Next Article in Journal
Filling Characteristics of Radiolarian Siliceous Shell Cavities at Wufeng-Longmaxi Shale in Sichuan Basin, Southwest China
Previous Article in Journal
Aridification in the Asian Interior Recorded by Mineral Assemblages in Tarim Basin since the Late Miocene and Its Link to Global Cooling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Geochemical Approach to Determine the Possible Precipitation Parameters of the Coniacian–Santonian Mazıdağı Phosphates, Mardin, Turkey

by
Derya Yildirim Gundogar
and
Ahmet Sasmaz
*
Department of Geological Engineering, Engineering Faculty, Firat University, Elazig 23119, Turkey
*
Author to whom correspondence should be addressed.
Minerals 2022, 12(12), 1544; https://doi.org/10.3390/min12121544
Submission received: 8 November 2022 / Revised: 22 November 2022 / Accepted: 28 November 2022 / Published: 30 November 2022
(This article belongs to the Section Mineral Deposits)

Abstract

:
The Tethyan phosphates were formed during the Upper Cretaceous and Eocene interval as a result of the collision of the African–Arabian and Eurasian plates and the closing of the Neo-Tethys Ocean. This study aimed to reveal the possible precipitation parameters of these phosphates by examining the main oxide, trace element, and rare earth element contents of the phosphates in the study region. The mean major oxide concentrations of the phosphates were found to be 51.6 wt.% CaO, 21.2 wt.% P2O5, 8.03 wt.% SiO2, 18.1 wt.% CO2, 0.51 wt.% K2O, 0.12 wt.% Fe2O3, 0.05 wt.% Al2O3, 0.18 wt.% MgO, and 0.02 wt.% MnO. The average trace element concentrations were 79 ppm Ba, 1087 ppm Sr, 0.23 ppm Rb, 14.7 ppm Ni, 108 ppm Cr, 262 ppm Zn, 27 ppm Cd, 21.6 ppm Y, 58 ppm V, 6.43 ppm As, 30.3 ppm Cu, 1.36 ppm Pb, 6.32 ppm Zr, 39 ppm U, 0.21 ppm Th, and 1.33 ppm Co. The average trace element contents were 1742 ppm, with this indicating an enrichment assemblage of Sr, Cd, As, and Zn in comparison to PAAS (The Post-Archean Australian Shale). The total REE concentrations in the Mazıdağı phosphates varied from 3.30 to 43.1 ppm, with a mean of 22.1 ppm recorded. All phosphates showed heavy REE (HREE) enrichments and had similar REE patterns to PAAS (The Post-Archean Australian Shale). All samples had strongly negative Ce and positive Eu, Pr, and Y anomalies. These anomalies indicate the existence of oxic and suboxic marine conditions during the formation of the phosphates. According to the proposed genetic model, the phosphates mostly formed in the oxic and suboxic zones of the Tethys Ocean and were precipitated on slopes that depended on strong upwelling from an organic-rich basin in anoxic/suboxic conditions from deeper seawater. The Pb isotope data obtained also indicate the existence of a deep-sea hydrothermal contribution to this phosphate formation.

Graphical Abstract

1. Introduction

Phosphate rock is a marine biochemical rock rich in phosphorus (P), with more than 18 wt.% P2O5 [1,2]. It is one of the most important raw materials, with it being applied, for instance, in agriculture for fertilizer production and in high-tech industries (e.g., REEs) due to the trace element concentrations extracted from it as a by-product [3,4]. Mazıdağı phosphates are part of the “Late Cretaceous–Eocene Giant Phosphate Belt” (Figure 1), which stretches across Jordan, Syria, and Turkey, and through North Africa from the Caribbean Sea in the west [5,6]. Phosphates are distributed widely in modern and ancient marine sediments. Today, phosphates are noted to form at continental margins and continental shelves in upwelling regions such as the Gulf of California, Namibia, Chile, and Peru [7]. Phosphates in this time period have been well studied in many palaeogeographic areas, including Australia [8], China [9,10,11], the West African craton [12,13,14], India [15,16], the Siberian Platform [17], and Mongolia [18]. Economic and subeconomic phosphate accumulations observed in the western margins of Gondwana and Baltica have been less studied [19]. Phosphates generally exhibit similar behavior to REE elements during precipitation. The REE composition of sedimentary phosphates is highly variable among different deposits. REE patterns can range from seawater patterns to shale and bell-shaped patterns [3]. The REE abundances and Ce and Eu anomalies in phosphates in different regions have been investigated by various researchers [4,20,21,22,23,24,25,26,27,28,29]. Phosphates generally have high REE contents and show negative Ce anomalies because of the redox conditions and composition of the depositional environment [13,30,31]. The Coniacian–Santonian Mazıdağı Sedimentary Basin is a region that can be studied in order to obtain the redox conditions and composition of the depositional environment. A number of studies have been conducted on the formation conditions and precipitation of phosphate in this basin [6,32,33,34,35,36,37,38]. The aim of this study was to investigate the major oxide, trace, and rare earth element contents of the studied phosphates and to reveal the physicochemical characteristics of the paleo environment where phosphates are formed.

2. Geological Setting

The studied phosphates are situated on the Arabian Platform, which is one of Turkey’s main tectonic units (Pontides, Anatolides, Taurides, Arabian Platform, and Border Folds) [39]. At the base of the region, the Precambrian Derik volcanics (Telbesmi Formation) crop out, which are composed of basaltic and andesitic pyroclastics, agglomerate, sandstones, and shale [32,33,34,35,36,38,39,40,41]. The Telbesmi Formation is represented around Derik by volcanogenic rocks in which red and grayish-burgundy colors are dominant in interlayers parallel to each other and inclined to the north. Volcanic rocks in the region consist of basalt and andesite lava products. Volcanogenic sandstones can be observed in the formation as interlayers with a thickness of 0.5–1 m. At some levels, andesitic rocks exhibit an agglomeratic/brecciated appearance and contain angular components with a diameter of 0.5–10 mm. Basalts are distinguished from red-colored andesite by their black-colored, fine-grained, and basically porous appearance. It has been noted that this formation was deposited in a shallow marine-coastal environment [39,42]. The predominance of volcanic lithology in the region indicates that volcanism is widespread, and the volcanogenic levels contain pyroclastic lithology concurrent with deposition, as well as lava-type products. The presence of clastic levels indicates that clastic sedimentation accompanies this volcanism. The clastic and carbonates with Cambrian, Ordovician, and Silurian ages in this region uncomfortably cover the Telbesmi Formation. Cretaceous carbonates with phosphate uncomfortably overlie both Precambrian series and Early Paleozoic passive-margin rocks in the study area (Figure 2). Cretaceous carbonates consist of three formations in the study area (Figure 2), the Sabunsuyu, Derdere, and Karababa formations from bottom to top [43]. The Sabunsuyu Formation mainly consists of coarse and thick-bedded crystalline dolomitic limestone. It is also overlain by the Derdere Formation, which consists of thick-bedded limestone, shale, and cherty limestone with Late Cenomanian macrofossils [6,37]. The Derdere Formation passes gradually to the Karababa Formation, which contains phosphate layers from the Late Coniacian to the Early Campanian [6,44]. The Karababa Formation includes a Tasit phosphate layer in its bottom section and has up to a 1.5 m thickness, including carbonates. Therefore, it also has secondary importance as a phosphate ore. The upper part of the Karababa Formation consists of cherty phosphate, limestone, and chert layers that are about 50 m in total thickness. Ghasemian et al. [38] reported that the Karababa Formation consists of limestone, white and gray chert, and reddish and cream phosphate layers.
The Mazıdağı phosphates from bottom to top occur on the gently dipping northern flank of the Derik anticline [6,32]. (1) Tasit phosphate (Turonian), (2) Kasrik (East Kasrik) phosphate, and (3) Şemikan (West Kasrık) phosphate (both Santonian–Coniacian) are all phosphorus-rich members of the Karababa Formation (Figure 3). Furthermore, (4) Akras phosphate (Maastrichtian), a phosphorus-rich member of the Kermav Formation, is located in the eastern part of the study area. The Şemikan (West Kasrik) member is the richest in phosphates, reaching 25%–32% P2O5 in content [32].
The Şemikan phosphate area is one of the areas with the highest economic potential among the Mazıdağı phosphates deposits. It consists of four different sectors. The average phosphate thickness in these sectors is around 0.75 cm. The phosphate level in the region is characterized by interlayers, alternating with chert and carbonate veins (Figure 4). While the oolitic phosphate level in this level is sometimes 20–30 cm thick, in some regions the total cherty, calcareous, and oolitic phosphate level can reach up to 4 m in places (Figure 4). Between these levels, white-cream-colored limestone and chert layers are frequently observed. Their persistence varies across the region, and their general slope is 3o-5o, which dips in the N-NE direction.
Light gray-white, high-grade phosphates and red clay phosphates show transitions in the horizontal direction. These transitions are usually gradual from one color to another. The cover layers on the phosphate in four different areas in the Şemikan region are generally composed of white and cream-colored limestone with layers, and their thickness is usually not less than 10 m. Increases in the thickness of the cover layers on the phosphate negatively affect ore production due to increased costs.

3. Samples and Analytical Methods

3.1. Samples

Thirteen oolitic phosphate ore, limestone, and chert samples with phosphate (nine phosphate ore samples, as well as two limestone with phosphate and two chert samples) were collected from different outcrops of the Şemikan phosphate deposits and subjected to several different analyses. In September–October 2020, different units were sampled from bottom to top, as seen in Figure 4. DY-14 and DY-21 were the chert samples, and DY-02 and DY-13 were the white-colored limestone samples. The other samples belonged to different phosphate levels of the Şemikan phosphate area.

