Journal Pre-proofs Fluid Inclusion and Stable Isotope Constraints (C, O, H) on the Dağbaş ı Fe– Cu–Zn Skarn Mineralization (Trabzon, NE Turkey) Yılmaz Demir, Ali Dişli PII: DOI: Reference:
S0169-1368(18)30302-0 https://doi.org/10.1016/j.oregeorev.2019.103235 OREGEO 103235
To appear in:
Ore Geology Reviews
Received Date: Revised Date: Accepted Date:
5 September 2018 30 October 2019 14 November 2019
Please cite this article as: Y. Demir, A. Dişli, Fluid Inclusion and Stable Isotope Constraints (C, O, H) on the Dağbaş ı Fe–Cu–Zn Skarn Mineralization (Trabzon, NE Turkey), Ore Geology Reviews (2019), doi: https://doi.org/ 10.1016/j.oregeorev.2019.103235
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2019 Published by Elsevier B.V.
1
Fluid Inclusion and Stable Isotope Constraints (C, O, H) on the Dağbaşı Fe–Cu–Zn
2
Skarn Mineralization (Trabzon, NE Turkey)
3 4 5 6 7 8 9
Yılmaz Demir1 and Ali Dişli1 1Recep
Tayyip Erdogan University, Department of Geological Engineering, 53100 Rize, Turkey (
[email protected])
Abstract
10
The Dağbaşı Fe–Cu–Zn skarn mineralization developed along the contact between the block
11
and lens shaped limestones of the Lower Cretaceous Berdiga Formation and the Upper Cretaceous
12
Dağbaşı Granitoid. The exoskarn-type mineralization is characterized by prograde stage garnet and
13
pyroxene, while the retrograde stage is characterized by epidote, tremolite, actinolite, and chlorite.
14
Quartz and calcites were observed in both stages of the skarn development. The ore minerals mainly
15
consist of magnetite and hematite, with a lesser amount of pyrrhotite, pyrite, chalcopyrite, sphalerite,
16
and minor galena. The homogenization temperatures (Th) and salinity values of the prograde stage
17
halite-bearing fluid inclusions are in the range of 412 to 514 °C and 48.8–61.8 wt% NaCl equ.,
18
respectively. The second stage liquid- and vapor-rich fluid inclusion assemblage reveals that boiling at
19
temperatures of 353–458 °C took place after the formation of halite-bearing fluid inclusions. Final
20
stage liquid-rich fluid inclusions were characterized by low Th (160 and 327 °C) and salinity values
21
(0.5 and 6.2 wt% NaCl equ.). The decreasing salinity trend of the fluid inclusions versus Th indicated
22
that meteoric water was involved in the hydrothermal solutions. Eutectic temperatures (Te) of the
23
prograde stage fluid inclusions were found to be CaCl2 dominated, while retrograde stage inclusions
24
contained different salt combinations rather than a specific salt type. The minimum trapping pressures
25
of the early stage brine fluid inclusions were calculated to be between 710 and 884 bar, while later
26
stage inclusions had much lower trapping pressures between ~195 and 445 bar. The δ18O isotopes of
27
prograde stage quartz, garnet, and pyroxenes are close to the composition of the hydrothermal
28
solutions of magmatic sources. Moreover, retrograde stage quartz, epidote, tremolite‒actinolite, and
29
calcite minerals and their equilibrated solutions were found to be highly depleted by δ18O isotopes.
30
Therefore, the fluid inclusion and stable isotope constraints suggest that the hydrothermal solutions of
31
magmatic origin were responsible for the prograde skarn stage, while a mixture of magmatic and
32
meteoric solutions were responsible for the ore formation in a shallow skarn environment.
33 34 35 36 37 38
Keywords: Dagbası Fe–Cu–Zn skarn, Skarn deposits, Fluid inclusion, Stable isotopes, Trabzon, NE Turkey
1
39
1. Introduction
40
The eastern Black Sea region is one of the most important mining provinces in
41
Turkey, located in the Alpine Metallogenic Belt, which extends approximately 500 km in
42
length and 100 km in width in an E–W direction. Many skarn-type mineralizations in this area
43
are accompanied by massive sulfide, porphyry, epithermal, and hydrothermal type ore
44
deposits. Some of the most well-known skarn deposits are Demirköy (Artvin), Kartiba,
45
Sivrikaya (Rize), Özdil, Ögene, Dağbaşı (Trabzon), Dereli, Kirazören (Giresun), Çambaşı
46
(Ordu), Eğrikar, Düzköy, Arnastal and Camiboğazı (Gümüşhane) (Fig. 1).
47 48
The geological setting of this region is especially appropriate for the formation of
49
skarn-type deposits, due to the large volume of Lower Cretaceous limestones of Berdiga
50
Formation and the limestone layers in an Upper Cretaceous volcano-sedimentary unit that is
51
in contact with younger granitic intrusions over an extensive area (Demir et al., 2017).
52
Therefore the skarn-type deposits in this region have attracted considerable academic research
53
and mining activities over two decades. Most of the previous research has focused on the
54
geological and mineralogical properties, skarn and granitoid geochemistry, and alteration
55
mineralogy of these skarn ores (Aslan, 1991; Hasançebi, 1993; Sipahi, 1997, 2011; Saraç,
56
2003; Çiftçi, 2011; Yılmaz, 2016; Sipahi et al., 2017; Demir et al., 2017; Demir, 2019).
57
According to these studies, some of the skarns in the region formed along the contact between
58
the limestones of the Lower Cretaceous Berdiga Formation and the Upper Cretaceous
59
granitoids (e.g., Dağbaşı, Özdil). However, some other skarns formed along the contact
60
between Upper Cretaceous volcano-sedimentary units and Upper Cretaceous‒Eocene
61
intrusions
62
mineralization has been reported from Özdil, Kartiba, Çambaşı, Kotana, Arnastal,
63
Camiboğazı, and Sivrikaya, whereas both exoskarn- and endoskarn-type mineralization have
64
been reported from Kirazören, Eğrikar, and Ögene (Demir et al., 2017; Demir, 2019).
65
Magnetite and hematite are the main ore minerals at Özdil, Kartiba, Camiboğazı, and
66
Sivrikaya, whereas sulfide minerals accompany magnetite and hematite at Arnastal, Kotana,
67
Kirazören, Eğrikar, and Dağbaşı skarn deposits (Demir et al., 2017). These studies have
68
shown that the skarn-type deposits in the region show differences in terms of the host rock
69
relationship, skarn type, and mineralogical properties. However, in these studies the reason
70
for these differences in skarn types and their mineralogical compositions are not sufficiently
71
explained.
(Kirazören,
Sivrikaya,
Arnastal-Camiboğazı,
72 2
Eğrikar).
Exoskarn-type
73
Fluid inclusion and stable isotope studies provide useful information on the formation
74
conditions of skarn deposits, the origin of hydrothermal solutions and fluid evolutions
75
(Shepherd et al., 1985; Wilkinson, 2001; Bodnar et al., 2014). With the fluid inclusion studies,
76
it is possible to determine the formation conditions (temperature, pressure, depth) of the
77
solutions, which were formed in the early and late skarn stages, and the compositions
78
(salinity, possible salt types) of these fluids. Using stable isotope studies, it is also possible to
79
determine the source of the solutions by measuring the isotope compositions of each mineral
80
phase. Isotope studies are also used to calculate the formation temperatures of equilibrated
81
mineral phases and to determine the isotope composition of the solutions equilibrated with
82
these minerals (Chiba et al., 1989; Zheng, 1993a, 1993b; Taylor, 1997).
83 84
The fluid inclusion and stable isotope characteristics of some of these skarns in the
85
northeastern region have also been investigated (Kurt, 2014; Sipahi et al., 2017; Demir et al.,
86
2017). Kurt (2014) indicated that the Kirazören skarn was formed at temperatures above 360–
87
400 °C by magmatic originated hydrothermal solutions, and the salinity values were higher
88
than 26.3 wt% NaCl equ. A mixture of magmatic and meteoric solutions was suggested for
89
the formation of the Eğrikar skarn deposit at a temperature range of 200–425 °C and salinity
90
values between 3.9–15.4 wt% NaCl equ. by Sipahi et al. (2017). A similar temperature
91
(between 166‒462 °C) and salinity range (between 0.35–14.3 wt% NaCl equ.) was also
92
determined for the Sivrikaya skarn deposit by Demir et al. (2017), and a mixture of magmatic
93
and meteoric solutions was suggested for the skarn formation. The skarns in these studies
94
were formed along the contact of Upper Cretaceous volcano-sedimentary units and Eocene
95
granitoids. However, the Dağbaşı skarn deposit is different from the others by the host rock
96
lithology, because it was formed along the Lower Cretaceous Berdiga limestones and Upper
97
Cretaceous Dağbaşı granitoid (Demir, 2019). Therefore, studies on the Kirazören, Eğrikar,
98
and Sivrikaya skarn deposits do not represent all of the skarn formations in the region, and do
99
not demonstrate the formation conditions, fluid compositions and genesis of all the skarn
100
deposits.
101 102
Four different skarn deposits were observed in the Dağbaşı area, including the İpekçili,
103
Köprüüstü, Kükürtlü and Dere Mahalle locations (Fig. 2). These deposits are not currently in
104
production, and there is currently no reserve record available. However, closed old adits and
105
an amount of mine waste around them indicate that there was once historical mining
106
production. In addition to ore-host rock relations and mineralogical properties, Demir (2019) 3
107
also investigated the composition of silicate, oxide, and sulfide mineral phase and
108
geochemical properties of the Dağbaşı granitoid and its relation with the skarn types.
109
However, this earlier research did not sufficiently clarify the composition and the source of
110
solutions and the formation conditions of Dağbaşı skarns. Therefore, in this present study
111
fluid inclusion and C–O–H isotopic characteristics of the Dağbaşı skarn mineralization are
112
described in detail in order to investigate the formation conditions, the source of hydrothermal
113
solutions and its evolutions.
114 115
2. Regional geology
116
The geological structure of northeastern Turkey, which lies in the Alpine‒Himalayan
117
Orogenic Belt, was formed as a result of the long-lived subduction of the Tethyan Ocean
118
(Okay and Şahintürk, 1997). Early to Middle Carboniferous high T/low to medium-P
119
metamorphic units which consist of quartz‒feldispathic gneiss, schist, amphibolite, phyllite,
120
chert, marble, and metaperidotites are the oldest units in the Hercynian basement (Topuz et
121
al., 2007; Dokuz, 2011; Dokuz et al., 2015). High-K, I-type granitoids of Carboniferous to
122
Early Permian age were emplaced into the Hercynian basement throughout the Sakarya Zone
123
(Topuz et al., 2010; Dokuz et al., 2017).
124 125
Paleozoic basement was overlain, transgressively, by Lower and Middle Jurassic
126
volcanics and interbedded carbonate sediments of the Şenköy Formation, which represent an
127
extensional arc environment (Okay and Şahintürk, 1997; Kandemir and Yılmaz, 2009). These
128
volcano-sedimentary sequences grade into the Upper Jurassic‒Lower Cretaceous carbonates
129
of the Berdiga Formation (Şengör et al., 2003; Karslı et al., 2010a). The Dağbaşı Skarn ore is
130
hosted by these platform carbonates.
131 132
The Upper Cretaceous was composed of an acidic and basic succession of volcano-
133
sedimentary units, with a thickness of more than 2 km. This volcano-sedimentary sequence
134
was separated into four different formations by Güven et al. (1998) – (from bottom to top)
135
Çatak, Kızılkaya, Çağlayan, and Tirebolu. The Çatak Formation, which was conformably
136
overlain by the platform carbonates of the Berdiga Formation, mainly consisted of andesite,
137
basalt and their pyroclastites, with interlayered sandstone, marl, shale, red biomicrite, and
138
micritic limestone. The Kızılkaya Formation is separated from the Çatak Formation by dacitic
139
and rhydacitic volcanites which contain clayey, sandstone, marl and red-biomicrite
140
intercalations (Güven, 1993). Prismatic columnar dacites are common in this unit, and host 4
141
numerous massive volcanogenic sulfide deposits across the region (Karslı et al., 2011). The
142
Kızılkaya Formation is conformably overlain by the Çağlayan Formation, which consists of
143
basic volcanic rocks and interlayered thin- to medium-bedded sandstones, red micritic
144
limestones, marls, clayey limestones, and tuffs. The Tirebolu Formation, which represents the
145
uppermost part of the Mesozoic sequence, conformably overlies the Çağlayan Formation, and
146
comprises rhyodacite and pyroclastites, with pelagic micritic limestone, sandstone, and clayey
147
intercalations (Güven, 1993; Güven et al., 1998).
148 149
Many different intrusions were emplaced into the northeastern region of Turkey from
150
Early Jurassic to Late Eocene times. These intrusions developed in various geodynamic
151
settings, and have different ages and compositions. The early Jurassic plutons (between
152
176.95±0.49 and 178.41±0.44 Ma) are mainly composed of gabbro and gabbroic diorite, and
153
characterized by relatively low SiO2 (47.09–57.15 wt%), moderate Na2O (1.19–3.92 wt%)
154
and high Mg# (46–75) (Karslı et al., 2017). These gabbroic plutons show metaluminous
155
geochemical character and belong to the slightly evolved I-type signature of continental
156
magmatic arc setting. Early Jurassic (188.0±4.3 Ma) plutons were also reported by Dokuz et
157
al. (2010) to exhibit a low-K, metaluminous to weakly peraluminous (ASI=0.94–1.11)
158
character and granodiorite and, to a lesser extent, tonalite composition. On the contrary, late
159
Jurassic plutons (153+3.4 Ma) consist of mostly quartz monzodiorite composition and show a
160
metaluminous (ASI=0.84–0.99) character with a medium to high K2O, and relatively high
161
MgO content (Dokuz et al., 2010). These early and late Jurassic intrusions were interpreted to
162
be the products of an arc continent collision event, in response to the closure of Paleotethys
163
(Dokuz et al., 2010).
