Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Asian Earth Sciences 73 (2013) 240–251

Contents lists available at SciVerse ScienceDirect

Journal of Asian Earth Sciences


journal homepage: www.elsevier.com/locate/jseaes

Fluid inclusion evidence for hydrothermal fluid evolution


in the Darreh-Zar porphyry copper deposit, Iran
Arash Nateghi ⇑, Ardeshir Hezarkhani
Department of Mining and Metallurgical Engineering, Amirkabir University of Technology (Tehran Polytechnic), Hafez Avenue, No. 424, Tehran, Iran

a r t i c l e i n f o a b s t r a c t

Article history: The Darreh-Zar porphyry copper deposit is associated with a quartz monzonitic–granodioritic–porphy-
Received 24 March 2012 ritic stock hosted by an Eocene volcanic sedimentary complex in which magmatic hydrothermal fluids
Received in revised form 22 April 2013 were introduced and formed veins and alteration. Within the deepest quartz-rich and chalcopyrite-poor
Accepted 23 April 2013
group A veins, LVHS2 inclusions trapped high salinity, high temperature aqueous fluids exsolved directly
Available online 13 May 2013
from a relatively shallow magma (0.5 kbar). These late fluids were enriched in NaCl and reached halite
saturation as a result of the low pressure of magma crystallization and fluid exsolution. These fluids
Keywords:
extracted Cu from the crystallizing melt and transported it to the hydrothermal system. As a result of
Porphyry copper deposit
Fluid evolution
ascent, the temperature and pressure of these fluids decreased from 600 to 415 °C, and approximately
Fluid inclusion 500–315 bars. At these conditions, K-feldspar and biotite were stabilized. Type A veins were formed at
Iran a depth of 1.2 km under conditions of lithostatic pressure and abrupt cooling. Upon cooling and decom-
pressing, the fluid intersected with the liquid–vapor field resulting in separation of immiscible liquid and
vapor. This stage was recorded by formation of LVHS1, LVHS3 and VL inclusions. These immiscible fluids
formed chalcopyrite–pyrite–quartz veins with sericitic alteration envelopes (B veins) under the lithostat-
ic–hydrostatic pressure regime at temperatures between 415 and 355 °C at 1.3 km below the paleowater
table. As the fluids ascended, copper contents decreased and these fluids were diluted by mixing with the
low salinity-external fluid. Therefore, pyrite-dominated quartz veins were formed in purely hydrostatic
conditions in which pressure decreased from 125 bars to 54 bars and temperature decreased from 355
to 298 °C. During the magmatic-hydrothermal evolution, the composition and P–T regime changed dras-
tically and caused various types of veins and alterations. The abundance of chalcopyrite precipitation in
group B veins suggests that boiling and cooling were important factors in copper mineralization in Dar-
reh-Zar.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction Cu deposit in Darreh-Zar on the basis of vein and alteration mineral


assemblages, vein crosscutting relations and fluid inclusion evi-
The Darreh-Zar porphyry copper deposit is located in the Uru- dence. Fluid evolution was derived from fluid compositions and
mieh–Dokhtar arc, 8 km southeast of the giant Sar-Cheshmeh de- P–T evaluation of this system, and a model was elaborated which
posit (Fig. 1), Iran, and contains 67 Mt of sulfide reserves with clarified vein formation, wall-rock alteration and metal
copper grade of 0.37 wt.% (unpublished data from Iranian Copper precipitation.
Company). Previous investigations in Darreh-Zar (Derakhshani
and Abdolzadeh, 2009; Grujicic and Volickovic, 1991; Hemmati,
2. Geological setting
1995; Maanijou, 1993; Ranjbar, 1997; Ranjbar et al., 2001) re-
vealed the lithological and alteration pattern in the area but did
The porphyry copper deposit in Darreh-Zar, the southwestern
not provide adequate data for evaluating the formation history of
part of Kerman, belongs to the Tertiary Urumieh–Dokhtar arc; this
the Darreh-Zar porphyry Cu deposit and physico-chemical condi-
is the southeastern segment of the Kerman Cenozoic magmatic arc
tions of the hydrothermal system responsible for alteration and
(KCMA; Shafiei et al., 2009), (Fig. 1) which was formed during the
mineralization in Darreh-Zar. The aim of this investigation is to
successive stages of Tethyan ocean closure during the Cretaceous–
evaluate the hydrothermal system which formed the porphyry
Oligocene subduction and late Paleogene continent–continent col-
lision (Berberian and King, 1981). Related in origin, porphyry cop-
⇑ Corresponding author. Tel.: +98 21 64542968; fax: +98 21 66405846. per deposits of the Kerman arc show a close temporal and spatial
E-mail addresses: nateghi.arash@yahoo.com (A. Nateghi), ardehez@aut.ac.ir (A. association with the Miocene adakite-like porphyries which in-
Hezarkhani). truded Eocene volcanic rocks (Shafiei et al., 2009).

1367-9120/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jseaes.2013.04.037
A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251 241

Fig. 1. Location of the Kerman Cenozoic magmatic arc in the Urumieh–Dokhtar volcanic belt and geologic map of KCMA with location of some porphyry copper deposits and
prospects (from Shafiei, 2010).

General geology and petrography are summarized here from tem (Ranjbar, 1997). Based on the mode of occurrence and mineral
more detailed descriptions by Hemmati (1995) and Maanijou associations, two main episodes of hydrothermal alteration have
(1993). The Darreh-Zar porphyry copper deposit is associated with been recognized. The first episode contains two pervasive alter-
quartz monzonitic–granodioritic–porphyritic stock hosted by the ation types, namely potassic and propylitic, and represents the ear-
Eocene volcanic sedimentary complex. The quartz monzonitic– lier stage of hydrothermal alteration. Potassic alteration is
granodioritic–porphyritic stock contains phenocrysts of Pl, Hbl, characterized by the conversion of mafic minerals to Bt, replace-
Qz and Bt (50–65 vol.%) and fine-grained groundmass of Qz, Pl, ment of Pl by Kfs + anhydrite (Ranjbar, 1997). The earliest potassic
Kfs, Bt and Amp with minor to trace apatite, zircon, titanite and ru- alteration zone occurs at the deep, central part of quartz monzon-
tile (rock-forming minerals are abbreviated according to Whitney itic–granodioritic–porphyritic stock and has relatively sharp
and Evans (2010)). After emplacement of the stock and develop- boundary with propylitic alteration zone (Derakhshani and Abdol-
ment of the porphyry system, the stock was intruded by a swarm zadeh, 2009). Propylitic alteration is represented by the replace-
of unmineralized dioritic to microdioritic dikes containing pheno- ment of Pl by epidote, conversion of Bt to chlorite and carbonate
crysts of highly altered Pl, Amp, Bt, Qz and minor apatite. Within veins. Minor minerals associated with propylitic alteration are al-
the central part of the stock, the youngest intrusive event is repre- bite, sericite, anhydrite, and pyrite. The second episode of alter-
sented by unmineralized and weakly altered small granodioritic ation comprises a phyllic alteration zone and argillic alteration
stocks and genetically related dikes. Most copper mineralization zone. Phyllic alteration is characterized by conversion of Pl pheno-
occurred during the stock formation (Fig. 2). Three main fault sys- crysts to sericite + Qz with up to 5 vol.% pyrite in veins and dissem-
tems have been recognized in the Darreh-Zar porphyry copper de- inations. The latest alteration stage is represented by an argillic
posit: An E–W striking premineralization fault, a NW–SE striking assemblage of kaolinite and montmorillonite after Fsp (Ranjbar,
mineralized fault and a N–S striking postmineralization fault sys- 1997), and overprints the underlying phyllic and propylitic alter-
242 A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251

