You are on page 1of 125

In the name of God

The compassionate, the merciful

a
second form of barite, Brt. 3 = the third form of barite (Whitney and Evans, 2010). All
photos are in crossed polarized light ................................... Error! Bookmark not defined.
Figure 3.15. Photomicrographs of different forms of barite mineralization in the eastern
part of the Qezel-Dagh deposit. Abbreviations: Brt. 1 = the first form of barite, Brt. 2 = the
second form of barite, Brt. 3 = the third form of barite (Whitney and Evans, 2010). All
photos are in crossed polarized light ................................................................................. 52
Figure 3.16. Photomicrographs of different forms of barite mineralization associated with
Fe in the eastern part of the Qezel-Dagh deposit. Photos of (a) and (c) are in crossed
polarized light and photos of (b) and (d) are in plane polarized light of (a) and (c).
Abbreviations: Brt. 1 = the first form of barite, Brt. 2 = the second form of barite, Brt. 3 =
the third form of barite, Opq = opaque mineral (Whitney and Evans, 2010) ...................... 54
Figure 3.17. Photomicrographs of Fe mineralization with replacement and semi-shaped
granular textures and pyrite as inclusions in Fe oxides and gangues in the eastern part of the
Qezel-Dagh deposit. Abbreviations: Gth = goethite, Hem = hematite, Lm = limonite, Py =
pyrite (Whitney and Evans, 2010). All photos are in reflected light ................................... 56
Figure 3.18. Photomicrographs of chalcopyrite mineralization with amorphous to
semishaped granular texture, pyrite with spherical and disseminated textures, and hematite
with replacement texture in the eastern part of the Qezel-Dagh deposit. Abbreviations: Ccp
= chalcopyrite, Hem = hematite, Py = pyrite (Whitney and Evans, 2010). All photos are in
reflected light ................................................................................................................... 58
Figure 3.19. Photomicrographs of Fe mineralization as vein, open-space fillings, and
colloform textures in the eastern part of the Qezel-Dagh deposit. Abbreviations: Gth =
goethite, Hem = hematite, Lm = limonite, Py = pyrite (Whitney and Evans, 2010). All
photos are in reflected light............................................................................................... 59
Figure 3.20. Photomicrographs of (a) specularite with a blady texture and (b) specularite
with a semi-shaped granular texture in the eastern part of the Qezel-Dagh deposit.
Abbreviations: Hem = hematite, Py = pyrite (Whitney and Evans, 2010). All photos are in
reflected light ................................................................................................................... 60
Figure 3.21. Paragenetic sequence of minerals in the Qezel-Dagh barite deposit........... 62

k
:
CHAPTER 1 Literature review

l
1.1 Introduction
Barium is one of the most common tracer elements of the alkaline earth elements in the

upper lithosphere with strong oxygen affinity but unlike the more abundant elements such

as calcium, it rarely forms independent minerals in igneous rocks. It forms an element in

certain rock-forming minerals, replacing potassium diodochically in feldspars and micas,

e.g., celsian, hyalophane, sanidine (in volcanic rocks) and kollacherite (barium muscovite).

Barium biotites in diorites and potash feldspar in alkaline rocks are the chief hosts of barium.

Finish granites rich in rare earth elements, as a general rule, contain relatively much barium,

but the converse is not true. Barium is sometimes considerable in alkaline rocks. In general,

in igneous rocks, the highest content of barium occurs in acidic rocks, syenites and nepheline

syenites (Karunakaran, 1973).

Barium is a common and quite ubiquitous element. Ba has about 0.05% abundance in

earth crust. and ranges from 250 to 584 ppm in the upper continental crust, according to

Krebs (2006), this makes it the 17th most common element in the crust of the earth. In the

geologic environment, barium is primarily found as the divalent cation Ba2+. Barium cannot

be readily incorporated into conventional rock-forming minerals because this ion is bigger

than the majority of other divalent cations. Barium concentrates in the leftover silicate liquid

as the silicate magmas fractionally crystallize. Additionally, silicate liquids generated by

partial melting are concentrated in it. According to Rudnick and Gao (2003), the weight

percentage of barium in the average upper continental crust is 0.0624. Ba has a lithophilic

affinity and is likely to concentrate in acid igneous and sedimentary rocks, ranging widely

in various rocks from 50 to 1200 ppm. Its ionic radius is fairly similar to the ionic radius of

K, and thus usually follows the K fate in geochemical processes (Kabata-Pendias, 2011).

Basaltic rocks normally have less barium than granitic rocks, whereas granitic rocks often

contain more barium than the typical continental crust. Shale barium concentrations often

fall within the same range as granitic rock barium concentrations.

1
Barite is not only produced from primary barite deposits, where barite is the primary

mineral, but there is a growing trend in processing the ores that contain barite in the form of

gangue or as the secondary mineral. Deposits having barite are the primary minerals

containing barite in a coarsely crystalline form. It has more or less an intergrown appearance

when associated with other ores and minerals. Since barite is a common mineral, it may be

found on every continent and is mined and processed in numerous nations (Hanor, 2000).

