Next Article in Journal
Petrogenesis and Tectonic Implications of Late Carboniferous Intrusions in the Tuwu-Yandong Porphyry Cu Belt (NW China): Constraints from Geochronology, Geochemistry and Sr–Nd–Hf Isotopes
Previous Article in Journal
Selective Recovery of Copper from Acid Leaching Solution through Slow Release Sulfide Precipitant
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Petrogenesis of Early Cenozoic Sarıcakaya–Nallıhan Volcanism in NW Turkey: Implications for the Geodynamic Setting and Source Characterization of the Balkanatolia Magmatic Realm

1
Department of History of Science, Istanbul University, Istanbul 34452, Turkey
2
Geological Engineering Department, Faculty of Mines, Istanbul Technical University, Istanbul 34467, Turkey
3
Geological Engineering Department, Istanbul University-Cerrahpaşa, Istanbul 34500, Turkey
*
Author to whom correspondence should be addressed.
Minerals 2022, 12(12), 1572; https://doi.org/10.3390/min12121572
Submission received: 10 October 2022 / Revised: 4 December 2022 / Accepted: 5 December 2022 / Published: 7 December 2022
(This article belongs to the Section Mineral Geochemistry and Geochronology)

Abstract

:
Sarıcakaya–Nallıhan Volcanism was generated within the Balkanatolia Magmatic Realm between 48 and 44 Ma (by 40Ar–39Ar age determination) and is represented by three different volcanic units all displaying subduction-related geochemical signatures, such as depletion in HFSE and enrichment in LREE and LILE. The first unit (V1) consists of nepheline-normative, olivine basalts with OIB-like affinity. The second (V2) and third (V3) units are represented by more evolved compositions such as basaltic-andesitic, andesitic, and dacitic-rhyolitic lavas. Even the most basic lavas have elevated Mg# values (62–69), and they are far from representing the true mantle melts. Source characterization of Sarıcakaya–Nallıhan Volcanism reveals that there might be two possible mantle sources for the primary melts of the lavas: (i) metasomatized peridotitic mantle fluxed by sedimentary melts, or (ii) accreted mélange. The direct melting of the mélange-like lithologies is a more favorable mechanism for the Middle Eocene (44–40 Ma) magmatism in Balkanatolia since the Hf–Nd trace element, Nd isotopic systematics and petrological modelling efforts supported the latter. Overall, Early Cenozoic magmatism within this realm was characterized, first (58–44 Ma) by contractional and later (44–40 Ma) by extensional tectonics and the late-stage magmatic phase in the area was possibly controlled by melting of accreted mélange-like lithologies. The presented data indicate that mélange melting might be much more common than envisaged for the magmatism in the Alpine–Himalayan orogenic belt.

1. Introduction

The Late Mesozoic–Early Cenozoic geodynamic evolution of northern Anatolia shares similarities with that of SE Europe, particularly with Balkan regions. This period is characterized by severe magmatism and coeval sedimentation events, except for the long magmatic cessation during the Paleocene in the northern part of Anatolia and present-day northern Greece–Bulgaria regions [1,2,3,4,5,6,7,8,9,10,11] (Figure 1). During the Early Cenozoic, northwestern Turkey constituted an important land bridge between Europe and Middle East/Asia and recently, together with the Balkans and Central Anatolia, it has been postulated as a paleo-biogeographic region called Balkanatolia [12,13]. This area contains widespread and continuous magmatism dispersed along the different parts of NW Turkey, Greece, Rhodope, and Thrace [7]. Hereafter, we term the region extending from the Balkans to NW Anatolia as Balkanatolia (Rhodope, Thrace, northern Greece, NW Biga Peninsula; Almacık–Armutlu, Sarıcakaya–Nallıhan, Sivrihisar, and the Tavşanlı Zone), considering the terminology outlined in [12].
Even though the Early Cenozoic magmatic activities can be traced to long distances from the Balkans, Anatolia, and even the Caucasus–Iran regions, the geodynamic evolution of the region is still under debate [3,11,14,15,16,17,18,19,20]. Along this long track of magmatic units, NW Anatolia marks an important nexus that records the changing conditions of the geodynamic setting during the orogenic development of the Balkanatolia, northern Turkey–Caucasus and Iran. There is a general agreement on the evolution of the Early Cenozoic magmatism in central and eastern Pontides (north of Anatolia): (i) a magmatic lull during the Early-Mid Paleocene collision of the Pontides vs. the Kırşehir Block and Anatolide–Tauride Block following Upper Cretaceous arc magmatism; (ii) emplacement of scarce adakite-like plutons of Late Paleocene–Early Eocene age due to crustal thickening, and (iii) severe post-collisional volcanism and plutonism during the Middle Eocene over a broad area along northern Turkey [3,5,14,15,16,18,19,21].
The proposed geodynamic models for the Early Cenozoic magmatism in the NW Turkey and Balkan regions in the literature share similarities with the models outlined above. The first model mainly implies that the Early Cenozoic magmatism was mostly governed by the subduction of the Pindos/Neotethyan Ocean beneath the Balkan peninsula and NW Anatolia, and many of the magmatic units display subduction-related geochemical features related to that event [2,11,22,23]. The second model suggests that this period of magmatism was generated long after the subduction of the Tethys slab, and mostly generated via late-stage slab break-off and/or delamination following final Anatolide–Tauride and Pontides amalgamation (particularly for Almacık–Armutlu–Nallıhan areas) [4,5,21,24,25]. In addition, for the Balkan Regions and the Biga Peninsula, different studies imply that the magmatism is generated via a back-arc extensional geodynamic setting due to the slab roll-back of the single subducted lithosphere of Pindos / Neotethyan Ocean [26,27,28].
The main difference between the geodynamic models mentioned above is whether the Early Cenozoic magmatism along Balkanatolia is related to the subduction of a single slab [23] or is related to the post-collisional re-organization of the plates after the detachment of the subduction slab [21]. Indeed, a more favorable interpretation of the magma genesis, which is called “the slab-break off”, does not contain any true sub-lithospheric or asthenospheric character at all [3].
Many of the interpretations mentioned above depend mostly on (i) the linearity of the volcano-plutonic units of Balkanatolia and the late Paleocene to Eocene marine to sub-aerial transition of the volcano-sedimentary basins [7] and (ii) possible relict slab images from seismic tomography [29,30]. However, as pointed out by [10] and [23], the proposed magma generation mechanisms did not support the generation from torn or detached slabs. Additionally, new studies even put into question the linkage of slab break-off and voluminous magmatism in general worldwide [31]. In addition, a recent review by [11] suggests generation above a subducted Neotethyan slab for the Late Paleocene–Middle Eocene magmatism from NW Anatolia to Iran.
To better understand the genesis and extent of the Early Cenozoic magmatism covering Balkanatolia and the Early Cenozoic geological evolution of the region (Figure 1), we focused on the volcanic units of Sarıcakaya–Nallıhan (NW Turkey) Basin in terms of whole-rock geochemical and isotopic systematics by in situ geochronology. In addition, we have also re-evaluated the previous isotopic age findings and suggest a different stratigraphic log for the volcano-sedimentary succession in the study area, which is supporting evidence for a different geodynamic evolution scenario [25].
Figure 1. Geological sketch map showing the distribution of Late Mesozoic to Early Cenozoic units around Greece and Western Anatolia. The data to generate the map were compiled from [2,4,5,22,23,32,33,34]. IPS: Intra-Pontides Suture, IAESZ: İzmir–Ankara–Erzincan Suture Zone, SMG: South Marmara Granitoids; AA: Almacık–Armutlu Volcanics; TZG: Tavşanlı Zone Granitoids; SG: Sivrihisar Granitoids; SN: Sarıcakaya–Nallıhan Volcanism.
Figure 1. Geological sketch map showing the distribution of Late Mesozoic to Early Cenozoic units around Greece and Western Anatolia. The data to generate the map were compiled from [2,4,5,22,23,32,33,34]. IPS: Intra-Pontides Suture, IAESZ: İzmir–Ankara–Erzincan Suture Zone, SMG: South Marmara Granitoids; AA: Almacık–Armutlu Volcanics; TZG: Tavşanlı Zone Granitoids; SG: Sivrihisar Granitoids; SN: Sarıcakaya–Nallıhan Volcanism.
Minerals 12 01572 g001