164 165
The composition of the Late Cretaceous plutons varies from low-K tholeiitic through
166
calc-alkaline (rarely high-K) metaluminous and peraluminous leucogranites to alkaline
167
syenites and monzonites (Okay and Şahintürk, 1997; Karslı et al., 2004; Karslı et al., 2007;
168
Topuz et al., 2007; Kaygusuz and Aydınçakır, 2011). The age of these plutons was
169
determined as between 88.1–86.0 million years by Kaygusuz et al. (2009) by the U–Pb
170
SHRIMP method conducted on the zircon minerals. Late Eocene intrusions, on the other
171
hand, are made up of granite‒granodiorite, with quartz‒monzonite‒tonalite porphyry
172
associations, and compositions varying from low-K tholeiitic to high-K calc-alkaline (Moore
173
et al., 1980; Boztuğ et al., 2004; Karslı et al., 2010b). The 40Ar-39Ar ages of hornblende and
174
biotite separates and U–Pb zircon SHRIMP ages from these rocks range between 48 Ma and 5
175
54 Ma (Karslı et al., 2010b; Topuz et al., 2011; Karslı et al., 2013). According to these
176
studies, Early Cretaceous to Late Eocene intrusions were formed by arc-collisional through
177
syncollisional crustal thickening to post-collisional extensional regimes.
178 179 180
3. Host rock geology
181
The basement around the Dağbaşı area is made up of Liassic volcanic rocks. This unit
182
corresponds to the Şenköy Formation, which crops out along the northeastern part of Turkey
183
(Kandemir and Yılmaz, 2009). This unit consists mainly of andesite, basalt, and their
184
pyroclastites and includes interlayered sandstones, marls, and red biomicrite lenses, which are
185
a few meters thick and several hundred meters in length. Based on the presence of
186
foraminifera (Involitina liassica, Trocholina, Lenticulina, Spirillina, Vidalina martana,
187
Lingulina, and members of the Lagenidea) found in the biomicrites, a Liassic age was
188
determined by Güven et al. (1998).
189 190
The andesites and basalts are macroscopically characterized by gas cavities, highly
191
fractured and altered. Brecciated, glassy, microlitic, microlitic porphyritic, amygdaloidal, and
192
void textures were commonly observed in the microscopic scale. This unit consists of
193
pyroxene, plagioclase, amphibole, biotite, augite and quartz. While the primary quartz content
194
does not exceed ~3%, this content increases due to silicification that has occured along the
195
fractures and gas cavities. The silicification is accompanied by epidote, chlorite, and calcite
196
along the contact with the granitoid.
197 198
The SiO2, MgO, and K2O values were reported between 55.74–57.57, 3.24–4.44, and
199
0.07–2.19 wt%, respectively in the andesites, and the same values were between 47.4–51.59,
200
2.62–8.52, and 0.13–1.58 wt%, respectively in basalts (Aydınçakır, 2006). These results
201
correspond to the trachyandesite, andesite, basaltic andesite and basalt composition on the
202
alkaline-silica diagram of Le Maitre et al. (1989). By using various trace element contents,
203
Aydınçakır (2006) reported that the samples show calc-alkaline basalt composition, and
204
volcanic arc related basalt features. The chondrite-normalized REE patterns of the samples
205
also indicated that both andesites and basalts were derived from the same sources (Kaygusuz
206
and Aydınçakır, 2009).
207
6
208
This volcano-sedimentary unit contains block- and lens-shaped massive limestone
209
layers that are up to 100 m thick and extend up to a few kilometers. These massive bedded
210
limestones are Early Cretaceous platform carbonates, identified as Berdiga Formation by
211
numerous researchers (Pelin, 1977; Kırmacı et al., 1996; Yılmaz and Kandemir, 2006;
212
Kırmacı et al., 2018). According to Yılmaz and Kandemir (2006), these platform carbonates
213
accumulated into the rift basin under the tectonically inactive conditions. Yılmaz et al. (2008)
214
also indicated that the Berdiga Formation formed in a shallow carbonate shelf environment.
215
These limestones show limited lateral extensions since they are formed in rift basins, and
216
display different lithofacies changing vertically. The lower parts of the formation consist of
217
thin to thick bedded dolomites, while the middle and upper parts consist of very thick bedded
218
to massive, locally laminated features displaying grainstone‒packstone and skeletal
219
wackestone lithofacies (Kırmacı et al., 2018). The limestones, with thicknesses varying
220
between 10 and 30 cm, are highly recrystallized close to the granitoid contact, and their initial
221
texture is unrecognizable. Crystallized limestones contain garnet, epidote, magnetite,
222
hematite, pyrite, pyrrhotite, and chalcopyrite in abundance throughout the skarn zones, and
223
silicification accompanies these minerals.
224 225
The Dağbaşı Granitoid was emplaced into both the volcano-sedimentary units of the
226
Senköy Formation and the massive banded limestone blocks of the Berdiga Formation (Fig.
227
2). The granitoid exhibits zonation from the central part towards the outer zone, in terms of
228
mineral size, abundance, and composition, with large crystals in the central part and smaller
229
crystals towards the edges. The granitoid is characterized by higher orthoclase abundances at
230
the central part, and these orthoclase abundances decrease towards the edge, while quartz and
231
plagioclase increase. Using the modal mineralogical abundances, the granitoid was
232
subdivided into four different zones by Aydınçakır (2006). In this study, Aydınçakır (2006)
233
indicated that monzogranite composition was dominant in the central parts, while the
234
granitoid became granodiorite, tonalite, and diorite composition towards the outer zone. The
235
granitoid is I-type, with a low to moderate K content, peraluminum and metaluminum
236
transitional, having properties similar to volcanic arc type calc-alkaline granitoids
237
(Aydınçakır, 2006; Demir, 2019). The granitoid contains some mafic microgranular enclaves
238
of surrounding volcanics, centimeters to 1 m in scale, along the contact with the volcanic
239
rocks. The age of pluton emplacement has been determined as between 82.9±1.3–88.1±1.7
240
Ma (Upper Cretaceous; Kaygusuz et al., 2009) using the U–Pb SHRIMP method conducted
241
on zircon minerals. 7
242 243
The youngest unit in the study area is the Uzuntepe Dacite, which consists of mainly
244
quartz megacrystals (up to cm in size) and a lesser amount of plagioclase, biotite, amphibole,
245
and fine-grained matrix. A number of Uzuntepe dacitic intrusions cut the basic volcano-
246
sedimentary units of Şenköy Formation. Their size ranges from a few hundred meters to 3 km.
247
The largest dacite outcrop is around the Osmanoğlu village and Demdemtaş hill, which
248
extends out to 3 km in size (Fig. 2). There was no contact between the Uzuntepe dacite and
249
the Berdiga limestones in the area. However the dacite dykes, which cut the Dağbaşı granitoid
250
in the vicinity of Sariot village, indicate that these dacites are younger than the Dağbaşı
251
granitoid (Aydınçakır and Kaygusuz, 2012).
252 253
4. Skarn mineralogy
254
Skarn type mineralizations have been observed along the border of Dağbaşı granitoid
255
and Berdiga limestones around the İpekçili, Köprüüstü, Kükürtlü and Dere Mahalle districts
256
(Fig. 2). Block and lens shaped Berdiga limestones are located in the host volcanites (Fig. 3).
257
Granitoid/limestone contact is not observable in the field, because the volcanic rocks outcrop
258
in a limited area between these units. All of these skarns developed as exoskarn type in the
259
closest border of the lens- and block-shaped limestones to the granitoid.
260 261
Both prograde and retrograde stage skarn mineral assemblages have been described in
262
the location of İpekçili and Köprüüstü ore. The prograde skarn mineral assemblage was
263
subdivided into early prograde and late prograde stages by Demir (2019). The early prograde
264
stage garnets and pyroxenes were described as having massive (Fig. 4a), rhythmically-banded
265
(Fig. 4b), nodular (Fig. 4c), and granular textures (Fig. 4d), while the late prograde stage
266
garnet and pyroxenes were described as having fracture-filling texture (Fig. 4e) in the
267
limestones. Demir (2019) indicated that these early prograde garnets were grossular
268
dominated, with the composition changing to andradite in the late prograde stage; whereas
269
early stage pyroxenes were hedenbergite in composition, becoming diopside in the late stage.
270
There was no relation between the early stage garnet and pyroxene with the ore minerals, but
271
the late stage garnets accompany magnetite and hematite in their growth zones (Fig. 4f). The
272
retrograde stage mineral assemblage in the İpekçili and Köprüüstü locations was represented
273
by hydrous silicate minerals, such as epidote, chlorite, and tremolite‒actinolite, accompanying
274
quartz and calcites minerals. Quartz, epidote, and calcites were formed along the fractures of
275
the volcanic host rocks (Fig. 4g) and limestones, while tremolite‒actinolite minerals were 8
276
commonly observed as massive textures (Fig. 4h). Retrograde stage tremolite and actinolite in
277
the Köprüüstü locations were also found to contain both macroscopic (Fig. 4i) and
278
microscopic (Fig. 4j) scale garnet and pyroxene inclusions.
279 280
Prograde stage minerals assemblages of garnet and pyroxenes were not determined in
281
the Kükürtlü and Dere Mahalle locations. In these locations the retrograde stage is
282
characterized by fracture-filling type (Fig. 4k) textures of epidote, quartz and calcite minerals
283
in the volcanic host rocks and limestones. In some cases, quartz inclusions, a few cm in size,
284
were also present in both the volcanic host rock and limestones (Fig. 4l).
285 286
In addition to magnetite and hematite, sulfide mineral assemblages were also
287
determined in the İpekçili and Köprüüstü locations, whereas the Kükürtlü and Dere Mahalle
288
locations were characterized by magnetite and hematite assemblage. Ore in the İpekçili and
289
Köprüüstü locations was characterized by irregularly shaped ore masses (Fig. 5a), and banded
290
ore textures along the weak zone of the limestone layers (Fig. 5b). Fracture-filling (Fig. 5c),
291
and breccia-filling (Fig. 5d) ore types were also common in the limestones and volcanic host
292
rocks. Taking into consideration the magnetite and hematite developments in the growth
293
zones of the late prograde stage garnets, Demir (2019) surmised that ore mineralization began
294
towards the end of the late prograde stage (Fig. 4f). However, the presence of these garnets as
295
inclusions in the magnetites and hematites indicates that the main ore development took place
296
after the development of garnets (Fig. 5e). Additionally, the presence of epidote, quartz, and
297
calcite accompanying the ore minerals (Fig. 5f) indicates that the main ore mineralization
298
formed during the retrograde skarn stage.
299 300
Magnetites and hematites also accompany sulfide ore minerals. In the polished
301
sections hematites were observed to be the filling between magnetite minerals (Fig. 5g).
302
Therefore, it is understood that magnetites were formed before hematites. These magnetites
303
and hematites are always surrounded by the sulfide minerals (Fig. 5h), which represent the
304
formation of oxide before sulfides. Chalcopyrite, and sphalerite formation along the pyrite
305
cracks also indicates the formation of pyrite before chalcopyrite minerals (Fig. 5i). Apart from
306
the presence of magnetite and hematite, there is some difference in terms of the sulfide
307
minerals between the İpekçili and Köprüüstü locations. Pyrrhotite is only present in the
308
İpekçili skarn zone, and accompany pyrite, chalcopyrite and sphalerites. This pyrrhotite, in
309
the İpekçili skarn, is characterized by a bird’s eye texture (Fig. 5j) and has been replaced by 9
310
late stage sphalerite. On the other hand, galena minerals, which were rarely observed in the
311
form of small inclusions in sphalerite (Fig. 5k), accompany pyrite, chalcopyrite, and
312
sphalerite in the Köprüüstü skarn ore.
313 314
The Kükürtlü and Dere Mahalle locations were characterized by magnetite and
315
hematite masses reaching up to a meter in size (Fig. 5l), and veins formed along the fractures
316
of the limestone and volcanic host rocks. In addition to the fracture-filling type veins, breccia-
317
filling type textures were also present. In this texture, breccias of magnetite and hematites
318
were filled up by later quartz and epidote minerals (Fig. 5m). In some cases, replacement
319
remnants of magnetite and hematite were also observed in the epidote- and quartz-bearing
320
samples. Disseminated magnetites and hematites were also observed in both macroscopic and
321
microscopic scale (Fig. 5n).
322 323
The presence of magnetite and hematites in the outer growth zone of late stage garnets
324
indicates that ore mineralization in the Dağbaşı area began at the end of the late prograde
325
stage. But the presence of garnet inclusions in the magnetite and hematite ores indicates that
326
the main ore mineralization was formed later than the prograde stage. The coexistence of
327
magnetite and hematites with the epidote-bearing samples and their contemporaneous
328
development (Fig. 5f) is consistent with the retrograde stage ore formation in the Dağbaşı
329
area. Ore mineralization in the Kükürtlü and Dere Mahalle district supports the retrograde
330
stage ore development because there was no evidence of prograde stage garnet and pyroxene
331
assemblage.
332 333
Magnetite and hematites were the main ore minerals in both the Kükürtlü and Dere
334
Mahalle locations. Limited amounts of pyrite and chalcopyrite minerals, a few microns in
335
size, were observed in these minerals (Fig. 5o). Therefore, the abundance of the sulfide
336
minerals in Kükürtlü and Dere Mahalle locations was much lower than in the İpekçili and
337
Köprüüstü ores. Similar findings were also reported by Demir (2019), and the generalized
338
mineral paragenesis and succession in the Dağbaşı Skarns are presented as shown in Figure 6.
339 340
5. Analytical techniques
341
Microthermometric measurements were performed on the approximately 200 µm-thick
342
double-polished sections using a Linkam THMG-600 heating‒freezing stage at the
343
Department of Geological Engineering of Recep Tayyip Erdoğan University, Turkey. This 10
344
equipment is suitable for temperature measurements between -180 and 600 °C. Liquid
345
nitrogen was used for cooling and freezing. The stage was calibrated using H2O– and H2O–
346
CO2 containing synthetic fluid inclusions (Sterner and Bodnar, 1984). The repeated
347
experiments indicate that the measurements were performed with an accuracy of ±0.2 °C for
348
the freezing, and ±0.5 °C for the heating processes. Salinity data, expressed as wt% NaCl
349
equ., was calculated from the melting temperature of the last ice crystal (Tm-ice) using the
350
equation reported by Bodnar and Vityk (1994). In this technique, the salinity is given as wt%
351
NaCl equ., because this calculation has an error margin of less than 5 wt%, even if there were
352
any other salt types. Isochores for the pressure estimations were also computed using the
353
equations reported by Bodnar and Vityk (1994). A total of 59 double-polished wafers were
354
prepared from garnet, epidote, quartz, and calcite samples. Prior to the microthermometric
355
measurements, fluid inclusions were petrographically investigated using the criteria of
356
Roedder (1984). The measurement process was performed according to Shepherd et al.