Some igneous rocks, including carbonatites, dacites, and rhyolites, contain barite as an

accessory mineral. According to (Hanor, 2000), it can also be found in a wide range of ore

deposits connected to continental igneous activity. Examples of these include porphyry Cu

and Cu-Mo deposits, polymetallic vein deposits, and W vein deposits in the western United

States. According to various geological settings and fluid sources, including magmatic

(WilliamsJones et al- 2000), metamorphic (Hanor, 2000), sedimentary basinal hydrothermal

fluids (Kontak et al - 2006), as well as ancient and modern oceanic water (Monnin and

Cividini, 2006), it is thought that barite forms in a wide range of P-T conditions (0 to 400

°C and 1 to 2000 bars) and Several mechanisms, including fluid cooling (Pfaff et al., 2010),

mixing of two or more fluids (Valenza et al 2000), intensive fluid-rock contact (Marchev et

al- 2002), and bacterial activities (Pfaff et al - 2010), might result in the formation of bauxite-

bearing deposits. Geologically, barite deposits are found in rocks ranging from Archaean

crystalline to late Tertiary sediments. Barite occurs most commonly as veins either singly

or associated with chalcopyrite, galena, pyrite, sphalerite, quartz, fluorite, siderite, calcite,

and dolomite, and also rarely with allanite, monazite, columbite, marcasite, magnetite,

tetraherdrite, etc.

Ba2+ and S in its oxidized state, SO42, are the two main components of barite. Most

naturally occurring fluids are undersaturated with respect to barite, despite the fact that Ba

and Sr are relatively abundant and widely distributed elements in Earth crustal rocks (Faure,

1998; Chow and Goldberg, 1960; Church and Wolgemuth, 1972; Monnin et al., 1999;

Rushdi et al., 2000; Ehya, 2012; Zhang et al., 2013). As a result, interaction between several

2
sources of Ba and SO4 is required for barite to precipitate. Saturation needs to be kept at a

certain level for barite in order to keep the mineral after precipitation. However,

economically significant barite and lead–zinc–barite deposits are confined to marine strata,

owing to the need for a large reservoir of dissolved sulfate (SO42−), which seawater supplies.

The barium ion (Ba2+) forms an extremely insoluble sulfate, so that, even at the low barium

content of seawater, the oceans are close to saturation with respect to the mineral barite

(Rushdi et al- 2000). Therefore it takes only a slight increase in Ba2+ concentration to induce

barite precipitation.

1.2 Applications
It is the ideal mineral for usage in industrial uses due to its unique qualities, which include

high density, low solubility, chemical inertness, great brightness and whiteness, softness,

and relative affordability (Bonel - 2005). Barite is primarily used in drilling fluids for oil

and gas exploration wells. About 90% of all barite produced goes into drilling muds used in

petroleum exploration. It is also used as a filler material and wetting agent in a range of

products from golf balls and bowling balls to paint, heavy printing paper, rope finishes and

rubber products (Harben, 1999). It is used in mould-release compounds in the metal casting

industry. X-rays and gamma rays are strongly blocked by barite and so it is used in cement

vessels used to contain radioactive material, and medicinal and medical applications. Barite

is also used in faceplates and funnel glass of cathode ray tubes of televisions and computer

monitors. Barium compounds will continue to be used for most of their current industrial

uses because they are less expensive than competing materials, and they will be used

increasingly in new, highly technical applications.

1.3 Classification of barite deposits


Barite deposits have been divided into groups depending on their form, place of origin, or

tectonic context (Brobst, 1958, 1970, Clark and Orris – 199, Maynard and Okita, 1991;

3
Jewell- 2000, Hanor - 2000). There are a number of different ways that barite can form in

the present marine environment, all of which depend on the interaction of Ba and SO4-rich

fluids to create supersaturation in the formation environment. The chemistry and form of the

barite crystals are influenced by the origin of the fluids (seawater, pore fluids, or

hydrothermal solution), as well as by where the precipitation occurred (water column, sea

bottom, sedimentary column, cold seep, or hydrothermal settings). Whereas barite and

igneous and magmatic-hydrothermal rocks are found in continent-wide contexts, the

majority of the world's barite mines are generated in sedimentary settings (Hanor, 2000;

Jewell, 2000, Clark et al - 2004, Johnson et al - 2017, Crockford et al - 2019). Paytan et al.

(2002) have identified four different modes of formation of barite in modern sediments:

(1) Authigenic – crystallization in the water column, making small crystals 0.5–5 mm; (2)

Diagenetic, diffusive – diffusion of Ba2+ to overlying seawater after release by bacterial

sulfate reduction (BSR) to pore waters of anoxic sediments;

(3) Diagenetic, advective – transport of barium by movement of water expelled to the

seafloor along faults or in seeps;

(4) Hydrothermal – advection of hot water to the seafloor.

The primary technique is a "from-above" source of barium for the ocean floor, however

the other three are "from-below" sources (Torres et al - 2003, for example). As as a

consequence of the above enhancement, which involves the incorporation of Ba2+ into

sinking organic matter (OM) (Dymond et al - 1992), and is characterized by a strong

correlation between organic carbon (Corg) and Ba, the Ba content of deep-sea sediments is

frequently used as a paleoproductivity measure (see, however, McManus et al - 1998). Due

to the frequent combination of such barites with bedded chert and phosphate-enriched silt,

several scholars (e.g. Jewell, 2000; Jewell and Stallard, 1991) have proposed that this

biological movement is the major source for Ba in bedded barites, such as those of Nevada.