Fig. 2. Detailed geological map of the Darreh-Zar area.

ation zones. On the basis of mineralogical data, five different min- 3.1. Type A veins
eralization zones have been distinguished (Unpublished data from
Geological Survey of Iran). In the leached cap rock, the most com- Quartz-dominated (80–95%) pyrite-bearing veins contain, small
mon minerals are hematite, limonite and jarosite. The leached cap amount of chalcopyrite, with minor bornite, molybdenite, anhy-
rock is very thin and is delimited by the present land surface to a drite and Kfs, and with little or no alteration envelopes. These veins
few meters below the surface. The most common minerals in the are discontinuous and show variable thickness of 0.5–3 mm. Type
oxide zone are malachite, azurite, tenorite, chrysocolla, goethite A veins occur mainly in the potassic alteration zone and also in
and chalcanthite. The oxide zone is also very thin and is bounded phyllic and propyllitic alteration zones and are the earliest recog-
by the leached capping zone (upper boundary) and supergene- nized vein type (Fig. 3A).
oxide zone (lower boundary). The supergene-oxide zone between
the oxide zone and supergene zone contains pyrite and chalcocite 3.2. Type B veins
along with minerals characteristic of the oxide zone. This zone is
underlain by the supergene zone and is mainly present within Type B veins contain Qz, chalcopyrite and pyrite with wide seri-
the phyllic zone of alteration. Within the supergene zone, pyrite citic alteration envelopes. These veins are generally more continu-
and chalcocite are the main minerals and occur in cracks or as dis- ous than Type A veins and appear in all alteration zones; however,
seminated sulfide grains. The supergene-hypogene zone between they are restricted mainly in the phyllic and potassic zones. Most
the supergene zone and hypogene zone contains pyrite, chalcocite, Type B veins crosscut and locally offset Type A. The main difference
covellite, chalcopyrite and rare bornite. The supergene-hypogene between these two type veins is in the well-developed sericitic
zone within the phyllic-potassic zone is marked by the conversion alteration envelopes of <few millimeters thick around Type B veins
of chalcopyrite to covellite and chalcocite; sulfides occur as dis- and lack of Kfs. Type B veins are relatively coarser grained and tend
seminated grains and in veins and veinlets. Toward the bottom to be oriented perpendicular to the vein walls. Most sulfide miner-
of this zone, chalcocite + covellite decrease whereas chalcopyrite als occur in a narrow discontinuous layer in the vein centers but
and bornite increase. The hypogene zone developed during the the sulfides are rarely disseminated in quartz (Fig. 3B). The volume
potassic–phyllic and potassic alteration stages and contains pyrite, ratio of chalcopyrite to pyrite is 2:1.
chalcopyrite, bornite, chalcocite and minor molybdenite. The
hypogene copper mineralization occurs as disseminated grains 3.3. Type C veins
and in veinlets.
Type C pyrite-dominated (60–80%) quartz-bearing veins con-
tain minor chalcopyrite, calcite and anhydrite and with pale
3. Veins and wall-rock alteration green–gray alteration envelopes. Most Type C veins are layers of
3–50 mm thick and are most abundant in the phyllic alteration
Based on vein mineral assemblages and vein cross-cutting rela- zone. Type C veins cut the above vein types and offset them in
tions, we have distinguished three types of veins associated with some cases. Quartz is present mainly near the vein margins but
three stages of vein formation. anhydrite, calcite and sulfide minerals form intergrowths in the
A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251 243

Fig. 3. Vein types in the Darreh-Zar porphyry copper deposit. (A) Quartz
dominated-pyrite vein (veins A). (B) Qz-chalcopyrite-pyrite vein with sericitic
alteration envelop. (C) Pyrite-dominated quartz vein with pale green–gray alter-
ation envelopes.

vein centers. Quartz is coarser grained. The only Cu mineral is chal-


copyrite as blebs and inclusions in pyrite (Fig. 3C).

4. Fluid inclusion studies

4.1. Fluid inclusion sampling


Fig. 4. Photomicrographs of fluid inclusions. (A) Liquid rich inclusion. (B) Vapor rich
inclusion. (C) Halite-bearing inclusion with a triangular daughter mineral (chalco-
430 samples from all drill cores were collected. Over 150 doubly pyrite). (D) Halite and sylvite-bearing inclusion with triangular daughter mineral
polished samples were prepared to characterize the types, sizes, (chalcopyrite). (E) Halite-bearing inclusion with an opaque daughter mineral. (F)
abundances and spatial distribution of fluid inclusions. Approxi- Halite-bearing inclusions with vapor rich inclusions. (G) A cluster of halite-bearing
mately 270 inclusions from all vein types were analyzed by mic- and vapor rich inclusions. (H) Primary LVHS1, LVHS3 and VL fluid inclusions
trapped along growth zone of quartz. (I) Primary LVHS2 fluid inclusions trapped in a
rothermometry. The studied samples contained fluid inclusions quartz crystal.
of 1–15 lm in diameter; most are between 4 and 12 lm.