3.2. Analytical Methods

First, in our geochemistry laboratory, the thirteen samples were crushed and ground to 200 mesh. Then, in Bureau Veritas Minerals, a 0.25 g split was heated in HNO3, HClO4, and HF to fuming and then dried. The residue was then dissolved in HCl. Finally, trace and rare earth elements in these samples were analyzed by ICP-MS, and major oxides were analyzed by ICP-AES at the Bureau Veritas Minerals in Canada (www.bvlabs.com accessed on 5 May 2021). Repeated analyses showed better than 5% reproducibility.
Pb isotopes (208Pb, 207Pb, 206Pb, and 204Pb) for phosphate ores were determined by ICP-MS at Bureau Veritas Minerals (Vancouver, Canada). The samples were dissolved with aqua regia (1 HNO3 + 3HCl mixture) and analyzed for 206Pb/204Pb, 207Pb/204Pb, and 208Pb/204Pb ratios. The NIST-SRM 981 standard was used, with the following results obtained: 206Pb/204Pb 16.9374; 207Pb/204Pb 15.4916; 208Pb/204Pb 36.7219 (www.bvlabs.com accessed on 5 May 2021).
The geochemical data were statistically subjected to the SNK (Student–Newman–Keul’s Procedure) using variance analysis (ANOVA) and SPSS 15.0 software. To find Ce, Eu, Y, and Pr anomalies in each sample, REE values were first divided by Post-Archean Australian Shale (PAAS) values (Taylor and McLennan, 1985) and then Eu, Ce, Pr, and Y anomalies were calculated with the following formulas: Ce/Ce* = Cen/√[Lan × Prn], Eu/Eu* = Eun/√[Smn × Gdn], Y/Y* = Yn/√[Dyn × Hon], Pr/Pr* = Prn/√[Cen × Ndn].

3.3. Quality Assurance

The Quality Assurance and ISO accreditations of all of the facilities were performed by recognized organizations. These accreditations and registrations satisfy the ISO standards. All Bureau Veritas Minerals are registered to ISO 9001 (www.bvlabs.com accessed on 5 May 2021).

4. Results and Discussion

4.1. Major Oxide Geochemistry

Among the major oxides in the phosphate ore samples, CaO was the most abundant major oxide, with an average of 51.6 ± 2.15 wt.%. The average P2O5 concentration in all of the the studied samples was 21.2 ± 1.88 wt.%, with this being the second most abundant major oxide in the studied phosphate samples, with high average contents varying from 2.22 to 31.7 wt.%. The average P2O5 content of the phosphate level in the study area was 26.3 ± 3.34 wt.%, except for the limestone (DY-02 and DY-13) and chert samples (DY-14 and DY-21). The P2O5 contents of the limestone samples (DY-02 and DY-13) in the Şemikan phosphate deposits were 4.29 and 10.1 wt.%, respectively (Table 1). The P2O5 contents of the chert samples (DY-14 and DY-21) were 2.22 and 22.3 wt.%, respectively. The average concentrations of SiO2 and Fe2O3 in the samples were 8.03 ± 0.23 wt.% and 0.11 ± 0.01 wt.%, respectively (Table 1). The average contents of Al2O3 and CO2 were 0.04 ± 0.01 and 18.1 ± 0.66 wt.%, respectively. The averages of the other major oxides (MnO: 0.02 wt.%, K2O: 0.01wt.%, MgO: 0.18 wt.%, Na2O: 0.51wt.%, TiO2: 0.01 wt.% and Cr2O3: 0.02 wt.%) were all lower than 1 wt.% (Table 1). The major oxide concentrations of the phosphate ores are listed as follows: CaO > P2O5 > CO2 > SiO2 > Na2O > MgO > Fe2O3 > Al2O3 > K2O > MnO > TiO2. The CaO/P2O5 ratios for the phosphate ores varied from 1.72 to 3.35, with a mean of 2.27 recorded. The average CaO/P2O5 ratio for the limestone and chert samples was 8.85. These phosphate ore ratios show the economic importance of these phosphate raw materials, as when these materials have CaO/P2O5 ratios lower than 1, they are considered more suitable and are therefore highly used in the industry, while CaO/P2O5 ratios higher than 1 increase sulfuric acid consumption during H3PO4 manufacturing, thus reducing their use [45]. The CaO/P2O5 ratio in the phosphate rocks was higher than 1.31, and this may have been related to either the occurrence of calcite or dolomite in whole rocks or the substitution of PO4 by CO3 [46,47]. P2O5 was found to have high positive correlation coefficients with Na2O (0.99), CaO (0.50), V (0.75), Cr (0.79), Sr (0.99), Y (0.91), Zr (0.79), Nb (0.69), Ba (0.93), Hf (0.69), U (0.95), Pb (0.81), As (0.89), and Zn (0.81). Total REEs (0.92) was found to show weak positive correlations with Cu (0.50), Rb (0.52), and Au (0.89), whereas P2O5 was found to show strong negative correlations with SiO2 (−0.48) and CO2 (−0.88) and weak negative correlations with Ni (−0.49 (Table 2; Figure 5).

4.2. Trace Element Geochemistry

The trace element analyses of the phosphate samples determined by the ICP-MS are shown in Table 3. The total trace element concentrations (∑TE) of the phosphate samples varied from 392 to 2684 ppm. The mean trace elements contents were 58 ppm for V, 108 ppm for Cr, 1.33 ppm for Co, 14.7 ppm for Ni, 30.3 ppm for Cu, 0.23 ppm for Rb, 1087 ppm for Sr, 6.32 ppm for Zr, 0.06 ppm for Nb, 79.2 ppm for Ba, 0.06 ppm for Hf, 6.43 ppm for As, 27 ppm for Cd, 39 ppm for U, 0.21 ppm for Th, 1.36 ppm for Pb, 262 ppm for Zn, and 0.04 ppm for Au (Table 3). The phosphates contained higher Cr, Sr, As, Cd, U, and Zn and lower V, Co, Ni, Rb, Zr, Hf, Nb, Ba, Th, and Pb in comparison to PAAS (Taylor and McLennan, 1985) (Figure 6). While the Sr from the large-ion lithophile elements were highly enriched, Rb and Ba were depleted compared to PAAS. Among the high field strength elements (HFSE; Nb, Hf, U, Th, and Zr), Hf, Zr, Nb and Th were significantly lower compared to PAAS, except for U (Figure 6).
The Th/U ratios in the studied phosphates varied from 0.002 to 0.025. The Ni/Co ratios changed from 1.20 to 118, with a mean of 28.2 recorded. The V/Sc and V/Cr ratios were 0.05 (0.03 to 0.11) and 0.54 (0.31 to 1.68), respectively (Table 4). These Ni/Co and V/Cr ratios indicate the oxic/suboxic association between redox conditions and the precipitation of phosphates, except for the Th/U and V/Sc values (Table 4). Furthermore, Algeo and Liu [48] have suggested that the use of these bimetallic indicators is not very reliable for the determination of past seawater conditions. The solubility of certain elements is influenced by the oxidation state of the water column. Marine sediments may therefore be enriched or depleted in these elements in accordance with the redox state of the water column. Suboxic environments are those with low levels of dissolved oxygen and intermediate H2S levels. Anoxic environments are characterized by the presence of both H2S and dissolved oxygen [49]. Redox-sensitive substitutes (Ni/Co, Th/U, V/Cr, V/Sc, and REEs) can show global or regional redox conditions, which can indicate the existence of a paleo-ocean redox environment [50].
As shown in Figure 7, the Nb/Ta and Zr/Hf ratios were similar with North and South Atlantic deep water and Pacific Ocean deep water [51,52,53,54,55]. The HFSE enrichments in phosphate ores result from their absorption in seawater [55,56,57,58]. This also indicates that the low Zr/Hf and Nb/Ta ratios in the phosphates had low HFSE values in seawater during the precipitation of the phosphate ores.

4.3. Rare Earth Element Geochemistry

The rare earth element contents of the phosphate samples are summarized in Table 5. The ΣREE concentrations of the samples ranged from 5.62 ppm to 89.9 ppm. The shale-normalized trend of the studied phosphates relative to PAAS [59] is shown in Figure 8. The REE concentrations of the phosphates were characterized by enriched heavy REEs and relatively depleted light and middle REEs. The mean rare earth element abundances (ppm) in the phosphates were as follows: heavy REEs (2.32) > middle REEs (1.01) > light REEs (0.44) (Figure 8). The Lan/Ybn ratios showed enrichment patterns between heavy REEs and light REEs. The Lan/Ybn values of the phosphates ranged between 0.51 and 0.76, with a mean of 0.60 ± 0.01 recorded (Table 5), which was comparable to the corresponding modern seawater values (0.2–0.5; [62]) of the phosphate samples, verifying the HREE enrichments [63]. The low (La/Yb)n ratios in the phosphate samples indicate that the REE concentrations were associated both with the adsorption of REEs during the evolution of these phosphates and through a substitution mechanism by recrystallization [63]. The Lan/Hon ratios changed from 0.51 to 0.72, with a mean of 0.63 ± 0.02 recorded; this also indicates that the phosphate samples were more enriched in HREEs than in LREEs. Furthermore, the (Sm/Pr)n and (Sm/Yb)n values of the phosphates indicate that they were enriched in HREEs (Figure 9). As shown in Figure 9, phosphates from Mazıdağı, Alborz [22], Tebessa [20], Seamount [64], Yangtze [65], and Hazm Al-Jalamaid [66] have been found to be enriched in HREEs and depleted in LREEs and MREEs. Sonrai [27] and Gorgan [25] deposits have been found to be enriched in MREEs and depleted in LREEs and HREEs.
The Ce/Ce* values of the studied phosphate samples changed from 0.16 to 0.22 and had strong negative Ce anomalies (Table 5). The Eu/Eu* and Ce/Ce* ratios were used to determine the precipitation conditions of the depositional environment, such as the temperature, pH, and ƒO2 [67,68]. Of all the rare earth elements, cerium is the most useful element for investigating the redox potentials of different sedimentary environments. Ce oxidizes readily and is irreversibly and continuously removed from seawater at the surface of oxide minerals; therefore, among all the REEs, it is best suited to being used to understand the genesis of Fe–Mn deposits [69]. The formation of tetravalent Ce is certainly not irreversible, and reduction does occur under anoxia, leading to positive Ce anomalies [70]. Negative Ce anomalies indicate both that there is excess oxygen in seawater and that the rapid precipitation of oxide minerals has occurred [71]. Ce is primarily Ce4+ under seawater pH and Eh conditions, and its solubility is very low. It precipitates as Ce4+ (CeO4) because of its strong depletion in seawater. Ce in oxidation states can occur in both the +3 and +4, although this only occurs in oxic environments [72].
The Eu/Eu* anomalies of the phosphate samples changed from 1.22 to 2.33 (Table 5). In comparison to PAAS, all the phosphate samples had strong positive Eu anomalies. Eu is usually found in a trivalent form, like other REEs. However, Eu+3 can be reduced to Eu+2 in reducing environments [74]. Therefore, according to Bau et al. [75], Eu can be used to investigate sedimentation conditions. Generally, a positive Eu anomaly is interpreted as a proxy for reducing conditions, high temperatures, and a hydrothermal REE supply [20,75]. An Eu anomaly can also be associated with other factors, such as a decrease in temperature or increases in ƒO2 and pH [76]. While positive Eu anomalies indicate the existence of sub-reduced conditions during the deposition of the sub-layer of phosphates [20], negative Eu anomalies are not indicative of oxic conditions. Under oxic conditions, Eu is strictly trivalent and behaves similarly to other trivalent REEs, leading to Eu/Eu* values of ~1 [20,75,77,78,79]. In this study, the oxic water samples had a REE+Y trend with heavier REE enrichment and a more negative Y anomaly in comparison to the anoxic conditions. The Y concentrations of the phosphates in the study area change from 3.0 ppm to 46.8 ppm, and these have stronger positive Y anomalies (from 3.11 to 4.20) compared to PAAS (Table 5). Yttrium is considered a REE element mainly because of its similar behavior to Ho. According to its ionic radius, it is usually inserted between Ho and Dy in the REE series [68,80]. Y content values can be used to assess post-deposition changes during fluid circulation [81]. These values show positive anomalies in seawater and are not affected by redox conditions, but the magnitude of these anomalies may decrease under diagenetic effects [68,74,81,82]. Kechiched et al. [20] suggested that the southern Algerian phosphates have low Y/Y* ratios and, therefore, that they reflect slight early diagenesis. However, in this study, while the phosphate samples were mostly affected by possible weathering at the top of the series, they were not affected by strong late diagenetic processes. The anoxic hypersaline brine waters have a negative Y anomaly in comparison to overlying oxic seawater in the Tyro sub-basin [83].