2. Geological Framework

The Sarıcakaya–Nallıhan Basin is situated within the NW portion of the Anatolia peninsula and in an imbricated and highly deformed accretionary wedge which comprises various tectonic units belonging to the İzmir–Ankara–Erzincan suture belt (Cretaceous ophiolitic mélange with various blocks), Karakaya Complex (Permo-Triassic accretionary complex with blocks) Mudurnu Formation (Jurassic age limestone platform carbonates) and Tavşanlı Zone (Late Cretaceous HP blueschist belt) (Figure 2. [10,25,35]). Stratigraphically, the pre-Cenozoic units serve as the basement rocks in the area which are unconformably overlain by the Eocene or much younger Neogene volcano-sedimentary successions along vast areas in the Sarıcakaya–Nallıhan Basin. This relation between basement rocks and the Eocene units is common for all the locations where the Eocene volcano-sedimentary units crop out in Balkanatolia (e.g., Almacık, Armutlu, the Biga Peninsula, Central and Eastern Pontides and the Rhodopes/Balkan Peninsula).
Eocene volcano-sedimentary succession in the Sarıcakaya–Nallıhan area has been studied by various researchers and attributed different names, such as Meyildere Volcanics [35,36,37], Bozaniç Volcanics [38], and Nallıhan Volcanics [25]. To simplify, we named the succession Sarıcakaya–Nallıhan Volcanics (SNV).
SNV resting unconformably over the amalgamated basement units are represented at the base by the lower Eocene fluvial Kızılçay Formation. The Kızılçay Formation contains red sandstone, mudstone, and conglomerate and displays vertical and lateral transitions to the Güvenç Formation which contains magmatic and sedimentary rocks. The sedimentary units are of two types: (i) sandstone–siltstone–marl alternation characteristics with abundant nummulite sp. fossils and (ii) volcaniclastics consisting of flow breccias and debris flow units. The magmatic rocks of the Güvenç Formation, contain massive lavas, dikes, plugs, and dome-like features. All those units of the Güvenç Formation display cross-cutting and interfingering relations with each other.
The volcanic units of the SNV can be grouped as (i) basaltic lavas (V1 unit), (ii) andesitic lavas (andesite, basaltic andesite, trachy-andesite, V2 unit), and (iii) dacitic and rhyolithic lavas (V3 unit, Figure 2), based on their modal mineralogy. A major part of the basalt, andesite, basaltic andesite, and trachy-andesites are represented by lava flows. On the other hand, dacitic and rhyolitic compositions are found to have built-up plugs and dome-like features rather than lava flows (Figure 3a–f).
Basaltic, and andesitic lavas display black to reddish colors, columnar jointing and flow bending in their outcrops and contain plagioclase, clinopyroxene, olivine, and hornblende with volcanic glass that can be detected by the naked eye in hand specimens. Dacitic and rhyolitic units show greyish-white colors and flow banding-columnar features, together with the vitreous glass that contains biotite phenocrysts.
The study of [25] mainly postulated that the volcanic sequence starts with basaltic lavas at the bottom, involving andesitic lavas in the middle and overlying andesitic and rhyolitic lavas at the top. However, we did not detect direct cross-cutting relationships between the different units. Hence, the main age variance and tectonic scheme of the lower-middle Eocene volcano-sedimentary units are in a chaotic fashion due to the collisional effects of Anatolia–Tauride and Pontides (e.g., [10,35,39]) and age distribution should be interpreted with caution. Miocene and younger sedimentary units uncomfortably rest on the Eocene volcano-sedimentary sequence.

3. Materials and Methods

Petrographic studies were conducted on 65 representative samples of the different volcanic units. Thin sections were prepared using the facilities of the Istanbul Technical University, Geological Engineering Department.
Whole rock geochemical analyses of 22 representative samples of the lavas and one sample of cumulate were performed by inductively coupled plasma atomic emission spectrometry (ICP-AES) at the ACME Analytical Laboratories Ltd. (Bureau Veritas, Edmonton, AB, Canada). Powdered sample (0.2 g) was fused with 1.5 g LiBO2 and then dissolved via four acid digestion steps. Loss on ignition was determined by weight difference after ignition at 1000 °C. Total iron concentration was expressed as Fe2O3. Detection limits are in the range of 0.001 to 0.1 weight percent (wt. %) for major element oxides, 0.1 to 10 ppm for trace elements, and 0.01 to 0.5 ppm for the rare earth elements (REE). Calibration and verification standards, together with reagent blanks, were added to the sample sequence.
Sr–Nd isotopic studies were conducted at Middle Eastern Technical University in Ankara, Turkey. Sr and Nd isotopic data were detected by using a Thermo Finnigan Triton thermal ionization mass spectrometer (TIMS) in static multi-collection mode. 87Sr/86Sr and 143 Nd/144 Nd data were normalized to 86Sr/88 Sr =0.1194 and 146 Nd/144 Nd= 0.7219, respectively. Sr NIST SRM 987 and Nd La Jolla standards were deduced as 0.710242 ± 10 (n = 3) and 0.511849 ± 10 (n = 2), respectively. No corrections were applied to Nd and Sr isotopic compositions for instrumental bias. Analytical uncertainties are given at the 2σ m level. A detailed description of the technique is given in [40].
40Ar/39Ar age determinations were performed on biotite, amphibole, and whole-rock matrix from seven samples at the University of Michigan. The methods for age determination have been described in [41]. The raw Ar–Ar data were previously published by [35] but we re-calculated the 40Ar–39Ar ages by using the Excel spreadsheet algorithm of [42]. We discuss the already published data to reveal the general outline of the volcano-sedimentary succession of the region.

4. Results

4.1. Petrography

Three different volcanic units have been identified petrographically: (i) basalt (V1 unit), (ii) andesite (V2 unit), and (iii) dacite and rhyolite (V3 unit).
V1 unit basalts mainly have clinopyroxene (% 20–60) and olivine (10–25) with abundant plagioclase microlites within the matrix, which all constitute hemicrystalline, porphyritic and interstitial textures. The main phenocryst assemblages detected are the volcanic glass (% 40–70) in modal mineralogy. Opaque phases might reach up to 5 percent in basalts.
Olivine crystals are mainly subhedral and display variations in size from 0.1 to 3.5 mm. Olivine crystals are found to be fresh and are occasionally replaced by iddingsite and serpentine in a few samples (Figure 4a). Clinopyroxene phenocrysts that have a diopsitic or augitic composition, identified by the extinction angle (35–45°) (Figure 4b), frequently have inclusions of olivine, plagioclase, and opaque minerals. A major portion of the feldspars is represented by plagioclase microlites frequently embedded in the glassy matrix. Plagioclase phenocrysts are rare in basalts, but when found they can reach up to 2.7 mm. Opaque phases are generally represented by ilmenite and magnetite.
The V2 unit andesitic lavas mainly show porphyritic and microlithic textures. Modally, they mainly contain feldspar (both plagioclase and sanidine; wt. % 25–60), amphibole (wt. % 15–20), and opaque phases (wt. % 5–10), clinopyroxene (wt. %. ̴5), and glassy matrix. Sericitization and epidotization are common. Mineral cloths are usually detected in the trachy-andesitic lavas (Figure 4c). Plagioclase (0.5–5 mm) display common zoning features and usually contain opaque mineral inclusions. Sanidine crystals (1–3 mm) are found in trachy-andesite. Amphibole (1–6 mm) commonly display greenish to brown colors (Figure 4d).
The V3 unit dacitic and rhyolitic lavas mainly contained low modal content of plagioclase, sanidine, quartz, biotite, and opaque phases (Figure 4e); they are dominated by the presence of flow banding volcanic glass. Quartz displays resorbed and embedded features (Figure 4f).
Overall, even though the studied units are Early Cenozoic in age, they mainly preserve their primary mineralogic compositions.

4.2. Geochronology

We re-evaluated the age data presented by [35] based on the total gas and plateau ages that have not been presented before. Our different stratigraphical interpretation of the SNV is addressed in the discussion section.
According to our calculations, the V1 unit (basalts) displayed 46.92–44.72 Ma (based on whole-rock matrix data) ages (Figure 5a–d), the V2 unit (andesitic lavas) gave 48.17–45.44 Ma (Figure 6a–d) and dacitic lavas (V3 unit) displayed 48.52–45.42 Ma ages (Figure 6e–h). Generally, presented age data usually display flat patterns, and only a small portion of the measurements of the last plateau steps displayed argon loss.

4.3. Whole Rock Geochemistry

The whole-rock major and trace element compositions of the lava samples of SNV are reported in Supplementary Table S1. The three units have been differentiated by their geochemical characteristics and by their petrographic characteristics as well. V1 unit lavas have silica-undersaturated compositions, with a range of 3.7%–13.6% content of normative nepheline, and they are represented by high Mg# values (62–69) in accordance with the presence of the modal olivine. They plotted along the basalt and slightly trachy-basaltic field and show mainly low and middle K and, more commonly, sodic character (Na2O/K2O= 0.9–5.9) (Figure 7a,b).
The V2 unit lavas cluster along the basaltic andesite, basaltic trachy-andesite, trachy-andesite, and andesite fields, have much lower Mg# values (27–56), and display much apparent middle to high K silica saturated characters (mainly quartz saturated).
The V3 unit lavas plotted along the dacite and rhyolite field, display high quartz normative values (24–27), middle-to-high K values, and show the most fractionated Mg# numbers (13–32). For comparison, data acquired by [25] and [38] are also plotted and they mostly overlapped with the data produced in this study (Figure 7a,b).
Harker variation diagrams for the volcanic rocks illustrate the evolutionary paths of the volcanic units (Figure 8 and Figure 9). All the studied lavas and acquired lava samples from the literature display negative correlations of TiO2, FeO (tot), MgO, CaO, and P2O5, relative to SiO2; whereas, Al2O3 shows inflections around wt. %. 55 SiO2, and Na2O displays scattered distribution relative to increasing SiO2 content. Trace elements, such as Ba, Rb, and Th, show general positive trends with increasing SiO2 content. Whereas Sr displays a general negative trend with increasing SiO2, other trace elements, such as Zr and Y (HFSEs), display flat-to-decreasing values with increasing SiO2.
The chondrite-normalized trace-element patterns of the samples from the three different lava units display apparent enrichments of LREE, compared to the M to HREE (Figure 10). For comparison, available data from the Sarıcakaya–Nallıhan Basin were also plotted. There is a clear decrease in the REE content from the V1 unit to the V3 unit lavas. In addition, the MREE content of the V3 unit samples from both this study and previous studies reveals apparent upward concave patterns. No significant Eu anomaly was found among the studied lavas (Figure 10).
The N-MORB normalized multi-element variation diagrams display apparent enrichments of Cs, Ba, Th, and LILEs. All units display distinct Nb–Ta negative anomalies. Interestingly, the V2 unit shows very clear Rb and Hf–Zr anomalies, whereas the other units show relatively smooth lines. In addition, contrary to the V1 and V3 units, the V2 unit does not show significant Ti and K anomalies (Figure 10)