4.2. Fluid inclusion types 4.2.1. LV fluid inclusions


LV fluid inclusions contain liquid and a vapor bubble which
Based on the room temperature phase assemblage, three types occupies less than 40% of inclusion volume (Fig. 4A). In a small
of fluid inclusion are categorized in Darreh-Zar as types LV, VL, and number of inclusions, a transparent and/or an opaque mineral
LVHS; the letter ‘‘L’’ refers to the liquid phase, ‘‘V’’ indicates the was observed, but these are thought to have been accidentally
presence of vapor, ‘‘H’’ specifies halite daughter mineral(s), and trapped. These fluid inclusions homogenize to liquid. Most LV
‘‘S’’ indicates other solid phase(s). inclusions have been identified as secondary, distributed along
244 A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251

healed fractures and rarely being recognized as primary along 4.2.3. LVHS fluid inclusions
euhedral growth zones of quartz. LVHS fluid inclusions consist of liquid, a vapor bubble and halite
at room temperature. Most LVHS inclusions contain up to four so-
4.2.2. VL fluid inclusions lid phases besides halite. Based on the type of solids, LVHS inclu-
VL fluid inclusions contain liquid and vapor bubbles which oc- sion is divided to three subtypes, which include
cupy more than 55% of the inclusion volume (Fig. 4B). Few VL liquid + vapor + halite + chalcopyrite ± anhydrite ± unknown phase
inclusions contain unidentified trapped solids. Most of the VL (type LVHS1, Fig. 4C), liquid + vapor + halite + sylvite + chalcopy-
inclusions occur in the growth zones of quartz and are primary rite ± anhydrite ± unknown phase ± opaque (type LVHS2, Fig. 4D)
in origin. They are elongated in shape; some have negative crystal and liquid + vapor + halite ± opaque (type LVHS3, Fig. 4E). The exis-
shapes. They homogenize to vapor and rarely to liquid. tence of LVHS inclusions in the growth zones of quartz indicates

Fig. 5. Histogram of salinity of LV, VL, LVHS1, LVHS2 and LVHS3 fluid inclusions.
A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251 245

Fig. 6. Histogram of homogenization temperature of LV, VL, LVHS1, LVHS2 and LVHS3 fluid inclusions.

their primary origin. LVHS1 and LVHS3 inclusions are almost al- C veins from phyllic and propylitic alterations zones and occur
ways in groups or clusters which also contain VL inclusions rarely in Type A veins.
(Fig. 4F, G, H and I). In contrast, most LVHS2 inclusions are scat- VL inclusions in Type A veins from potassic alteration zone oc-
tered in groups or clusters which contain only one inclusion type. cur along micro-fractures; they are also present in clusters with
LVHS1 fluid inclusions. In addition, they are common in Type B
veins from potassic alteration zone and Type B and C veins from
4.3. Distribution of fluid inclusion types phyllic alteration zone at shallower levels.
LVHS inclusions occur in all vein types from the potassic alter-
LV inclusions are common in all vein groups but are found in ation zone (the deepest part of the intrusive) to the shallow level
inconsistent proportions. They are most abundant in Type B and veins. LVHS1 and LVHS2 inclusions are dominant in veins from
246 A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251

deeper parts but LVHS3 inclusions commonly occur at shallower 315 °C (371 °C in average) in B veins and 520–381 °C (453 °C in
levels. average) in C veins (Fig. 6B). Although all type VL inclusions
homogenized to vapor, in rare cases, no changes were observed
4.4. Fluid inclusions microthermometry even when heated to 600 °C. The assemblage with the highest
salinity (14 ± 2 wt.% NaCl equiv.) homogenized at 364° ± 18 °C.
Temperature of phase changes in fluid inclusions were mea- The assemblage with the lowest Th L-V (360° ± 3 °C) had the lowest
sured with Fluid Inc. U.S.G.S. type gas-flow stage which operated salinity of 3 ± 2 wt.% NaCl equiv., and the assemblage with the
by passing preheated or precooled N2 gas around the sample highest Th L-V (515° ± 31 °C) had a salinity of 11 ± 2 wt.% NaCl
(Werre et al., 1979). Stage calibration was performed using syn- equivalent.
thetic fluid inclusions. Temperature accuracy at the standard refer-
ences was ±0.2 °C at 56.6 °C (melting point of CO2), ±0.1 °C at 0 °C 4.4.3. LVHS inclusions
(triple point of H2O), ±2 °C at 374.1 °C (critical homogenization of The liquid–vapor homogenization temperature was between
H2O) and ± 9 °C at 573 °C (quartz a/b transition). The heating rate 435 and 273.5 °C (317 °C in average) and 487 and 237 °C (398 °C
was approximately 1 °C/min near the temperatures of phase tran- in average) for LVHS1 and LVHS2 inclusions in A veins, respec-
sitions (Table 1 and Fig. 7). tively. In B veins, the Th L-V (°C) values varied between 390 and
258 °C (321 °C in average) for type LVHS1 inclusions and 485 and
4.4.1. LV inclusions 320 °C (415 °C in average) for type LVH2 inclusions. The liquid
Clear evidence of first melting was observed at temperature less and vapor phases in type LVHS3 inclusions homogenized to liquid
than 21.2 °C (eutectic temperature for system H2O–NaCl), indi- at the temperature between 470 and 211 °C in A veins, 485 and
cating the presence of dissolved salts in addition NaCl and KCl 244 °C in B veins and 345 and 225 °C in C veins. In type LVHS1
(Sterner and Bodnar, 1984). Clathrate was not observed in fluid inclusions, halite dissolution temperatures ranged between 453
inclusions, which indicated that these inclusions do not contain and 291 °C in A veins and 450 and 223 °C in B veins. In type LVHS2
significant CO2. Crushing of quartz under anhydrous glycerin con- inclusions, halite and sylvite were dissolved at T from 456 to 259
firmed this conclusion. The final ice melting temperatures (Tm ice) and 98 to 50.3 °C and also ranged from 499 to 255 and 119 to
for these inclusions were between 1.8 and 8.1 °C in A veins, 55.3 °C in A veins and B veins, respectively (Fig. 6C, D and E). Con-
1.2 and 9.7 °C in B veins and 1.3 and 12.3 °C in C veins, cor- sequently, estimates of salinity based on these temperatures varied
responding to salinities of 11.8–3.1 wt.% NaCl equiv., 13.6–2.1 wt.% between 52.9 and 37.5 wt.% NaCl equiv. in type LVHS1 inclusions
NaCl equiv. and 16.2–2.2 wt.% NaCl equiv., respectively (Bodnar, and 54.2–35.5 wt.% NaCl equiv. in type LVHS2 inclusions (A veins)
1993) (Fig. 5A). All types LV inclusions homogenized to liquid at (Sterner et al., 1988; Lecumberri-Sanchez et al., 2012). Note that
range of 447.9–245 °C (mostly 250–325 °C) in A veins, 514– the equation of Sterner et al. (1988) is only valid under vapor-sat-
245 °C (mostly 245–375 °C) in B veins and 350–215 °C in C veins urated condition (Th L-V > Tm halite). Therefore, salinities of fluid
(Fig. 6A). The assemblage with the highest Th L-V (423° ± 25 °C) also inclusions that homogenize by halite dissolution (Tm halite > Th L-V)
had the highest salinity of 11 ± 0.9 wt.% NaCl equiv., and the were calculated using the equation of Lecumberri-Sanchez et al.
assemblage with the lowest Th L-V (246° ± 7 °C) had salinity of (2012) and the Microsoft Excel spreadsheet HOKIEFLINC_H2O–
7 ± 0.7 wt.% NaCl equiv. The assemblage with the lowest salinity NACL of Steele-MacInnis et al. (2012).
(5 ± 0.8 wt.% NaCl equiv.) homogenized at 332° ± 25 °C. The salinity of fluid inclusions in B veins was between 52.8 and
33.1 wt.% NaCl equiv. for type LVHS1 inclusions and 59.8–35 wt.%
4.4.2. VL inclusions NaCl equiv. in type LVHS2 inclusions (Fig. 5C and D). The temper-
Evidence of first melting temperature less than eutectic temper- atures of halite dissolving in type LVHS3 inclusions ranged from
ature for system H2O–NaCl indicates a presence of dissolved salts 450 to 283 °C, corresponding to the salinity of 53.4–36.9 wt.% NaCl
besides NaCl and KCl (Sterner and Bodnar, 1984). The Tm ice value equiv. (A veins), and 510–280 °C, corresponding to the salinity of
varied between 1.1 and 13.4 °C in A veins, 0.5 and 14.9 °C 61–37.2 wt.% NaCl equiv. in B veins. In C veins, Tm halite values ran-
in B veins and 0.9 and 13.2 °C in C veins, corresponding to the ged between 381 and 211 °C, corresponding to the salinity of 46.1–
salinity from 17.3 to 1.9 wt.% NaCl equiv., 18.6–0.9 wt.% NaCl 32.4 wt.% NaCl equiv. (Fig. 5E). The assemblage of LVHS2 inclusions
equiv. and 17.1–1.6 wt.% NaCl equiv. (Bodnar, 1993) (Fig. 5B). Type with the highest salinity (55 ± 2 wt.% NaCl equiv.) had the highest
VL inclusions showed variable homogenization temperatures rang- homogenization temperature of 463° ± 19 °C, and the assemblage
ing between 575 and 383 °C (456 °C in average) in A veins, 455 and with the lowest salinity (40 ± 2 wt.% NaCl equiv.) had the lowest