According to Schenau et al. (2001), the Arabian Sea is a current example of barium

enrichment through this method.

4
Barite is first dissolved deep within the sediment mass, then transported in sulfate-free

water to the sediment-water interface where it combines with saltwater sulfate to redeposit

barite (from below enrichment) (Dickens, 2001). A "barite front" forms inside the shallow

portion of the sediment when diffusion through pore fluids is the only transport mode

involved (Riedinger et al., 2006; Torres et al., 1996b). However, given the close distances

that diffusive transport can work over, it is doubtful that this mechanism would result in

significant bedded barite deposits. The likely outcome is a coating of barite nodules. The

more likely scenario for massive barite accumulations is fluid movement along faults, which

would allow for a long-term transfer of barium to the surface, as is the case throughout the

Peru subduction zone (Dia et al., 1993; Torres et al., 1996a).

This advective transport of Ba2+ can be in cold water, producing ‘cold-seep barites’ (e.g.,

Aharon and Fu, 2000; Aquilina et al., 1997; Greinert et al., 2002) or in hot water, producing

‘hydrothermal barite’ (e.g. Canet et al - 2005, Koski et al 1985 to 1988).

The downpour of Ba-enriched OM that generates the enrichments that come "from above"

is regarded to be the ultimate cause of the cold-seep Ba. According to Aloisi et al. (2004),

all examples of cold-seep barites that have been documented so far come from regions

with increased biological productivity in surface waters. The most plausible mechanism

for the release of this barium that was previously biogenic is sulfate reduction in

combination with methane oxidation at depth in the sediment pile. As a result, Murchey

et al, (1987) noted that one may anticipate a black shale- phosphate-chert relationship for

historic cold-seep barite deposits. The same defense may be made for historic

hydrothermal barites, as those found at Red Dog, which likewise exhibit a strong

preference for silica-rich black shales.

Based on the different fluid sources and formation mechanisms, sedimentary barite can

be divided into the following categories: (1) marine or pelagic barite, (2) cold seep barite,

(3) hydrothermal barite, and (4) diagenetic barite (Hanor, 2000; Paytan et al., 2002; Hein et

al., 2007; Griffith and Paytan, 2012; Zhou et al., 2015). They are discussed as following:

5
1.3.1 Barite deposits that are seafloor or pelagic
According to Bishop (1988), Bernstn (1992, 1998), Ganeshram (2003), and others, barite

can precipitate in the water column when barium, which is produced during the

decomposition of organic material, creates microenvironments that are supersaturated with

barite. Marine or pelagic barite is the name given to this particular form of authigenic barite.

Numerous marine organisms have Ba concentrations that are significantly higher than those

of seawater (Fisher, 1991). As a result, these organisms may serve as a source of Ba for the

microenvironments in the water column where marine barite is believed to precipitate

(Goldbergen and Arrhenius, 1958; Bishop, 1988; Berns, 1992–1998). (Dehairs, 1980;

Bishop, 1988; Bertram and Coweny, 1997) have reported the discovery of marine barite in

clusters of sub-micrometer grains (usually 5 m) with or without a crystalline habit, sub-

spherical particles, and particles with a clearly defined crystalline habit.

1.3.2 Barite deposits in cold seeps


In a variety of dynamic tectonic environments, including as convergent zones, marginal

basins, and chimneys, massive quantities of barite develop near cold seeps. as well as rifted

and strike-slip continent-wide their profit margins, accretionary prisms are emerging in the

forearc of continental-margin arcs. Cold seep barite is the name given to barite that forms at

the interface between sediment and water in conjunction with fluid movement and ejection.

Cold seep barite refers to barite precipitation at the sediment-water interface when tectonic

and hydrological factors unrelated to volcanic or hydrothermal activity force Ba-rich fluids

out of the sediment. Fluids migrate vertically under excessive pressure as a result of

sedimentary loading brought on by lateral compressive tectonics and/or high sedimentation

rates. Barite can form at or close to the sea floor where fluids interact with sulfate-rich

saltwater as they rise from depth. This kind of barite is present at both passive and active

margins and along transform faults. Along the inactive continental Nile edge, for example,

salt tectonics can regulate fluid discharge that causes barite precipitation (Gon-tharet 2007).