357
(1985).
358 359
Raman spectroscopic studies were carried out at the WITec GmbH (Ulm, Germany).
360
A 514 nm (green) laser mounted on the Alpha 300 confocal Raman spectrometer was used for
361
the measurements. The entire area of each of the fluid inclusions was scanned, using the 1 µm
362
intervals of the laser beam. Data was collected over a spectral range of 100 to 4000 cm-1.
363 364
A total of 22 samples from the skarn calcites were subjected to δ13C and δ18O isotope
365
analyses at the Cornell Stable Isotope Laboratory, USA, using a Thermo Scientific Delta V
366
Advantage mass spectrometer coupled to a Thermo Scientific Gas Bench II. A sample of
367
approximately 0.5–1 mg of calcite was reacted with phosphoric acid at 25 °C for 18 hours to
368
produce CO2. The produced CO2 gas was transferred into the isotope ratio mass spectrometer
369
and analyzed for its 13C/12C, and 18O/16O ratios. The laboratory stated that the precision of the
370
analyses was less than 0.20‰ in this method. The δ13C and δ18O isotope results were reported
371
relative to the Pee Dee Belemnite (PDB) standard. The results of the PDB standard were
372
converted to the Standard Mean Ocean Water (V–SMOW) standard using the empirical
373
relationship (Δ18OSMOW =1.03086*Δ18OPDB+30.86) of Friedman and O’Neil (1977).
374 375
The δ18O and δD analyses were carried out on the selected garnet, pyroxene, epidote,
376
quartz and tremolite samples. Before analysis, each sample was crushed to approximately 1
377
mm in size, and pure minerals were hand-picked using a binocular microscope. The carbonate 11
378
was removed with a treatment of 1 N HCl solution. The samples were cleaned with pure
379
water, and the purity was approved by X-ray diffraction.
380 381
The isotope analyses were performed at the Scottish Universities Environmental
382
Research Centre. In this laboratory, oxygen is extracted from a 1–2 mg sample by the CIF3
383
fluorination technique. The extracted oxygen was converted to CO2 with a reaction by hot
384
graphite rod. The isotope composition of the evolved CO2 was measured by a VG PRISM III
385
isotope ratio spectrometer. The results were reported relative to the V–SMOW standard. The
386
precision was ±0.2‰ (1σ) based on repeated analyses.
387 388
6. Fluid inclusion studies
389
Fluid inclusion studies were conducted on garnet, epidote, quartz, sphalerite, and
390
calcite minerals to determine their formation temperatures and the conditions under which
391
mineralization
392
microthermometric measurements, were commonly observed in quartz, calcite, and sphalerite
393
minerals, whereas fluid inclusions in garnet and epidote minerals were limited. Prior to the
394
freezing‒heating experiments, a genetic classification of the fluid inclusions in each crystal
395
was performed, according to the criteria suggested by Roedder (1984) and Shepherd et al.
396
(1985). Fluid inclusions randomly distributed in the host crystal or aligned parallel to the
397
grain boundaries are considered as primary, and inclusion trails occurring as planar groups
398
and askew to crystal boundaries are regarded as of secondary origin. Microthermometric
399
measurements were carried out on inclusions with a primary origin so as to avoid potential
400
post-mineralization effects in the secondary inclusions.
occurred
in
the
Dağbaşı
Skarns.
Fluid
inclusions,
suitable
for
401 402
Liquid-, solid-, and vapor-bearing (L+S+V) three-phase, and liquid- and vapor-bearing
403
(L+V) two-phase fluid inclusions were identified in the researched minerals. The fluid
404
inclusions were commonly polygonal and irregular in shape, while lesser amounts of rounded,
405
elliptical, tube- and pear-shaped morphologies were also present. In some cases, tabular-
406
shaped inclusions were also observed in calcite minerals. The size of the investigated fluid
407
inclusions generally varied between 10 and 60 µm. Inclusions larger than 30 µm in size were
408
rarely observed in the investigated minerals. None of the fluid inclusions contain separated
409
liquid CO2 phase at room temperature.
410
12
411
Although there are four different skarn locations around the Dağbaşı granitoid, no
412
decisive diversity was observed between these locations, in terms of the inclusion shape, size,
413
or microthermometric results. The results are presented with consideration to the inclusion
414
type and vapor/liquid ratio in each mineral phase, since the main differences vary according to
415
these properties. In terms of the number and volumetric proportions of the phases present at
416
room temperature, and the homogenization type of the fluid inclusions (to a liquid or vapor
417
phase), three major types of fluid inclusions were recognized from the Dağbaşı skarn ore. The
418
detailed properties of each type of inclusion are described below.
419 420
6.1.
Classification
421
Type I fluid inclusions are three-phase (L+S+V) inclusions. This type of inclusion was
422
identified in the early prograde stage garnet (Fig. 7a), quartz (Fig. 7b), and calcite (Fig. 7c)
423
minerals. Two-phase fluid inclusions (L+V), with vapor ratios higher than liquid phase
424
(V>L), were classified as Type II. Homogenization always occur to vapor phase in this type
425
of inclusions. These were identified only in late prograde stage garnet (Fig. 7d) and quartz
426
(Fig. 7e) minerals. Type III inclusions are two-phase (L+V), with vapor ratios always lower
427
than unity. These inclusions comprised three subgroups. Type IIIa have vapor bubbles that
428
occupy 30–40 vol% of the inclusions. Homogenization to the liquid phase occurs in these
429
types of inclusions. Type IIIa fluid inclusions were identified in the late prograde stage of
430
garnet and quartz minerals and were found to accompany Type II inclusions, even in the same
431
area of host crystals (Fig. 7f, g). Therefore, both Type II and Type IIIa inclusions were
432
grouped as a fluid inclusion assemblage. Type IIIa inclusions, in garnet and quartz minerals,
433
homogenized to liquid phase, whereas the Type II inclusions homogenized to the vapor phase
434
in the same minerals.
435 436
Type IIIb inclusions were observed in the retrograde stage epidote, calcite, sphalerite,
437
and quartz minerals (Fig. 7h‒j). Type IIIa and Type IIIb inclusions have similar shape, size
438
and liquid/vapor ratios. However, there was no coexistence of Type IIIb and Type II fluid
439
inclusions. Type IIIb inclusions were also characterized by their lower Th and salinity values
440
compared to the Type IIIa inclusions. In some cases, Type IIIb inclusions were observed to
441
participate in Type IIIc inclusions in the quartz minerals (Fig. 7k). Type IIIc inclusions are
442
characterized by a volumetrically smaller vapor/liquid ratio than in Type IIIa and Type IIIb
443
(Fig. 7k, l) that accounts for ~5–20 vol% of the inclusions, and this type of inclusion was
13
444
observed in all the investigated minerals of retrograde stage. Type IIIc fluid inclusions are
445
characterized by lower homogenization temperature (Th) and salinity values.
446 447
6.2.
Microthermometric results from the fluid inclusions
448
6.2.1. Eutectic temperatures
449
The eutectic temperatures (Te) of different types of inclusions in all the investigated
450
minerals were measured, and the results are presented in Table 1. The Te of the three-phase
451
Type I inclusions in the garnet, quartz, and calcite minerals was between -53.2 °C and -46.6
452
°C. For the gas-rich Type II and liquid-rich Type IIIa inclusions in late prograde stage garnet,
453
and quartz, Te temperatures were in the range of -57.4 °C to -34.6 °C, and -57.2 °C to -32.4
454
°C, respectively. Similar Te temperatures were measured in the fluid-rich Type IIIb inclusions
455
between -58.6 °C to -29.0 °C in the quartz, epidote and calcite minerals (Fig. 8).
456 457
When the Te of the Type I inclusions are compared to the eutectic temperatures of
458
various water–salt systems, this temperature range corresponds to the H2O–CaCl2 dominated
459
fluids. On the contrary, Te of the Type II, Type IIIa, and Type IIIb inclusions corresponded to
460
mixtures of different salt combinations rather than specific salt solutions. For instance, Te
461
temperatures close to -49.8 °C corresponded to the H2O–CaCl2 dominated solutions, while the
462
Te temperature range of -32.4 °C to -35 °C corresponded to the H2O–FeCl2–MgCl2 system.
463
Moreover, eutectic temperatures close to -56.6 °C corresponded to the melting temperature of
464
CO2 phase, rather than specific water–salt combinations. Although there were no separated
465
gas phases and clathrate occurrences in the fluid inclusions at room temperature, the values
466
approximating -56.6 °C indicated that a limited amount of dissolved CO2 may be present in
467
the composition of the gas- and fluid-rich inclusions.
468 469
The Te temperatures of the Type IIIc inclusions were measured between -35.2 °C and
470
-20.6 °C. Most of the composition of these inclusions, ranging between -28 °C and -20.6 °C,
471
was close to the eutectic temperatures of the H2O–NaCl system. Conversely, the composition
472
of a very few Type IIIc inclusions corresponded to the mixture of a H2O–FeCl2–MgCl2
473
system, with a Te range of -29 °C to -35.2 °C (Fig. 8).
474 475
6.2.2. Last ice melting, and halite melting temperatures and salinity values
476
The ice melting temperature (Tm-ice), which was used to calculate the salinity values
477
of the inclusions, was measured in the two-phase (L+V) Type II, Type IIIa, Type IIIb, and 14
478
Type IIIc inclusions. The halite melting temperature (Tm-halite) was measured in halite
479
bearing Type I inclusions to calculate salinity values. The Tm-ice and Tm-halite values,
480
measured from the different inclusion types, are presented in Table 1.
481 482
The halite melting temperatures of the brine (Type I) fluid inclusions were measured
483
in the range of 412 °C to 514 °C, and the salinity values of these inclusions was calculated
484
between 48.8 and 61.8 wt% NaCl equ. according to Bodnar and Vitky (1994) (Fig. 9). Tm-ice
485
temperatures of the Type II fluid inclusions were measured between -6.8 °C and -9.6 °C in
486
garnet, and -5.3 °C and -10.5 °C in quartz minerals. Accordingly, the salinity of these
487
inclusions was calculated between 10.3 and 13.7 wt% NaCl equ. in garnet, and 8.3 and 14.7
488
wt% NaCl equ. in quartz minerals. The salinity values in the Type II and Type IIIa fluid
489
inclusions in the garnet and quartz minerals were similar. But the salinity values of Type IIIb
490
inclusions in quartz, calcite, and epidote minerals were distinctively lower than the Type II
491
inclusions, with calculated values in the range of 4.8–10.1, 2.9–6.9, and 4.1–9.8 wt% NaCl
492
equ., respectively. The Type IIIc inclusions, observed in the retrograde stage of all
493
investigated minerals, had lower salinity values, in the range of 0.9–6.2, 0.5–4.8, and 1.1–5.1
494
wt% NaCl equ. in epidote, calcite, and quartz minerals, respectively. The salinity values of the
495
Type IIIb and Type IIIc inclusions in sphalerite minerals were very close to each other,
496
ranging between 0.7 and 5.1 wt% NaCl equ. (Fig. 9).
497 498
Taking into consideration the measured Tm-ice and Tm-halite temperatures, the
499
highest salinity values occurred in the Type I inclusions (61.8 wt% NaCl equ.), whereas the
500
lowest values were in the Type IIIc inclusions (0.3 wt% NaCl equ.). The variations in Figure
501
9 indicate a clearly decreasing trend between salinity and Th for all measured inclusions.
502 503
6.2.3. Homogenization temperature measurements
504
The homogenization temperatures (Th) of different types of fluid inclusions are given
505
in Table 1. Two different homogenization styles were observed in halite-bearing Type I
506
inclusions – melting temperatures of halite crystals (Tm-halite) and homogenization to liquid
507
phase by vapor disappearance (Th). Tm-halite was always higher than Th in these inclusions.
508
Homogenization of the vapor into the liquid phase occurred at between 352.5 °C and 368.6 °C
509
in garnet, 357 °C and 402 °C in quartz, and 348 °C and 446 °C in calcite minerals, but the
510
Tm-halite in these minerals ranged between 412 °C and 418 °C in garnet, 438 °C and 461 °C
511
in quartz, and 492 °C and 514 °C in calcite minerals (Table 1). 15
512 513
The homogenization temperatures of the Type II fluid inclusions, that homogenized to
514
vapor phase, were in the range of 355 °C–449 °C in garnet and 355.2 °C–452 °C in quartz
515
minerals. The Th of Type IIIa inclusions, which coexist with Type II inclusions in both
516
minerals, are in the range of 353 °C–458 °C in garnet, and 368 °C–451.2 °C in quartz. The
517
coexistence of both types of inclusions in the same fluid inclusion assemblage, with their
518
similar Th and salinity values, clearly indicates a boiling assemblage for both types (Bodnar et
519
al., 1985). The Th of Type IIIb inclusions in quartz, epidote, and calcite, which have no
520
coexistence with Type II inclusions, have a slightly lower Th range of 318 °C–412 °C.
521
The homogenization temperatures of Type IIIc inclusions in quartz, calcite, and
522
epidote minerals were in the range of 160 °C–327 °C, whilst Th of the Type IIIc inclusions in
523
all these minerals were much lower than the other fluid inclusion types. The homogenization
524
temperatures of the Type IIIb and Type IIIc inclusions in sphalerites were also close, between
525
231 °C and 327 °C. Although these two types of fluid inclusions occur in sphalerite, the Th
526
histogram of both types of fluid inclusions were represented by a unimodal distribution (Fig.
527
10) due to the close Th ranges of both types of inclusions.
528 529
6.3.
Pressure estimates
530
Estimates of the trapping pressure of the halite-bearing Type I fluid inclusions, which
531
homogenized by halite dissolution, can be calculated as a function of the liquid‒vapor Th and
532
Tm-halite, according to Sanchez et al. (2012). In this technique, after the calculation of the
533
salinity value of halite-bearing fluid inclusions, the halite liquidus line is drawn for the
534
calculated salinity value. The Th isochore for the measured Th temperature is also drawn by
535
the calculated salinity. The intersection of Th isochore with the halite liquidus gives the
536
minimum trapping pressure of each fluid inclusion. Based on this technique, an 884 bar
537
trapping pressure was calculated from the intersection of the halite liquid line, drawn from the
538
highest Tm-halite (514 °C) and Th isochore (446 °C) (Fig. 11). Accordingly, a 710 bar
539
trapping pressure was calculated from the intersection of the halite liquid line, from the lowest
540
Tm-halite (412 °C) and Th isochore (352 °C).