Table 1
Microthermometric data of fluid inclusions.

Group Fluid inclusions type Te (°C) Tm ice (°C) Tm halite (°C) Tm sylvite (°C) Th L-V Salinity (wt.% NaCl)
Range Avg. Range Avg. Range Avg. Range Avg. Range Avg. Range Avg.
A LV 15 to 27.5 (16)a 22.9 1.8 to 8.1 4.7 – – – – 447.9 to 245 305.2 11.8 to 3.1 7.4
A VL 17 to 48 (31) 29.2 1.1 to 13.4 7.5 – – – – 575 to 383 456.4 17.3 to 1.9 10.7
A LVHS1 40 to 65 (17) 49.2 6.4 to 24 14.9 453 to 291 360.3 – – 435 to 273.5 317.3 53.6 to 37.3 43.7
A LVHS2 44 to 64 (9) 55.4 8 to 25.4 16.8 456.6 to 259 345.6 98 to 50.3 72.7 487 to 237 398.8 54.1 to 35.2 42.4
A LVHS3 37 to 67.4 (20) 52.5 3.6 to 28 11.7 450 to 283 372.9 – – 470 to 211 336.2 53.2 to 36.8 44.8
B LV 16 to 48.5 (16) 23.3 1.2 to 9.7 3.8 – – – – 514 to 245 311.8 13.6 to 2.1 6.0
B VL 19 to 56.5 (27) 27.1 0.5 to 14.9 5.4 – – – – 455 to 315 371.9 18.6 to 0.9 7.8
B LVHS1 39.2 to 65.2 (22) 49.3 5 to 28 15.1 450 to 223 375.9 – – 390 to 258 321.9 53.2 to 33.1 45.4
B LVHS2 30.4 to 56 (11) 50.1 3.5 to 17.5 11.2 499 to 255.5 418.5 119 to 55.3 88.1 485 to 320 415.3 59.7 to 35 50.1
B LVHS3 38 to 62 (14) 49.7 3.7 to 33 12.8 510 to 280 385.0 – – 485 to 244 331.3 61.2 to 36.6 46.6
C LV 19 to 31.1 (28) 22.0 1.3 to 12.3 5.7 – – – – 350 to 215 278.3 16.2 to 2.2 8.5
C VL 23.5 to 30.6 (9) 25.8 0.9 to 13.2 6.6 – – – – 520 to 381 453.7 17.1 to 1.6 9.5
C LVHS3 26 to 43 (27) 35.9 4.1 to 12.1 7.6 381 to 211 275.8 – – 345 to 225 274.5 45.3 to 32.5 36.6
a
Number of fluid inclusions.
A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251 247

homogenization temperature of 392° ± 67 °C. Among the LVHS1


assemblages, the assemblage with the highest homogenization
temperature (421° ± 21 °C) had the highest salinity of 50 ± 2 wt.%
NaCl equiv., and the assemblage with the lowest homogenization
temperature (319° ± 30 °C) had the lowest salinity of 38 ± 4 wt.%
NaCl equivalent. The highest and the lowest homogenization tem-
perature were measured in the assemblages of LVHS3 inclusions at
474° ± 11 °C and 266° ± 4 °C, respectively.
Within most LVHS assemblages, fluid inclusions homogenized
both by halite dissolution and by vapor disappearance. Most LVHS1
and LVHS3 assemblages homogenized to liquid by halite dissolu-
tion and the rest of them homogenized by vapor disappearance
over a small temperature interval (Fig. 8A and C), whereas most
LVHS2 fluid inclusions exhibited near simultaneous halite and va-
por disappearance temperature (Fig. 8B).