6
Pelagic barite deposited in sulfate-reducing sediments along the continental edge has been

remobilized, which has been suggested as a source of Ba for fluids connected to the

development of cold seep barite, among other possibilities (Torres 1996b), leaching of Ba

from primarily continental sources (Aquilina., 1997) and outflow of brines rich in Ba,

obtained from the passage of fluids via salt deposits at depth (Manheim and Bischoff, 1969;

Kharaka 1987; Macperson, 1989; Sassen 2004). High concentrations of methane and

hydrogen sulfide characterize the vent fluids at some vent sites (Barry et al., 1996; Greinert

et al., 2002), and cold seep support unique biologic communities. Cold-seep barite deposits

are characteristically porous, with a banded or layered structure indicating multiple episodes

episodes or an intermittent style of venting. Although sedimentation rates are commonly

high in areas of cold seep, barite deposits generally are nearly pure (>90 weight percent

BaSO4). Minor amounts of pyrite are associated with some deposits (Fu et al., 1994; Greinert

et al., 2002). Examples of contemporary locations where cold seep barite has been found

include the San Clemente Fault (Torres, 2002), the Gulf of Mexico (Fuet 1994), the Peru

Margin (Dia - 1993; Torres, 1996b), Monterrey Bay (Naehr et 2000), the Sea of Okhotsk

(Greinert - 2002), the Nile deep-sea fan (Gon-tharet 2007), and the Gulf of Guinea.

1.3.3 Barite deposits that are diagenetic


Diagenetic barite deposits can be found on the ocean floor, along the edges of sulfate-

reducing zones in bottom waters, and in regions where cold-seep exhalations have occurred.

During post-deposition diagenetic processes, diagenetic barite is precipitated from pore

fluids within the sediments. According to Hein (2007), in some systems, tectonic advection

can result in the overpressured rock units and sediment dewatering that can cause fluid flow

that is richer in Ba or SO4 and precipitate diagenetic barite within sediments. As an

alternative, a decrease in solubility brought on by variations in fluid pressure and

temperature might cause barite to precipitate inside sediments (Hanory - 2000). According

to Griffith and Paytan (2012), the most frequent mechanism for diagenetic barite to develop

is by barite dissolution caused by sulfate reduction in sediments, followed by re-

7
precipitation of barite at redox boundaries where Ba-rich pore fluids interact with pore water

rich in sulfate. Barite will not be preserved if SO4 concentrations fall below those needed

for barite saturation when pore waters turn anoxic and SO4 reduction rates outpace the rate

of SO4 replenishment into pore fluids. According to Breheret and Brumsack (2000) and

Dickens (2001), this dissolution will often lead to an increase in Ba concentrations in the

pore water due to a reduction in SO4 and loss of particle barite. The precipitation of

diagenetic barite within the sedimentary column, frequently at the oxic-anoxic boundary,

may then occur as a result of barium diffusing within the sediment and interacting with

solutions containing SO4 (Bolze - 1974; Dean and Schreiber, 1977; Brumsack and Gieskes,

1983; Cecile, 1983; Breheret and Brumsack, 2000). This process may also take place in

some sub-oxic sediments where SO4 reduction occurs but pore fluids still contain SO4

(McManus 1998). Dean and Schreiber (1977) and Breheret and Brumsack (2000) both

discovered that diagenetic barite generated in the sedimentary column often appears as flat,

tabular-shaped crystals in barite-rich beds or as nodules in sedimentary layers.

1.3.4 Hydrothermal barite deposits


Since there is more heat movement in rift basins, hydrothermal barites are only found

there. In this environment, it is typical to detect enrichment in heavy metals like lead and

zinc in addition to the barite. Based on the underlying crust, rifts were classified as either

continental or oceanic. Only the oceanic context, which can be a back-arc spreading center

or a mid-ocean ridge, is known to contain modern occurrences of barite, although the

majority of ancient lead-zinc-barium deposits fall into the category of continental deposits.

The hydrothermal deposits forming in various tectonic settings are related to fluid

circulation, volcanism, and emplacement of magma bodies at depth. Hydrothermal volcanic

activity that rises from the depths and mixes with saltwater close to the bottom results in the

formation of Ba-rich fluids that precipitate as barite. Hydrothermal fluids may be directed

by extensional faults and cracks upward onto the seafloor, where they mix with saltwater

and serve as the main source of SO4 for barite precipitation. Oceanic or continental rocks

8
leach Ba into hydrothermal fluids under the influence of heat from magmatic activity.

Another potential source of Ba for these fluids is hydrothermal leaching of pelagic sediments

high in Ba (Murchey, 1987). Barite solubility decreases during the lowering of pressure at

any temperature, and with decreasing temperature, below 100 °C (Hanor, 2000). It further

makes barite precipitation from hydrothermal solutions possible. The geochemistry of the

hydrothermal fluid, as well as size and composition of barite deposit, is determined by the

type and amount of host volcanic rocks and the sediments through which the fluid has passed

(Hanor, 2000). Hydrothermal fluid temperature is further distinguished in different

environments where hydrothermal barite forms. Barite can precipitate from low-temperature

(<120 °C) hydrothermal fluids around warm springs at the seafloor. Barite may also

precipitate at average temperatures (150-250 °C) in hydrothermal settings of continental

margins where fluid circulation due to high heat flow is common (Hein et al., 2007, and

references therein). This mode of precipitation occurs at the seafloor near hydrothermal

plumes and forms chimneys and mounds. It might likewise happen inside the residue as

scattered precious stones in storm cellar breaks at maritime back-bend bowl spreading

focuses, crack zones, and volcanic curves (e.g., East Pacific Ascent 1 N, Huheey et al., 1993;

Tonga curve, south-west Pacific, Stoffers et al., 2006; the Kurile and western Aleutian Island

circular segments, north-west Pacific, Glasby et al., 2006; Okinawa and Mariana Box, Japan,

Noguchi, 2011).