4.4. Sr–Nd Isotopic Systematics

143Nd/144Nd(i) vs. 87Sr/86Sr(i) values of the studied lavas are usually plotted within the mantle array field and display higher 143Nd/144Nd(i) values relative to the bulk silicate earth (Figure 11a). 87Sr/86Sr(i) values of the samples range from 0.704448 to 0.70545 and ꞓ143Nd/144Nd(i) values are between 0.512630 and 0.512757, respectively. Some portions of the dacitic and rhyolitic lavas display elevated Sr isotope values that might be related to the seawater alteration. єNd values range from +1.1 to +3.5.
Many of the samples overlap with the compilation of the Eocene mafic lavas of [21] (Figure 11b). On the other hand, they often coincide with the existing Sr–Nd data produced by [25,38]. Most of the granitoid samples from the Balkanatolia Magmatic Realm (i.e., granitoids of the Tavşanlı Zone, Sivrihisar, the Rhodope area, and the South Marmara region) generally overlap and show lower єNd values compared to the mafic volcanic members of Balkanatolia.

5. Discussion

5.1. Geochronological Constraints and Stratigraphy

Kasapoğlu et al. [25] contended that the U–Pb zircon age dating conducted on samples of SNV yielded ages of 51.7 ± 4.7 Ma on basalt and 47.8 ± 2.4 Mya on rhyolite samples from Nallıhan. However, interestingly, the study’s appendix [25] presents U–Pb ages as highly variable and spanning a wide range, even in the same sample (i.e., for the basalt sample (N5) dispersing from 58.7 ± 2.7 to 48.8 ± 2.1, for the andesite sample (NB-33) from 55.5 ± 2.0 to 47.3 ± 1.5, and for the rhyolite sample (NB-8) from 58.9 ± 1.6 to 47.6 ± 1.3). These findings put into question the stratigraphic and genetic meaning of the isotopic ages. On the other hand, [38] presented much more uniform results from Sarıcakaya, those revealed 40Ar–39Ar plagioclase and biotite ages ranging between 48.13 ± 0.15 Ma and 48.78 ± 0.23 Ma but used only 3 different samples.
In this study, we re-evaluated the isotopic age data presented by [35] based on the total gas ages and plateau ages, which were not discussed within the context of the stratigraphy of the volcano-sedimentary succession in the Sarıcakaya–Nallıhan Basin.
Interestingly, the presented ages contradict the published ages and stratigraphic scheme outlined by [25], which classified the stratigraphy depending on basalt–andesite–rhyolite order. However, our field observations and new 40Ar–39Ar ages presented in this study do not support that generic order. According to our field studies, we did not detect a direct cross-cutting relationship between the basalt and andesite dacite/rhyolite. Such dispersions in U–Pb zircon ages might be related to the protracted and prolonged magma chamber processes [53]. We favor that the Eocene volcanism around the Sarıcakaya–Nallıhan Basin possibly occurred in the 48 to 44 Ma range.

5.2. Source Characteristics

Early Cenozoic magmatic units of the Anatolian segment of the Alpine–Himalayan orogenic belt often show subduction-related signatures, such as (i) negative Hf–Zr and Nb–Ta anomalies, (ii) relatively enriched LREE patterns, and (iii) LILE enrichment relative to the HFSEs mainly associated with a derivation from a mantle source metasomatized by subduction components. Moreover, several studies indicate that during the course of the magmatism, volcanic rocks that display OIB-like signatures, such as TiO2/Yb values, high Na, and nepheline-normative composition, crop out along the upper parts of the Eocene volcano-sedimentary basins [3,18]. These features are interpreted as the involvement of the asthenospheric mantle in the partial melting processes [7] (or sub-lithospheric mantle according to [54]), melting of the metasomatic veins in the peridotitic mantle [8], or more commonly, slab break-off processes [3,5].
All our samples plot above the mantle array on the classical discriminator diagram of Th/Yb vs. Nb/Yb that is occasionally used for evaluating the degree of the derivation of material from the subducting slab ([55], Figure 12a). Interestingly, the most primitive samples of the V1 unit are plotted near the OIB field in this diagram. This is in accordance with the observations from the TiO2/Yb vs. Nb/Yb diagram, on which V1 unit lavas are plotted along the OIB-like field, whereas the more fractionated, high SiO2 samples plot along the MORB field (Figure 12b). This implies that the most basic and nepheline-normative V1 unit displays an OIB-like signature.
Even though the V1 unit has OIB-like characteristics and slightly high Mg# values (62–69), along with high Ni content (72–211 µg/g) values, they still represent fractionated melts that were modified by a subducted slab with the possible contribution of melts and fluids resulting from an altered oceanic crustal (AOC) and/or sedimentary pile on top of the slab. To assess the share of the various agents in the mantle source of Eocene volcanics, we used the Ba/Th vs. La/Sm(n) ratios of the samples belonging to the V1 and V2 units (Figure 12c), eliminating the felsic samples to prevent misinterpretations. The basic and intermediate samples display high La/Sm(n), ratios ranging between 5 and 11, which imply that the addition of sedimentary melts was the primary reason for mantle metasomatism.
Interestingly, even though the Early Cenozoic magmatism around Balkanatolia has been interpreted as generated via the melting of the different source regions (e.g., subduction-modified lithosphere, thickened continental crust, or even asthenosphere) the geochemical-isotopic features of the different magma series only support derivation from a fractionated magma source.
For the worldwide examples of orogenic magmas produced in collisional and post-collisional tectonic settings, the melting of the mélange material has become a frequently suggested model to explain the geochemical and isotopic fingerprints of subduction-related magmatism (mixed metabasaltic and serpentinitic materials in various orders [56,57,58]). The Nd isotopic values vs. Hf/Nd ratios of the magmatic rocks effectively help to discriminate the primary melts sourced either from the subduction-modified mantle source or from the mélange-like materials dominated by the various sedimentary components [59]. The general distribution of our samples on the plot supports the second option, the mélange melting mechanism (experimental melts of volcanoclastic-derived sedimentary material) was viable for the generation of the SNV (Figure 12d). However, additional data, such as Mg, and Ba isotopic values, must be used to better characterize the mantle source of Eocene magmatism.
Figure 12. (a) Th/Yb vs. Nb/Yb. (b) TiO2/Yb vs. Nb/Yb plots of [51] The fields for Ocean Island Basalts (OIB), Enriched- and Normal-Mid Ocean Ridge Basalts (E- and N-MORB) are from [60,61]. (c) La/Sm(n) vs. Ba/Th diagram that differentiates altered oceanic related fluid melts vs. subducted sedimentary related inputs. (d) 143Nd/144Nd vs. Hf/Nd diagram that tracks the partial melting of mélange material in accretionary/collisional settings [55]. For comparison, data from [25,38] were also plotted.
Figure 12. (a) Th/Yb vs. Nb/Yb. (b) TiO2/Yb vs. Nb/Yb plots of [51] The fields for Ocean Island Basalts (OIB), Enriched- and Normal-Mid Ocean Ridge Basalts (E- and N-MORB) are from [60,61]. (c) La/Sm(n) vs. Ba/Th diagram that differentiates altered oceanic related fluid melts vs. subducted sedimentary related inputs. (d) 143Nd/144Nd vs. Hf/Nd diagram that tracks the partial melting of mélange material in accretionary/collisional settings [55]. For comparison, data from [25,38] were also plotted.
Minerals 12 01572 g012
In general, the origin of OIB-like alkaline magma generation in various settings has been attributed to the low degree of melting of garnet peridotite (or pyroxenite), usually occupying the deep portion of the lithospheric mantle [62,63] or sometimes referred to as sub-lithospheric mantle or asthenosphere [54]. Hence, within the context of the Early Cenozoic magmatism in northern Turkey, the magmatic rocks with OIB-like, alkaline character and relatively high #Mg content, yet enriched in LILE, have often been attributed to the derivation from the deep lithosphere or even from the asthenosphere by slab tear or delamination processes [3,5,64].
If we assume the source region is represented by the homogenous peridotitic mantle, the La/Sm vs. Sm/Yb values of the magmatic rocks can be used to assess the relative abundance of the spinel and/or garnet phases during the batch melting of the mantle. To better characterize the magmatism in the whole Balkanatolia range scale, we also plotted Early Cenozoic basic units from Biga, Armutlu–Almacık, and the Central to Eastern Pontides regions (Figure 13a).
Non-modal batch-melting models of spinel and garnet peridotites show that the contribution of the melts derived from a garnet-bearing peridotitic mantle is much more plausible for the generation of SNV. Similar rocks along the Central Pontides, studied by [3], also have high Mg# (up to 70) and display mildly alkaline and OIB-like characteristics. These samples are much more scattered in terms of the melting behavior of their mantle source, which is evident from a much higher contribution from spinel-peridotite and much higher melting degrees. Volcanic units from Biga and Armutlu–Almacık also display a high apparent contribution from the spinel peridotite-sourced regions. However, in a broader aspect, all these basic units in those areas display mildly alkaline features, negative Nb–Ta normalized values, and relatively high Mg# values, but the available data were not sufficient to discriminate the source areas and it was not possible to obtain the geodynamic settings because of the absence of detailed isotopic knowledge.
The presented modeling indicates that the SNV might be sourced from the deeper-garnet-bearing mantle areas of the lithospheric mantle; however, as pointed out by [58], alkaline lavas with a clear arc signature might also be generated via melting of the mélange diapirs themselves without the necessity for ancient episodes of mantle metasomatism by slab-derived additions.
In summary, the possible mantle source for the Saricakaya–Nallihan Volcanism has a metasomatized character and has OIB-like nepheline-normative character that is highly fluxed by sedimentary melts. This source region might be either a subduction metasomatized region or represent direct melting of the mélange-like accreted material that is quite common in the Tethyan accretionary complexes along the northern portion of Anatolia, and even in the Balkan Range [69]. Even though, traditionally, these geochemical features are attributed to deeply sourced magma generation (such as the garnet stability field 80–100 Km in depth), they can be generated via re-melting of the accreted mélange material along the ancient suture lines. Here, we speculate that this situation might not only be applicable to the Sarıcakaya–Nallıhan area and could be common in all of the Balkanatolia Magmatic Realm, since this region also contains many volcanic suites that show similar Sr–Nd characteristics and enrichment in LREE–LILE, with depletion in HFSE, and employ samples with OIB-like alkaline features.