4.5. K/Na ratios of high salinity fluids

Based on the dissolution temperatures of halite and sylvite in Fig. 7. Homogenization temperature vs. salinity of LV, VL, LVHS1, LVHS2 and LVHS3
the LVHS2 inclusions containing both daughter minerals, the K/ fluid inclusions.
Na ratios of high salinity fluids were calculated using the computer
program SALTY (Bodnar et al., 1989). The content of NaCl ranged
other hand, the first fluids exsolved at moderate to low pressures
between 52.68 and 27.98 wt.% and that of KCl was from 16.95 to
have low salinities and the salinity of the exsolving fluid increases
12.34 wt.%, corresponding to K/Na atomic ratios ranging from
as crystallization proceeds. While evidence for direct exsolution of
0.39 to 0.20 (0.28 in average). Although first melting temperatures
a high salinity fluid under a deep (>1.3 kbar) conditions from a
indicated that the significant amount of Fe, Ca, Mn and other ele-
magma has been observed in some porphyry copper deposits
ments were present in LVHS2 inclusions, the bulk composition of
(e.g., Bajo de la Alumbrera, Argentina), in which moderately high
LVHS2 inclusions was estimated by ignoring the effect(s) of other
salinity (50–60 wt.% NaCl equiv.) inclusions are interpreted as the
species on the solubility of NaCl and KCl. Therefore, LVHS2 inclu-
most parental ore fluids (Ulrich et al., 2002), Cline and Bodnar
sions most likely had higher total salinities and lower NaCl and
(1991) asserted that the generation of extremely high salinity flu-
KCl contents than the estimated ones (Kwak et al., 1986; Quan
ids in this environment is difficult and the highest salinity they
et al., 1987). The absence of sylvite daughters in LVHS1 and LVHS3
might expect would be about 30–40 wt.% NaCl equivalent. Follow-
inclusions at room temperature limited the K/Na ratios lower than
ing the assumptions of Cline and Bodnar (1991) for a crystallizing
0.2 (Roedder, 1984).
silicic magma (initial water content and initial chlorine/water
parameters), we inferred that LVHS2 inclusions trapped aqueous
5. Discussion fluids exsolved directly from a relatively shallow magma
(0.5 kbar) and occurred late in the crystallization history. This al-
5.1. Magmatic origin of high salinity fluids lowed efficient extraction of Cu into the latest fluids to exsolve
from magma. Therefore, the high salinities and high metal contents
High salinity LVHS2 inclusions occur primarily in the deepest of fluids recorded by the LVHS2 inclusions suggest that these fluids
part of the stock from the A and B veins. Assuming that the LVHS2 may be the most likely parental ore fluid at Darreh-Zar. As the sin-
inclusion composition is in the NaCl–KCl–H2O system (see Sec- gle-phase magmatic fluid ascended to shallower levels (deposit
tion 4.5), the atomic ratios varied from 0.39–0.20 (0.28 in average). area), the P–T conditions intersected the liquid–vapor curve and
In general, these atomic ratios are comparable with porphyry cop- the fluids separated into immiscible liquid and vapor.
per mineralization in many localities including Granisle, British
Columbia (Wilson et al., 1980), Bell, British Columbia (Wilson
et al., 1980), Calabona, Italy (Stefanini and Williams-Jones, 1996), 5.2. Entrapment pressure for saline inclusions
Sungun, Iran (Hezarkhani and Williams-Jones, 1998). Furthermore,
the K/Na atomic ratios suggested that high salinity fluids were in Among the LVHS2 assemblages, one assemblage showed consis-
equilibrium with two feldspars (or granitic magma) only at the tent values of Th L-V (400 °C) with a wide range of Tm halite (50 °C).
temperature above 600 °C (Lagache and Weisbrod, 1977). This Therefore, assuming a lithostatic pressure regime and average rock
temperature overlapped with lower T range in which the biotite– density of 2.7 g/cm3, the lowest Tm halite in this assemblage (415 °C)
feldspar porphyry was molten (Wilson et al., 1980). Based on these corresponds to the actual trapping temperature and pressure of
observations and high minimum pressure (see Section 5.2), we 315 bars and depth of approximately 1.2 km (see Lecumberri-San-
concluded that LVHS2 inclusions may represent the most primitive chez et al., 2012).
magmatic fluid recorded at the level of the deposit. VL inclusions are observed closely associated with saline inclu-
Thermodynamic modeling (Cline and Bodnar, 1991) indicated sions (occurring in clusters with LVHS1 inclusions and in growth
that boiling is not necessary for producing high salinity fluids rec- zones or along healed fractures with LVHS3 inclusions) and this
ognized in many porphyry copper deposits, and depending on tem- may indicate that some of these inclusions are formed in an immis-
perature, confining pressure, initial water content and extent of cible system or boiling. Therefore, the pressure and temperature of
crystallization of the melt, the bulk salinity of aqueous fluids ex- homogenization are equal to the pressure and temperature of for-
solved from the melt may vary from <2 wt.% NaCl to nearly pure mation (Roedder and Bodnar, 1980). On the other hand, Roedder
NaCl melts. The first fluid exsolved from a melt under a deep con- and Bodnar (1980) mentioned that if petrographic data suggests
dition (>1.3 kbar) has a high salinity and fluids exsolved as crystal- boiling and the halite-bearing inclusions homogenize by halite dis-
lization progresses have lower chlorine concentrations. On the solution then the saline and vapor rich inclusions cannot signify an
248 A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251

immiscible pair because phase equilibrium constraints require that of boiling, the homogenization temperature of boiling assemblages
the inclusions which homogenize by halite dissolution trapped in (coexistence of LVHS1 and LVHS3 inclusions that homogenize by
the liquid stable, vapor absent field. Here, despite the microther- vapor disappearance and VL inclusions) were used for estimating
mometric behavior of halite-bearing inclusions, we propose that the entrapment P and trapping depth of the fluids related to the
LVHS1, LVHS3 and VL inclusions were generated by boiling (in Sec- formation of different vein types.
tion 5.3 we discuss the possibility that these inclusions represent To estimate trapping pressures for the inclusions that homoge-
overprinting of boiling assemblages by halite-saturated brines nize by vapor disappearance and are associated with vapor rich
resulting from segregation and overpressuring of the high-salinity inclusions, the procedure of Bodnar and Vityk (1994) and correla-
fluids formed by phase separation), and based on the assumption tion formula for phase relations in the PTX space of Driesner and
Heinrich (2007) (SoWat programs) were used. The boiling assem-
blage in deep A veins with the highest homogenization tempera-
ture (400° ± 33 °C) and salinities of 42 ± 2 wt.% NaCl equiv.
indicated the trapping P of approximately up to 193 bars (Fig. 9).
Other boiling assemblages in A veins indicated pressures of 138–
165 bars at temperatures of 369.5° ± 65.5 °C to 382° ± 31 °C. The
lowest pressure (64 bars) in A veins is represented by one boiling
assemblage (304 °C). Crustal rocks at T > 400 °C were sufficiently
hot to behave plastically at normal strain in which lithostatic pres-
sure prevails (Fournier, 1999). Assuming an average rock density of
2.7 g/cm3, and a lithostatic pressure regime, these inclusions in A
veins were formed at depths of approximately 1 km. The hottest
boiling assemblages in B veins, with salinities of 39 ± 5 wt.% NaCl
equiv. and Th L-V of 355.5° ± 25.5 °C, correspond to a pressure of
125 bars (Fig. 9). A boiling assemblage with the lowest homogeni-
zation temperature (approximately 320 °C) in B veins indicates
pressure of 86 bars. At temperatures below 400 °C, a hydrostatic
pressure regime dominates because rocks behaved in a brittle
manner (Fournier, 1999). Therefore, assuming a fluid density near
1 g/cm3, this pressure corresponded to 1.3 km below the paleo-
water table. In C veins, cooler boiling assemblages showed vapor
disappearance temperatures of up to 289° ± 10 °C and salinities
of up to 35 ± 0.9 wt.% NaCl equiv., corresponding to hydrostatic
pressure of 54 bars (Fig. 9) and depth of 0.6 km below the paleo-
water table.