1.4 Barite deposit types


Barite deposits are divided into three morphologic types by Clark and Orris (1991),

including bedded (stratiform), veins and cavity fills, and residual deposits. They are

discussed as following:

1.4.1 Bedded-sedimentary deposits


Bedded barite deposits are commercially the most significant because of their large size

and amenability to modern large-scale mining methods (Jewell, 2000) and constitute one of

the key elements of several exhalative base metal sulfide mines (Hanor, 2000). The barite

9
deposits can extend for kilometers along strike and range in thickness from millimeters to

100 meters or more. Barite can have a dark tint, which indicates the presence of organic

carbon. Among the gangue minerals are clays, iron oxides (such as hematite, limonite, or

magnetite), quartz, and pyrite or other sulfide minerals. Barite occurs in bedded deposits

either as a principal mineral or cementing agent, or in close proximity to stratiform massive

sulfide deposits of sufficient number and quality to make the areas more famous as sources

of lead, zinc, or other metals. In marine basins, bedded-sedimentary barite deposits develop

in conjunction with sediments that exhibit significant biological productivity (Jewell, 2000).

Most bedded deposits occur in sequences of sedimentary rocks characterized by abundant

chert and black siliceous shale and siltstone, known as ‘black bedded barite’. They range in

age from Precambrian to Tertiary (Pooly-1988; de Brodt korb, 1989; Clark, 1990;

Clark and Orris, 1991; Jewell, 2000; Hanor, 2000), but are usually early to Mid-Paleozoic

(Ramos and de Brodtkorb, 1989). Individual beds are massive to laminated, fine-grained

and may contain from 50% to 95% barite that can be used with little or no treatment other

than grinding. The deposits, which are also referred to as SEDEX barite, are thought to have

formed by mixing of barium-bearing fluids and marine waters (Hanor, 2000). SEDEX barite

is most abundant in Phanerozoic sedimentary sequences but is scarce in Precambrian

sequences, which is thought by most workers to reflect the transition from early anoxic

sulfate poor oceans to later oxygenated sulfate-richer oceans (Jewell, 2000; Lyons et al.,

2006). Bedded barite deposits were divided by Clark et al. (1990) into five types:

(1) With base metal sulfides (cratonic rift type), associated with alkali volcanic rocks,

e.g. Meggen and Rammelsberg, Germany; Ballynoe and Silvermines, Ireland; Selwyn

Basin, Canada; and Red Dog, Alaska, USA.

(2) Without base metal sulfides (continental margin type), e.g. Arkansas and Nevada,

USA; Quinling District, China.

(3) In volcanic sequences, e.g. Kuroko, Japan.

(4) Stratiform barite deposits, e.g. Sardinia, Italy; Andhra Pradesh, India.

(5) Exogenetic barite deposits, e.g. Krakow area, Poland.


10
There is dispute on the conditions under which the fluids combined, despite the fact that

the formation of barite is generally recognized. For instance, according to evolutionary

theories put out by Shawe (1969), Jewell and Stallard (1991), and Jewell (1994, 2000), barite

is confined by high primary production in areas of ocean upwelling. This model proposed

that the settling of large amounts of organic matter resulted in anoxia and sulfate

consumption in the deep waters below, favoring the dissolving of biologically generated

barite particles and the buildup of dissolved barium. Massive barite deposits developed

along the boundaries between the sulfate- and barium-rich seawaters that were present above

the anoxic deep waters at the borders of basins (Cecileat 1983). Another hypothesis is that

the barite was formed at seeps or vents when chemically reduced, barium-containing fluids

were released from sediments along the continental edge. Numerous geological findings

(Cecile, 1983; Lydon, 1985; Poole, 1988; Dube', 1988; Clarket, 2004; Johnson, 2004) and a

similarity to the barite mineralization that has recently been found on contemporary

continental edges (Torres et al., 2003) provide credence to this hypothesis.

There are theories that the barium-bearing fluids at some Paleozoic sites also included

methane or other hydrocarbons (Torresal, 2003; Johnson, 2004). This would be significant

for barite formation because in most marine sediments methane is consumed by an anaerobic

microbial process that employs sulfate as an oxidant (anaerobic oxidation of methane, or

AOM) (Jørgensen and Kasten, 2006, and references therein). In detail, AOM may be a two-

step reaction that takes place within microbial aggregates in which electrons are transferred

from methane oxidizing archaea to sulfate-reducing bacteria by intermediates chemical

species that may include hydrogen, acetate, or formate.

Although methane oxidation and sulfate reduction occur in distinct steps, the sulfate

reducing bacteria are critical to the energetics of the overall process (Jørgensen and Kasten,

2006), so in effect AOM and barite precipitation could compete for sulfate where supplies

of the anion are limited. This effect has been noted in reaction-transport models of modern

barite-depositing seeps (Aloisi et al., 2004), which suggest that high barium:methane ratios

11
in seep fluids favor the precipitation of barite whereas low barium:methane ratios can

suppress barite and favor methanogenic calcite.