541 542
The presence of both vapor- and fluid-rich inclusions in the same FIA, the
543
homogenization of vapor-rich inclusions to vapor phase, and fluid-rich inclusions to liquid
544
phase, and their similar Th and salinity ranges (Table 1), are accepted as evidence of boiling
545
(Wilkinson, 2001; Bodnar, 2003). If boiling occurred, the homogenization of Type II and 16
546
Type IIIa inclusions would show on the two-phase critical curve (Fig. 11) according to Baker
547
and Lang (2003) and Sanchez et al. (2012). In this case, the measured pressure is equal to the
548
actual pressure, and no pressure correction was required (Roedder and Bodnar, 1980;
549
Wilkinson, 2001; Sanchez et al., 2012). Therefore, the minimum and maximum Th
550
temperatures of both the Type II and Type IIIa fluid inclusions, between 353 °C and 458 °C,
551
corresponded to the trapping pressure between 195 and 445 bar on the two-phase critical
552
curve (Fig. 11), respectively.
553 554
Type IIIb inclusions are characterized by moderate Th (318–412 °C, average 362 °C)
555
and salinity values (2.9–10.1 wt% NaCl equ., average 6.1 wt% NaCl equ.). The trapping
556
pressure for this type of inclusion can only be obtained where independent temperatures are
557
available. However, there were no independent temperatures for the calculation of the
558
trapping pressure of this kind of inclusion. Instead, the trapping pressure of Type IIIc fluid
559
inclusions, which were characterized by lower Th (160 °C–327 °C, average 240 °C) and
560
salinity values (0.5–6.2 wt% NaCl equ., average 3.2 wt% NaCl equ.), were calculated by the
561
independent temperatures. In this technique, the trapping pressure of the fluid inclusions was
562
defined by the intersection of independent temperatures with the fluid inclusion isochore,
563
which were drawn by using the average salinity and Th of fluid inclusions (Roedder and
564
Bodnar, 1980; Brown and Lamb, 1989). In this study, independent temperatures were
565
calculated between 135 °C and 281 °C from the retrograde stage quartz‒epidote, and
566
quartz‒tremolite mineral pairs. The highest trapping pressure of 310 bar was determined by
567
the intersection of the P–T isochron with the highest independent geothermometer of 281 °C.
568 569
6.4.
Raman studies
570
None of the fluid inclusions were found to contain separated gas phases, and clathrate
571
occurrences were also not observed during the microthermometric study. Although this
572
suggests that CO2 was not present in the composition of the fluid inclusions, the Te
573
temperatures of some fluid inclusions were very close to the melting temperature of CO2,
574
indicating that these inclusions may contain some dissolved CO2. Hedenquist and Henley
575
(1985) and Rosso and Bodnar (1995) indicated that gas species may not be detected by
576
microthermometric study if the bulk density is lower than 2 mol.%. Therefore, Raman
577
spectroscopy measurements were performed on representative fluid inclusions to confirm
578
whether there was evidence of separated CH4, CO2, or any other volatiles. No significant
579
spectra were obtained from the fluid inclusions in sphalerite, due to the absorption of the 514 17
580
nm laser wavelength, or in calcite, due to the fluorescence effect; however, meaningful
581
Raman measurements were achieved on the fluid inclusions of quartz, epidote, and garnet
582
minerals. The measurements were performed by the whole-area scanning technique because
583
there were no separated gas phases in the fluid inclusions.
584 585
According to the Raman measurements, the majority of the fluid inclusions had an
586
H2O composition that contained various dissolved salts, based on the 3100–3650 cm-1 band
587
range of the spectrum. It was not possible to determine the dissolved salt species in the fluid
588
inclusions, because the dissolved salt species in the solutions did not have characteristic peaks
589
in the Raman spectrum. They only affected the distribution range or intensity of the Raman
590
signals (Burke, 2001; Bakker, 2004; Baumgartner and Bakker, 2009; Frezzotti et al., 2012).
591
Despite that, Type IIIa inclusions in garnet (Fig. 12a), epidote (Fig. 12b), and quartz (Fig.
592
12c, d) were found to rarely contain CH4 according to the characteristic CH4 peak of spectra
593
at around 2917 cm-1. It has been confirmed from other studies that the given peak values
594
correspond to CH4 species in the fluid inclusions (Burke, 2001; Bakker, 2004; Lin et al.,
595
2007; Lin and Bodnar, 2010). These CH4 species were only detected in three of 18 measured
596
samples.
597 598
7. Carbon and oxygen isotope studies
599
In this study, δ13C and δ18O isotope analyses were performed on the prograde and
600
retrograde stage skarn calcites from the Köprüüstü, İpekçili, and Dere districts. These results
601
are listed in Table 2. The δ13C isotopic values of the prograde stage skarn calcites were
602
measured between -1.93 and 2.88‰, while retrograde stage skarn calcites were between -0.81
603
and 1.95‰ (Fig. 13a). The δ18O isotopic compositions of the prograde and retrograde stage
604
skarn calcites were measured between 9.02–16.18‰, and 7.33–9.72‰, respectively.
605 606
The δ13C isotope values of both prograde and retrograde stage skarn calcites, in the
607
range of -1.93 and 2.88‰, overlap both magmatic (0 to -10‰) and marine fields (-3 to 3‰,
608
Clarke and Fritz, 1997; Bowman, 1998; Hoefs, 2009). On the other hand, the δ18O isotope
609
compositions of these calcites range between the magmatic rocks and marine carbonates (Fig.
610
13a). The range of these values in Fig. 13a indicates that skarn calcites are considerably
611
depleted compared to the average δ18O and δ13C isotope compositions of the Berdiga
612
limestones (28.4‰SMOW and 0.6‰PDB, respectively, Kırmacı et al., 2018).
613 18
614
There is a depletion trend in the variations in the δ13C and δ18O isotopes of skarn
615
calcites, which follows a typical fluid‒rock interaction pattern (Bowman et al., 1985; Shin and
616
Lee, 2003). Therefore, the effect of fluid‒rock interaction on the decreasing δ13C and δ18O
617
isotopes was evaluated with the model of Taylor (1976). In this model the fluid is assumed to
618
be isotopically equilibrated with the rock in the closed system. The isotopic fractionation
619
factor was calculated according to the mineral‒water equation of Zheng (1993a). The initial
620
δ18O and δ13C isotope compositions of Berdiga limestones reported by Kırmacı et al. (2012)
621
were used for the calculation of the final δ18O and δ13C isotopes of skarn calcites. The M–A
622
mixing curves were calculated for varying X(CO2) and fluid rock ratios of 0.05, 0.2, 0.5 and
623
0.5, 2, 10, 40, 80, 100, respectively (Fig. 13b). The highest Th-halite temperature of halite-
624
bearing skarn calcites (514 °C) was used for the calculations.
625 626
The δ13C and δ18O isotopes measured from the prograde and retrograde stage skarn
627
calcites of Dağbaşı deposits rather follow the depletion curves of 0.05 and 0.1 X(CO2). The
628
isotopic depletion is shown in both oxygen and carbon isotopes. The fluid‒rock ratios of
629
prograde stage skarn calcites are highly variable and they range between 0.5 and 80. On the
630
contrary, fluid‒rock ratios of retrograde stage skarn calcites are in the range of 10–40.
631 632
8. Oxygen and hydrogen isotope studies
633
In order to clarify the metasomatic processes between the Dağbaşı granitoid and the
634
limestones, representative garnet, pyroxene, quartz, epidote, and tremolite minerals, from the
635
prograde and retrograde skarn stages of Köprüüstü and İpekçili, were analyzed for δ18O
636
isotopes. The results are given in Table 3. In addition, δ18O analysis of quartz and plagioclase
637
from the central part of the granitoid that was not affected by the skarn zones was performed
638
in order to compare with the skarn zones.
639 640
The δ18O isotope results of the quartz and plagioclase minerals from the central part of
641
the Dağbaşı Granitoid ranged from 11.8 to 12.4‰. These values are compatible with the
642
isotopic composition of magmatic waters given by published studies (Taylor, 1979; Hoefs,
643
2009). When the δ18O values of the skarn minerals are considered, the results of the prograde
644
skarn stage of quartz, pyroxene, and garnet range between 8.2 and 10.3‰, 1.9 and 5.3‰, and
645
3.0 and 4.8‰, respectively. The δ18O values of the retrograde stage were between 5.8–7.2‰
646
for quartz, 0.5–0.6‰ for tremolite, and -1.9 and -0.6‰ for epidote (Table 3). These values
647
indicate that the prograde stage skarn minerals were depleted when compared to the center of 19
648
the granitoid, and retrograde stage skarn minerals were also further depleted in comparison to
649
those at the center of the granitoid and the prograde stage skarn minerals. Therefore, the δ18O
650
isotope values show a gradually decreasing trend from the center of the granitoid to the
651
retrograde skarn stage (Fig. 14).
652 653
The δD isotope analyses were carried out on the retrograde stage epidote and tremolite
654
minerals from İpekçili and Köprüüstü in order to determine the origins of the hydrothermal
655
solutions that were effective in the formation of the skarns in the Dağbaşı area. The results are
656
given in Table 3. According to this, the δD isotope values for the epidote ranged between -52
657
and -84‰, and tremolite minerals between -49 and -63‰. The variation in the δD isotope
658
results are similar to those of magmatic water, but the variation in δ18O isotopes measured
659
from these minerals was highly depleted when compared to magmatic sources. This data,
660
which provides a distribution between the magmatic and the meteoric water line in Fig. 15,
661
can be explained by dilution with meteoric water, as found in various studies (Craig, 1961;
662
Taylor, 1974; Sheppard, 1984; Hoefs, 2009).
663 664
8.1.
Isotopic equilibration and thermometer
665
In this present study, the equilibrium temperatures of quartz‒plagioclase,
666
quartz‒pyroxene, quartz‒garnet, garnet‒pyroxene, quartz‒epidote and quartz‒tremolite
667
mineral pairs were calculated, and the results are given in Table 3. Fluid inclusion results and
668
calculated temperatures which were reported in previous studies (Kaygusuz and Aydınçakır,
669
2011) were used as independent thermometers to provide the isotopic composition of the
670
equilibrated solutions.
671 672
The equilibrium temperature of quartz and plagioclase, measured from the center of
673
the granitoid, was 998.3 °C, using the equilibrium equations of Zheng (1993a). By using the
674
mineral‒water equilibrium equations for quartz‒water and plagioclase‒water, the δ18O
675
composition of the solutions in equilibrium with quartz and plagioclase were calculated to be
676
10.11‰ and 11.44‰, respectively. The average temperature (827 °C) given by Kaygusuz and
677
Aydınçakır (2011) was used as an independent thermometer for the calculation of the isotopic
678
composition of the solutions. Both the measured isotopic compositions of the minerals
679
(between 11.8 and 12.4‰) and the calculated isotope compositions of the equilibrated
680
solutions (10.11 and 11.44‰) are characteristically distributed in the range of primary
681
magmatic water (Hoefs, 2009). 20
682 683
The δ18O isotope composition of the hydrothermal solutions, equilibrated with the
684
prograde stage garnet, pyroxene, and quartz, was calculated to be between 3.68 and 7.73‰ for
685
the average Th temperatures (456 °C) of the Type I fluid inclusions (equation of Zheng,
686
1993a). The δ18O isotope composition of the retrograde stage solutions, equilibrated with the
687
epidote, tremolite and quartz minerals, was also calculated between -0.68 and 2.53‰ for the
688
average Th temperatures (~321 °C) of the Type IIIa and Type IIIb fluid inclusions in all
689
investigated minerals, according to the equation provided by Zheng (1993b).
690 691
The equilibrium temperatures calculated between the quartz‒pyroxene, quartz‒garnet
692
and garnet‒pyroxene mineral pairs of the prograde stage from both Köprüüstü and İpekçili
693
skarns ranged from 374.9 to 459.2 °C (Table 3), with the exception of a single lower
694
equilibrium temperature of 163.8 °C. Additionally, the equilibrium temperatures between
695
quartz‒epidote, and quartz‒tremolite mineral pairs, representing the retrograde stage at the
696
same locations, were calculated at between 135.8 and 281.5 °C. These calculated equilibrium
697
temperatures are very close to the Th range of fluid inclusions measured from retrograde stage
698
quartz and epidotes.
699 700
9. Discussion
701
9.1.
702
Skarn type deposits typically occur along the contact between igneous intrusions and
703
carbonate host rocks. Therefore a variety of metasomatic processes involving fluids of
704
magmatic, metamorphic, meteoric, and marine origin were suggested for the skarn formations
705
(Einaudi and Burt, 1982; Kwak, 1986; Baker and Lang, 2003; Meinert et al., 2005). In these
706
studies, high salinity fluid inclusions (>50 wt% NaCl equ.) were generally linked to magmatic
707
sources. Moreover, Wilkinson (2001) indicated that these salinity values reach up to 70 wt%
708
NaCl equ. for the skarn type deposits. Therefore, the high salinity Type I fluid inclusions (in
709
the range of 48.8 to 61.8 wt% NaCl equ.), measured from early prograde stage garnet, quartz
710
and calcites, may refer to magmatic sources.
Origin of the hydrothermal fluids
711 712
However, it is necessary to evaluate the fluid inclusion results together with the
713
isotopic data in order to determine the source of hydrothermal solutions. The δ18O isotope
714
compositions of the prograde stage garnet, pyroxene, and accompanying quartz were
715
measured at between 1.9 and 10.3‰. The isotopic compositions of the hydrothermal solutions 21
716
equilibrated with these early stage minerals were calculated at between 3.68 and 7.73‰
717
according to the mineral‒water equation of Zheng (1993a), based on the independent
718
thermometer of 456 °C, which is the average Ths of the prograde stage (Type I) fluid
719
inclusions (Table 1). Both the δ18O isotope compositions of the skarn minerals and the
720
equilibrated hydrothermal solutions are predominantly consistent with magmatic sources
721
(Taylor and Shepherd, 1986). This is because δ18O isotopes of the garnet, pyroxene and
722
accompanying quartz in the range of 4–9‰ are accepted as an indication of their derivation
723
from magmatic sources (Bowman, 1998; Meinert et al., 2005).