5.3. Capability of boiling for producing the observed saline fluids

The coexistence of vapor rich inclusions with saline inclusions


that homogenize by vapor disappearance in addition to inclusions
that homogenize by halite dissolution suggests a significant shift in
trapping P–T conditions during the magmatic hydrothermal evolu-
tion. There are two theoretical P–T paths for the origin of this

Fig. 8. Plot of Th L-V vs. Tm halite saline fluid inclusions. Dashed line indicated the Fig. 9. Pressure estimates for fluid inclusions homogenizing by vapor disappear-
difference between halite dissolution and vapor disappearance temperature for ance using the SoWat programs. Phase diagram is from correlation formula of
fluid inclusions homogenizing by halite dissolution (see text). Driesner and Heinrich (2007).
A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251 249

behavior: (1) an isobaric path and (2) an isothermal path. If these had trapping pressures ranging between 225 and 3036 bars in
features were the result of isobaric evolution, then the earliest fluid veins A and B. LVHS3 fluid inclusions homogenized by halite disso-
inclusions are those that homogenize by vapor disappearance, lution demonstrated the formation pressure between 131 and
whereas, if the features result from isothermal evolution, then 1909 bars in veins C. A number of saline inclusions homogenized
the earliest fluid inclusions are those that homogenize by halite by halite dissolution showed high pressure (up to 1400 bars).
dissolution (Bodnar, 1994; Becker et al., 2008). Therefore, if we These inclusions may reflect the over-pressuring condition before
propose that boiling occurred at the earliest stage, then the earlier the cracking of the rock (Bouzari and Clark, 2006; Cline and Bodnar,
inclusions would be those homogenizing by vapor disappearance. 1994). Fluid inclusions with higher pressures (>1400 bars) could be
Cline and Bodnar (1994) evaluated the possibility of boiling to gen- produced by halite saturation of fluid, which poisons the growing
erate halite-bearing inclusions that homogenize by halite dissolu- crystal surface and, after that, grows around the foreign halite crys-
tion at the Questa deposit. Their observations and assumptions tal (see Becker et al., 2008), or experienced stretching or leakage
concluded that generation of observed saline fluids by boiling (Lecumberri-Sanchez et al., 2012). If Th L-V is plotted vs. Tm halite
was improbable. Here, we apply their procedure to assess the op- at constant Th L-V, these inclusions will show higher Tm halite, or var-
tion of boiling to generate halite-bearing inclusions that homoge- iable Th L-V and Tm halite within an assemblage. Therefore, the pres-
nize by halite dissolution. sure estimated from these inclusions will be higher than the actual
The densest cluster of Th L-V vs. Tm halite values showed that Th L-V pressure (Becker et al., 2008; Lecumberri-Sanchez et al., 2012).
occurred between 372 and 258 °C, and Tm halite occurred between
431 and 345 °C. Using the experimental PVTX data of Becker
5.4. Evolution of the magmatic hydrothermal system
et al. (2008) for the system H2O–NaCl, the pressures in which
inclusions left the halite liquidus and, consequently, the isochore
It is important to note that the Darreh-Zar magmatic hydrother-
through the L + H field were determined (Fig. 10). Based on the
mal system is a closed system in which the salinity and copper
experimental data of Bodnar et al. (1985), a liquid with the salinity
contents of the exsolved aqueous fluid decrease during the mag-
of 45 wt.% NaCl equiv. (mean saline inclusions at Darreh-Zar) at P
matic crystallization. The idea of this being a closed system is sup-
of approximately 200 bars is produced by boiling at 410 °C. As a re-
ported by the lack of occurrence of contemporary volcanic rocks
sult, the temperature of the boiled fluid must decrease from 410 °C
with mineralized porphyry intrusions within the Kerman Cenozoic
to 380 °C, (Tm halite producing the salinity of 45 wt.% NaCl equiv.) in
magmatic arc (Hassanzadeh, 1993; Shafiei, 2010; Waterman and
order to produce inclusions which homogenize by halite dissolu-
Hamilton, 1975). High salinity and high-T fluids which exsolved di-
tion (Fig. 10). The existence of assemblages homogenizing by vapor
rectly from crystallizing melt were responsible for quartz-rich,
disappearance (boiling assemblage) up to 400 °C was consistent
chalcopyrite-poor, group A veins. As fluids ascended, the tempera-
with this behavior.
ture and pressure of magmatic-hydrothermal fluids decreased
Therefore, according to the above discussion and the existence
from approximately above 600–415 °C and 500–315 bars, respec-
of vapor rich inclusions with saline inclusions, we propose that
tively. Therefore, cooling and decompression moved the fluids
fluid inclusions with Th L-V > Tm halite are produced by boiling and
from the single-phase P–T field into the field of immiscible liquid
fluid inclusions with Tm halite > Th L-V are trapped at a later time,
and vapor. Group B veins (Qz + chalcopyrite + pyrite) with sersitic
after the vapor phase has been removed (perhaps owing to its
alteration envelope were formed when hydrothermal fluids were
greater buoyancy and lower viscosity compared to the liquid),
cooled from 415 to approximately 355 °C and decompressed from
and after continuous T decrease and P increase of the boiled fluid.
315 to 125 bars (the brittle-plastic transition). At shallower levels
In addition, the full data set was used for LVHS1 and LVHS3 fluid
and in pyrite-dominated group C veins, homogenization tempera-
inclusions homogenized by halite dissolution in order to estimate
ture and salinity of LVHS inclusions significantly decreased. These
the minimum formation pressure (Fig. 11). Therefore, using the
may point out that LVHS inclusions prior to trapping in veins C
model of Lecumberri-Sanchez et al. (2012) (based on the experi-
were cooled and diluted with a low salinity-external fluid. The sec-
mental data for the system H2O–NaCl in Becker et al. (2008)) and
ondary origin of liquid rich inclusions and the abundance of these
the Microsoft Excel spreadsheet HOKIEFLINC_H2O–NACL of
inclusions at shallower levels suggest an external origin, and the
Steele-MacInnis et al. (2012), LVHS1 and LVHS3 fluid inclusions