Bedded barite deposits are classified into two main groups based on the tectonic context

(Maynard and Okita, 1991). Continental-margin deposits, which include barite but lack

economically significant Pb and Zn mineralizations (e.g., Arkansas and Nevada in the

United States; Quinling District in China), are the first category. The second type, cratonic-

rift deposits, may or may not be connected to stratiform Pb and Zn mineralization (such as

in the Selwyn Basin, Jason and Tom, Canada; Meggen and Ramm elsberg, Germany;

Ballynoe and Silver mines, Ireland; and Red Dog, Alaska, USA). Similarly according to

Jewell (2000), bedded barites formed on the seafloor can be assigned to three types of basins:

marine rift basins on continental crust; transtensional basins in strike-slip settings (pull-

aparts); and accretionary prisms of subduction zones. In rift settings, deeply penetrating

normal faults provide pathways for the transfer of heat and metals to the seafloor, making

these particularly favorable settings for ore deposit formation, for both barite and base

metals. Marine rifts come in two major varieties: spreading-center rift valleys such as

Guaymas, Juan de Fuca, and the Red Sea, which are dominated by oceanic crust (oceanic

rifts); and seawater-flooded rifts on continental crust such as the North Sea (continental

rifts). Strike-slip settings also produce steeply dipping tensional faults but combine these

with compressional features. That is, both transpressional and transtensional regions will be

present in the fault zone. The resultant basins undergo fast fault movement, and the near

proximity of the uplifts and basins tends to cause rapid clastic deposition, which reduces

seabed mineralization. Deep thrust faults in accretionary prisms provide relatively lowangle

channels for fluid movement, and compression predominates there. Significant amounts of

water and solutes are transferred, but no heat. Low temperatures as a result limit metal

mobility while yet allowing significant Ba2+ flows to the bottom.

12
1.4.2 Vein and cavity-fill deposits

Barite deposited from hydrothermal fluids or deep-seated brines occurs in faults, joints,

bedding planes, breccia zones, solution channels and cavities. The resulting veins are

characterized by sharp contacts, extensive pinching and swelling, and extreme variations in

length, depth and attitude. Depositions are typically tabular and provide along constructions,

often intermittently, that can be identified for hundreds or thousands of meters. The vein

minerals may reach thicknesses of several tens of meters. Gangue minerals include

carbonates (calcite and siderite), fluorite, quartz, and pyrite or other sulfide minerals. Cavity-

fill deposits are more inconsistent in form; they tend to be limited to a specific sedimentary

stratum or a series of strata. Because of complex geometry, mining vein deposits can be

difficult and expensive. These deposits are generally smaller than the stratiform deposits.

Vein barite is usually extracted as a by- or co-product of lead–zinc mining. This deposit

form has served as a source of exceptionally pure barite, which is typically required to have

a BaSO4 grade of at least 95% for use in fillers and ceramics.

1.4.5 Residual deposits

The weathering of previous deposits results in the formation of weakly consolidated

materials known as residual barite deposits. Residual barite occurs in surficial deposits in

which the barite is present as loose fragments embedded in residual clay. The barite and the

clay are derived from weathering of the underlying rock, generally dolomite. Barite

fragments range in size from sand size to lumps weighing 100 kg or more. The deposits

often contain unweathered rock pieces, quartz, and clay minerals. There may also be metal

sulfides, iron oxides, and sulfide weathering products.

13
1.5 Deposits of barite in Iran

Barite deposits may be found all over the world. Iran has seen the discovery of over 100

barite deposits and potential with a cumulative resource of 10 Mt (Ghorbanifar, 2013). Iran

has seen an increase in barite output during the past ten years. Iran is the fifth-largest

producer of barite in the world, with an annual output of 330,000 t. The majority of barite

produced worldwide is utilized as a weighting ingredient in drilling fluids for oil and gas

wells. Barite mineralization occurs in host rocks ranging in age from Late Precambrian to

Miocene (Ganji, 2015). A wide variety of barite mineralizations in Tertiary volcanic rocks

occur in association with volcano-sedimentary successions in the southern Central Alborz

Mountain Range (Azerbaijan) and in the Urumieh-Dokhtar magmatic arc (Ehya, 2012).

Barite mineralization in dolomitic rocks is widespread in central Iran, the Alborz Mountain

Range and the Sanandaj–Sirjan Zone (SSZ) (Ghorbani, 2002). In accordance with (Ehya,

2012), the greatest Iranian barite reserves are in the Central Alborz and Central Iran zones,

where dolomitic and volcano-sedimentary rocks predominate. Figure 1.1 shows where some

of Iran's barite reserves are located on the country's topographical map.