5.3. Geochemical Evolution of the Magmatism in Crustal Levels

As stated in the sections above, even the most basic lavas of SNV show an evolved nature. In addition, a major portion of the lavas also has intermediated to acidic composition and low Mg# number (V1 lavas: 62–69; V2 lavas: 36–56 and V3 lavas: 12–25). Overall, a major portion of the elements, such as TiO2, Fe2O3, MgO, CaO, and P2O5, shows a negative correlation with increasing silica content, which supports the fractionation of the ferromagnesian minerals and minor amounts of feldspar. Flat-to-negative trends of the Al2O3 values also support the fractionation of the plagioclase phases. Slightly positive scattering and flat distribution of Ba and Rb can be interpreted as the fractionation of the biotite and feldspar and alteration-assimilation-related processes. The apparent negative and negative-to-flat trend of Sr and Y agree with the feldspar and amphibole fractionations, respectively. Scattered patterns of Zr and Th can be seen with ilmenite, zircon, monazite, or allanite crystallization, or accompanied assimilation.
The chondrite-normalized REE patterns of all the units display LREE enrichment relative to M-HREE. The V1 unit shows the most significant REE enrichment relative to chondrite composition among all units. From V1 to V3, the content of the REE enrichment decreases, and the V3 unit shows concave-shaped M-HREEs, which support amphibole fractionation.
Overall, most of the major oxide and trace element data supported the fractionation of common ferromagnesian minerals and feldspars, with or without crustal assimilation. However, since the 87Sr/86Sr values display flat trends with increasing SiO2 content (Figure 14a), the assimilation-related processes can be negligible, as also pointed out by [25].
To better reveal the fractionation-related processes, we applied Rayleigh fractionation modeling by using incompatible and compatible Sc and Th elements in Figure 14b. Compiled Kd values are given in Supplementary Table S3. Modeled curves mostly intersected with the data compilation gathered in this study. Calculated curves that primarily used clinopyroxene, amphibole, and feldspar from an intermediate magma composition could manifest the geochemical characteristics of most of the V1–V2–V3 units. Interestingly, some samples of the V1 unit show flat trends on Harker diagrams that can be related to magma mixing or source control. Overall, most of the major element variation diagrams of the lavas are controlled by fractional crystallization-related processes without any apparent assimilation-related modification.

5.4. Geodynamic Implications

One of the key aspects of the geodynamics controlling the Balkanatolia segment of the prominent belt extending from Rhodopes to Iranides is the presence of a different evolutionary history for the Neotethyan Ocean in the Izmir–Ankara–Erzincan zone (NW Anatolia) and the Pindos Ocean (Balkan regions) between the Late Cretaceous and Eocene. During the Late Cretaceous to Paleocene, HP metamorphism and accretion of the different blocks of the Afyon–Tavşanlı Zone and the Izmir–Ankara–Erzincan zone took place, and Eocene magmatism in Balkan regions and NW Anatolia was attributed to the breaking off of the Neotethyan slab sometime around Latest Cretaceous–Early Cenozoic [3,6]. However, it is not clear where, and at what depth, the Neotethyan slab break-off occurred, how it propagated through the Pindos Oceanic realm, and why the Central-Eastern Pontides (Northern Anatolia) Eocene magmatism is mostly attributed to the post-collisional magmatism [14], whereas the NW Anatolia and Balkan peninsula is rather associated with the slab break-off and/or arc-related magmatism ([7] and references therein).
To reveal the time-dependent evolution of the Early Cenozoic magmatism along the Balkanatolia region and to track the extent of the Early Cenozoic magmatic phase, we gathered all the available isotopic ages obtained from the plutonic and volcanic rocks. In addition to this, we also plotted the data regarding the metamorphism age that marks the continental subduction that is developed coevally along the southern border of the Balkanatolia realm, which is known as the Cycladic Blueschist Belt (Figure 15).
Overall, even though the whole Early Cenozoic magmatic phase shows similar evolution, as revealed by overlapping geochronological data, the Biga Peninsula and Rhodope–Thrace–Greece regions have much younger and protracted ages of magma generation that usually persist through the Eo-Oligocene and younger. In addition, in the Armutlu–Almacık and Sarıcakaya–Nallıhan regions, the age of magmatism is almost identical and not likely to extend beyond ̴40 Ma.
The ancient subduction interface, the Cycladic Blueschist Belt, has a distributed metamorphism age pattern that spans from the Early to Middle Eocene [34]. The presence of this blueschist belt might be conventionally interpreted as the anatomical part of the convergent margin setting processes, such as the HP subduction belt and corresponding arc magmatism situated within the Balkanatolia Magmatic Realm. However, the extent of the HP belt in the Aegean Sea and Anatolia diminishes within the western tip of Anatolia [34]
Hence, the relative position of the Balkanatolia Magmatic Realm represents a magmatic arc front along the northern portion of the Pindos–Neotethyan Ocean subduction that is marked by the apparent presence of HP blueschist belts. Thus, it is reasonable to attribute the Early Cenozoic magmatism in NW Turkey and Balkan regions to having been generated because of the subduction, as postulated by several different studies [11] and references therein]. However, the distance of the arc from its trench and the relationships between the presumable arc and the ancient trench itself have still not been convincingly debated.
Here, we propose a two-stage evolutionary model for the generation of the Early Cenozoic magmatism along Balkanatolia. The first phase, roughly between 58 to 45 Ma, is dominated by magmas which were produced in a compressional tectonic setting, as evidenced by the development of syn-tectonic deformation in the granitoid bodies (e.g., Kapıdağ, Sivrihisar, and South Marmara granitoids), as well as the different volcano-sedimentary complexes marked by the melting of the subduction-modified mantle sources (Figure 16a). During that phase, the Sarıcakaya–Nallıhan Volcanism was also produced over the recently welded accretionary margin of the Sakarya Continent of Eurasia plate, via the melting of mélange material and/or peridotitic mantle fluxed by sediment melts.
For the generation of the Sarıcakaya–Nallıhan Volcanism, we favor the mélange-melting model due to the Nd isotopic ratios vs. Hf–Nd systematics of the representative samples of the V1 and V2 units. Following the first phase around major portions of Balkanatolia, the second phase of magmatism was controlled by extensional tectonics along the Biga and Rhodope–Thrace portion of the Balkanatolia Magmatic Realm, which was indicated by the younger ages of the magmatism and its proximity to the subduction interface of the Cycladic Blueschist Belt, possibly due to slab roll-back (44 Ma and beyond). These magmatic episodes display protracted age ranges and are continuously developed within the Oligocene to Miocene in NW and western Anatolia due to control of the preceding “Hellenic Slab” [70,71].

6. Conclusions

Sarıcakaya–Nallıhan Volcanism (48 to 44 Ma) gives important insights into the extent and the source characteristics of the Early Cenozoic magmatism along the Balkanatolia paleo-archipelago.
The most primitive units of the Middle-Eocene volcanism, the OIB-like, nepheline-normative basalts with arc characteristics, can be demonstrated to have been formed, not by the melting of the metasomatized peridotitic mantle, but by mélange material itself. Overall, intermediate and acidic units in the Sarıcakaya–Nallıhan Volcanics are mostly produced by the fractional crystallization of the primary magmas.
Geodynamically, the Balkanatolia Magmatic Realm was possibly generated under the control of accretionary tectonics by the melting of the thickened and delaminated lithospheric mantle between ̴ 58 and 44 Ma. Following this period, extensional tectonics became much more pronounced, and the loci of the magmatism shifted toward the southern portions of NW Anatolia and the southern Balkans. Sarıcakaya–Nallıhan Volcanism occurred within the deformed fold-and-thrust belt of NW Anatolia and it was possibly generated via the melting of the accreted mélange-like material within the thickened continental crust/lithosphere. This volcanism produced the primitive (Mg# 62–69), yet evolved, lavas. The already fractionated nature of the magmas along that region resembles the magmas which were produced in many different accretionary plate boundaries worldwide and they might indicate that magmatism related to the melting of the accreted mélange material in the newly welded, lithospheric crustal domains is much more common than envisaged.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min12121572/s1, Table S1: Whole Rock Geochemistry Results; Table S2: Partition Coefficient used for partial melting model [65,67,72]; Table S3: Kd’s have been compiled from GERM Partition Coefficient (Kd) Database of Earth Reference Data and Models website (http://earthref.org/, accessed on 5 October 2022).