724 725
A magmatic origin for the early stage skarn development is further supported by the
726
δ13C and δ18O isotope ratios of the skarn calcites. This is because the range of the δ18O
727
isotopes for the skarn calcites between magmatic rocks and marine carbonates suggests
728
magmatic originated hydrothermal solutions (O’Neil, 1977; Sheppard, 1984; Hoefs, 2009,
729
Fig. 13a). Besides, both the δ18O and δ13C isotopes show an isotopic depletion trend, which is
730
very consistent with the water‒rock interaction model (Taylor and O’Neil, 1977; Bowman et
731
al., 1985; Holness, 1997; Bowman, 1998; Shin and Lee, 2003). On this model, higher
732
fluid‒rock ratios of the prograde stage calcites (up to 80; Fig. 13b) were associated with
733
magmatic solutions in similar studies (Bowman, 1998; Shin and Lee, 2003; Vallance et al.,
734
2009; Oyman, 2010; Orhan et al., 2011). In these studies, on the other hand, fluid‒rock ratios
735
of retrograde stage skarn calcites at around 10–40 (Fig. 13b) were attributed to a mixture of
736
magmatic and meteoric solutions.
737 738
The lower salinity values of the retrograde stage Type IIIc fluid inclusions and their
739
decreasing salinity trend versus Th are clearly associated with the mixture of magmatic and
740
meteoric solutions (Fig. 9). The δ18O isotope compositions of epidote, tremolite, and
741
accompanying quartz which belonged to the retrograde stage were extensively depleted,
742
varying between -1.9 and 7.2‰, when compared to the magmatic-dominated prograde stage.
743
Accordingly, water in equilibrium with these late stage skarn minerals at temperatures of 321
744
°C has δ18O isotope values between -0.68 and 2.53‰ (Table 3), which is consistent with
745
predominantly meteoric solutions (Taylor, 1979).
746 747
9.2.
Formation conditions
748
The Ths of the prograde stage fluid inclusions in the garnet, quartz, and calcite
749
minerals were in the range of 412 and 514 °C (Table 1). The equilibrium temperatures of the 22
750
garnet–pyroxene, garnet–quartz, and pyroxene–quartz mineral pairs were also in the range of
751
375–451 °C, 399–459 °C, and 388 °C, respectively. Both the equilibrium temperatures of
752
stable isotope data and the Ths from the fluid inclusions range in close proximity to the
753
prograde skarn stage. The trapping pressure of this stage was calculated to be between 710
754
and 884 bar, based on the minimum and maximum salinity values of 48.8 wt% NaCl equ. and
755
61.8 wt% NaCl equ., respectively (Fig. 11).
756 757
The presence of brine-rich (Type I) and vapor-rich (Type II) fluid inclusions may
758
suggest immiscible entrapment; however, there is some evidence against immiscible fluids.
759
This evidence includes: 1) there are none coexisting within the same fluid inclusion
760
assemblage; 2) the Ths for Type I and Type II inclusion types differ greatly; and 3)
761
homogenization of the brine inclusions occurred via halite disappearance (Roedder and
762
Bodnar, 1980; Shepherd et al., 1985). In any case, coexisting vapor-rich inclusions with the
763
halite-bearing inclusions, and homogenization by halite dissolution, cannot represent an
764
immiscible pair because phase equilibrium does not permit the coexistence of both types of
765
inclusions (Bodnar, 2003). In addition, salinity values of the vapor-rich inclusions of up to
766
14.7 wt% NaCl equ. indicate that these inclusions cannot be immiscible pairs, but rather they
767
are a product of different stages. Bodnar et al. (1985) indicated that, in the case of the
768
coexistence of brine- and vapor-rich inclusions, the salinity of these vapor-rich inclusions
769
never exceeds 1 wt% NaCl equ.
770 771
On the other hand, the presence of both vapor-rich (Type II) and liquid-rich (Type
772
IIIa) inclusions in the same FIA likely indicates boiling. Similar Ths from vapor-rich and
773
liquid-rich inclusions of 355–452 °C, and 353–458 °C, respectively may support localized
774
boiling (Roedder and Bodnar, 1997). In the case of boiling, the homogenization of Type II
775
and Type IIIa inclusions would present on the two-phase critical curve. Taking into
776
consideration a minimum (353 °C) and maximum (458 °C) Th of this boiling fluid inclusion
777
assemblage, the trapping pressure is estimated to be between 195 and 445 bar, respectively
778
(Fig. 11).
779 780
The average salinity and Th temperatures of the Type IIIc fluid inclusions in the
781
retrograde stage quartz, epidote, and calcite minerals were calculated as 3.2 wt% NaCl equ.,
782
and 240 °C, respectively. The trapping pressure of these retrograde stage skarn minerals was
783
calculated as 310 bar based on the highest independent thermometer of 281 °C (Fig. 11). 23
784
However, the lower limit of the trapping pressure could not be calculated because the
785
minimum independent temperatures of 136 °C were lower than the average Th isochore of
786
this type of fluid inclusion in Fig. 11.
787 788
The calculated pressure values for the prograde skarn stage were around 884 bar, but
789
decrease as low as 195 bar through later stages. The calculated maximum pressure of the fluid
790
inclusions corresponded to the 3.3 km thick stratigraphic units overlying the mineral deposit
791
(assuming that the density of overlying units was 2.7 gr/cm3). On the other hand, a thickness
792
of 0.7 km was calculated based on the minimum pressure. Such a sudden change in the skarn
793
environment is unreasonable. However, Yardley and Lloyd (1995) indicated that a
794
metasomatic process in the skarn front caused the increasing hydraulic pressure and resulted
795
in the collapse of the skarn system when the hydraulic pressure exceeded the overlying
796
stratigraphic units. Such a collapse of the skarn system is a response to hydrofracturing and
797
brecciation which changes in permeability and porosity (Dipple and Gerdes, 1998).
798
Additionally, Clechenko and Valley (2003) stated that hydrofracturing and brecciation can
799
cause the mixture of meteoric waters in the skarn system. Such a mixture of meteoric water
800
may explain the boiling event in Dağbaşı skarn. Meinert et al. (2005) also pointed out that
801
fractured and brecciated host rocks in the skarn front refer to the shallow emplacement of
802
intrusion and skarn formation.
803
These statements indicate that the prograde stage of Dağbaşı skarn occurred in higher
804
hydraulic pressure conditions. Increasing hydrofracturing and brecciation resulted from the
805
mixture of meteoric water and boiling through later stages. The lower limit of the pressure
806
values calculated for the retrograde skarn stage indicates that the Dağbaşı skarns were formed
807
under shallow conditions. The similar emplacement depth of the Dağbaşı granitoid, which
808
was calculated at between 0.3 and 8 kbar by Kaygusuz and Aydınçakır (2011), is quite
809
compatible with the Dağbaşı skarns. Besides, the commonly observed ore textures of the
810
stockwork veining, brittle fracturing, and brecciation, as well as intensive hydrothermal
811
alteration in the Dağbaşı skarns, were considered as indicators of the shallow depths in
812
pioneering studies (Einaudi and Burt, 1982; Meinert et al., 2005), and highly compatible with
813
the shallow skarn formation.
814 815
9.3.
Source of methane-bearing fluid inclusions
816
On the basis of the petrographic and microthermometric study, there were no methane
817
and CO2 gas phases in the fluid inclusions. However, a minor amount of CH4 was detected in 24
818
the Type IIIa fluid inclusions by the Raman spectroscopy. Similar findings were also reported
819
by Hedenquist and Henley (1985) and Rosso and Bodnar (1995), that if the bulk density is
820
less than 2 mol% in the fluid inclusions these gas species may not be detected. The CH4 is not
821
present in all Type IIIa inclusions, because only three of the 18 measured Type IIIa
822
inclusions, which always coexist with Type II fluid inclusions of the boiling assemblage, were
823
found to contain CH4 phases.
824 825
Methane and trace concentrations of other hydrocarbons have been widely reported in
826
metamorphic, ultramafic, and igneous hosted ore deposits and hydrothermal systems (Welhan,
827
1988; Polito, 1999; Whiticar, 1999; Charlou et al., 2002; Fan et al., 2004; Phillips and Powel,
828
2010). According to Welhan (1988), the possible sources of methane in hydrothermal systems
829
are: (1) thermal degradation of hydrocarbons; (2) biological production; (3) outgassing of
830
juvenile carbons as CH4; and (4) inorganic synthesis in reactions at higher temperatures.
831 832
High salinity fluid inclusions of prograde stage skarn minerals do not contain CH4.
833
Therefore, the distinct lack of CH4 in the prograde stage was interpreted to suggest that the
834
CH4 in the fluid inclusions was not magmatic in origin. This is because, if this CH4 were
835
magmatic in origin, it would have been detected in the prograde stage Type I fluid inclusions.
836
The decarbonization process of fluid‒rock interaction, between low salinity fluids and
837
carbonaceous host rock, was also suggested in some studies as the basis for the origin of CH4-
838
bearing fluid inclusions in hydrothermal systems (Fu et al., 2014). However, this process is
839
also inappropriate for the explanation of CH4 in Dağbaşı skarns, because, if there was a
840
decarbonization process between hydrothermal solutions and carbonates, CO2 also would
841
have been detected in these fluid inclusions. However, the microthermometric study and
842
Raman measurements confirm that there were no CO2 phases. Additionally, if there was a
843
decarbonization process, the decarbonization trend of the data in Fig. 13a would have been
844
observed.
845 846
Biogenic origin was also suggested in some other studies (Whiticar, 1999; Ueno et al.,
847
2006; Potter and Longstaffe, 2007) by the incorporation of organic material into the high
848
temperature hydrothermal systems. Similarly, Shen et al. (2016) also reported that the organic
849
material of the host rocks produced the CH4 gas by thermal decomposition during the
850
emplacement of granitoid intrusions. Higher organic compounds in the limestones were also
851
indicated by Demir et al. (2017), and thermal degradation of organic materials was interpreted 25
852
as the most likely source of CH4 in the Sivrikaya skarn deposit. Therefore, it is suggested that
853
the CH4 at Dağbaşı skarn may have derived from the organic compounds of the limestone.
854
Therefore, the thermogenic origin of the CH4 at Dağbaşı skarn should be considered.
855 856
9.4. Comparisons with other skarn deposits
857
Skarn deposits in the northeastern region of Turkey are spatially associated with both
858
Early Cretaceous Berdiga limestones and Late Cretaceous carbonate layers in volcano-
859
sedimentary units (Saraç, 2003; Çiftçi, 2011; Demir et al., 2017; Sipahi, 2011; Sipahi et al.,
860
2017). The skarns associated with the Early Cretaceous Berdiga Formation are always
861
characterized by a considerable amount of sulfide phases (Aslan, 1991; Çiftçi 2011; Saraç,
862
2003). On the other hand, minor amounts of pyrite and chalcopyrite were accompanied by
863
magnetite and hematite along the skarn formation of the Upper Cretaceous carbonates (Saraç,
864
2003; Sipahi, 201; Demir et al., 2017). Reported ore reserves for these deposits are: three
865
hundred sixty thousand ton for Dereli (55 wt% Fe2O3), seven hundred fifty thousand ton for
866
Çambaşı (65.21 wt% Fe2O3), and five million ton for Kartiba (77.65 wt% Fe2O3), while there
867
are no reserve estimates for the other skarn locations (Saraç, 2003). However, old adits and
868
mine waste around these locations indicate that these skarns may also have significant ore
869
reserves to be mined. All these reserve estimates indicate that these skarns are relatively
870
small, but the polymetallic nature of these sulfide-bearing skarns makes them economically
871
significant because these metals needs to be recovered together with the main ore minerals of
872
magnetite and hematites.
873 874
Comparing the skarn type deposits of the eastern and western regions of Turkey by the
875
host rock lithology, they are related to Lower Cretaceous Berdiga limestones and Upper
876
Cretaceous volcano sedimentary units in the northeastern region (Demir et al., 2017).
877
However, in the western region, the Ayazmant and Evciler deposits were hosted by Early
878
Triassic metapelites‒metabasites and marbles (Öztürk et al., 2008; Oyman, 2010), and the
879
Kozbudaklar and Susurluk deposits were associated with the Triassic and Mesozoic
880
carbonates, respectively (Orhan et al., 2010; Orhan, 2017). Based on the skarn-related
881
intrusions, skarns in the northeastern region are related to the Late Cretaceous‒Eocene
882
granitoids, whereas skarn type deposits in the western region (i.e., Ayazmant, Kozbudaklar,
883
Susurluk, and Evciler) are related to the Eocene‒Miocene aged granitic intrusions (Öztürk et
884
al., 2008; Oyman, 2010; Orhan et al., 2010; Orhan, 2017). There were no porphyry systems in
885
close proximity to the skarn deposits of the northeastern region (Demir et al., 2017); however, 26
886
porphyry-bearing intrusions related to skarn deposits have been reported from the Ayazmant
887
and Kozbudaklar deposits of western Turkey (Oyman, 2010; Orhan, 2017). These
888
explanations indicate that the skarn-related granitoids in the eastern region are related to older
889
granitic intrusions compared to the western region. Therefore the absence of a porphyry
890
system in the eastern Black Sea region can be explained by the erosion of the older granitic
891
intrusions and accompanying porphyry deposits.