According to Ghorbani (2013), barite mineralization in Iran can be subdivided in four

mineralization phases: (1) late Precambrian-early Cambrian, (2) Permian-Triassic, (3)

Cretaceous, and (4) Tertiary. These barite mineralization phases occurred in different

geostructural zones and cause various types of barite deposits are found throughout Iran in

a variety of geo-structural zones. In the late Precambrian-early Cambrian barite

mineralization phase, the barite mineralization is mainly hosted in ancient dolomitic and

also volcanic (especially rhyolitic) rocks and occurs in Alborz Mountains, Northwest of the

Sanandaj–Sirjan and Central Iran geo-structural zones. In the Permian–Triassic barite

mineralization phase, The Alborz Mountains, Sanandaj-Sirjan geostructural areas, and

Central Iran are common spots for the barite deposit, which is predominantly contained in

Triassic dolomitic rocks. This barite mineralization phase is mainly associated with fluorite

and lead-zinc mineralization. The most famous barite resources of Iran are formed at this

14
phase. In the Cretaceous barite mineralization phase, the barite mineralization is mainly

hosted in the Lower Cretaceous carbonate rocks and occurs in Central Iran and Sanandaj–

Sirjan geo-structural zones. This barite mineralization phase is associated with lead and zinc

mineralization. In Tertiary barite mineralization phase, the barite mineralization is related

to Eocene and Oligocene volcanic to volcano-sedimentary rocks which frequently occur in

Urumieh–Dokhtar volcanic zone, Central Iran and South of Alborz Mountains.

Figure 1.1. Map showing the location of Ba-(Pb-Cu-Zn) deposits in the Alborz zone of Iran's structural map; Central
Iranian Geological and Structural Progressive Zone (CIGS); E, ranges in East Iran; L, Lut block; M, Makran zone; PB,
Posht-e-Badam block; SSZ, Sanandaj-Sirjan zone; K, Kopeh-Dagh; KT, Khazar-Talesh-structural zone; (Tectonic and
structural map of Iran updated from Aghanabati, 2004; Alavi, 1991) T stands for the Tabas block, TM for Tertiary
magmatic rocks, UD for the Urumieh-Dokhtar magmatic arc, Y for the Yazd block, Z for the Zabol region, and Z for
the Zagros mountains. information on barite deposits in Rahimpour-Bonab and Shekarifard, (2002), Nazemi (2005),
Ehya (2012), Vazifeh (2013), Tajeddin et al. (2019), Rajabi et al. (2012), Hashemi et al. (2014), Zarasvandi et al. (2014),
Zolfi and Simmonds (2015), Maghfouri (2017), Lotfi et al. (2017), and Mir-mohammadi(2017).

15
According to abovementioned evidences, it can be concluded that the major geo-structural

zones of Iran in which the barite deposits were formed are as following:

I. Metamorphic and volcanic domains of Sanandaj–Sirjan geo-structural zone

II. Volcano-sedimentary domain of Alborz geo-structural zone

III. Sedimentary domain of Central Iran geo-structural zone

IV. Volcanic and sedimentary domains in the micro-continent of Central Iran

geostructural zone.

Finally from the viewpoint of deposit types, the Iranian barite deposits mostly include in

four types: (1) vein-type barite deposits, (2) epigenetic carbonate-hosted (MVT) barite

deposits, (3) stratiform of sedimentary exhalative (SEDEX) type barite deposits, and (4)

stratiform of volcanogenic massive sulfide (VMS) type barite deposits (Ganji, 2015).

1.6 Mines of stratiform barite (Cu-Zn-Pb) at Mahabad,

northwest Iran, supported by sediment


One of Iran's earliest mining regions is thought to be the sediment-hosted barite-base

metal deposits in Mahabad. This region is home to numerous economically significant

mineral deposits, including SEDEX. (e.g., Hosein-abad: Mahmoodi, 2018), VMS (e.g.,

Barika: Tajeddin, 2019, Bavanat: Mousivand, 2007, Chahgaz: Mousivand, 2011, and

Sargaz: Badrzadeh, 2011), orogenic gold (e.g., Kharapeh: Niroomand, 2011), Saqqez-

Sardasht gold zone including Qolqoleh (Aliyari, 2007), Kervian (Almasi - 2014),

Qabaqloujeh (Nosratpoor, 2008), and Mirgehnaghshineh (Asghari - 2018)), and carbonate

hosted Zn–Pb deposits (e.g., Emarat: Ehya - 2010, Darrehnoghreh: Momenzadeh, 1976,

Irankuh: Rastad, 1981; Ghazban et al., 1994; Boveiri and Rastad, 2016, and Angouran: Gilg

et al., 2006; Boni - 2007; Daliran - 2013). There are only a few published data on the barite-

base metal mineralization in northwestern Iran. There are several other barite deposits in

SW and NE Mahabad, showing different host rocks (Behyari - 2019).

16
1.7 Purpose of the study
Many aspects of the Qezel-Dagh barite deposit are poorly understood, including the ore

geology, ore facies, oxygen and sulfur isotope composition of the mineral, and the kind of

deposits. The objective of this study is to record the geological and geochemical

characteristics of the Qezel-Dagh barite deposit in northeast Mahabad, NW Iran. The

objective of this thesis is also to identify the type of deposit and the origin of the ore-forming

process based on a thorough description of the mineralogical characterization, host rock

geochemistry, and oxygen and sulfur isotopes of barite at the regional level in the Qezel-

Dagh region. No previous literature has determined the origin or the oreforming process. It

is anticipated that this study will contribute to an improved comprehension of the

mechanisms behind the formation of barite-base metal deposits in Iran, which share

comparable host rocks.