Author Contributions

Conceptualization, G.G. and Z.K.; methodology, G.G., F.G., Z.K. and N.A.; data curation, G.G., F.G., Z.K. and N.A.; writing—original draft preparation, G.G., F.G., Z.K. and N.A.; writing—review and editing, G.G., F.G., Z.K. and N.A.; visualization, G.G. and F.G.; supervision, Z.K. and N.A. All authors have read and agreed to the published version of the manuscript.

Funding

This study is supported by the Istanbul Technical University (ITU-BAP) by grant number 37498.

Data Availability Statement

Not applicable.

Acknowledgments

G.G., F.G., Z.K. and N.A. would like to acknowledge the careful reviews, suggestions, and support of the editors and two anonymous reviewers. G.G. would also like to thank Cansu Doğa Topbay for her help in field studies and mineral separation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Okay, A.I.; Tüysüz, O. Tethyan sutures of northern Turkey. Geol. Soc. Lond. Spec. Publ. 1999, 156, 475–515. [Google Scholar] [CrossRef] [Green Version]
  2. Christofides, G.; Perugini, D.; Koroneos, A.; Soldatos, T.; Poli, G.; Eleftheriadis, G.; Del Moro, A.; Neiva, A.M. Interplay between Geochemistry and Magma Dynamics during Magma Interaction: An Example from the Sithonia Plutonic Complex (NE Greece). Lithos 2007, 95, 243–266. [Google Scholar] [CrossRef] [Green Version]
  3. Keskin, M.; Genç, Ş.C.; Tüysüz, O. Petrology and Geochemistry of Post-Collisional Middle Eocene Volcanic Units in North-Central Turkey: Evidence for Magma Generation by Slab Breakoff Following the Closure of the Northern Neotethys Ocean. Lithos 2008, 104, 267–305. [Google Scholar] [CrossRef]
  4. Marchev, P.; Georgiev, S.; Raicheva, R.; Peytcheva, I.; von Quadt, A.; Ovtcharova, M.; Bonev, N. Adakitic Magmatism in Post-Collisional Setting: An Example from the Early–Middle Eocene Magmatic Belt in Southern Bulgaria and Northern Greece. Lithos 2013, 180–181, 159–180. [Google Scholar] [CrossRef]
  5. Gülmez, F.; Genç, Ş.C.; Keskİn, M.; Tüysüz, O. A Post-Collision Slab-Breakoff Model for the Orgin of the Middle Eocene Magmatic Rocks of the Armutlu–Almacık Belt, NW Turkey and Its Regional Implications. Geol. Soc. Lond. Spec. Publ. 2013, 372, 107–139. [Google Scholar] [CrossRef]
  6. Gülmez, F.; Genç, Ş.C.; Prelević, D.; Tüysüz, O.; Karacik, Z.; Roden, M.F.; Billor, Z. Ultrapotassic Volcanism from the Waning Stage of the Neotethyan Subduction: A Key Study from the Izmir–Ankara–Erzincan Suture Belt, Central Northern Turkey. J. Petrol. 2016, 57, 561–593. [Google Scholar] [CrossRef]
  7. Ersoy, E.Y.; Akal, C.; Genç, Ş.C.; Candan, O.; Palmer, M.R.; Prelević, D.; Uysal, İ.; Mertz-Kraus, R. U-Pb Zircon Geochronology of the Paleogene—Neogene Volcanism in the NW Anatolia: Its Implications for the Late Mesozoic-Cenozoic Geodynamic Evolution of the Aegean. Tectonophysics 2017, 717, 284–301. [Google Scholar] [CrossRef] [Green Version]
  8. Göçmengil, G.; Karacik, Z.; Genç, Ş.C.; Billor, M.Z. 40Ar–39Ar Geochronology and Petrogenesis of Postcollisional Trachytic Volcanism along the İzmir-Ankara-Erzincan Suture Zone (NE, Turkey). Turk. J. Earth Sci. 2018, 27, 1–31. [Google Scholar] [CrossRef]
  9. van Hinsbergen, D.J.J.; Torsvik, T.H.; Schmid, S.M.; Maţenco, L.C.; Maffione, M.; Vissers, R.L.M.; Gürer, D.; Spakman, W. Orogenic Architecture of the Mediterranean Region and Kinematic Reconstruction of Its Tectonic Evolution since the Triassic. Gondwana Res. 2020, 81, 79–229. [Google Scholar]
  10. Mueller, M.A.; Licht, A.; Campbell, C.; Ocakoğlu, F.; Akşit, G.G.; Métais, G.; Coster, P.M.C.; Beard, K.C.; Taylor, M.H. Sedimentary Provenance From the Evolving Forearc-to-Foreland Central Sakarya Basin, Western Anatolia Reveals Multi-Phase Intercontinental Collision. Geochem. Geophys. Geosyst. 2022, 23. [Google Scholar] [CrossRef]
  11. Okay, A.I.; Topuz, G.; Kylander-Clark, A.R.C.; Sherlock, S.; Zattin, M. Late Paleocene—Middle Eocene Magmatic Flare-up in Western Anatolia. Lithos 2022, 428–429, 106816. [Google Scholar] [CrossRef]
  12. Licht, A.; Métais, G.; Coster, P.; İbilioğlu, D.; Ocakoğlu, F.; Westerweel, J.; Mueller, M.; Campbell, C.; Mattingly, S.; Wood, M.C.; et al. Balkanatolia: The Insular Mammalian Biogeographic Province That Partly Paved the Way to the Grande Coupure. Earth-Sci. Rev. 2022, 226, 103929. [Google Scholar] [CrossRef]
  13. Böhme, M.; Aiglstorfer, M.; Antoine, P.; Appel, E.; Havlik, P.; Métais, G.; Prieto, J. Na Duong (Northern Vietnam)—An exceptional Window into Eocene Ecosystems from Southeast Asia. Zitteliana Reihe A Mitt. Bayer. Staatssamml. Palaontol. Geol. 2013, 53, 121–167. [Google Scholar]
  14. Topuz, G.; Okay, A.I.; Altherr, R.; Schwarz, W.H.; Siebel, W.; Zack, T.; Satır, M.; Şen, C. Post-Collisional Adakite-like Magmatism in the Ağvanis Massif and Implications for the Evolution of the Eocene Magmatism in the Eastern Pontides (NE Turkey). Lithos 2011, 125, 131–150. [Google Scholar] [CrossRef]
  15. Arslan, M.; Temizel, İ.; Abdioğlu, E.; Kolaylı, H.; Yücel, C.; Boztuğ, D.; Şen, C. 40Ar–39Ar Dating, Whole-Rock and Sr–Nd–Pb Isotope Geochemistry of Post-Collisional Eocene Volcanic Rocks in the Southern Part of the Eastern Pontides (NE Turkey): Implications for Magma Evolution in Extension-Induced Origin. Contrib. Miner. Pet. 2013, 166, 113–142. [Google Scholar] [CrossRef]
  16. Eyuboglu, Y.; Dudas, F.O.; Thorkelson, D.; Zhu, D.-C.; Liu, Z.; Chatterjee, N.; Yi, K.; Santash, M. Eocene Granitoids of Northern Turkey: Polybaric Magmatism in an Evolving Arc–Slab Window System. Gondwana Res. 2017, 50, 311–345. [Google Scholar] [CrossRef]
  17. Schleiffarth, W.K.; Darin, M.H.; Reid, M.R.; Umhoefer, P.J. Dynamics of Episodic Late Cretaceous–Cenozoic Magmatism across Central to Eastern Anatolia: New Insights from an Extensive Geochronology Compilation. Geosphere 2018, 14, 1990–2008. [Google Scholar] [CrossRef]
  18. Göçmengil, G.; Karacık, Z.; Genç, Ş.C.; Prelević, D.; Billor, Z. 40Ar–39Ar Ages and Petrogenesis of Middle Eocene Post-Collisional Volcanic Rocks along the Izmir-Ankara-Erzincan Suture Zone, NE Turkey. J. Asian Earth Sci. 2019, 173, 121–142. [Google Scholar] [CrossRef]
  19. Dokuz, A.; Aydin, F.; Karslı, O. Postcollisional Transition from Subduction- to Intraplate-Type Magmatism in the Eastern Sakarya Zone, Turkey: Indicators of Northern Neotethyan Slab Breakoff. GSA Bull. 2019, 131, 1623–1642. [Google Scholar] [CrossRef]
  20. Moritz, R.; Popkhadze, N.; Hässig, M.; Golay, T.; Lavoie, J.; Gugushvili, V.; Ulianov, A.; Ovtcharova, M.; Grosjean, M.; Chiaradia, M.; et al. At the Crossroads of the Lesser Caucasus and the Eastern Pontides: Late Cretaceous to Early Eocene Magmatic and Geodynamic Evolution of the Bolnisi District, Georgia. Lithos 2020, 378–379, 105872. [Google Scholar] [CrossRef]
  21. Altunkaynak, Ş.; Sual, G.; Aldanmaz, E.; Genç, C.Ş.; Dilek, Y.; Furnes, H.; Foland, K.A.; Yang, J.; Yıldız, M. Eocene Granitic Magmatism in NW Anatolia (Turkey) Revisited: New Implications from Comparative Zircon SHRIMP U–Pb and 40Ar–39Ar Geochronology and Isotope Geochemistry on Magma Genesis and Emplacement. Lithos 2012, 155, 289–309. [Google Scholar] [CrossRef]
  22. Okay, A.; Satir, M. Geochronology of Eocene plutonism and metamorphism in northwest. Geodinamica Acta 2006, 19, 251–266. [Google Scholar] [CrossRef]
  23. Ustaömer, P.A.; Ustaömer, T.; Collins, A.S.; Reischpeitsch, J. Lutetian Arc-Type Magmatism along the Southern Eurasian Margin: New U-Pb LA-ICPMS and Whole-Rock Geochemical Data from Marmara Island, NW Turkey. Miner. Pet. 2009, 96, 177–196. [Google Scholar] [CrossRef]