4.4. Source of REE Enrichment in Phosphate Samples

Emsbo et al. [3] have reported that REE contents in phosphates are generally homogeneous within a specific geologic time but show variety among different geologic times. Modern phosphate rocks have low REE contents and similar patterns to modern seawater, showing that rare earth elements were fractionated minimally [84]. In this study, the Mazıdağı phosphates (Figure 10) were observed to have lower REE contents than phosphates of similar geologic ages, such as Tunisian, Moroccan, and Algerian phosphates [22]. Furthermore, the REE, Lan/Ybn, Ce/Ce*, Gdn/Ybn, and Eu/Eu* values of the Mazıdağı phosphates were lower than those of these Tunisian, Moroccan, and Algerian phosphates [85], except for their Eu/Eu* values. As seen in Figure 10, the Algerian, Tunisian, and Moroccan phosphates were found to have negative Eu anomalies of 0.70, 0.71, and 0.66, respectively. The REE contents of phosphates are intensely controlled by geologic processes, and many researchers have suggested that they resemble modern seawater [30,86]. Differences in REE abundances have been related to the depth and duration of precipitation, as well as facies and grain size [74,87]. REE concentrations in phosphate samples vary due to differences in ocean chemistry, the variability of clastic phases, basin configuration, and the degree of influence of hydrothermal fluids [3,30,88]. Alternatively, extreme REE concentrations in phosphates are considered to be caused by the redistribution of REEs between authigenic and detrital phase forms during the diagenetic equilibration of phosphate with pore water [87,89].

4.5. Pb Isotope Geochemistry

The average concentrations (ppm) of 204Pb, 206Pb, 207Pb, and 208Pb isotopes in the Mazıdağı phosphates were 0.028 ppm, 0,54 ppm, 0.447 ppm, and 1.102 ppm, respectively (Table 6). The average 206Pb/204Pb, 207Pb/204Pb, and 208Pb/204Pb ratios were 19.18, 15.70, and 39.04, respectively (Table 6). According to Wang et al. [90], these Pb isotope results show that the lead in the Mazıdağı phosphates was of a lower crust origin (Figure 11a,b). In particular, the 207Pb/204Pb isotope ratios and their potential sources show that the Mazıdağı phosphates were sourced mostly from fluids or metals with a lower crust origin [91]. These ratios indicate that Pb isotope data can be related to oceanic crust areas, such as Atlantic and Pacific crust areas and pelagic sediments (Figure 11c,d). Dissolved Pb in seawater mostly derives from terrestrial input to the oceans, young continental margins, Pb derived from mid-ocean ridge basalts (MORB), the dissolution of wind-blown debris, and island arcs [92]. Highly radiogenic Pb isotope results in the North Atlantic have been associated with intense weathering processes in the former cratonic regions of northern Canada and Greenland [92]. In the present study, the Pb isotope results observed in the Fe–Mn crusts in deep water show that the highest 206Pb/204Pb radiogenic compositions occurred in the NW Atlantic (>19.1) whereas the lowest 206Pb/204Pb ratios occurred in the Southern Ocean (<18.9) and central North Pacific (<18.7). The Fe–Mn crusts spanned the water depth range of ~700 and 4600 m, including within the present-day NE Atlantic deep water and Mediterranean outflow water [93]. At the same time, the studied Pb isotope data also covered the same area as terrigenous, pelagic, biogenic sediments, and Mn nodules [92]. The Pb isotope ratios obtained indicate that the potential source of Pb in Mazıdağı phosphate ores was seawater from the shallow marine basin of the Tethys Ocean during the Upper Cretaceous period. This appears to be consistent with the geological and geochemical evidence [94,95]. Further, the Pb isotope ratios obtained indicate that hydrothermal activity at mid-ocean ridges contributes to the formation of phosphate deposits, such as the early Oligocene Paratethyan manganese deposits in Nikopol (Ukraine) [96] and Chiatura (Georgia) [97].

4.6. Origin of the Phosphate and Its Genetic Model

Marine phosphates form as irregular masses, layers, sands, oolites, and pellets along the continental margins in the different periods [98]. While some scientists have proposed that phosphates formed directly as a result of the inorganic precipitation of the phosphates from seawater, Burnett [98] reported that the phosphate concentrations in the pore waters are higher than in the bottom waters. Therefore, it has been suggested that marine phosphates can be formed by precipitation of phosphate in the pore waters of anoxic sediments. The presence of phosphates in pore waters is necessary for phosphate formation. REE concentrations in phosphates are up to 50–100 times higher than both shale-normalized values and bulk REE content in modern seawater [99]. Therefore, it is unlikely that phosphates were precipitated directly from seawater. Felitsyn and Morad [100] have indicated that the most important source of REEs in phosphate ores is organic material, and they suggested that the most important mechanism to explain REE enrichment in authigenic phosphate is the direct ejection of REEs from marine pore waters. In addition, it has been suggested that hydrothermal fluids may be sources of phosphate and REEs [101]. Besides seawater and hydrothermal fluids, diagenetic pore water is considered the main source of REEs in sedimentary phosphates [3].
In order to explain the formation of Mazıdağı phosphates, a genetic model was developed, which is shown in Figure 12. According to this model, seawater can be separated into several different layers, such as the upper oxic layer (oxygen content: 2.0–8.0 mL/L), the dysoxic layer (oxygen content: 0.2–2.0 mL/L), the suboxic layer (oxygen content: 0.0–0.2 mL/L), and the bottom anoxic layer (oxygen content: 0.0 mL/L), based on redox conditions [102]. A previous study found that phosphates were precipitated on slopes during the strong upwelling stage in an organic-rich basin of deeper anoxic/suboxic conditions [103]. Most of the settled organic material in the sediment–water interface of the seafloor is, firstly, oxidized; then, phosphate moves to the water above, along with the sediment pore water. At the same time, phosphate is adsorbed by the iron/manganese oxides and moves up to the slopes together with benthic flux containing large amounts of trace elements and REEs [104]. The Fe–Mn oxides in benthic flux enter pore water together with REEs and PO43−. PO43− combines with Ca2+ to form phosphates, and eventually, PO43−, trace elements, and many REE phases (with most of these elements likely originating from hydrothermal vents (Figure 1 and Figure 12) collect in ores [10,58,96,97].
The phosphate deposits of the eastern Mediterranean region are almost always associated with chert, porcelanite, and oyster-bearing limestone, in addition to minor marl, chalk, and sandstone [105,106,107,108,109]. In this study, within the Mazıdağı phosphate deposits, chert levels were observed with a thickness of 25–40 cm in certain places (Figure 4). The development of chert levels is thought to have come to an end during the precipitation of carbonates and phosphates, or otherwise to have been precipitated simultaneously with the carbonates. Chert levels have also been observed in Moroccan, Jordanian, Tunisian, and Syrian phosphate deposits on the same belt [20,106,109,110,111,112]. Tlili et al. [112] claim that chert levels may be related to the bacterial activity observed in chert beds, which induces the dissolution of silica, which, in turn, saturates the depositional environment through silicic acid. Another mechanism that enhances this process is oceanic upwelling, which brings P and Si from the deep ocean (which is rich in these and other elements) to the surface. Organic matter preserved in shallow marine sediments and containing P and Si is then released and enriched in the pore water of the sediments through microbial mediation and diagenesis, leading to the precipitation of phosphate and siliceous deposits [6,102,103,104,105,106,107,113,114]. In addition, the significant enrichment of some trace elements (As, Cd, Cu, U, and Zn) in phosphate ores, according to PAAS, indicates the presence of hydrothermal contributions (Figure 12).