Fig. 11. Pressure estimates for fluid inclusions homogenizing by halite dissolution
Fig. 10. P–T diagram for a solution containing 45 wt.% NaCl equiv. (see text). temperature with isobars from Becker et al. (2008).
250 A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251

variation in salinity and homogenization T of these inclusions re- diluting the high temperature fluids and forming the propylitic and
flects the variable degree of mixing with saline fluids. Therefore, argillic alteration zones.
pyrite-dominated veins were formed in purely hydrostatic condi-
tions in which P and T decreased from 125 bars to 54 bars and from
Acknowledgements
355 to 298 °C, respectively. The regular decrease for the difference
between halite dissolution and vapor disappearance temperature
This research was funded by Amirkabir University research
for fluid inclusions homogenizing by halite dissolution from the
grant to Ardeshir Hezarkhani. Special thanks to Matthew Steele-
deep veins to a shallower level veins, (DT = 165 °C in A veins,
MacInnis and David Lentz for their suggestions and reviews. This
DT = 150 °C in B veins and DT = 115 °C in C veins) reflected the ef-
manuscript was greatly improved by the reviews of Robert J. Bod-
fect of greater pressure and prompts cooling at deep veins, which
nar, and an anonymous reviewer and the editorial suggestions of
stabilizes the Kfs and Bt (potassic alteration) and the slight cooling
Juhn G. Liou.
and convergence of the temperature of fluid and rock at shallower
levels, as proposed for the Cerro Colorado porphyry Cu deposit by
Bouzari and Clark (2006). The obvious variations in fluids compo- Appendix A. Supplementary material
sition, different cooling paths (abrupt or gradual) and different
pressure regimes (lithostatic, lithostatic–hydrostatic, hydrostatic) Supplementary data associated with this article can be found, in
from the potassic alteration to phyllic alteration revealed the roles the online version, at http://dx.doi.org/10.1016/j.jse-
of composition, temperature and pressure in order to figure alter- aes.2013.04.037. These data include Google maps of the most
ation mineral assemblages. Contemporaneous with the main important areas described in this article.
hydrothermal stage, low salinity and low-T meteoric water (re-
corded by LV fluid inclusions in A and B veins) caused propylitic
References
alteration of the peripheral part of the stock. The presence of chal-
copyrite as the daughter mineral in LVHS2 and LVHS1 inclusions Becker, S.P., Fall, A., Bodnar, R.J., 2008. Synthetic fluid inclusions. XVII. PVTX
indicated that copper was transported by these fluids. In the Dar- properties of high salinity H2O–NaCl solutions (>30 wt.% NaCl): application to
reh-Zar porphyry copper deposit, the main chalcopyrite precipita- fluid inclusions that homogenize by halite disappearance from porphyry copper
and other hydrothermal deposits. Economic Geology 103, 539–554.
tion occurred in group B veins. The experimental data (Cline and Berberian, M., King, G.C., 1981. Toward a paleogeography and tectonic evolution of
Bodnar, 1991) showed that copper precipitates at higher tempera- Iran. Canadian Journal of Earth Sciences 18, 210–265.
tures (>350 °C) in highly saline fluids with high Cu concentrations. Bodnar, R.J., 1993. Revised equation and table for determining the freezing point
depression of H2O–NaCl solution. Geochimica et Cosmochimica Acta 57, 683–
We argue that, in Darreh-Zar, copper was precipitated mainly at T
684.
between 400 and 350 °C. Therefore, based on the low chalcopyrite Bodnar, R.J., 1994. Synthetic fluid inclusions: XII. The system H2O–NaCl.
content of A veins and the abundance of chalcopyrite in B veins, Experimental determination of the halite liquidus and isochores for a 40 wt.%
which was contemporary with boiling, we infer that boiling is NaCl solution. Geochimica et Cosmochimica Acta 58, 1053–1063.
Bodnar, R.J., Vityk, M.O., 1994. Interpretation of microthermometric data for H2O–
the major factor for copper precipitation in Darreh-Zar. By copper NaCl fluid inclusions. In: DeVivo, B., Frezzotti, M.L. (Eds.), Fluid inclusions in
precipitation, the copper content of parental magmatic fluids re- minerals: Methods and Applications: Blacksburg. Virginia Polytechnic Institute
duces and the absence of chalcopyrite daughter mineral in LVHS3 and State University, VA, pp. 117–130.
Bodnar, R.J., Burnham, C.W., Sterner, S.M., 1985. Synthetic fluid inclusions in natural
inclusions happens. quartz. III. Determination of phase equilibrium properties in the system H2O–
NaCl to 1000 °C and 1500 bars. Geochimica et Cosmochimica Acta 49, 1861–
1873.
Bodnar, R.J., Sterner, S.M., Hall, D.L., 1989. SALTY: a FORTRAN program to calculate
6. Conclusions composition of fluid inclusions in the system NaCl–KCl–H2O. Computers &
Geosciences 15, 19–41.
The temperature, fluid composition and pressure evolution in Bouzari, F., Clark, A.H., 2006. Prograde evolution and geothermal affinities of a
major porphyry copper deposit: The Cerro Colorado hypogene protore, I Region,
the Darreh-Zar porphyry copper deposit were similar to other por- northern Chile. Economic Geology 101, 95–134.
phyry copper deposits in the Urumieh–Dokhtar arc (Etminan, Cline, J.S., Bodnar, R.J., 1991. Can economic porphyry copper mineralization be
1978; Hezarkhani, 2006, 2009; Hezarkhani and Williams-Jones, generated by a typical calc-alkaline melt? Journal of Geophysical Research 96,
8113–8126.
1998). However, some homogenization behaviors were different
Cline, J.S., Bodnar, R.J., 1994. Direct evolution of brine from a crystallizing silicic
in comparison, reflecting the local P–T fluctuation during the melt at the Questa, New Mexico, Molybdenum deposit. Economic Geolology 89,
hydrothermal evolution in Darreh-Zar. The present study showed 1780–1802.
that LVHS2 inclusions trapped high salinity and high-T aqueous Derakhshani, R., Abdolzadeh, M., 2009. Geochemistry, mineralization and alteration
zones of Darrehzar porphyry copper deposit, Kerman, Iran. Journal of Applied
fluids exsolved directly from a crystallizing magma, as a single Sciences 9, 1628–1646.
phase, at relatively low pressures. These late fluids become satu- Driesner, T., Heinrich, C.A., 2007. The system H2O–NaCl. Part I: Correlations formula
rated in halite as a result of the low pressure of magma crystalliza- for phase relation in temperature–pressure–composition space from 0 to
1000 °C, 0 to 5000 bars, and 0 to 1 XNaCl. Geochimica et Cosmochimica Acta 71,
tion, and they caused depletion of copper from crystallizing melt 4880–4901.
and transportation of copper to the hydrothermal system. The Etminan, H., 1978. Fluid inclusion studies of the porphyry copper ore bodies at Sar-
abundance of LVHS1, LVHS3 and VL inclusions in chalcopyrite-pyr- Cheshmeh, Darreh Zar and Mieduk (Kerman region, southeastern Iran) and
porphyry copper discoveries at Miduk, Gozan, and Kighal, Azarbaijan region
ite-Qz veins (B veins) indicated the important role of boiling in (northwestern Iran). International Association on the Genesis of Ore Deposits
addition to cooling in chalcopyrite precipitation in the Darreh- (IAGOD), 5th, p. 88.
Zar porphyry copper deposit. Different cooling paths and different Fournier, R.O., 1999. Hydrothermal process related to movement of fluid from
plastic into brittle rock in the magmatic-epithermal environment. Economic
pressure regimes in addition to variation in the composition of Geology 94, 1193–1211.
fluid inclusions in different stages of vein formation were distin- Grujicic, B., Volickovic, S., 1991. Copper deposit Darrehzar mineral inventories
guished. An abrupt cooling path under the lithostatic pressure re- computation. National Iranian Copper Industries, Exploration Department,
Internal Report.
gime stabilized Kfs and Bt and formed quartz-rich, chalcopyrite-
Hassanzadeh, J., 1993. Metallogenic and tectono-magmatic events in the SE sector
poor group A veins, at deep levels. During fluid ascent, cooling of the Cenozoic active continental margin of Iran (shahr e Babak area, Kerman
via the convergence of fluid and rock temperature, as well as province). Unpublished Ph.D. dissertation, Los Angeles, University of California,
decompression in the hydrostatic regime, formed the pyrite-dom- p. 204.
Hemmati, J., 1995. Economic Geology of Darrehzar porphyry copper deposit.
inated, quartz veins (C veins) and phyllic alteration. Also, data from Unpublished M.Sc. thesis (in Persian), Tehran, Shahid Beheshti University, p.
fluid inclusions indicated the role of meteoric water in cooling and 150.
A. Nateghi, A. Hezarkhani / Journal of Asian Earth Sciences 73 (2013) 240–251 251