  24. Genç, S.; Yılmaz, Y. An example of post-collisional magmatism in northwestern Anatolia: The Kızderbent volcanics (Armutlu peninsula, Turkey). Turk. J. Earth Sci. 1997, 6, 33–42. [Google Scholar] [CrossRef]
  25. Kasapoğlu, B.; Ersoy, Y.E.; Uysal, İ.; Palmer, M.R.; Zack, T.; Koralay, E.O.; Karlsson, A. The Petrology of Paleogene Volcanism in the Central Sakarya, Nallıhan Region: Implications for the Initiation and Evolution of Post-Collisional, Slab Break-off-Related Magmatic Activity. Lithos 2016, 246–247, 81–98. [Google Scholar] [CrossRef]
  26. van Hinsbergen, D.J.J.; Hafkenscheid, E.; Spakman, W.; Meulenkamp, J.E.; Wortel, R. Nappe Stacking Resulting from Subduction of Oceanic and Continental Lithosphere below Greece. Geol 2005, 33, 325. [Google Scholar] [CrossRef] [Green Version]
  27. Jolivet, L.; Brun, J.-P. Cenozoic Geodynamic Evolution of the Aegean. Int. J. Earth Sci. 2010, 99, 109–138. [Google Scholar] [CrossRef]
  28. Brun, J.-P.; Sokoutis, D. 45 m.y. of Aegean Crust and Mantle Flow Driven by Trench Retreat. Geology 2010, 38, 815–818. [Google Scholar] [CrossRef]
  29. van Hinsbergen, D.J.J.; Kaymakci, N.; Spakman, W.; Torsvik, T.H. Reconciling the Geological History of Western Turkey with Plate Circuits and Mantle Tomography. Earth Planet. Sci. Lett. 2010, 297, 674–686. [Google Scholar] [CrossRef]
  30. Okay, A.I.; Zattin, M.; Özcan, E.; Sunal, G. Uplift of Anatolia. Turk. J. Earth Sci. 2020, 29, 696–713. [Google Scholar] [CrossRef]
  31. Garzanti, E.; Radeff, G.; Malusà, M.G. Slab Breakoff: A Critical Appraisal of a Geological Theory as Applied in Space and Time. Earth-Sci. Rev. 2018, 177, 303–319. [Google Scholar] [CrossRef]
  32. Karacık, Z.; Yılmaz, Y.; Pearce, J.A.; Ece, Ö.I. Petrochemistry of the South Marmara Granitoids, Northwest Anatolia, Turkey. Int. J. Earth Sci. 2008, 97, 1181–1200. [Google Scholar] [CrossRef]
  33. Tomaschek, F.; Kennedy, A.K.; Villa, I.V.; Lagos, M.; Ballhaus, C. Zircons from Syros, Cyclades, Greece--Recrystallization and Mobilization of Zircon During High-Pressure Metamorphism. J. Petrol. 2003, 44, 1977–2002. [Google Scholar] [CrossRef] [Green Version]
  34. Çetinkaplan, M.; Candan, O.; Oberhänsli, R.; Sudo, M.; Cenki-Tok, B. P–T–t Evolution of the Cycladic Blueschist Unit in Western Anatolia/Turkey: Geodynamic Implications for the Aegean Region. J. Metamorph. Geol. 2020, 38, 379–419. [Google Scholar] [CrossRef]
  35. Şahin, M.; Yaltırak, C.; Karacık, Z. A Case Study of Compression to Escape Tectonic Transition: Tectonic Evolution of the Nallıhan Wedge and Comparison with the Tercan Wedge (Eastern Mediterranean, Turkey). J. Asian Earth Sci. 2019, 174, 311–331. [Google Scholar] [CrossRef]
  36. Altınlı, İ.E. Geology of the Eastern Territory of Nallihan (Ankara Province); Istanbul Üniversitesi Fen Fakültesi Mecmuasi: Istanbul, Turkey, 1977; p. 42. [Google Scholar]
  37. Göncüoğlu, M.C.; Turhan, N.; Şentürk, K.; Özcan, A.; Uysal, Ş.; Yaliniz, M.K. A geotraverse across northwestern Turkey: Tectonic units of the Central Sakarya region and their tectonic evolution. Geol. Soc. Lond. Spec. Publ. 2000, 173, 139–161. [Google Scholar] [CrossRef]
  38. Büyükkahraman, G. Petrology of Eocene Volcanic Rocks from the Central Sakarya Zone (Northwestern Anatolia, Turkey): New Evidence from Ar-Ar and Sr-Nd Isotope Determinations. Arab. J. Geosci. 2016, 9, 675. [Google Scholar] [CrossRef]
  39. Göçmengil, G.; Karacık, Z.; Genç, Ş.C. Volcano Stratigraphic Investigation of the Post-Collisional Middle Eocene Magmatism Around İzmir-Ankara-Erzincan Suture Zone (NE, Turkey). Türkiye Jeol. Bülteni/Geol. Bull. Turk. 2018, 61, 131–162. [Google Scholar] [CrossRef] [Green Version]
  40. Köksal, S.; Göncüoğlu, M.C. Sr and Nd Isotopic Characteristics of Some S-, I- and A-Type Granitoids from Central Anatolia. T. J. Earth Sci. 2008, 17, 111–127. [Google Scholar]
  41. Spell, T.L.; McDougall, I. Characterization and Calibration of 40Ar/39Ar Dating Standards. Chem. Geol. 2003, 198, 189–211. [Google Scholar] [CrossRef]
  42. Koppers, A.A. ArArCALC—Software for 40Ar/39Ar age calculations. Comput. Geosci. 2002, 28, 605–619. [Google Scholar] [CrossRef]
  43. Bas, M.J.L.; Maitre, R.W.L.; Streckeisen, A.; Zanettin, B.; IUGS Subcommission on the Systematics of Igneous Rocks. A Chemical Classification of Volcanic Rocks Based on the Total Alkali-Silica Diagram. J. Petrol. 1986, 27, 745–750. [Google Scholar] [CrossRef] [Green Version]
  44. Irvine, T.N.; Baragar, W.R.A. A Guide to the Chemical Classification of the Common Volcanic Rocks. Can. J. Earth Sci. 1971, 8, 523–548. [Google Scholar] [CrossRef]
  45. Kuno, H. Lateral Variation of Basalt Magma Type across Continental Margins and Island Arcs. Bull. Volcanol. 1966, 29, 195–222. [Google Scholar] [CrossRef]
  46. Peccerillo, A.; Taylor, S.R. Geochemistry of Eocene Calc-Alkaline Volcanic Rocks from the Kastamonu Area, Northern Turkey. Contrib. Mineral. Petrol. 1976, 58, 63–81. [Google Scholar] [CrossRef]
  47. Boynton, W.V. Geochemistry of Rare Earth Elements: Meteorite Studies. In Rare Earth Element Geochemistry, Henderson, P., Ed.; Elsevier: New York, NY, USA, 1984; pp. 63–114. [Google Scholar]
  48. Sun, S.; McDonough, W.F. Chemical and Isotopic Systematics of Oceanic Basalts: Implications for Mantle Composition and Processes. SP 1989, 42, 313–345. [Google Scholar] [CrossRef]
  49. Gill, J.B. Orogenic Andesites and Plate Tectonics; Springer: Berlin/Heidelberg, Germany, 1981; p. 390. [Google Scholar]
  50. Zindler, A.; Hart, S. Chemical geodynamics. Annu. Rev. Earth Planet. Sci. 1986, 14, 493–571. [Google Scholar] [CrossRef]
  51. Demirbilek, M.; Mutlu, H.; Fallick, A.E.; Sarıiz, K.; Kibici, Y. Petrogenetic Evolution of the Eocene Granitoids in Eastern Part of the Tavşanlı Zone in Northwestern Anatolia, Turkey. Lithos 2018, 314–315, 236–259. [Google Scholar] [CrossRef]
  52. Özdamar, Ş.; Sunal, G.; DemİRoğlu, M.; Yaltirak, C.; Bİllor, M.Z.; Georgiev, S.; Hames, W.; Dunkl, I.; Aydin, H.C. Metamorphism, Magmatism, and Exhumation History of the Tavşanlı Zone, NW Turkey: New Petrological Constraints. Turk. J. Earth Sci. 2018, 27, 269–293. [Google Scholar] [CrossRef]
  53. Szymanowski, D.; Wotzlaw, J.-F.; Ellis, B.S.; Bachmann, O.; Guillong, M.; von Quadt, A. Protracted Near-Solidus Storage and Pre-Eruptive Rejuvenation of Large Magma Reservoirs. Nat. Geosci. 2017, 10, 777–782. [Google Scholar] [CrossRef]
  54. Lustrino, M.; Anderson, D.L. The Mantle Isotopic Printer: Basic Mantle Plume Geochemistry for Seismologists and Geodynamicists. In The Interdisciplinary Earth: A Volume in Honor of Don L. Anderson; Foulger, G.R., Lustrino, M., King, S.D., Eds.; Geological Society of America: Boulder, CO, USA, 2015; Volume 514. [Google Scholar]
  55. Pearce, J.A. Geochemical Fingerprinting of Oceanic Basalts with Applications to Ophiolite Classification and the Search for Archean Oceanic Crust. Lithos 2008, 100, 14–48. [Google Scholar] [CrossRef]
  56. Marschall, H.R.; Schumacher, J.C. Arc Magmas Sourced from Mélange Diapirs in Subduction Zones. Nat. Geosci. 2012, 5, 862–867. [Google Scholar] [CrossRef]
  57. Codillo, E.A.; Le Roux, V.; Marschall, H.R. Arc-like Magmas Generated by Mélange-Peridotite Interaction in the Mantle Wedge. Nat. Commun. 2018, 9, 2864. [Google Scholar] [CrossRef] [PubMed]