5. Conclusions

The Mazıdağı phosphates were formed on a shallow marine platform located in the south of the Neo-Tethys Ocean between the African–Arabic and Eurasian plates during the Upper Cretaceous period. The Mazıdağı phosphate deposits from bottom to top occur on the gently dipping northern flank of the Derik anticline. (1) Tasit phosphate (Turonian), (2) Kasrik (East Kasrik) phosphate, and (3) Şemikan (West Kasrık) phosphate (Santonian–Coniacian) are all phosphorus-rich members of the Karababa Formation. (4) Akras phosphate (Maastrichtian) is a phosphorus-rich member of the Kermav Formation, which is located in the eastern part of the study area. The Şemikan (West Kasrik) member was found to be the richest in phosphate, reaching 25–32% P2O5 in content. This Şemikan phosphate is the phosphate with the highest economic potential among the Mazıdağı phosphates. It consists of four different sectors. The average phosphate thickness in these sectors was found to be around 0.75 cm. The phosphate levels in the region were characterized by interlayers, alternating with chert, clay, and carbonate veins. While the oolitic phosphate levels were sometimes 20–30 cm thick, in some regions the total cherty, calcareous, and oolitic phosphate levels reached up to 4 m in places. The major oxide concentrations of the phosphate ores were CaO > P2O5 > LOI > SiO2 > Na2O > MgO > Fe2O3 > Al2O3 > K2O > MnO > TiO2. The CaO/P2O5 ratios for phosphate ores varied from 1.72 to 3.35, with a mean of 2.27 recorded. The average CaO/P2O5 ratio for limestone and chert samples was 8.85. The total trace element contents in the studied phosphate nodules varied from 395 to 2712 ppm, with a mean of 1742 ppm obtained. The phosphates contained more Cr, Sr, As, Cd, U, and Zn and lower amounts of V, Co, Ni, Rb, Y, Zr, Hf, Nb, Ba, Th, and Pb in comparison to PAAS. While the Sr from the large-ion lithophile elements was highly enriched, Rb and Ba were depleted compared to PAAS. Among the high field strength elements (HFSE; Nb, Hf, U, Th, Zr, and Y) Hf, Zr, Nb, Y, and Th were significantly lower compared to PAAS, except for U. Samples DY-24 and DY-25, taken from the Şemikan 3 region, contained high Au concentrations (0.14 and 0.15 ppm). Some trace element ratios indicated that the precipitation of the phosphate ores took place mainly in oxic and suboxic zones. The phosphate ores had low total REE concentrations, ranging between 3.30 and 22.1 ppm, and were found to have been enriched with mostly heavy REEs; they had lower amounts of middle REEs and light REEs when compared to PAAS. All of the the phosphate samples had negative Ce and positive Eu, Y, and Pr anomalies. The studied Pb isotope data covered the same area as terrigenous, pelagic, biogenic sediments, and Mn nodules. These ratios indicate that Pb isotope data can be related to oceanic crust areas such as the Atlantic and Pacific crust areas and pelagic sediments. The Pb isotope ratios obtained are also consistent with the existence of a hydrothermal contribution to ore deposits.
The source of phosphorus in shallow marine environments is related to the transport of phosphorus to oceans as a result of the weathering of continental rocks or the uplift of phosphorus and silica carried by cold water from deep-sea environments. According to the proposed model, phosphorus and silica in shallow marine environments are derived from both elements being recycled between the deep oceanic reservoir and the surface water by upwelling currents, followed by sedimentation. Upwelling currents spread deep, cold marine water to sea surfaces rich in Si and P. These elements are the basic nutrients for phytoplankton, the lowest tier in the marine food chain, which inhabit the photic zone or the upper 100 to 200 m of the seawater column. Phytoplankton follows zooplankton and other higher marine life, and bioproductivity is thus greatly enhanced by this process. Most of the oxygen in the water column below the photic zone is depleted by aerobic bacteria feeding on the rain of dead organic matter, thus creating an oxygen minimum zone. In shallow-water epicontinental shelves, such as those of the Neo-Tethys Afro-Arabian margins, the oxygen minimum zone coincides with the sea floor, thus producing a favorable anaerobic environment for the preservation of organic matter. The continuation of this regime maintains a high rate of sedimentation and, consequently, a high rate of burial, as well as the formation of organic-rich sediments.

Author Contributions

Conceptualization, D.Y.G.; methodology, D.Y.G.; investigation, D.Y.G.; resources, D.Y.G.; writing—original draft preparation, D.Y.G. and A.S.; writing—review and editing, D.Y.G. and A.S.; visualization, D.Y.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Cengiz Holding-Eti Gübre mining company.

