrossRef Full Text | Google Scholar

Sengor, A. M. C., Satir, M., and Akkok, R. (1984b). Timing of tectonic events in the Menderes massif, Western Turkey: implications for tectonic evolution and evidence for Pan-African basement in Turkey. Tectonics 3, 693–707. doi:10.1029/tc003i007p00693

CrossRef Full Text | Google Scholar

Shao, J., Yang, S., and Li, C. (2012). Chemical indices (CIA and WIP) as proxies for integrated chemical weathering in China: ınferences from analysis of fluvial sediments. Sedimen Geol. 266, 110–120. doi:10.1016/j.sedgeo.2012.03.020

CrossRef Full Text | Google Scholar

Siby, K., Nath, B. N., Ramaswamy, V., Naman, D., Gnaneshwar Rao, T., Kamesh Raju, K. A., et al. (2008). Possible detrital, diagenetic, and hydrothermal sources for Holocene sediments of the Andaman back arc basin. Mar. Geol. 247, 178–193. doi:10.1016/j.margeo.2007.09.006

CrossRef Full Text | Google Scholar

Sidibe, M. F., and Yalcin, M. G. (2019). Petrography, mineralogy, geochemistry, and Genesis of the Balaya bauxite deposits in Kindia region, Maritime Guinea, West Africa. J. Afr. Earth Sci. 149, 348–366. doi:10.1016/j.jafrearsci.2018.08.017

CrossRef Full Text | Google Scholar

Sinisi, R. (2018). Mineralogical and Geochemical Features of Cretaceous Bauxite from San Giovanni Rotondo (Apulia, Southern Italy): A Provenance Tool. Minerals 8 (12), 567. doi:10.3390/min8120567

CrossRef Full Text | Google Scholar

Taylor, S. R., and McLennan, S. M. (1981). The composition and evolution of the continental crust: rare Earth element evidence from sedimentary rocks. Philosophical Transactions of the Royal Society of London. Ser. A, Math. Phys. Sci. 301 (1461), 381–399. doi:10.1098/rsta.1981.0119

CrossRef Full Text | Google Scholar

Temur, S. (2001). Industrial Raw Materials. Konya (Turkey): Line Bookstore, 386.

Google Scholar

Temur, S., Kansun, G., Karadag, M, M., Arık, F., and Ozturk, A. (2003). Stratigraphic location of diasporitic bauxites of Masatdagi (Alanya-Antalya), 109–122.

Google Scholar

Viers, J., and Wasserburg, G. J. (2004). Behavior of Sm and Nd in a lateritic soil profile. Geochimica Cosmochimica Acta 68 (9), 2043–2054. doi:10.1016/j.gca.2003.10.034

CrossRef Full Text | Google Scholar

Wright, J., Schrader, H., and Holser, W. (1987). Paleoredox variations in ancient oceans recorded by rare Earth elements in fossil apatite. Geochim. Cosmochim. Acta 51, 631–644. doi:10.1016/0016-7037(87)90075-5

CrossRef Full Text | Google Scholar

Yalcin, M. G., and Ilhan, S. (2013). Major and Trace Element Geochemistry of Bauxites of Ayranci, Karaman, Central Bolkardag, Turkey. Asian J. Chem. 25, 2893–2904. doi:10.14233/ajchem.2013.14275

CrossRef Full Text | Google Scholar

Yusoff, Z. M., Ngwenya, B. T., and Parsons, I. (2013). Mobility and fractionation of REEs during deep weathering of geochemically contrasting granites in a tropical setting, Malaysia. Chem. Geol. 349, 71–86. doi:10.1016/j.chemgeo.2013.04.016

CrossRef Full Text | Google Scholar

Zamanian, H., Ahmadnejad, F., and Zarasvandi, A. (2016). Mineralogical and geochemical investigations of the Mombi bauxite deposit, Zagros Mountains, Iran. Chem. Erde Geochem. 76, 13–37. doi:10.1016/j.chemer.2015.10.001

CrossRef Full Text | Google Scholar

Keywords: gemology, metabauxite, Raman spectroscopy, Sodra, diaspore (AlOOH)

Citation: Hikmet Kinaci E and Ozturk A (2023) Gemological features of diaspore in sodra—Milas (mugla) region. Front. Earth Sci. 11:1145674. doi: 10.3389/feart.2023.1145674

Received: 16 January 2023; Accepted: 13 March 2023;
Published: 27 March 2023.

Edited by:

Mustafa Gurhan Yalcin, Akdeniz University, Türkiye

Reviewed by:

Javier Garcia-Guinea, Spanish National Research Council (CSIC), Spain
Nurdane Ilbeyli, Akdeniz University, Türkiye
Ender Sarifakioglu, Cankiri Karatekin University, Türkiye

Copyright © 2023 Hikmet Kinaci and Ozturk. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Alican Ozturk, aozturk@ktun.edu.tr

These authors have contributed equally to this work

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.