  58. Cruz-Uribe, A.M.; Marschall, H.R.; Gaetani, G.A.; Le Roux, V. Generation of Alkaline Magmas in Subduction Zones by Partial Melting of Mélange Diapirs—An Experimental Study. Geology 2018, 46, 343–346. [Google Scholar] [CrossRef]
  59. Nielsen, S.G.; Marschall, H.R. Geochemical Evidence for Mélange Melting in Global Arcs. Sci. Adv. 2017, 3, e1602402. [Google Scholar] [CrossRef] [Green Version]
  60. Workman, R.K.; Hart, S.R.; Jackson, M.; Regelous, M.; Farley, K.A.; Blusztajn, J.; Kurz, M.; Staudigel, H. Recycled Metasomatized Lithosphere as the Origin of the Enriched Mantle II (EM2) End-Member: Evidence from the Samoan Volcanic Chain: Recycled Metasomatized Lithosphere. Geochem. Geophys. Geosyst. 2004, 5. [Google Scholar] [CrossRef]
  61. White, W.M.; Klein, E.M. Composition of the Oceanic Crust. In Treatise on Geochemistry; Elsevier: Amsterdam, The Netherlands, 2014; pp. 457–496. [Google Scholar]
  62. Hirschmann, M.M.; Kogiso, T.; Baker, M.B.; Stolper, E.M. Alkalic Magmas Generated by Partial Melting of Garnet Pyroxenite. Geol 2003, 31, 481. [Google Scholar] [CrossRef]
  63. Niu, Y.; Wilson, M.; Humphreys, E.R.; O’Hara, M.J. The Origin of Intra-Plate Ocean Island Basalts (OIB): The Lid Effect and Its Geodynamic Implications. J. Petrol. 2011, 52, 1443–1468. [Google Scholar] [CrossRef] [Green Version]
  64. Karsli, O.; Uysal, İ.; Dilek, Y.; Aydin, F.; Kandenir, R. Geochemical Modelling of Early Eocene Adakitic Magmatism in the Eastern Pontides, NE Anatolia: Continental Crust or Subducted Oceanic Slab Origin? Int. Geol. Rev. 2013, 55, 2083–2095. [Google Scholar] [CrossRef]
  65. McKenzie, D.; O’Nions, R.K. Partial melt distributions from inversion of rare Earth element concentrations. J. Petrol. 1991, 32, 1021–1091. [Google Scholar] [CrossRef]
  66. Thirlwall, M.F.; Smith, T.E.; Graham, A.M.; Theodorou, N.; Hollings, P.; Davidson, J.P.; Arculus, R.J. High Field Strength Element Anomalies in Arc Lavas: Source or Process? J. Petrol. 1994, 35, 819–838. [Google Scholar] [CrossRef]
  67. Adam, J.; Green, T. Trace Element Partitioning between Mica- and Amphibole-Bearing Garnet Lherzolite and Hydrous Basanitic Melt: 1. Experimental Results and the Investigation of Controls on Partitioning Behaviour. Contrib. Mineral. Petrol. 2006, 152, 1–17. [Google Scholar] [CrossRef]
  68. Erenoglu, O.; Bozcu, M.; Billor, M.Z. Age and Petrology of Eocene-Oligocene Calc-Alkaline Volcanism in Biga Peninsula (NW Turkey): Implications for Magma Origin and Geodynamic Evolution. J. Afr. Earth Sci. 2022, 192, 104559. [Google Scholar] [CrossRef]
  69. Göçmengil, G.; Altıntaş, İ.E.; Topuz, G.; Çelik, Ö.F.; Özkan, M. Diverse Tectonic Settings of Formation of the Metaigneous Rocks in the Jurassic Metamorphic Accretionary Complexes (Refahiye, NE Turkey) and Their Geodynamic Implications. Geodin. Acta 2013, 26, 294–310. [Google Scholar] [CrossRef] [Green Version]
  70. Aysal, N. Mineral Chemistry, Crystallization Conditions and Geodynamic Implications of the Oligo–Miocene Granitoids in the Biga Peninsula, Northwest Turkey. J. Asian Earth Sci. 2015, 105, 68–84. [Google Scholar] [CrossRef]
  71. Karacık, Z.; Genç, Ş.C.; Gülmez, F. Petrochemical Features of Miocene Volcanism around the Çubukludağ Graben and Karaburun Peninsula, Western Turkey: Implications for Crustal Melting Related Silicic Volcanism. J. Asian Earth Sci. 2013, 73, 199–217. [Google Scholar] [CrossRef]
  72. Green, T.H.; Blundy, J.D.; Adam, J.; Yaxley, G.M. SIMS determination of trace element partition coefficients between garnet, clinopyroxene and hydrous basaltic liquids at 2–7.5 GPa and 1080–1200 °C. J. Lithos. 2000, 53, 165–187. [Google Scholar] [CrossRef]
Figure 2. Simplified sketch map showing the distribution of the basaltic (V1 unit), andesitic (V2 unit), and dacitic–rhyolitic lavas (V3 unit) in the Sarıcakaya–Nallıhan Basin.
Figure 2. Simplified sketch map showing the distribution of the basaltic (V1 unit), andesitic (V2 unit), and dacitic–rhyolitic lavas (V3 unit) in the Sarıcakaya–Nallıhan Basin.
Minerals 12 01572 g002
Figure 3. (a) Continental red beds of the Kızılçay Formation which unconformably overlay the basement units. (b) Nummulite-bearing sandstone levels of the Güvenç Formation. (c) Columnar joints in basaltic lavas. (d) General view of the flow deposits and lava flow alternations. (e) General view of the rhyolitic-dacitic plug of the SNV. (f) Close-up view of the rhyolitic-dacitic unit in the Sarıcakaya–Nallıhan Basin. The rhyolitic–dacitic unit also shows common columnar jointing.
Figure 3. (a) Continental red beds of the Kızılçay Formation which unconformably overlay the basement units. (b) Nummulite-bearing sandstone levels of the Güvenç Formation. (c) Columnar joints in basaltic lavas. (d) General view of the flow deposits and lava flow alternations. (e) General view of the rhyolitic-dacitic plug of the SNV. (f) Close-up view of the rhyolitic-dacitic unit in the Sarıcakaya–Nallıhan Basin. The rhyolitic–dacitic unit also shows common columnar jointing.
Minerals 12 01572 g003
Figure 4. (a) Iddingisitization of the partly fresh olivine in the basalt. (b) Clinopyroxene phenocrysts in the V1 unit. (c) Mineral cloths in the andesitic V2 units. (d) Apparent pleochroism of the amphibole phenocrysts in the V2 unit. (e) General view of the phenocryst-rich V3 unit. (f) Resorbed quartz phenocrysts embedded in the V3 unit. Ol: olivine, Cpx: clinopyroxene, Hbl: hornblende, Sn: sanidine, Bt: biotite, Qtz: quartz.
Figure 4. (a) Iddingisitization of the partly fresh olivine in the basalt. (b) Clinopyroxene phenocrysts in the V1 unit. (c) Mineral cloths in the andesitic V2 units. (d) Apparent pleochroism of the amphibole phenocrysts in the V2 unit. (e) General view of the phenocryst-rich V3 unit. (f) Resorbed quartz phenocrysts embedded in the V3 unit. Ol: olivine, Cpx: clinopyroxene, Hbl: hornblende, Sn: sanidine, Bt: biotite, Qtz: quartz.
Minerals 12 01572 g004
Figure 5. (af) 40Ar–39Ar age spectra and plateau ages of basaltic units (V1 unit) of SNV.
Figure 5. (af) 40Ar–39Ar age spectra and plateau ages of basaltic units (V1 unit) of SNV.
Minerals 12 01572 g005
Figure 6. (ah) 40Ar–39Ar age spectra and plateau ages from andesitic and dacitic units of SNV (ad figures belonging to the V2 unit and eh figures belonging to the V3 unit).
Figure 6. (ah) 40Ar–39Ar age spectra and plateau ages from andesitic and dacitic units of SNV (ad figures belonging to the V2 unit and eh figures belonging to the V3 unit).
Minerals 12 01572 g006
Figure 7. (a) Classification of the SNV (total alkali vs. silica) on diagram of [25,38,43] (TB: trachybasalt; BTA: basaltic trachyandesite; TA: trachy-andesite; T: trachyte; B: basalt; BA: basaltic andesite; A: andesite; D: dacite; IB: [44] K: [45]). (b) Classification of lavas on K2O vs. SiO2 diagram of [46]. The lava samples from the other studies examining SNV were also plotted for comparison.
Figure 7. (a) Classification of the SNV (total alkali vs. silica) on diagram of [25,38,43] (TB: trachybasalt; BTA: basaltic trachyandesite; TA: trachy-andesite; T: trachyte; B: basalt; BA: basaltic andesite; A: andesite; D: dacite; IB: [44] K: [45]). (b) Classification of lavas on K2O vs. SiO2 diagram of [46]. The lava samples from the other studies examining SNV were also plotted for comparison.
Minerals 12 01572 g007
Figure 8. (ag) Harker major oxide diagrams of SNV (The symbols as Figure 7).
Figure 8. (ag) Harker major oxide diagrams of SNV (The symbols as Figure 7).
Minerals 12 01572 g008