Acknowledgments

We thank the Editor-in-Chief, Paul Sylvester and Assistant Editor, Charisma Lu, and the anonymous reviewers for their very constructive and helpful comments. This study was financially supported by Cengiz Holding-Eti Gübre mining company (Turkey). We thank Cengiz Holding-Eti Gübre for its support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nelson, G.J.; Pufahl, P.K.; Hiatt, E.E. Paleoceanographic constraints on Precambrian phosphorite accumulation, Baraga Group, Michigan, USA. Sediment. Geol. 2010, 226, 9–21. [Google Scholar] [CrossRef]
  2. Dassamiour, M.; Mezghache, H.; Raji, O.; Bodinier, J.L. Depositional environment of the Kef Essennoun phosphorites (northeastern Algeria) as revealed by P2O5 modeling and sedimentary data. Arab. J. Geosci. 2021, 14, 1078. [Google Scholar] [CrossRef]
  3. Embso, P.; McLaughlin, P.I.; Breit, G.N.; Bray, E.A.; Koenig, A.E. Rare earth element in sedimentary phosphate deposits: Solution to the global REE crisis? Gondwana Res. 2015, 27, 776–785. [Google Scholar]
  4. Kechiched, R.; Laouar, R.; Bruguier, O.; Salmi-Laouar, S.; Kocsis, L. Glauconite-bearing sedimentary phosphorites from the Tébessa region (eastern Algeria): Evidence of REE enrichment and geo-chemical constraints on their origin. J. Afr. Earth Sci. 2018, 145, 190–200. [Google Scholar] [CrossRef]
  5. Soudry, D.; Glenn, C.R.; Nathan, Y.; Segal, I.; Vonder Haar, D. Evolution of Tethyan phosphogenesis along the northern edges of the Arabian–African shield during the Cretaceous-Eocene as deduced from temporal variations of Ca and Nd isotopes and rates of P accumulation. Earth Sci. Rev. 2006, 78, 27–57. [Google Scholar] [CrossRef]
  6. Ghasemian, S.; Ozturk, H.; Cansu, Z. Geochemistry of red and cream phosphorites from the Şemikan phosphorite deposit, SE Turkey: Implication for phosphorite deposition conditions in the Upper Cretaceous. J. Afr. Earth Sci. 2022, 185, 104398. [Google Scholar] [CrossRef]
  7. Schulz, H.N.; Schulz, H.D. Large sulfur bacteria and the formation of phosphorite. Science 2005, 307, 416–418. [Google Scholar] [CrossRef] [PubMed]
  8. Creveling, J.R.; Johnston, D.T.; Poulton, S.W.; Kotrc, B.; Marz, C. Phosphorus sources for phosphatic Cambrian carbonates. Geol. Soc. Am. Bull. 2013, 126, 145–163. [Google Scholar] [CrossRef]
  9. Ling, H.F.; Chen, X.; Li, D.; Wang, D.; Shields-Zhou, G.A. Cerium anomaly variations in Ediacaran–earliest Cambrian carbonates from the Yangtze Gorges area, South China: Implications for oxygenation of coeval shallow seawater. Precambrian Res. 2013, 225, 110–127. [Google Scholar] [CrossRef]
  10. Jiang, S.Y.; Zhao, H.X.; Chen, Y.-Q.; Yang, T.; Yang, J.-H. Trace and rare earth element geochemistry of phosphate nodules from the lower Cambrian black shale sequence in the Mufu Mountain of Nanjing, Jiangsu province, China. Chem. Geol. 2007, 244, 584–604. [Google Scholar] [CrossRef]
  11. Chen, D.F.; Dong, W.Q.; Qi, L.; Chen, G.Q.; Chen, X.P. Possible REE constraints on the depositional and diagenetic environment of Doushantuo Formation phosphorites containing the earliest metazoan fauna. Chem. Geol. 2003, 201, 103–118. [Google Scholar] [CrossRef]
  12. Flicoteaux, R.; Trompette, R. Cratonic and foreland Early Cambrian phosphorites of West Africa: Palaeoceanographical and climatical contexts. Palaeogeogr. Palaeoclimatol. Palaeoecol. 1998, 139, 107–120. [Google Scholar] [CrossRef]
  13. Bertrand-Sarfati, J.; Flicoteaux, R.; Moussine-Pouchkine, A.; Ait Kaci, A.A. Lower Cambrian apatiticstromatolites and phosphoarenites related to the glacioeustatic cratonic rebound (Sahara, Algeria). J. Sed. Res. 1997, 67, 957–974. [Google Scholar]
  14. Blot, A.; Affaton, P.; Seddoh, K.F.; Aregba, A.P.; Godonou, S.K. Phosphates du Proterozo€ıque superieur dans la chaıne des Dahomeyides (circa 600 Ma) de la region de Bassar (Nord-Togo, Afrique de l’Ouest). J. Afr. Earth Sci. 1988, 7, 159–166. [Google Scholar] [CrossRef]
  15. Banerjee, D.M.; Schidlowski, M.; Siebert, F.; Brasier, M.D. Geochemical changes across the Proterozoic-Cambrian transition in the Durmala phosphorite mine section, Mussoorie Hills, Garhwal Himalaya, India. Palaeogeogr. Palaeoclimatol. Palaeoecol. 1997, 132, 183–194. [Google Scholar] [CrossRef]
  16. Mazumdar, A.; Banerjee, D.M.; Schidlowski, M.; Balaram, V. Rare-earth elements and stable isotope geochemistry of early Cambrian chert-phosphorite assemblage from the Lower Tal Formation of the Krol Belt (Lesser Himalaya, India). Chem. Geol. 1999, 156, 275–297. [Google Scholar] [CrossRef]
  17. Ilyin, A.V. Phosphorites of the Russian Craton. Earth-Sci. Rev. 1998, 45, 89–101. [Google Scholar] [CrossRef]
  18. Anderson, R.P.; Macdonald, F.A.; Jones, D.S.; McMahon, S.; Briggs, D.E.G. Dosuhantuo-type microfossils from latest Ediacaran phosphorites of northern Mongolia. GEOL 2017, 45, 1079–1082. [Google Scholar] [CrossRef]
  19. Wallin, B. Origin of the Lower Cambrian phosphate bed at Vassbo, Sweden. Terra Nova 1989, 1, 274–279. [Google Scholar] [CrossRef]
  20. Kechiched, R.; Laouar, R.; Bruhuier, O.; Kocsis, L.; Salmi-Laouar, S. Comprehensive REE + Y and sensitive redox trace elements of Algerian phosphorites (Tébessa, eastern Algeria): A geochemical study and depositional environments tracking. J. Geochem. Expl. 2020, 208, 106396. [Google Scholar] [CrossRef]
  21. Li, M. Evaluation of paired carbon isotopic signals from the Ediacaran Doushantuo phosphorites: Diagenetic or primary? Precambrian Res. 2020, 349, 105502. [Google Scholar] [CrossRef]
  22. Salama, W.; Khirekesh, Z.; Amini, A.; Bafti, B.S. Diagenetic evolution of the upper Devonian phosphorites, Alborz mountain range, northern Iran. Sediment. Geol. 2018, 376, 90–112. [Google Scholar] [CrossRef]
  23. Zhu, B.; Jiang, S.Y. A LA-ICP-MS analysis of rare earth elements on phosphatic grains of the Ediacaran Doushantuo phosphorite at Weng’an, South China: Implication for depositional conditions and diagenetic processes. Geol. Mag. 2017, 154, 1391–1397. [Google Scholar] [CrossRef]
  24. Abou El-Anwar, E.A.; Mekky, H.S.; Abd El Rahim, S.H.; Aita, S.K. Mineralogical, geochemical characteristics and origin of late cretaceous phosphate in Duwi formation (Geble Duwi mine), Red Sea region, Egypt. Egypt. J. Pet. 2017, 26, 157–169. [Google Scholar] [CrossRef] [Green Version]
  25. Abedini, A.; Calagari, A.A. REEs geochemical characteristics of lower Cambrian phosphatic rocks in the Gorgan-Rasht Zone, northern Iran: Implications for diagenetic effects and depositional conditions. J. Afr. Earth Sci. 2017, 135, 115–124. [Google Scholar] [CrossRef]
  26. Abed, A.M.; Jaber, O.; Alkuisi, M.; Sadaqah, R. Rare earth elements and uranium geochemistry in the Al-Kora phosphate province, Late Cretaceous, Northwestern Jordan. Arab. J. Geosci. 2016, 9, 187. [Google Scholar] [CrossRef]
  27. Khan, K.F.; Dar Shamim, A.; Khan Saif, A. Geochemistry of phosphate bearing sedimentary rocks in parts of Sonrai block, Lalitpur district, Uttar Pradesh, India. Chem. Erde. 2012, 72, 117–125. [Google Scholar] [CrossRef]
  28. Khan, S.A.; Khan, K.F.; Dar, S.A. REE geochemistry of early Cambrian phosphorites of Masrana and Kimoi blocks, Uttarakhand, India. Arab. J. Geosci. 2016, 9, 1–10. [Google Scholar] [CrossRef]
  29. Xin, H.; Jiang, S.; Yang, J.; Wu, H.; Pi, D. Rare earth element geochemistry of phosphatic rocks in Neoproterozoic Ediacaran Doushantuo Formation inhushan section from the Yangtze Gorges area, South China. J. Earth. Sci. 2016, 27, 204–210. [Google Scholar] [CrossRef]
  30. Picard, S.; Lecuyer, C.; Bahat, J.; Garciad, J.; Dromart, G. Rare-earth element contents of Jurassic fish and reptile teeth and their potential relation to seawater composition (Anglo-Paris Basin, France and England). Chem Geol. 2002, 18, 1–16. [Google Scholar] [CrossRef]
  31. Yaomin, Y.; Guangzhi, T.; Ruizhong, H. REE and trace element geochemistry of Yinachang Fe-Ca-REE deposit, Yunuan Province, China. Chin. J. Geochem. 2004, 23, 265–274. [Google Scholar] [CrossRef]
  32. Berker, E. The Mardin–Mazıdağı–Derik phosphate deposits, south east Turkey. In Phosphate Deposits of the World. Phosphate Rock Resources, 2; Notholt, A.J.G., Sheldon, R.P., Davidson, D.F., Eds.; Cambridge University Press: Cambridge, MA, USA, 1989; pp. 380–386. [Google Scholar]
  33. Aydin, I.; Imamoglu, S.; Aydin, F.; Saydut, A.; Hamamci, C. Determination of mineral phosphate species in sedimentary phosphate rock in Mardin, SE Anatolia, Turkey by sequential extraction. Microchem. J. 2009, 91, 63–69. [Google Scholar] [CrossRef]
  34. Varol, B. Sedimentary petrography and origin of phosphate peloids of the Mazıda_g-Derik Area (Mardin, Southeast Turkey). Min. Res. Expl. Bull. 1989, 109, 65–73. [Google Scholar]
  35. Varol, B.; Shafiri, J.G.J. Üst Kretase glakonilerinin (Kilis, GD Türkiye) sedimenter petrografisi ve çokel ortamları. Turk. J. Earth Sci. 1995, 4, 13–26. [Google Scholar]
  36. Imamoglu, M.S.; Nathan, Y.; Çoban, H.; Soudry, D.; Glenn, C.R. Geochemical, mineralogical and isotopic signatures of the Şemikan, west Kasrık ‘‘Turkish’’ phosphorites from the derik–mazıdağı–mardin area, SE Anatolia. Int. J. Earth Sci. 2009, 98, 1679–1690. [Google Scholar] [CrossRef]
  37. Yilmaz, I.O.; Cook, T.D.; Hosgör, I.; Wagreich, M.; Rebman, K.; Murray, A.M. The upper Coniacian to upper Santonian drowned Arabian carbonate platform, the Mardin–Mazıdağı area, SE Turkey: Sedimentological, stratigraphic, and ichthyofaunal records. Cretac. Res. 2018, 84, 153–167. [Google Scholar] [CrossRef]