892 893
The δ18O and δD isotope results indicate that magmatic dominated solutions were
894
effective at the prograde stage of the Dağbaşı skarn deposits. On the other hand, the highly
895
depleted δ18O isotopes of the retrograde stage skarn minerals and their equilibrated solutions
896
indicate a dilution of magmatic solutions by the meteoric fluids. Highly depleted δ18O
897
isotopes of the skarn calcites and equilibrated solutions are also consistent with the mixture of
898
meteoric fluids. The δ18O isotopes, reported from Eğrikar skarn, are also consistent with the
899
magmatic source at the prograde stage, while the mixture of magmatic and meteoric water is
900
suggested for the retrograde skarn stage (Sipahi et al., 2017). According to the sulfur isotopes,
901
magmatic sources were demonstrated for the prograde stage of Kotana skarn deposit, and the
902
involvement of meteoric water was suggested at the retrograde stage (Çiftçi, 2011). Similar
903
findings were also reported from the Susurluk, Ayazmant and Evciler deposits of western
904
Turkey, where δ18O isotopes at these deposits were clearly linked to the magmatic source at
905
the early skarn stage, while the mixture of magmatic and meteoric solutions was suggested for
906
the later stages (Öztürk et al., 2008; Oyman, 2010; Orhan et al., 2017). When comparing the
907
skarn deposits of eastern and western Turkey, both of them are characterized by magmatic
908
dominated hydrothermal solutions at the early skarn stage, while later stages are characterized
909
by a mixture of magmatic and meteoric solutions.
910 911
The increasing meteoric effect on the Dağbaşı skarn deposit was also confirmed by the
912
fluid inclusion studies. The higher salinity values of the prograde stage fluid (up to 61.8 wt%
913
NaCl equ.) indicate that magmatic dominated fluids were effective on the skarn environment.
914
However, the decreasing salinity and Th trend of the retrograde stage fluid inclusions are
915
evidence of dilution with meteoric water. Salinity values reported from the Kotana (Çiftçi,
916
2011), Sivrikaya (Demir et al., 2017) and Eğrikar deposits (Sipahi et al., 2017) are somewhat
917
lower than the Dağbaşı skarn ore, up to 15, 15.4 and 14.3 wt% NaCl equ., respectively.
918
However, the decreasing salinity trend of the fluid inclusions in Sivrikaya and low salinity
919
values in the Eğrikar and Kotana deposits were accepted as evidence of dilution by meteoric 27
920
water. Moreover, higher salinity values of the prograde stage of the Susurluk and
921
Kozbudaklar skarn deposits of western Turkey, which reach up to 70 wt% NaCl equ. (Orhan
922
et al., 2010; Orhan, 2017), were also accepted as evidence of magmatic-related solutions. On
923
the other hand, decreasing salinity values of the retrograde stage were also accepted as
924
evidence of a meteoric effect on these deposits.
925 926
The homogenization temperatures of the fluid inclusions measured from the different
927
stages of skarn minerals were between 160–514 °C in the Dağbaşı skarn ore. These values
928
were slightly higher than the temperature ranges of the Eğrikar (between 200 and 425 °C;
929
Sipahi et al., 2017), Sivrikaya (166–462 °C; Demir et al., 2017), and Kotana (380–460 °C;
930
Çiftçi, 2011) skarn deposits. Much higher temperature values were reported from the Susurluk
931
(371–>600 °C; Orhan et al., 2010), Ayazmant (300–576 °C; Oyman, 2010), and Kozbudaklar
932
(308–>600 °C; Orhan et al., 2017) skarn deposits of western Turkey. When comparing these
933
temperature ranges of skarn deposits from eastern and western Turkey, it is apparent that the
934
western skarn deposits formed at higher temperature conditions. The much deeper formation
935
depths of the western skarn deposits may be one of the reasons for the higher temperature
936
conditions (Meinert et al., 2005) because the trapping pressure of the early stage fluid
937
inclusions in the Dağbaşı skarns was less than 884 bar, and there is no pressure estimate
938
reported from the other skarn locations of eastern Turkey. However, pressure estimates
939
reported from the Susurluk (~1000 bar) and Kozbudaklar (~2000 bar) skarns of western
940
Turkey are higher (Orhan et al., 2011; Orhan, 2017) than that of the Dağbaşı skarn ore.
941 942
10. Conclusions
943
Dağbaşı skarn deposits formed along the contact between the Lower Cretaceous
944
Berdiga limestone and the Upper Cretaceous Dağbaşı granitoid. The prograde stage was
945
represented by garnet and pyroxenes, while the retrograde stage was represented by epidote,
946
tremolite, actinolite, and chlorite. Quartz and calcites accompany both stages of the skarn
947
development. The ore minerals which were formed at the retrograde skarn stage mainly
948
consist of magnetite and hematite. Pyrrhotite, pyrite, chalcopyrite, sphalerite, and minor
949
galena are the accompanying sulfide phases in the ore.
950 951
The salinity values of the prograde stage fluid inclusions were measured to be between
952
48.8 and 61.8 wt%, while retrograde stage fluid inclusions were characterized by lower
953
salinity values. Higher salinity values of the prograde stage indicate magmatic dominated 28
954
solutions, whereas the lower salinity values of the retrograde stage and well-defined
955
decreasing salinity and Th trends of the fluid inclusions indicated a mixture of meteoric water
956
at the retrograde stage. The mixture of magmatic and meteoric solutions was further
957
supported by the boiling evidence of fluid inclusions.
958 959
The homogenization temperatures of the fluid inclusions in the prograde stage range
960
between 412 and 514 °C, and these measured temperatures were well within the appropriate
961
range for the calculated equilibrium temperatures of the prograde stage. On the other hand,
962
the equilibrium temperatures of the retrograde mineral assemblage (between 135 and 281 °C)
963
were slightly lower than the measured Th of the fluid inclusions (160–327 °C).
964 965
Thermal degradation of organic materials in carbonates was suggested as the source of
966
CH4 in fluid inclusions. This is because, if the CH4 were magmatic in origin, it would have
967
been determined in the first stage fluid inclusions. Moreover, if there was a decarbonization
968
process between the hydrothermal solutions and carbonates, CO2 would have been observed
969
in these inclusions.
970 971
The δ18O and δD isotopes of the prograde stage minerals and their equilibrated
972
solutions indicate that magmatic dominated solutions were effective at this stage. On the other
973
hand, δ18O and δD isotopes of the retrograde stage minerals and their equilibrated solutions
974
were found to be compatible with the meteoric origin.
975 976
Pressure estimates from the fluid inclusions indicate that early stage skarns were
977
formed in the range of 884 to 710 bar, whereas the formation pressure decreased to as low as
978
195 bar at the late skarn stage. Taking into consideration that ore formation takes place at the
979
retrograde stage, shallow skarn environments were suggested for the Dağbaşı skarn ore.
980 981
Acknowledgements
982
This study was financially supported by TÜBİTAK through project number 112Y331.
983
Special thanks are due to Andrea Jauss for providing the confocal Raman measurements at
984
WITec GmbH in Ulm, Germany. We would like to thank Serkan Şenkaya, Mustafa Aksu,
985
Kadir Bayraktar, and Mehdi İlhan for their assistance during the fieldwork and laboratory
986
studies. We are also grateful to anonymous reviewers whose valuable suggestions greatly
987
improved the earlier version of the manuscript. 29
988
References
989 990
Aslan, Z., 1991. Özdil (Yomra-Trabzon) Yöresinin Petrografisi Skarn Oluşukları ve GranatPiroksen Ritmikleri. KTÜ Fen Bilimleri Enstitüsü, Master Thesis, Trabzon, 72p.
991 992 993
Aydınçakır, E., 2006. Dağbaşı (Araklı-Trabzon) granitoyidi ve çevresinin petrografik, jeokimyasal ve petrolojik özelliklerinin incelenmesi. KTU, Fen Bilimleri Enstitüsü, Master Thesis, Trabzon, 120p.
994 995
Aydınçakır, E., Kaygusuz, A., 2012. Geç Kretase Yaşlı Dağbaşı (Araklı, Trabzon) volkanitlerinin petrografik ve jeokimyasal özellikleri. GÜFBED, 2, 2, 123-142.
996 997 998
Baker, T., Lang, J.R., 2003. Reconciling fluid inclusion types, fluid processes, and fluid sources in skarns: an example from the Bismark Deposit Mexico. Miner. Deposita 38, 474–495.
999 1000 1001
Bakker, R.J., 2004. Raman Spectra of fluid and crystal mixtures in the systems H2O, H2ONaCl and H2O-MgCl2 at low temperatures: Applications to fluid inclusion research. Can. Mineral. 42, 1283–1314.
1002 1003 1004
Baumgartner, M., Bakker, R.J., 2009. Raman spectroscopy of püre H2O and NaCl-H2O containing syntetic fluid inclusions in quartz-astudy of polarization effects. Miner. Petrol. 95, 1–15.
1005 1006 1007
Bodnar, R.J., 2003. Introduction to aqueous fluid systems. In I. Samson, A. Anderson, D. Marshall, Eds. Fluid Inclusions: Analysis and interpretation. Mineral. Assoc. Canada, Short Course Series 32, 81–99.
1008 1009 1010
Bodnar, R.J., Reynolds, T.J., Kuehn, C.A., 1985. Fluid inclusion systematics in epithermal systems. In Geology and Geochemistry of Epithermal Systems (B.R. Berger and P.M. Bethke, eds) Society of Economic Geologist, Rev. Econ. Geol. pp. 2, 73–98.
1011 1012 1013
Bodnar, R.J., Vityk, M.O., 1994. Interpretation of microthermometric data for H2O-NaCl fluid inclusions. In Fluid Inclusions in Minerals, Methods and Appl. (ed. B. De Vivo and M.L. Frezotti), Virginia Tech 117–130.
1014 1015 1016
Bodnar, R.J., Lecumberri-Sanchez, P., Moncada, D., Steele-MacInnis, M., 2014. Fluid inclusions in hydrothermal ore deposits. In: Holland, H.D., Turekian, K.K. (Eds.) Treatise on Geochemistry (2th edition, 13) Elsevier, Oxford, pp. 119–142.
1017 1018
Bowman, J.R., O’Neil, J.R., Essene, J.R., 1985. Contact skarn formation et Elkhorn, Montana, II: Origin and evolution of C-O-H skarns fluids. Am. J. Sci. 285, 621–660.
1019 1020 1021
Bowman, J.R., 1998. Stable isotope systematics of skarns, mineralized intrusion related skarn system. D.R., Lentz (Ed.), Mineralogical Assoc. Canada, Short Course Series 26, pp. 99–114.
1022 1023 1024
Boztuğ, D., Jonckheere, R., Wagner, G.A., Yeğingil, Z., 2004. Slow Senonian and fast Paleocene-Early Eocene uplift of the granitoides in the cetral eastern Pontides, Turkey: Apatite fission-track results. Tectonophysics 382, 213–228.
1025 1026 1027
Brown, P.E., Lamb, W.M., 1989. PVT propeties of flluids in the system H2O-CO2-NaCl: new graphical presentations and implications for fluid inclusion studies. Geochim. Cosmochim. Ac. 53, 1209–1221.
1028
Burke, E.A.J., 2001. Raman microspectrometry of fluid inclusions. Lithos 55, 139–158.
30
1029 1030 1031
Charlou, J.L., Donval, J.P., Fouquet, Y., Jean-Baptiste, P., Holm, N., 2002. Geochemistry of high H2 and CH4 vent fluid issuing from ultramafic rocks at the Rainbow hydrothermal field (36°14’ N, Mar). Chem. Geol., 191, 345–359.
1032 1033 1034
Chiba, H., Chacko, T., Clayton, R.N., Goldsmith, J.R., 1989. Oxygen isotope fractionations involving diopside, forsterite, magnetite and calcite: application to geothermometry. Geochim. Cosmochim. Ac. 53, 2985–2995.
1035 1036
Clarke, I.D., Fritz, P., 1997. Environmental Isotopes in Hydrogeology. Lewis Publishers, CRC Press, Newyork, 352pp.
1037 1038 1039
Clechenko, C.C., Valley, J.W., 2003. Oscillatory zoning in garnet from the Willsboro wollastonite skarn, Adirondacks mts, Newyork: a record of shallow hydrothermal processes preserved in granülite facies terane. J. Metamorph. Geol. 21, 771–784.
1040 1041
Craig, H., 1961. Standad for reporting concentrations of deuterium and oxygen-18 in natural waters. Science 133, 1833–1834.
1042 1043 1044
Çiftçi, E., 2011. Sphalerite associated with pyrrhotite-chalcopyrite ore occurring in the Kotana Fe-skarn deposit (Giresun, NE Turkey): Exolutions or replacement. Turk. J. Earth Sci. 20, 307–320.
1045 1046
Demir, Y., 2019. Geological, mineralogical, and geochemical properties of the Dağbaşı skarn ores (Araklı-Trabzon, NE-Turkey). Bull. Min. Res. Exp. 158, 165–194.
1047 1048 1049
Demir, Y., Uysal, I., Kandemir, R., Jauss, A., 2017. Geochemistry, fluid inclusion and stable isotope constraints (C and O) of the Sivrikaya Fe-skarn mineralization (Rize, NE Turkey). Ore Geol. Rev. 91, 153–172.
1050 1051
Dipple, G.M., Gerdes, M.L., 1998. Reaction-infiltration feedback and hydrotherdynamics at the skarn front. Mineral. Assoc. Can., Short Course Series 26, 71–97.
1052 1053 1054
Dokuz, A., Aydınçakır, E., Kandemir, R., Karslı, O., Siebel, W., Derman, A.S., 2017. Late Jurassic magmatism and stratigraphy in the eastern Sakarya zone Turkey; Evidence for the slab berakoff of Paleotethyyan Ocean Lithosphere. J. Geol. 125, 1, 1–31.
1055 1056 1057
Dokuz, A., 2011. A slab detachment and delamination model for the generation of Carboniferous high-potassium I-type magmatism in the Eastern Pontides, NE Turkey: Köse composite pluton. Gondwana Res. 19, 926–944.
1058 1059 1060 1061
Dokuz, A., Karslı, O., Chen, B., Uysal, I., 2010. Sources and petrogenesis of Jurassic granitoids in the Yusufeli area, Northeastern Turkey: implications for pre- and postcollisional lithospheric thinning of the Eastern Pontides. Tectonophysics 480, 259– 279.
1062 1063 1064 1065
Dokuz, A., Uysal, I., Yıldırım, D., Karslı, O., Meisel, T., Kandemir, R., 2015. Geochemistry Re-Os isotopes and highly siderophil element abundances in the Eastern Pontide peridotites NE Turkey, Multiple episodes of melt extraction depletion melt rock interaction and fertilization of the Rheic Ocean Mantle. Gondwana Res. 27, 2, 612–628.
1066 1067
Einaudi, M.T., Burt, D.M., 1982. Introduction-terminology, classification,and composition of skarn deposits. Econ. Geol. 77, 745–754.