Figure 9. (af) Harker minor element diagrams of SNV (the symbols as Figure 7).
Figure 9. (af) Harker minor element diagrams of SNV (the symbols as Figure 7).
Minerals 12 01572 g009
Figure 10. (ac) Chondrite-normalized REE diagrams of SNV. Normalization values are taken from [47]. (df) Primitive mantle-normalized [48] trace-element diagrams of SNV.
Figure 10. (ac) Chondrite-normalized REE diagrams of SNV. Normalization values are taken from [47]. (df) Primitive mantle-normalized [48] trace-element diagrams of SNV.
Minerals 12 01572 g010
Figure 11. (a) Initial 87Sr/86Sr and 143Nd/144Nd compositions of three different volcanic units from the Sarıcakaya–Nallıhan Basin. Compositions of MORB and Mantle Array from [49]; EMI (enriched mantle type I), BSE (Bulk silicate earth), HIMU, CHUR (Chondritic Uniform Reservoir) Sr and -Nd reference lines after [50]. (b) Comparison of studied lava flows with the different reservoirs from the Balkanatolia Magmatic Realm. NW Anatolia Eocene mafic lavas data from [3,7]; Nallıhan–Sarıcakaya data from [25,38]; Rhodope–Thrace–Greece magmatics from [4]; and Tavşanlı Zone granitoids from [51,52].
Figure 11. (a) Initial 87Sr/86Sr and 143Nd/144Nd compositions of three different volcanic units from the Sarıcakaya–Nallıhan Basin. Compositions of MORB and Mantle Array from [49]; EMI (enriched mantle type I), BSE (Bulk silicate earth), HIMU, CHUR (Chondritic Uniform Reservoir) Sr and -Nd reference lines after [50]. (b) Comparison of studied lava flows with the different reservoirs from the Balkanatolia Magmatic Realm. NW Anatolia Eocene mafic lavas data from [3,7]; Nallıhan–Sarıcakaya data from [25,38]; Rhodope–Thrace–Greece magmatics from [4]; and Tavşanlı Zone granitoids from [51,52].
Minerals 12 01572 g011
Figure 13. Non-modal batch melting curves for the Early Cenozoic volcanic units from the Balkanatolia Magmatic Realm and different units along northern Turkey on a La/Sm vs. Sm/Yb diagram. Curves were calculated based on the melting of garnet and spinel peridotite sources. Only samples with Mg# >60 are plotted on the diagram. Source modes from [65]. Source modes and melt modes as follows [66,67]: garnet peridotite, Olv0.598 (0.05) + Opx0.211 (0.2) + Cpx0.076 (0.3) + Gt0.115 (0.45), and spinel peridotite, Olv0.578 (0.1) + Opx0.27 (0.27) + Cpx0.119 (0.5) + Sp0.033 (0.13). Distribution coefficients also from [65,67]. Model implemented from [5]. Armutlu–Almacık data from [5]; Biga values from [7,68]; Central Pontides values from [3]; and Eastern Pontides values from [18], references therein. Grey diamonds in the Sarıcakaya–Nallıhan section taken from [25]. Additional information is shown in Supplementary Table S2.
Figure 13. Non-modal batch melting curves for the Early Cenozoic volcanic units from the Balkanatolia Magmatic Realm and different units along northern Turkey on a La/Sm vs. Sm/Yb diagram. Curves were calculated based on the melting of garnet and spinel peridotite sources. Only samples with Mg# >60 are plotted on the diagram. Source modes from [65]. Source modes and melt modes as follows [66,67]: garnet peridotite, Olv0.598 (0.05) + Opx0.211 (0.2) + Cpx0.076 (0.3) + Gt0.115 (0.45), and spinel peridotite, Olv0.578 (0.1) + Opx0.27 (0.27) + Cpx0.119 (0.5) + Sp0.033 (0.13). Distribution coefficients also from [65,67]. Model implemented from [5]. Armutlu–Almacık data from [5]; Biga values from [7,68]; Central Pontides values from [3]; and Eastern Pontides values from [18], references therein. Grey diamonds in the Sarıcakaya–Nallıhan section taken from [25]. Additional information is shown in Supplementary Table S2.
Minerals 12 01572 g013
Figure 14. (a) 87Sr/86Sr vs. SiO2 contents of the SNV; (b) Th vs. Sc bivariate diagram displaying theoretical Rayleigh fractional crystallization curves depending on phase combinations inserted. Kd values for different minerals were taken from http://earthref.org (accessed on 5 October 2022) and are given in Supplementary Data. Tick marks on each curve correspond to 5% crystallization intervals.
Figure 14. (a) 87Sr/86Sr vs. SiO2 contents of the SNV; (b) Th vs. Sc bivariate diagram displaying theoretical Rayleigh fractional crystallization curves depending on phase combinations inserted. Kd values for different minerals were taken from http://earthref.org (accessed on 5 October 2022) and are given in Supplementary Data. Tick marks on each curve correspond to 5% crystallization intervals.
Minerals 12 01572 g014
Figure 15. Age compilation of the magmatic and HP metamorphic events that shaped the Balkanatolia Magmatic Realm. Figure implemented from the study of Ersoy et al. [7]. Data compiled from the studies of [4,5,7] and references therein, [23,25,32,33,34,35,51,52,69].
Figure 15. Age compilation of the magmatic and HP metamorphic events that shaped the Balkanatolia Magmatic Realm. Figure implemented from the study of Ersoy et al. [7]. Data compiled from the studies of [4,5,7] and references therein, [23,25,32,33,34,35,51,52,69].
Minerals 12 01572 g015
Figure 16. The geodynamic scenario proposed for the generation of the Balkanatolia Magmatic Realm. (a) From 58 to 44 Ma, magmatism was mostly dominated by the driving compressional deformation in Anatolide–Tauride and Pontides and was marked by the melting of the modified mantle sources. (b) After 44 Ma, Cycladic HP metamorphism became more pronounced and was possibly due to the roll-back of the subducting Pindos–Neotethyan slab giving rise to extensional deformation and prolonged magmatism in the Biga Peninsula and Rhodope–Thrace. In contrast, magmatism along NW Anatolia and the main portion of western Anatolia continued with cessations and did not persist during the Early Oligocene.
Figure 16. The geodynamic scenario proposed for the generation of the Balkanatolia Magmatic Realm. (a) From 58 to 44 Ma, magmatism was mostly dominated by the driving compressional deformation in Anatolide–Tauride and Pontides and was marked by the melting of the modified mantle sources. (b) After 44 Ma, Cycladic HP metamorphism became more pronounced and was possibly due to the roll-back of the subducting Pindos–Neotethyan slab giving rise to extensional deformation and prolonged magmatism in the Biga Peninsula and Rhodope–Thrace. In contrast, magmatism along NW Anatolia and the main portion of western Anatolia continued with cessations and did not persist during the Early Oligocene.
Minerals 12 01572 g016
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Göçmengil, G.; Gülmez, F.; Karacik, Z.; Aysal, N. Petrogenesis of Early Cenozoic Sarıcakaya–Nallıhan Volcanism in NW Turkey: Implications for the Geodynamic Setting and Source Characterization of the Balkanatolia Magmatic Realm. Minerals 2022, 12, 1572. https://doi.org/10.3390/min12121572

AMA Style

Göçmengil G, Gülmez F, Karacik Z, Aysal N. Petrogenesis of Early Cenozoic Sarıcakaya–Nallıhan Volcanism in NW Turkey: Implications for the Geodynamic Setting and Source Characterization of the Balkanatolia Magmatic Realm. Minerals. 2022; 12(12):1572. https://doi.org/10.3390/min12121572

Chicago/Turabian Style

Göçmengil, Gönenç, Fatma Gülmez, Zekiye Karacik, and Namık Aysal. 2022. "Petrogenesis of Early Cenozoic Sarıcakaya–Nallıhan Volcanism in NW Turkey: Implications for the Geodynamic Setting and Source Characterization of the Balkanatolia Magmatic Realm" Minerals 12, no. 12: 1572. https://doi.org/10.3390/min12121572

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