  38. Zhu, B.; Jiang, S.; Yang, J.; Pi, D.; Ling, H.; Chen, Y. Rare earth element and Sr–Nd isotope geochemistry of phosphate nodules from the lower Cambrian Niutitang Formation, NW Hunan Province, South China. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2014, 398, 132–14339. [Google Scholar] [CrossRef]
  39. Şengör, A.M.C.; Yılmaz, Y. Tethyan evolution of Turkey: A plate tectonic approach. Tectonophysics 1981, 75, 181–241. [Google Scholar] [CrossRef]
  40. Perinçek, D. Sedimentation on the Arabian Shelf under the Control of Tectonic Activity in Taurid Belt. In Proceedings of the 5th Petroleum Geologists Congress, Ankara, Turkey, 1980; pp. 77–93. [Google Scholar]
  41. Gürsu, S.; Moller, A.; Goncüoglu, M.C.; Koksal, S.; Demircan, H.; Koksal, F.T.; Kozlu, H.; Sunal, G. Neoproterozoic continental arc volcanism at the northern edge of the Arabian Plate, SE Turkey. Precambrian Res. 2015, 258, 208–233. [Google Scholar] [CrossRef]
  42. Tetiker, S.; Yalçın, H.; Bozkaya, O. Evidence of the diagenetic history of sediment composition in Precambrian-Early Paleozoic rocks: A study from the Southeast Anatolian Autochthon, Mardin (Derik-Kızıltepe), Turkey. Arab. J. Geosci. 2015, 5, 11261–11278. [Google Scholar] [CrossRef]
  43. Yılmaz, E.; Duran, O. Güneydoğu Anadolu Bölgesi otokton ve allokton birimler stratigrafi adlama sözlüğü ‘Lexicon’; Türkiye Petrolleri Anonim Ortaklığı: Çankaya, Turkey, 1997; p. 460. [Google Scholar]
  44. Cater, J.M.L.; Gillcrist, J.R. Karstic reservoirs of the mid-Cretaceous Mardin Group, SE Turkey: Tectonic and eustatic controls on their genesis, distribution and preservation. J. Petrol. Geol. 1994, 17, 253–278. [Google Scholar] [CrossRef]
  45. Lauriente, D.H. Phosphate industry overview. In Chemical Economics Handbook; SRI International: Zurich, Switzerland, 1996; 94p. [Google Scholar]
  46. Slansky, M. Geology of Sedimentary Phosphate; North Oxford Academic Publishing Company Limited: Oxford, UK, 1986. [Google Scholar]
  47. Awadalla, G.S. Geochemistry and microprobe investigations of Abu Tartur REE-bearing phosphorite, Western Desert. Egypt. J. Afr. Earth Sci. 2010, 57, 431–443. [Google Scholar] [CrossRef]
  48. Algeo, T.J.; Liu, J. A re-assessment of elemental proxies for paleoredox analysis. Chem. Geol. 2020, 540, 119549. [Google Scholar] [CrossRef]
  49. Yao, C.; Guo, W.; Liu, J.; Li, H. Multiple Proxies on the Paleoenvironment of the Early Cambrian Marine Black Rock Series in the Tarim Basin, NW China: Molybdenum Isotope and Trace Element Evidence. Int. J. Geosci. 2017, 8, 965–983. [Google Scholar] [CrossRef]
  50. Jones, B.J.; Manning, A.C. Comparison of geochemical indices used for the interpretation of palaeoredox conditions in ancient mudstones. Chem. Geol. 1994, 111, 111–129. [Google Scholar] [CrossRef]
  51. Schmidt, K.; Bau, M.; Hein, J.R.; Koschinsky, A. Fractionation of the geochemical twins Zr/Hf and Nb/Ta during scavenging from sea-water by hydrogenetic ferromanganese crusts. Geochim. Cosmochim. Acta 2014, 140, 468–487. [Google Scholar] [CrossRef]
  52. Censi, P.; Randazzo, L.A.; Zuddas, P.; Saiano, F.; Aricò, P. Trace element behavior in seawater during Etna’s pyroclastic activity in 2001: Concurrent effects of nutrients and formation of alteration minerals. J. Volcanol. Geotherm. Res. 2010, 193, 106–116. [Google Scholar] [CrossRef]
  53. Censi, P.; Inguaggiato, C.; Chiavetta, S.; Schembri, C.; Sposito, F. The behavior of zirconium, hafnium and rare earth elements during the crystallization of halite and other salt minerals. Chem. Geol. 2017, 453, 80–91. [Google Scholar] [CrossRef]
  54. Godfrey, L.V.; Zimmermann, B.; Lee, D.C.; King, R.L.; Vervoort, J.D. Hafnium and neodymium isotope variations in NE Atlantic seawater. Geochem. Geophys. Geosyst. 2009, 10, 1–13. [Google Scholar] [CrossRef]
  55. Firdaus, M.L.; Minami, T.; Norisuye, K.; Sohrin, Y. Strong elemental fractionation of Zr–Hf and Nb–Ta across the Pacific Ocean. Nat. Geosci. 2011, 4, 227–230. [Google Scholar] [CrossRef]
  56. Sasmaz, A. The Atbara porphyry gold–copper systems in the Red Sea Hills, Neoproterozoic Arabian–Nubian Shield, NE Sudan. J. Geochem. Expl. 2020, 214, 106539. [Google Scholar] [CrossRef]
  57. Censi, P.; Raso, M.; Saiano, F.; Zuddas, P.; Oliveri, E. Zr/Hf ratio and REE behavior: A coupled indication of lithogenic input in marginal basins and deep-sea brines. Deep-Sea Res. Part II 2019, 164, 216–223. [Google Scholar] [CrossRef]
  58. Sasmaz, A.; Zuddas, P.; Cangemi, M.; Piazzese, D.; Ozek, G.; Censi, P. Zirconium and hafnium fractionation and distribution of Rare Earth Elements in neutral–alkaline waters: Case study of Lake Van hydrothermal system, Turkey. J. Geochem. Expl. 2021, 226, 106784. [Google Scholar] [CrossRef]
  59. Taylor, S.R.; McLennan, S.M. The Continental Crust: Its Composition and Evolution; Blackwell Scientific: Oxford, UK, 1985; p. 312. [Google Scholar]
  60. Godfrey, L.V.; Mills, R.; Elderfield, H.; Gurvich, E. Lead behavior at TAG hydrothermal vent field, 26°N, Mid-Atlantic ridge. Mar. Chem. 1994, 46, 237–254. [Google Scholar] [CrossRef]
  61. Godfrey, L.V.; White, W.M.; Salters, V.J.M. Dissolved zirconium and hafnium distributions across a shelf break in the northeastern Atlantic Ocean. Geochim. Cosmochim. Acta 1996, 60, 3995–4006. [Google Scholar] [CrossRef]
  62. Reynard, B.; Lécuyer, C.; Grandjean, P. Crystal-chemical controls on rare earth element concentrations in fossil biogenic apatite and implications for paleoenvironmental reconstructions. Chem. Geol. 1999, 155, 233–241. [Google Scholar] [CrossRef]
  63. Shuaib, M.; Khan, K.F.; Hussain, A.; Jawad, M.; Dar, S.A. Rare-earth element (REE) geochemistry of late Paleocene–middle Eocene phosphate nodules in the Subathu Basin of Solan District, Himachal Pradesh, India. J. Sedimen Environ. 2021, 6, 1–12. [Google Scholar] [CrossRef]
  64. Hein, J.R.; Koschinsky, A.; Mikesell, M.; Mizell, K.; Glenn, C.R.; Wood, R. Marine Phosphorites as Potential Resources for Heavy Rare Earth Elements and Yttrium. Minerals 2016, 6, 88. [Google Scholar] [CrossRef] [Green Version]
  65. Chen, J.; Yang, R.; Wei, H.; Gao, J. Rare earth element geochemistry of Cambrian phosphorites from Yangtze Region (China). J. Rare Earth Elem. 2013, 31, 101–112. [Google Scholar] [CrossRef]
  66. Hamelin, B.; Dupre, B.; Allegre, C.J. Lead-strontium isotopic variations along the East Pacific Rise and Mid-Atlantic Ridge: A comparative study. Earth Planet. Sci. Lett. 1984, 67, 340–350. [Google Scholar] [CrossRef]
  67. Bau, M.; Möller, P. Rare-earth element fractionation in metamorphogenic hydrothermal calcite, magnesite and siderite. Mineral. Petrol. 1992, 45, 231–246. [Google Scholar] [CrossRef]
  68. Bau, M. Controls on the fractionation of isovalent trace elements in magmatic and aqueous systems: Evidence from Y/Ho, Zr/Hf, and lanthanide tetrad effect. Contrib. Mineral. Petrol. 1996, 123, 323–333. [Google Scholar] [CrossRef]
  69. Takahashi, Y.; Hirata, T.; Shimizu, H.; Ozaki, T.; Fortin, D. A rare earth element signature of bacteria in natural waters? Chem. Geol. 2007, 244, 569–583. [Google Scholar] [CrossRef]
  70. Takahashi, Y.; Hayasaka, Y.; Morita, K.; Kashiwabara, T.; Nakada, R.; Marcus, M.A.; Kato, K.; Tanaka, K.; Shimizu, H. Transfer of rare earth elements (REE) from manganese oxides to phosphates during early diagenesis in pelagic sediments inferred from REE patterns, X-ray absorption spectroscopy, and chemical leaching method. Geochem. J. 2015, 49, 653–674. [Google Scholar] [CrossRef]
  71. Tostevin, R.; Shields, G.A.; Tarbuck, G.M.; He, T.; Clarkson, M.O. Effective use of the cerium anomalies as a redox proxy in carbonate-dominated marine settings. Chem. Geol. 2016, 438, 146–162. [Google Scholar] [CrossRef] [Green Version]
  72. German, C.R.; Holliday, B.P.; Elderfield, H. Redox cycling of rare earth elements in the suboxic zone of the Black Sea. Geochim. Cosmochim. Acta 1991, 55, 3553–3558. [Google Scholar] [CrossRef]
  73. Al-Hobaib, A.S.; Baioumy, H.M.; Al-Ateeq, M.A. Geochemistry and origin of the Paleocene phosphorites from the Hazm Al-Jalamid area, northern Saudi Arabia. J. Geochem. Expl. 2013, 132, 15–25. [Google Scholar] [CrossRef]
  74. Shields, G.A.; Stille, P. Diagenetic constraints on the use of cerium anomalies as palaeoseawater redox proxies: An isotopic and REE study of Cambrian phosphorites. Chem. Geol. 2001, 175, 29–48. [Google Scholar] [CrossRef]
  75. Bau, M.; Balan, S.; Schmidt, K.; Koschinsky, A. Rare earth elements in mussel shells of the Mytilidae family as tracers for hidden and fossil high–temperature hydrothermal systems. Earth Planet. Sci. Lett. 2010, 299, 310–316. [Google Scholar] [CrossRef]
  76. Deng, X.H.; Chen, Y.J.; Yao, J.M.; Bagas, L.; Tang, H.S. Fluorite REE-Y (REY) geochemistry of the ca. 850 Ma Tumen molybdenite-fluorite deposit, eastern Qinling, China: Constraints on ore genesis. Ore Geol. Rev. 2014, 63, 532–543. [Google Scholar] [CrossRef]
  77. Ehya, F. Variation of mineralizing fluids and fractionation of REE during the emplacement of the vein-type fluorite deposit at Bozijan, Markazi Province, Iran. J. Geochem. Explor. 2012, 112, 93–106. [Google Scholar] [CrossRef]
  78. Roy, A.; Chakrabarti, G.; Shome, D. Geochemistry of the Neoproterozoic Narji limestone, Cuddapah Basin, Andhra Pradesh, India: Implication on palaeoenvironment. Arab. J. Geosci. 2018, 11, 784–796. [Google Scholar] [CrossRef]
  79. Tobia, F.H. Stable isotope and rare earth element geochemistry of the Baluti carbonates (Upper Triassic), Northern Iraq. Geosci. J. 2018, 22, 975–987. [Google Scholar] [CrossRef]