1068 1069 1070
Fan, H.R., Xie, Y.H., Wang, K.Y., Wilde, S.A., 2004. Methane-rich fluid inclusions in skarn nearthe giant REE-Nb-Fe deposit at Bayan Obo, Northern China. Ore Geol. Rev. 25, 301–309.
1071 1072
Frezzotti, M.L., Tecce, F., Casagli, A., 2012. Raman spectroscopy for fluid inclusion analysis. Geochem. Explor. 112, 1–20. 31
1073 1074 1075
Friedman, I., O’Neil, J.R., 1977. Compilation of stable isotope fractionation factors of geochemical interest, Data of Geochemistry. United States goverment printing Office, Washington, 117pp.
1076 1077
Fu, B., Terrence, P.M., Alison, M.F., David, P., Mark, A.K., 2014. CH4-N2 in the Maldon gold deposit, central Victoria, Australia. Ore Geol. Rev. 58, 225–237.
1078 1079
Güven, İ.H., Nalbantoğlu, A.K., Takaoğlu, S., 1998. 1/100.000 Ölçekli Açınsama Nitelikli Türkiye Jeoloji Harita Serisi. MTA Yayınları.
1080 1081
Güven, İ.H., 1993. 1/250.000 scaled geological and metallogenical map of the Eastern Black Sea Region. General Directorate of Mineral Research and Exploration (in Turkish).
1082 1083
Hasançebi, N., 1993. Dağbaşı (Araklı-Trabzon) granitoyidine bağlı cevherleşmelerin incelenmesi. KTU, Fen Bilimleri Enstitüsü, Master Thesis, Trabzon, 65p.
1084 1085 1086
Hedenquist, J.W., Henley, R.W., 1985. The importance of CO2 on freezing point measurements of fluid inclusions: Evidence from active geothermal systems and implications for epithermal ore deposit. Econ. Geol. 80, 1379–1406.
1087
Hoefs, J., 2009. Stable isotope geochemistry (6th Edition). Springer Verlag, 285p.
1088 1089 1090
Holness, M.B., 1997. Fluid flow paths and mechanism of fluid infiltration in carbonates during contact metamorphism: the Beinn an Dubhaich aureole, Skye. J. Metamorph. Geol. 15, 59–70.
1091 1092 1093 1094
Kandemir, R., Yılmaz, C., 2009. Lithostratigraphy, facies, and deposition environment of the lower Jurassic Ammonitico Rosso type sediments (ARTS) in the Gümüşhane area, NE Turkey: Implications for the opening of the northern branch of the Neo-Tethys Ocean. J. Asian Earth Sci. 34, 4, 586–598.
1095 1096 1097 1098
Karslı, O., Chen, B., Aydın, F., Şen, C., 2007. Geochemical and Sr-Nd-Pb isotopic compositions of the Eocene Dölek and Sariçiçek plutons, eastern Turkey: implications for magma interaction in the genesis of High-K Calc-Alkaline granitoids in a postcollision extensional setting. Lithos 98, 67–96.
1099 1100 1101 1102
Karslı, O., Dokuz, A., Uysal, I., Aydin, F., Kandemir, R., Wijbrans, R.J., 2010b. Generation of the Early Cenozoic adakitic volcanism by partial melting of mafic lower crust, Eastern Turkey: implications for crustal thickening to delamination. Lithos 114, 109– 120.
1103 1104 1105 1106 1107
Karslı, O., Dokuz, A., Kandemir, R., 2017. Zircon Lu-Hf isotope systematics and U-Pb geochronology, whole-rock Sr-Nd isotopes and geochemistry of the early Jurassic Gökçedere pluton, Sakarya zone-Ne Turkey: a magmatic response to roll back of the paleo tethyan oceanic lithosphere. Conrib. Mineral. Petr. 172, 31. DOI 10.1007/s00410017-1346-0.
1108 1109 1110
Karslı, O., Uysal, İ., Ketenci, M., Dokuz, A., Aydın, F., Kandemir, R., Wijbrans, J., 2011. Adakite-like granitoid porphries in eastern pontides, NE Turkey: potencial parental melts and geodynamic implications. Lithos 127, 354–372.
1111 1112 1113
Karslı, O., Aydin, F., Sadiklar, M.B., 2004. The morphology and chemistry of K-feldspar megacrysts from Ikizdere Pluton: evidence for acid and basic magma interactions in granitoid rocks, NE Turkey. Chem. Erde-Geochem. 64, 155–170.
1114 1115
Karslı, O., Dokuz, A., Uysal, I., Aydin, F., Bin, C., Kandemir, R., Wijbrans, R.J., 2010a. Relative contributions of crust and mantle to generation of Campanian high-K calc-
32
1116 1117
alkaline I-type granitoids in a subdution setting, with spezial reference to the Harşit pluton, Eastern Turkey. Contrib. Mineral. Petr. 160, 467–487.
1118 1119 1120
Karslı, O., Uysal, İ., Dilek, Y., Aydın, F., Kandemir, R., 2013. Geochemical modelling of early Eocene adacitic magmatism in the Eastern Pontides, NE Anatolia: continental crust or subducted oceanic slab origin?, Int. Geol. Rev. 55, 16, 2083–2095.
1121 1122 1123 1124
Kaygusuz, A., Aydınçakır, E., 2009. Mineralogy, whole-rock and Sr-Nd isotope geochemistry of mafic microgranular enclaves in Cretaceous Dağbaşı granitoids, Eastern Pontides, NE Turkey: Evidence of magma mixing, mingling and chemical equilibration. Chem. Erde-Geochem. 69, 247–277.
1125 1126 1127
Kaygusuz, A., Aydınçakır, E., 2011. Petrogenesis of a Late Cretaceous composite pluton from the eastern Pontides. The Dağbaşı Pluton, (NE-Turkey). Neues Jb. Mineral. Abh. 188, 211–233.
1128 1129 1130
Kaygusuz, A., Chen, B., Aslan, Z., Siebel, W., Şen, C., 2009. U-Pb SHRIMP zircon ages, geochemical and Sr-Nd isotopic compositions of the Late Cretaceous I-type Sarıosman Pluton, Eastern Pontides NE Turkey. Turk. J. Earth Sci. 18, 4, 549–581.
1131 1132 1133
Kırmacı, M.Z., Koch, R., Bucur, J.I., 1996. An Early Cretaceous section in the Kırcaova area (Berdiga Limestone, NE-Turkey) and its correlation with platform carbonates in WSlovenia. Facies 34, 1–22.
1134 1135 1136 1137
Kırmacı, M.Z., Yıldız, M., Kandemir, R., Eroğlu-Gümrük, T., 2018. Multistage dolomitization in late Jurassic-Early Cretaceous platform carbonates (Berdiga Formation), Başoba Yayla (Trabzon), NE Turkey, Implication of the generation of magmatic arc on dolomitization. Mar. Petrol. Geol. 89, 515–529.
1138 1139 1140
Kurt, Y., 2014. Giresun, Bulancak Kirazören bölgesi skarn tipi demir yataklarının jeolojik ve jeokimyasal incelenmesi. İstanbul Üniversitesi, Fen Bilimleri Enstitüsü, Master Thesis, İstanbul, 111p.
1141 1142
Kwak, T.A.P., 1986. Fluid inclusions in skarns (carbonate replacement deposits). J. Metamorph. Geol. 4, 363–384.
1143 1144 1145 1146
Le Maitre, R.W., Bateman, P., Dudek, A., Keller, J., Lameyre Le Bes, M.J., Sabine, P.A., Schmid, R., Sorenson, H., Streckeisen, A., Woolley, A.R., Zanettin, B., 1989. A classification of igneous rocks and glossary of terms. Blackwell Scientific Publications, Oxford, U.K.
1147 1148 1149
Lin, F., Bodnar, R.J., 2010. Synthetic fluid inclusions XVIII: experimental determination of the PVTX properties of H2O-CH4 to 500 °C, 3 kbar and XCH4 4 mol%. Geochim. Cosmochim. Ac. 74, 3260–3273.
1150 1151 1152
Lin, F., Bodnar, R.J., Becker, S.P., 2007. Experimental determination of the Raman CH4 symmetric streching (VI) band position from 1–650 bar and 0.3–22 °C: application to fluid inclusion studies. Geochim. Cosmochim. Ac. 71, 3746–3756.
1153 1154
Meinert, L.D., Diple, G.M., Nicolescu, S., 2005. World skarn deposits, society of economic geologist, Inc. Econ. Geol., 100th Anniversary Volume 299–336.
1155 1156 1157
Moore, W.J., McKee, E.H., Akınci, Ö.T., 1980. Chemistry and chronology of plutonic rocks in the Pontid mountains, northern Turkey. In: Jankovic, S., Sillitoe, R.H. (Eds.), Eur. Copper Deposits, Belgrade, pp. 209–216.
33
1158 1159 1160
O’Neil, J.R., 1986. Theoretical and experimental aspects of isotopic fractionation. In stable isotopes in high temperature geological processes (Ed. J.W. Valley, et al.); Rev. Mineral. 16, 1–40.
1161 1162 1163
Okay, A.I., Şahintürk, Ö., 1997. Geology of the Eastern Pontides, In: A.G. Robinson, (ed.), Regional and Petroleum Geology of the Black Sea and Surrounding Region. AAPG Mem. 68, 291–311.
1164 1165 1166
Orhan, A., 2017. Evolution of the Mo-rich scheelite skarn mineralization at Kozbudaklar, Western Anatolia, Turkey: Evidence from mineral chemistry and fluid inclusions. Ore Geol. Rev. 80, 141–165.
1167 1168 1169
Orhan, A., Demirbilek, M., Mutlu, H., 2017. Geochemistry of the Topuk Pluton associated with the Kozbudaklar W-skarn deposit (Western Anatolia, Turkey): Implication for crystallization conditions. J. Afr. Earth Sci. 130, 141–160.
1170 1171
Orhan, A., Mutlu, H., Fallick, A., 2011. Fluid infiltration effects on stable isotope systematics of the Susurluk skarn deposit, NW Turkey. J. Asian Earth Sci. 40, 550–568.
1172 1173
Orhan, A., Mutlu, H., Hanilçi, N., 2010. Microthermometric characteristics of the oxidized type W-Skarn, in Susurluk, Balıkesir, Turkey. Mineral Res. Expl. Bull. 141, 53–68.
1174 1175
Oyman, T., 2010. Geochemistry, mineralogy and genesis of the Ayazmant Fe-Cu skarn deposit in Ayvalık, (Balıkesir), Turkey. Ore Geol. Rev. 37, 175–201.
1176 1177 1178
Öztürk, Y.Y., Helvacı, C., Satır, M., 2008. The influence of meteoric water on skarn formation and late-stage hydrothermal alteration at the Evciler skarn occurrance, Kazda, NW Turkey. Ore Geol. Rev. 34, 271–284.
1179 1180
Pelin, S., 1977. Geologic investigation of southeast part of Alucra (Giresun) area for petrolium potential. Publication of Karadeniz Technical University, No: 87.
1181 1182
Phillips, G.N., Powell, R., 2010. Formation of gold deposits: a metamorphic devolatization model. J. Metamorph. Geol. 28, 689–718.
1183 1184 1185
Polito, P.A., 1999. Exploration implications predicted by the distribution of Carbon-OxygenHydrogen gases above and within the junction gold deposit, Kambalda, western Australia. The University of Adelaide, PhD. Thesis, South Australia, 312p.
1186 1187 1188 1189
Potter, J., Longstaffe, F.J., 2007. A gas-chromatograph, continuous flow-isotope ratio massspectrometry method for delta C-13 and delta D measurement of complex fluid inclusion volatiles: examples from the Khibina alkaline igneous complex. Chem. Geol. 244, 186–201.
1190 1191
Roedder, E., 1984. Fluid inclusions. Reviews in Mineralogy, Mineralogical Society of America 12, pp. 12–45.
1192 1193
Roedder, E., Bodnar, R.J., 1980. Geological pressure determinations from fluid inclusion studies. Ann Rev. Earth Planet Sci. 8, 263–301.
1194 1195 1196
Roedder, E., Bodnar, R.J., 1997. Fluid inclusion studies of hydrothermal ore deposits. In geochemistry of hydrothermal ore deposits (Ed. H.L. Barnes). Wiley and Sons Inc. Newyork, pp. 657–698.
1197 1198 1199
Rosso, K.M., Bodnar, R.J., 1995. Microthermometric and Raman spectroscopic detection limits of CO2 in fluid inclusions and the raman spectroscopic characterization of CO2. Geochim. Cosmochim. Ac. 59, 19, 3961–3975.
34
1200 1201 1202
Sanchez, P.L., MacInnis, S.M., Bodnar, R.J., 2012. A numerical model to estimate trapping conditions of fluid inclusions that homogenize by halite disappearance. Geochim. Cosmochim. Ac. 92, 14–22.
1203 1204
Saraç, S., 2003. Doğu Karadeniz Fe-Skarn Yataklarının karşılaştırmalı mineralojik ve jeokimyasal özellikleri. KTÜ Fen Bilimleri Enstitüsü, PhD. Thesis, Trabzon, 259p.
1205 1206
Shen, P., Pan, H., Zhu, H., 2016. Two fluid sources and genetic implications for the Hatu gold deposit, Xinjiang, China. Ore Geol. Rev. 73, 298–312.
1207 1208
Shepherd, T.J., Rankin, A.H., Alderton, D.H.M., 1985. A practical guide to fluid inclusion studies. Blackie, Glasgow, 239p.
1209 1210
Sheppard, S.M.F., 1984. Isotope geothermometry. In Thermometrie et Barometrie Geologiques (ed. M. Lagache), Soc. Fr. Mineral. Crystallogr. pp. 349–412.
1211 1212 1213
Shin, D., Lee, I., 2003. Evaluation of the volatization and infiltration effect on the stable isotopic and mineralogical variations in carbonate rocks adjacent to the Cretaceous Muamsa Granite, South Korea. J. Asian Earth Sci. 22, 227-243.
1214 1215 1216
Sipahi, F., 1997. Camiboğazı ve Sarıtaş yaylaları arasındaki bölgenin petrografi ve maden yatakları açısından incelenmesi, KTÜ, Fen Bilimleri Enstitüsü, Master Thesis, Trabzon, 81p.