Hezarkhani, A., 2006. Hydrothermal evolutions at the Sar-Cheshmeh porphyry Cu– Shafiei, B., 2010. Lead isotope signature of the igneous rocks and porphyry copper
Mo deposit, Iran: evidence from fluid inclusions. Journal of Asian Earth Sciences deposits from the Kerman Cenozoic magmatic arc (SE Iran), and their
28, 408–422. magmatic-metallogenetic implications. Ore Geology Reviews 28, 27–36.
Hezarkhani, A., 2009. Hydrothermal fluid geochemistry at Chah-Firuzeh porphyry Shafiei, B., Haschke, M., Shahabpour, J., 2009. Recycling of orogenic arc crust triggers
copper deposit, Iran: evidence from fluid inclusions. Journal of Geochemical porphyry Cu mineralization in Kerman Cenizoic arc rocks, southeastern Iran.
Exploration 101, 254–264. Mineralium Deposita 44, 265–283.
Hezarkhani, A., Williams-Jones, A.E., 1998. Controls of alteration and mineralization Steele-MacInnis, M., Lecumberri-Sanchez, P., Bodnar, R.J., 2012. HOKIEFLINCS_H2O–
in the Sungun porphyry copper deposit, Iran: evidence from fluid inclusions and NACL: a Microsoft Excel spreadsheet for interpreting microthermometric data
stable isotopes. Economic Geology 93, 651–670. from fluid inclusions based on the PVTX properties of H2O–NaCl. Computers &
Kwak, T.A.P., Brown, W.M., Abeysinghe, P.B., Tan, T.H., 1986. Fe solubilities in very Geoscinces. http://dx.doi.org/10.1016/j.cageo.2012.01.022.
saline hydrothermal fluids: Their relation to zoning in some deposits. Economic Stefanini, B., Williams-Jones, A.E., 1996. Hydrothermal evolution in the Calabona
Geology 81, 447–465. porphyry copper system (Sardinia, Italy): The path to an uneconomic deposit.
Lagache, M., Weisbrod, A., 1977. The system two alkali feldspars-KCl–NaCl–H2O at Economic Geology 91, 774–791.
moderate to high temperatures and pressures. Contributions to Mineralogy and Sterner, S.M., Bodnar, R.J., 1984. Synthetic fluid inclusions in natural quartz I.
Petrology 54, 261–271. Compositional types synthesized and applications to experimental
Lecumberri-Sanchez, P., Steele-MacInnis, M., Bodnar, R.J., 2012. A numerical model geochemistry. Geochimica et Cosmochimica Acta 48, 2659–2668.
to estimate trapping conditions of fluid inclusions that homogenize by halite Sterner, S.M., Hall, D.L., Bodnar, R.J., 1988. Synthetic fluid inclusions. V. Solubility of
disappearance. Geochimica et Cosmochimica Acta 92, 14–22. the system NaCl–KCl–H2O under vapor-saturated conditions. Geochimica et
Maanijou, M., 1993. Alteration halos and their connection to mineralization of Cosmochimica Acta 52, 989–1005.
Darrehzar porphyry Cu deposit and its geochemical zoning, Pariz area, Kerman, Ulrich, T., Günther, D., Heinrich, C.A., 2002. The evolution of a porphyry Cu–Au
Iran. Unpublished M.Sc. thesis (in Persian), Tehran, Shahid Beheshti University, deposit, based on LA–ICP–MS analysis of fluid inclusions: Bajo de la Alumbrera,
p. 200. Argentina. Economic Geology 97, 1889–1920.
Quan, R.A., Clock, P.L., Kesler, S.E., 1987. Chemical analyses of halite trend inclusions Waterman, G.C., Hamilton, R.L., 1975. The Sar Cheshmeh porphyry copper deposit.
from the Granisle porphyry copper deposit, British Columbia. Economic Economic Geology 70, 568–576.
Geology 82, 1912–1930. Werre, R.W., Jr., Bodnar, R.J., Bethke, P.M., Barton, P.B., 1979. A novel gas-flow fluid
Ranjbar, H., 1997. Lithogeochemical patterns associated with the Darrehzar inclusion heating-freezing stage. Geological Society of America Abstract with
porphyry Cu deposit, Pariz area, Iran. CIM Bulletin 90, 85–90. Programs 11, p. 539.
Ranjbar, H., Hassanzadeh, H., Torbati, M., Ilaghi, O., 2001. Integration and analysis of Whitney, D.L., Evans, B.W., 2010. Abbreviations for names of rock-forming minerals.
airborne geophysical data of the Darrehzar area, Kerman Province, Iran, using American Mineralogist 95, 185–187.
principal component analysis. Journal of Applied Geophysics 48, 33–41. Wilson, J.W.J., Kesler, S.E., Cloke, P.L., Kelly, W.C., 1980. Fluid inclusions
Roedder, E., 1984. Fluid inclusions. Reviews in Mineralogy 12, 644. geochemistry of the Granisle and Bell porphyry copper deposits, British
Roedder, E., Bodnar, R.J., 1980. Geologic pressure determinations from fluid Columbia. Economic Geology 75, 45–61.
inclusion studies. Annual Review of Earth and Planetary Sciences 8, 371–406.

You might also like