  80. Bau, M.; Dulski, P. Distribution of yttrium and rare-earth elements in the Penge and Kuruman iron-formations, Transvaal Supergroup, South Africa. Precambrian Res. 1996, 79, 37–55. [Google Scholar] [CrossRef]
  81. Fazio, A.M.; Scasso, R.A.; Castro, L.N.; Carey, S. Geochemistry of rare earth elements in early-diagenetic miocenephosphatic concretions of Patagonia, Argentina:Phosphogenetic implications. Deep-Sea Res. II 2007, 54, 1414–1432. [Google Scholar] [CrossRef]
  82. Zhang, J.; Amakawa, H.; Nosaky, Y. The comparative behavior of Y and lanthanides in seawater on the North Pacific. Geophys. Res. Lett. 1994, 21, 2677–2680. [Google Scholar] [CrossRef]
  83. Bau, M.; Möller, P.; Dulski, P. Yttrium and lanthanides in eastern Mediterranean seawater and their fractionation during redox-cycling. Mar. Chem. 1997, 56, 123–131. [Google Scholar] [CrossRef]
  84. Elderfield, H.; Pagett, R. Rare earth elements in ichthyoliths: Variations with redox conditions and depositional environment. Sci. Total Environ. 1986, 49, 175–197. [Google Scholar] [CrossRef]
  85. Buccione, R.; Kechiched, R.; Mongelli, G.; Sinisi, R. REEs in the North Africa P-Bearing deposits, paleoenvironments, and economic perspectives: A review. Minerals 2021, 11, 214. [Google Scholar] [CrossRef]
  86. Martin, E.E.; Scher, H.D. Preservation of seawater Sr and Nd isotopes in fossil fish teeth: Bad news and good news. Earth Planet. Sci. Lett. 2004, 220, 25–39. [Google Scholar] [CrossRef]
  87. Shields, G.A.; Webb, G.E. Has the REE composition of seawater changed over geological time? Chem. Geol. 2004, 204, 103–107. [Google Scholar] [CrossRef]
  88. Lécuyer, C.; Reynard, B.; Grandjean, P. Rare earth element evolution of Phanerozoic seawater recorded in biogenic apatites. Chem. Geol. 2004, 204, 63–102. [Google Scholar] [CrossRef]
  89. Bright, C.A.; Cruse, A.M.; Lyons, T.W.; MacLeod, K.G.; Glascock, M.D. Seawater rare-earth element patterns preserved in apatite of Pennsylvanian conodonts? Geochim. Cosmochim. Act 2009, 73, 1609–1624. [Google Scholar] [CrossRef]
  90. Wang, C.M.; Deng, J.; Carranza, E.J.M.; Lai, X.R. Nature, diversity and temporal–spatial distributions of sediment-hosted Pb–Zn deposits in China. Ore Geol. Rev. 2014, 56, 327–351. [Google Scholar] [CrossRef]
  91. Zartman, R.E.; Doe, B.R. Plumbotectonics—The model. Tectonophysics 1981, 75, 135–162. [Google Scholar] [CrossRef]
  92. von Blanckenburg, F.; O’Nions, R.K.; Hein, J.R. Distribution and sources of pre-anthropogenic lead isotopes in deep ocean water from Fe-Mn crusts, Geochim. Cosmochim. Acta 1996, 60, 4936–4957. [Google Scholar] [CrossRef]
  93. Muiños, S.B.; Frank, M.; Made, C.; Hein, J.R.; van de Flierdt, T. New constraints on the Pb and Nd isotopic evolution of NE Atlantic water masses. Geochem. Geophys. Geosystems 2008, 9, 18. [Google Scholar] [CrossRef] [Green Version]
  94. Andrieu, A.S.; Honnorez, J.; Lancelot, J. Lead isotope compositions of the TAG mineralization, Mid-Atlantic Ridge, 26°08′N. In Proceedings of the Ocean Drilling Program, Scientific Results, Charleston, SC, USA, 8 January–14 February 1997; Herzig, P.M., Humphris, S.E., Miller, D.J., Zierenberg, R.A., Eds.; Texas: Charleston, SC, USA, 1998. [Google Scholar]
  95. Bolton, B.R.; Frakes, L.A. Geology and genesis of manganese oolite, Chiatura, Georgia, U.S.S.R. Geol. Soc. Am. Bull. 1985, 96, 1398–1406. [Google Scholar] [CrossRef]
  96. Sasmaz, A.; Zagnitko, V.M.; Sasmaz, B. Major, trace and rare earth element (REE) geochemistry of the Oligocene stratiform manganese oxide-hydroxide deposits in the Nikopol, Ukraine. Ore Geol. Rev. 2020, 126, 103772. [Google Scholar] [CrossRef]
  97. Sasmaz, A.; Sasmaz, B.; Hein, J.R. Geochemical approach to the genesis of the Oligocene-stratiform manganese-oxide deposit, Chiatura (Georgia). Ore Geol. Rev. 2021, 128, 103910. [Google Scholar] [CrossRef]
  98. Burnett, W.C. Geochemistry and origin of phosphorite deposits from off Peru and Chile. Geol. Soc. Am. Bull. 1977, 88, 813–823. [Google Scholar] [CrossRef]
  99. Elderfield, H.; Greaves, M.J. The rare earth elements in seawater. Nature 1982, 296, 214–219. [Google Scholar] [CrossRef]
  100. Felitsyn, S.; Morad, S. REE patterns in latest Neoproterozoic early Cambrian phosphate concretions and associated organic matter. Chem. Geol. 2002, 187, 257–265. [Google Scholar] [CrossRef]
  101. Stalder, M.; Rozendaal, A. Apatite nodules as an indicator of depositional environment and ore genesis for the Mesoproterozoic Broken Hill-type Gamsberg Zn–Pb deposit, Namaqua Province, South Africa. Miner. Dep. 2004, 39, 189–203. [Google Scholar] [CrossRef]
  102. Tyson, R.V.; Pearson, T.H. Modern and Ancient Continental Shelf Anoxia: An Overview; Geological Society Publications: London, UK, 1991; Volume 58, pp. 1–24. [Google Scholar]
  103. Trappe, J. Phanerozoic phosphorite depositional systems. In Lecture Notes in Earth Sciences; Springer: Berlin, Germany, 1998; Volume 76. [Google Scholar]
  104. Slomp, C.P.; Van Capellen, P. The global marine phosphorus cycle: Sensitivity to oceanic circulation. Biogeosciences 2007, 4, 155–171. [Google Scholar] [CrossRef] [Green Version]
  105. Sasmaz, A.; Onal, A.; Sagiroglu, A.; Onal, M. Origin and nature of the mineralizing fluids of thrust zone fluorites in Çelikhan (Adıyaman, Eastern Turkey): A geochemical approach. Geochem. J. 2005, 39, 131–139. [Google Scholar] [CrossRef]
  106. Baioumy, H.; Farouk, S. The geochemical and economic significance of REE in the Upper Cretaceous-Eocene Tethyan phosphorites. J. Afr. Earth Sci. 2022, 194, 104635. [Google Scholar] [CrossRef]
  107. Almogi-Labin, A.; Bein, A.; Sass, E. Late Cretaceous upwelling system along the southern Tethys margin (Israel): Interrelationship between productivity, bottom water environments, and organic matter preservation. Paleoceanography 1993, 8, 671–690. [Google Scholar] [CrossRef]
  108. Abed, A.M. The eastern Mediterranean phosphorite giants: An interplay between tectonics and upwelling. GeoArabia 2013, 18, 67–94. [Google Scholar] [CrossRef]
  109. Abed, A.M. Review of Uranium in the Jordanian Phosphorites: Distribution, Genesis and Industry. Jordan J. Earth Environ. Sci. 2011, 4, 35–45. [Google Scholar]
  110. El Bamiki, R.; Raji, O.; Ouabid, M.; Elghali, A.; Yazami, O.K. Phosphate rocks: A review of sedimentary and igneous occurrences in Morocco. Minerals 2021, 11, 1137. [Google Scholar] [CrossRef]
  111. Shlewit, H. Treatment of phosphate rocks with hydrochloric acid. J. Radioanal. Nucl. Chem. 2011, 287, 49–54. [Google Scholar] [CrossRef]
  112. Tlili, A.; Felhi, M.; Montacer, M. Origin and Depositional Environment of Palygorskite and Sepiolite from the Ypresian Phosphatic Series, Southwestern Tunisia. Clays Clay Miner. 2010, 58, 573–584. [Google Scholar] [CrossRef]
  113. Powell, J.H.; Moh’d, B.K. Early diagenesis of Late Cretaceous chalk-chert-phosphorite hardgrounds in Jordan; Implications for sedimentation on a Coniacian–Campanian pelagic ramp. GeoArabia 2012, 17, 17–38. [Google Scholar] [CrossRef]
  114. Sasmaz, A.; Kilic, A.D.; Akgul, B.; Sasmaz, B. A spectral approach on mineralogy and geochemistry of garnet skarns in Arc-Type granitoids. Spectrochim. Acta Part A Mol. Biomol. Spectr. 2023, 286, 122037. [Google Scholar] [CrossRef]
Figure 1. Paleogeographic map of the European, SE Turkey, and Middle Eastern regions in the Upper Cretaceous at 75 MA (modified from Ron Blakey, NAU Geolgy-http://www2.nau.edu/rcb7. Accessed on 1 January 2011). Data from [37].
Figure 1. Paleogeographic map of the European, SE Turkey, and Middle Eastern regions in the Upper Cretaceous at 75 MA (modified from Ron Blakey, NAU Geolgy-http://www2.nau.edu/rcb7. Accessed on 1 January 2011). Data from [37].
Minerals 12 01544 g001
Figure 2. Generalized geological and location map of Mazıdağı phosphate deposit and its surrounding area. Data from [37].
Figure 2. Generalized geological and location map of Mazıdağı phosphate deposit and its surrounding area. Data from [37].
Minerals 12 01544 g002
Figure 3. Stratigraphic columnar section of Derik–Mazıdağı region (taken from [6]).
Figure 3. Stratigraphic columnar section of Derik–Mazıdağı region (taken from [6]).
Minerals 12 01544 g003
Figure 4. Typical phosphate and chert outcrops of the Mazıdağı phosphates.
Figure 4. Typical phosphate and chert outcrops of the Mazıdağı phosphates.
Minerals 12 01544 g004
Figure 5. Binary major oxide plots; (a) P2O5% vs. SiO2; (b) P2O5 vs. MgO; (c) P2O5 vs. Fe2O3; (d) P2O5 vs. Na2O; (e) P2O5 vs. CaO; (f) P2O5 vs. ∑REE.
Figure 5. Binary major oxide plots; (a) P2O5% vs. SiO2; (b) P2O5 vs. MgO; (c) P2O5 vs. Fe2O3; (d) P2O5 vs. Na2O; (e) P2O5 vs. CaO; (f) P2O5 vs. ∑REE.
Minerals 12 01544 g005aMinerals 12 01544 g005b
Figure 6. PAAS (Post Archean Australian Shale) normalized according to Taylor and McLennan [59] trace element pattern of the phosphate samples.
Figure 6. PAAS (Post Archean Australian Shale) normalized according to Taylor and McLennan [59] trace element pattern of the phosphate samples.
Minerals 12 01544 g006
Figure 7. Nb/Ta and Zr/Hf ratios in Mazıdağı phosphates compared to Pacific and Arctic deep water and North Atlantic deep water. Data from [50,54,55,60,61].
Figure 7. Nb/Ta and Zr/Hf ratios in Mazıdağı phosphates compared to Pacific and Arctic deep water and North Atlantic deep water. Data from [50,54,55,60,61].
Minerals 12 01544 g007
Figure 8. PAAS-normalized [59] REE pattern of the phosphate samples.