1217 1218
Sipahi, F., 2011. Formation of skarns at Gümüşhane (Northeastern Turkey). Neues Jb. Miner. Abh. 188–2, 169–190.
1219 1220 1221 1222
Sipahi, F., Akpınar, İ., Eker, Ç.S., Kaygusuz, A., Vural, A., 2017. Formation of the Eğrikar (Gümüşhane) Fe-Cu skarn type mineralization in NE Turkey: U-Pb zircon age, lithogeochemistry, mineral chemistry, fluid inclusion, and O-H-C-S isotopic compositions. J. Geochem. Explor. 182, 32–52.
1223 1224 1225
Sterner, S.M., Bodnar, R.J., 1984. Synthetic fluid inclusions in natural quartz I. Compositional types synthesized and applications to experimental geochemistry. Geochim. Cosmochim. Ac. 48, 2659–2668.
1226 1227
Şengör, A.M.C., Özeren, S., Genç, T., Zor, E., 2003. East Anatolian high plateau as a mantlesupported, North-south shortened domal structure. Geophys. Res. Lett. 30, 24, 1–4.
1228 1229 1230
Taylor, B.E., 1976. Origin and significance of C-O-H fluids in the formation of Ca-Fe-Si skarn, Osgood Mountains, Humboldt country, Nevada. Stanford University, PhD. Thesis, Stanford, California, 149p.
1231 1232 1233
Taylor, B.E., O’Neil, J.R., 1977. Stable isotope of metasomatic Ca-Fe-Al-Si skarns and associated metamorphic and igneous rocks. Osgood Mountains, Nevada. Contrib. Mineral. Petr. 63, 1–49.
1234 1235
Taylor, H.P., 1974. The application of oxygen and hydrogen isotope studies to problem of hydrothermal alteration and ore deposition. Econ. Geol. 69, 843–883.
1236 1237 1238
Taylor, H.P., 1979. Oxygen and hydrogen isotope relationships in hydrothermal mineral deposits. In: Barnes, H.L., (Ed), Geochemistry of hydrothermal ore deposits, 2th edition. Wiley Interscience, Newyork, pp. 236–277.
1239 1240 1241
Taylor, H.P., Shepherd, S.M.F., 1986. Igneous rocks. I. Processes of isotopic fractionation and isotope systematics. In Valley, J.W., Taylor, H.P., O’Neil, J.R.(Eds.), Stable isotopes in high temperature geological processes. Rev. Mineral. 16, pp. 227–272.
35
1242 1243 1244
Taylor, H.P., 1997. Oxygen and Hydrogen isotope relationships in hydrothermal mineral deposits, In: Barnes, H.L. (Ed.), Geochemistry of hydrothermal ore deposits, 3th edition. John Wiley and Sons, New York, 99, 229–302.
1245 1246 1247
Topuz, G., Alther, R, Schwarz, W.H., Dokuz, A., Meyer, H.P., 2007. Variscan amphibolite facies rocks from the Kurtoğlu metamorpic complex. Gümüşhane area, Eastern Pontides, Turkey. Int. J. Earth Sci. 96, 861–873.
1248 1249 1250
Topuz, G., Altherr, R., Siebel, W., Schwarz, W.H., Zack, T., Hasözbek, A., Barth, M., Satır, M., Şen, C., 2010. Carboniferous high-potassium I-type granitoid magmatism in the Eastern Pontides: The Gümüşhane pluton (NE Turkey). Lithos 116, 1–2, 92–110.
1251 1252 1253 1254
Topuz, G., Okay, A.I., Altherr, R., Schwarz, W.H., Siebel, W., Zack, T., Satir, M., Sen, C., 2011. Post-collisional adakite-like magmatism in the A˘gvanis massif and implications for the evolution of the Eocene magmatism in the Eastern Pontides (NE Turkey). Lithos 125, 131–150.
1255 1256
Ueno, Y., Yamada, K., Yoshida, N., Maruyama, S., Isozaki, Y., 2006. Evidence from fluid inclusions for microbial methanogenesis in the early Archean era. Nature 440, 516–519.
1257 1258 1259
Vallance, J., Fontbote, L., Chiaradia, M., Markowski, A., Schmidt, S., 2009. Magmatic dominated fluid evolution in the Jurassic Nambija gold skarn deposits (southerneastern Ecuador). Miner. Deposita 44, 389–413.
1260
Welhan, J.A., 1988. Origin of methane in hydrothermal systems. Chem. Geol. 71, 183–198.
1261 1262
Whiticar, M., 1999. Carbon and hydrogen isotope systematics of bacterial formation and oxidation of methane. Chem. Geol. 161, 291–314.
1263
Wilkinson, J.J., 2001. Fluid inclusions in hydrothermal ore deposits. Lithos 55, 229–272.
1264 1265
Yardley, B.W.D., Lioyd, G.E., 1995. Why metasomatic fronts are really metasomatic sides. Geology 23, 53–56.
1266 1267
Yılmaz, C., Carannante, G., Kandemir, R., 2008. The rift-related Late Cretaceous drowning of the Gümüşhane carbonate platform (NE Turkey). Ital. J. Geosci. 127, 1, 37–50.
1268 1269 1270
Yılmaz, C., Kandemir, R., 2006. Sedimentary records of the extensional tectonic regime with temporal cessation: Gümüşhane Mesozoic basin (NE Turkey). Geol. Carpath. 57, 1, 3– 13.
1271 1272 1273
Yılmaz, M., 2016. Eğrikar (Torul, Gümüşhane) Fe-Cu Skarn cevherleşmesinin petrografik ve jeokimyasal açıdan incelenmesi. Gümüşhane Üniversitesi, Fen Bilimleri Enstitüsü, Master Thesis. Gümüşhane, 87p.
1274 1275
Zheng, Y.F., 1993a. Calculation of oxygen isotope fractionation in anhydrous silicate minerals. Geochim. Cosmochim. Ac. 57, 1079–1091.
1276 1277
Zheng, Y.F., 1993b. Calculation of oxygen isotope fractionation in hydroxyl-bearing silicates. Earth Planet. Sc. Lett. 120, 247–263.
1278 1279 1280 1281 1282 1283 1284 36
1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 1327 1328 1329 1330 1331 1332 1333 1334
Figure Captions: Figure 1. Major skarn occurrences and associated lithological units along the northeastern region of Turkey (modified from Güven, 1993). Figure 2. Geological map (modified from Kaygusuz and Aydınçakır, 2011) and skarn locations around the Dağbaşı (Trabzon) region. Figure 3. Geological map and skarn locations around the a) İpekçili, b) Köprüüstü, c) Kükürtlü, and d) Dere Mahalle districts (modified from Kaygusuz and Aydınçakır, 2011 and Demir, 2019). Figure 4. Macroscopic and microscopic scale ore texture in Dağbaşı skarn ore: a) early stage massive garnet in İpekçili skarn; b) rhythmically banded intergrowth of garnet and pyroxenes; c) nodular shaped garnet and pyroxenes in the carbonates; d) granular intergrowth of garnet and pyroxenes; e) late stage garnet developments along the cracks in the limestones (rarely including pyroxene); f) magnetite and hematite developments along the growth zone of late stage garnets; g) quartz, epidote and calcite developments in the volcanic rocks and carbonates during the retrograde skarn stage; h) massive tremolite‒actinolite intergrowth; i, j) macroscopic and microscopic scale garnet inclusions in the retrograde stage tremolite‒actinolite minerals; k) fracture-filling type quartz, epidote, and calcites in the volcanic host rocks; l) macroscopic scale quartz inclusions in the epidote-bearing samples (Grt: garnet; Pr: pyroxene; Mag: magnetite; Hem: hematite; Qz: quartz; Cal: calcite; Tr: tremolite; Act: actinolite; Ep: epidote). Figure 5. Macroscopic and microscopic scale ore texture in Dağbaşı skarn ore: a) magnetiteand hematite-bearing massive ore; b) banded ore along the carbonate layers; c) fracture-filling type magnetite veins in volcanites; d) breccia-filling type magnetite ore in volcanic host rocks; e) garnet inclusions in hematites represent later stage ore development; f) simultaneous ore formation with the retrograde stage epidotes; g) hematite formation around the magnetite minerals; g, h) magnetite and hematites were observed to be enclosed in chalcopyrites; i) chalcopyrite and sphalerite formation along the pyrite fractures; j) characteristic bird’s eye texture of pyrrhotites; k) galena inclusions in sphalerites; l) magnetite- and hematitedominated ore mass around the Kükürtlü and Dere Mahalle skarns; m) magnetite and hematite breccias were filled up by later stage quartz and epidotes; n) disseminated magnetite intergrowth in tremolites; o) minor amount of chalcopyrite inclusions in the hematites. Figure 6. Paragenetic diagram showing the mineralization stages in the Dağbaşı skarn (modified from Demir, 2019). Figure 7. Photomicrographs of primary fluid inclusions in different minerals: halite-bearing Type I inclusions in garnet (a), quartz (b) and calcite (c) minerals; vapor-rich Type II inclusions in garnet (d) and quartz (e) minerals; liquid-rich Type IIIa inclusions in garnet (f), and quartz (g) minerals; liquid-rich Type IIIb inclusions in epidote (h), calcite (i), sphalerite (j), and quartz (k) minerals; coexistence of Type IIIb and Type IIIc inclusions in quartz (k), Type IIIc inclusions in quartz (l). (S: solid, V: vapor, L: liquid phases.) 37
1335 1336 1337 1338 1339 1340 1341 1342 1343 1344 1345 1346 1347 1348 1349 1350 1351 1352 1353 1354 1355 1356 1357 1358 1359 1360 1361 1362 1363 1364 1365 1366 1367 1368 1369 1370 1371 1372 1373 1374 1375 1376 1377 1378 1379 1380 1381 1382 1383 1384
Figure 8. Comparision of calculated salinity values (wt% NaCl equ.) of different types of fluid inclusions with the eutectic temperatures (Te, °C) of specific salt solutions. Figure 9. Distribution of Th temperatures of the different types of fluid inclusions versus salinity data from the Dağbaşı skarn ore (gray lines represent the surface fluid dilution trend of the fluid inclusion data). Figure 10. Histogram of homogenization temperature of fluid inclusions in garnet, epidote, quartz, calcite, and sphalerite minerals. Figure 11. Estimated pressure conditions of different fluid inclusion types. Type I inclusions, homogenized via halite dissolution, provide trapping pressure estimates between 884 and 710 bar based on isochores projected according to Sanchez et al. (2012) (lower limits of shaded area). Type II and Type IIIa inclusions are plotted on the liquid‒vapor curve according to the minimum and maximum Th ranges. Pressure estimates for these types of inclusions are between 445 and 195 bar. The isochores for Type IIIc inclusions were plotted using the equation of state of Brown and Lamb (1989) for average Th temperature (240 °C). The highest pressure estimates for Type IIIc inclusions were 310 bar for the highest independent geothermometer of quartz‒epidote mineral pairs (281 °C). The lower limit of pressure estimates cannot be plotted on the diagram because the independent geothermometer is much lower than the average Type IIIb isochore (gray arrows represent the pressure changes through the skarn development). Figure 12. Raman spectra of the CH4-bearing fluid inclusions in garnet (a), epidote (b) and quartz (c, d) minerals. Figure 13. a) Diagram of δ13CPDB vs. δ18OSMOW for the Dağbaşı skarn calcites, and comparison with the isotope composition of mostly known rock types (Sheppard, 1984; Hoefs, 2009). (Red empty circles represent the isotope composition of solutions equilibrated with prograde stage calcites for the Th temperatures of 514 °C. Blue empty symbols represent the isotope composition of solutions equilibrated with skarn calcites for the Ths of fluid inclusions of 311 °C). b) Plots of δ13C vs. δ18O isotopes of skarn calcites on the fluid‒rock interaction model of Taylor (1977). The curves calculated for different X(CO2) and fluid‒rock ratios of 0.01, 0.1, 05 and 0.5, 2, 10, 40, 80 and 100, respectively. Figure 14. δ18O composition of prograde and retrograde stage minerals from the Dağbaşı skarn ore (gray arrows represent decreasing isotopic trend from prograde to retrograde stage; plagioclase and two quartz minerals represent the center of the granitoid. Figure 15. δ18O and δD isotope compositions of retrograde stage epidote and tremolite minerals and the δ18O and δD isotopes of equilibrated solutions. The colored areas on the figure represent the isotope composition of different geological environments (modified from Taylor, 1974; Sheppard, 1984; Hoefs, 2009). (I: igneous, S: sedimentary, SMOW: standard mean ocean water.)
38
1385 1386 1387 1388 1389 1390 1391 1392 1393 1394 1395 1396 1397 1398 1399 1400 1401 1402 1403
Table Captions: Table 1. Microthermometric fluid inclusion data from garnet, epidote, calcite, quartz and sphalerite minerals of Dağbaşı skarn ore. Table 2. Carbon and oxygen isotope composition of skarn calcites from the Dağbaşı Fe–Cu– Zn skarn mineralization. The isotope composition of the water, equilibrated with calcites, was also calculated for the highest Th temperature of fluid inclusions in prograde stage calcite (514 °C) and the average Th temperatures of the retrograde stage calcites (311 °C). The equation of Friedman and O’Neil (1977) was used for the calculations. Table 3. Oxygen and hydrogen isotope composition of the prograde and retrograde stage minerals from the Dağbaşı skarn ore. Plagioclase and two quartz minerals represent the center of granitoid (*The oxygen isotope composition of equilibrated solutions was calculated from different mineral‒water equations of Zheng (1993a, b).
Highlights: -
Both isotope and fluid inclusion data indicate that magmatic sources were responsible for the early stage skarn mineralization
1404 1405 1406
-
Highly depleted isotope data and the decreasing salinity trend of the fluid
1407
inclusions correspond to the mixture of meteoric water during the later stage
1408
skarn development
1409 1410
-
Decreases in the pressure conditions resulted in the boiling of hydrothermal solutions between prograde and retrograde stage skarn development
1411 1412 1413
-
source of CH4 in the fluid inclusions.
1414 1415 1416 1417 1418 1419 1420 1421 1422 1423 1424 1425
Thermal degradation of organic materials in carbonates was suggested as the
-
Declaration of interests ☒ The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. ☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests:
39
1426
1427 1428 1429 1430 1431
40