Figure 8. PAAS-normalized [59] REE pattern of the phosphate samples.
Minerals 12 01544 g008
Figure 9. (a) The (La/Sm)n vs. (La/Yb)n binary plot showing the influence of diagenesis to phosphate samples [22] and (b) the (Sm/Pr)n vs. (Sm/Yb)n diagram displaying the HREE enrichments recorded in the phosphate deposits. Data from [20,22,25,27,64,65,73].
Figure 9. (a) The (La/Sm)n vs. (La/Yb)n binary plot showing the influence of diagenesis to phosphate samples [22] and (b) the (Sm/Pr)n vs. (Sm/Yb)n diagram displaying the HREE enrichments recorded in the phosphate deposits. Data from [20,22,25,27,64,65,73].
Minerals 12 01544 g009
Figure 10. Comparison of the REE, (La/Yb)n, (Ce/Ce*)n, (Gd/Yb)n, and (Eu/Eu)n parameters in the Tunisian, Algerian, Moroccan, and Mazıdağı phosphates. Data from [85]. The horizontal bar in the box refers to the median value; the ends of the whiskers are the maximum and minimum values of variables; the top and bottom of the boxes are the values of the first and third quartiles; circles represent the outlier values of the data set.
Figure 10. Comparison of the REE, (La/Yb)n, (Ce/Ce*)n, (Gd/Yb)n, and (Eu/Eu)n parameters in the Tunisian, Algerian, Moroccan, and Mazıdağı phosphates. Data from [85]. The horizontal bar in the box refers to the median value; the ends of the whiskers are the maximum and minimum values of variables; the top and bottom of the boxes are the values of the first and third quartiles; circles represent the outlier values of the data set.
Minerals 12 01544 g010
Figure 11. (a,b) Comparison of Pb isotopic compositions of the Mazıdağı phosphates with selected Chinese sediment-hosted Pb-Zn deposits, plotted in the plubotectonic model diagram of Zartman and Doe [90], and (c,d) comparison of Pb isotope compositions of the Mazıdağı phosphates with MORB, ferromanganese crusts, and pelagic sediments, and Pacific, Atlantic, and Indian oceans pelagic sediment. Data taken from [91,92].
Figure 11. (a,b) Comparison of Pb isotopic compositions of the Mazıdağı phosphates with selected Chinese sediment-hosted Pb-Zn deposits, plotted in the plubotectonic model diagram of Zartman and Doe [90], and (c,d) comparison of Pb isotope compositions of the Mazıdağı phosphates with MORB, ferromanganese crusts, and pelagic sediments, and Pacific, Atlantic, and Indian oceans pelagic sediment. Data taken from [91,92].
Minerals 12 01544 g011
Figure 12. Simplified model of the depositional environment of Mazıdağı phosphates.
Figure 12. Simplified model of the depositional environment of Mazıdağı phosphates.
Minerals 12 01544 g012
Table 1. Major oxide contents (%) of the phosphate samples.
Table 1. Major oxide contents (%) of the phosphate samples.
SiO2Al2O3Fe2O3MgOCaONa2OK2OTiO2P2O5MnOCr2O3CO2SumCaO/P2O5
DL0.010.010.040.010.010.010.010.010.010.010.002−5.1
DY-012.050.040.110.2254.90.710.01<0.0127.50.020.0214.299.92.00
DY-020.910.01<0.040.2155.30.110.01<0.014.29<0.010.0139.110012.88
DY-031.450.010,070.2155.40.430.01<0.0116.5<0.010.0225.799.93.35
DY-114.630,110.250.1753.70.710.020.0131.3<0.010.038.899.91.72
DY-121.050.050.100.2055.50.740.01<0.0130.2<0.010.0311.899.81.84
DY-1312.800.020.100.1748.80.260.01<0.0110.1<0.010.0127.599.94.81
DY-1433.500.140.220.1337.00.460.02<0.0122.3<0.010.015.999.91.66
DY-200.530.010.010.1756.30.740.01<0.0131.7<0.010.0210.299.91.77
DY-2136.10.030.160.1635.60.060.01<0.012.22<0.010.0225.610016.05
DY-221.250.040.060.2055.50.730.01<0.0129.5<0.010.0312.399.91.88
DY-236.920.030.090.1952.00.330.01<0.0113.5<0.010.0226.799.93.86
DY-240.730.020.010.1756.10.690.01<0.0130.6<0.010.0211.299.91.83
DY-252.410.010.080.2055.00.600.01<0.0125.6<0.010.0215.899.92.15
Avrg8.030.040.110.1851.60.510.010.0121.20.020.0218.199.94.29
Table 2. Correlation relationships with major oxides and trace elements of P2O5 in phosphate samples (Negative correlations are green color, positive correlations are yellow color).
Table 2. Correlation relationships with major oxides and trace elements of P2O5 in phosphate samples (Negative correlations are green color, positive correlations are yellow color).
SiO2Al2O3Fe2O3MgOCaONa2OK2OTiO2MnOCO2VCrCoNiCu
P2O5−0.480.27−0.040.030.500.990.230.06−0.05−0.880.750.790.22−0.490.50
RbSrYZrNbBaHfAsCdUThPbZnAu∑REE
P2O50.520.990.910.790.690.930.690.890.370.950.400.870.810.490.92
Table 3. Trace element concentrations (ppm) of the phosphate samples.
Table 3. Trace element concentrations (ppm) of the phosphate samples.
VCrCoNiCuRbSrZrNbBaHfAsCdUThPbZnAuSc∑TETh/UNi/CoV/ScV/Cr
DY-01631231.87.228.00.3014186.40.111120.115.813.648.60.71.912370.0891.720620.0144.00370.51
DY-0212300.511.14.60.104181.40.02200.021.93.28.20.10.39620.0850.25740.01222.2600.40
DY-0346860.610.016.60.209516.50.05770.055.532.931.80.21.162090.0651.114750.00616.7420.53
DY-11761890.223.643.90.50144315.30.14950.1413.823.451.60.42.584070.0563.523960.008118220.40
DY-12921550.615.394.40.3015319.80.081120.087.546.660.20.21.785570.0981.726840.00325.5540.59
DY-1324771.218.817.90.105673.60.02380.022.633.917.40.10.651960.0470.79980.00615.7340.31
DY-14121720.710.529.20.5010647.00.061130.066.757.357.30.22.202060.0761.917480.00315.0641.68
DY-20721230.67.819.10.2015357.40.06980.069.06.351.70.11.562960.0661.222280.00213.0600.58
DY-2111760.632.32.70.101980.80.02100.021.35.414.00.10.21490.0230.23920.02553.8550.14
DY-22911588.810.659.70.2015177.70.08960.089.668.351.00.31.562490.0611.623270.0061.20570.58
DY-23441010.820.448.10.207784.60.04560.044.019.225.70.10.772430.0371.013460.00425.5440.44
DY-24641140.58.115.40.2014147.60.051030.057.925.947.50.11.693790.1441.521900.00216.2430.56
DY-2533960.415.913.90.1012944.10.021010.028.019.646.60.11.273130.1481.619460.00239.8210.35
Avrg581081.3314.730.30.2310876.320.0679.20.066.4327.4390.211.362620.081.3817200.00728.245.50.54
Table 4. Redox classification of the depositional environment. Data from [48,49].
Table 4. Redox classification of the depositional environment. Data from [48,49].
IndicatorOxicSuboxicAnoxicEuxinicThe Studied Phoshates
H2S No free H2S in the water column Free H2S present in the water column
O2 concentration in bottom waters O2 > 20.2 < O2 < 2O2 < 0.2O2 = 0
Th/U>72–70–2-0.88
V/Cr<22–4.25>4.25>4.250.49
Ni/Co<55–7>7-3.69
V/Sc--->2445.5
Table 5. Rare earth element contents of the phosphate samples (ppm).
Table 5. Rare earth element contents of the phosphate samples (ppm).
Sample NoDedection LimitDY-01DY-02DY-03DY-11DY-12DY-13DY-14DY-20DY-21DY-22DY-23DY-24DY-25
La0.510.11.306.6015.49.104.105.9011.71.0010.54.4012.87.60
Ce0.13.100.401.905.102.801.302.203.600.303.301.403.902.40
Pr0.021.170.130.671.740.980.460.611.290.111.090.491.390.78
Nd0.024.160.512.577.634.011.852.375.310.454.811.925.673.02
Sm0.020.800.070.591.390.740.340.480.840.090.860.421.090.62
Eu0.020.270.030.130.440.220.110.150.300.020.240.100.340.21
Gd0.021.500.220.822.071.370.490.751.600.131.400.581.990.94
Tb0.020.250.030.130.380.210.090.120.250.030.200.090.270.16
Dy0.021.580.170.972.641.470.680.871.750.141.560.691.781.30
Y0.0227.44.1218.746.827.912.016.529.93.0027.813.131.622.1
Ho0.020.420.050.290.770.400.170.240.500.040.440.170.530.33
Er0.021.330.180.902.451.390.600.881.600.141.330.651.771.11
Tm0.020.200.020.140.370.230.090.120.240.020.250.080.250.16
Lu0.020.240.030.140.420.220.100.150.250.020.240.100.260.19
Yb0.021.240.140.852.201.230.550.741.320.131.190.531.511.02
∑REE-53.87.4235.489.957.322.932.160.55.6255.224.735.141.9
Y/Ho-65826561707169607463776067
(Ce/Ce*)n-0.210.160.180.200.200.180.210.220.390.190.170.220.19
(Eu/Eu*)n-1.902.231.221.851.671.721.692.011.251.581.291.811.84
(Y/Y*)n-3.224.203.493.463.493.143.413.223.703.313.373.343.11
(Pr/Pr)n-3.462.943.223.193.213.222.533.311.253.053.183.302.97
(La/Yb)n-0.600.690.570.510.550.550.590.650.570.760.640.630.55
((Gd/Yb)n-0.730.950.580.570.670.540.610.730.610.710.660.770.68
(Sm/Pr)n-1.000.060.060.050.060.050.060.060.060.040.050.050.06
(Sm/Yb)n-0.330.250.350.320.310.310.330.320.350.370.400.370.31
Table 6. Pb isotope ratios of the Mazıdağı phosphates.
Table 6. Pb isotope ratios of the Mazıdağı phosphates.
204Pb206Pb207Pb208Pb206Pb/204Pb207Pb/204Pb208Pb/204Pb
DL0.010.010.010.01
DY-010.0350.680.5501.3619.4015.7138.86
DY-020.0200.390.3150.7819.5315.7539.00
DY-030.0250.480.3930.9819.2015.7239.20
DY-110.0330.620.5161.2818.8015.6438.80
DY-120.0280.540.4681.1019.3015.7039.30
DY-130.0160.310.2660.6219.1015.6338.75
DY-140.0380.730.6021.4819.2015.8438.90
DY-200.0360.680.5641.4118.9015.6739.17
DY-210.0280.540.4401.1019.3015.7139.30
DY-220.0200.390.3150.7919.2515.7539.35
DY-230.0180.350.2820.7219.4415.6738.80
DY-240.0400.760.6281.5518.9015.7038.75
DY-250.0300.570.4701.1819.0015.6739.33
Avrg0.0280.540.4471.10219.1815.7039.04
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gundogar, D.Y.; Sasmaz, A. Geochemical Approach to Determine the Possible Precipitation Parameters of the Coniacian–Santonian Mazıdağı Phosphates, Mardin, Turkey. Minerals 2022, 12, 1544. https://doi.org/10.3390/min12121544

AMA Style

Gundogar DY, Sasmaz A. Geochemical Approach to Determine the Possible Precipitation Parameters of the Coniacian–Santonian Mazıdağı Phosphates, Mardin, Turkey. Minerals. 2022; 12(12):1544. https://doi.org/10.3390/min12121544

Chicago/Turabian Style

Gundogar, Derya Yildirim, and Ahmet Sasmaz. 2022. "Geochemical Approach to Determine the Possible Precipitation Parameters of the Coniacian–Santonian Mazıdağı Phosphates, Mardin, Turkey" Minerals 12, no. 12: 1544. https://doi.org/10.3390/min12121544

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