1.8 Research questions


In this thesis, we investigate several questions that are fundamental to understanding the

genesis of barite mineralization in the Qezel-Dagh barite deposit:

A. How was the distribution of major, trace, and rare earth elements (REE) in ores and

host rocks ?

B. What were the source(s) of sulfur and the mechanism(s) of sulfate reduction involved

in the formation of the sulfide mineralization?

These questions are addressed in four chapters.

In Chapter 4, the geochemical compositions of host rocks have been used to distinguish

the tectonic setting, provenance, redox state of sediment rocks that host mineralization

(Question A).

17
In Chapter 4, we present isotopic data (δ34S and δ18O) for barite. The S and O isotopic

data reveal important new insights into sulfur cycling during host sediment deposition,

diagenesis, and hydrothermal mineralization (Question B).

1.9 Material and methods

1.9.1 Previous work and mining history


Having a total of 22% of the country's barite mines, the Mahabad area has long been

considered as great potential for barite exploration (Bahrami et al., 2018). In addition to the

Qezel-Dagh deposit (northeast Mahabad), there are several other barite deposits in the east

and northeast of the Mahabad area, showing different host rocks (Behyari et al., 2019). With

a total reserve of 3.5 Mt., barite mines in this area produce 40% of the total consumption for

diamond core drilling operations in Iran (Bahrami et al., 2018). New exploration activities

also show a high potential for undiscovered barite deposits in this area. The current

contribution assists in a better understanding of the genesis of barite(base metal)

mineralization in this area with important implications for the exploration of new deposits

in NW Iran through geological studied, isotopic composition, and geochemistry of REE.

Laboratory studies
a. Preparation of ** thin sections for petrographic studies of host rocks and ** polished

sections from ore samples for petrographic and mineralogical studies at Urmia University.

b. Preparation of 11 samples of host rocks and 10 samples of barite samples for

geochemical studies.

c. Trace-element concentrations of 11 samples of host rocks by Inductively Coupled

Plasma-Mass Spectrometry (ICP-MS) at Zar-Azma Mineral studies Co. in Tehran.

d. Major- and trace-element concentrations (including REE) of 10 samples of barite

mineralization by X-ray Fluorescence (XRF) and Inductively Coupled Plasma-Mass

18
Spectrometry (ICP -MS), respectively, at the Activation Laboratories Ltd. (ALS Chemex,

Turkey).

e. Separation of 8 barite samples for δ 34S and δ18O isotopic analysis.

f. Sulfur and oxygen isotopic analysis on 8 barite samples at the Stable Isotope

Research Laboratory of Arak University, Iran using an Isoprime Isotope Ratio Mass

Spectrometer (Elementar).

19
CHAPTER 2: Geological setting of the

Study area

2.1 Regional setting

According to geological data from Iran, geologic processes resulted in diverse

chronological and geographical consequences, which led to a complicated tectono-

magmatic history for Iran (Aghanabati, 2004). In the center of the Alpine-Himalayan

mountain range is Iran. The growth of the Tethys area had a major impact on Iran's geology,

notably its tectonic growth. The geology and especially the tectonic style of Iran are highly

influenced by the history and evolution of the Tethyan oceans. The tectonic events that

affected the Iranian plate were caused by the opening and closing of the Paleo- and Neo-

Tethys Oceans (1977, Stöcklin). The opening of the PaleoTethys Ocean in northern Iran is

thought to have taken place during the Ordovician-Silurian period, followed by the

subduction of its oceanic crust beneath the Turan plate in the north in the Late Devonian and

the collision of the Iranian microcontinent and Turan plates in the Late Triassic and Jurassic

periods (Stampfli, 2000; Natalin and Sengör, 2005). The development of the Neo-Tethys

oceanic crust in southern Iran and its subsequent sinking into the central Iran plate took

occurred throughout the late Jurassic to Cretaceous following the closing of the Paleo-
20
Tethys Ocean (Mohajjel et al., 2003; Richards et al., 2006). Following the collision of the

Iranian and Arabian plates in the Oligo-Miocene, continuous subduction led to the

consumption of the Neo-Tethys oceanic basin (Stöcklin, 1977; Stampfli, 2000).

Iran has been divided into several geo-structural zones, each characterized by a relatively

unique record of stratigraphy, magmatic activities, metamorphism, orogenic events,

tectonics, and overall geological style. The tectonic and structural setting of Iran in the

Alpine–Himalayan orogenic belt, and the structural evolution of Iran, has been the focus of

many studies. The first study divided Iran into 10 structural zones based on certain

geological features. This structural division remained a reference for the Iranian geologists

for almost three decades. However, the new observations and findings require a revision to

this structural scheme. The newer structural schemes are mostly derived and inspired by the

first structural division presented by Stöcklin (1977). In recent years, new interpretations

and models have been offered regarding the geological setting of Iran by different the Iranian

geologists. On the basis of these new structural divisions, Iran subdivided to 11 main geo-

structural zones, including Khouzestan plain, Zagros fold belt, Zagros thrust, Sanandaj–

Sirjan zone, Urumieh–Dokhtar zone, Alborz Mountains, Central Iran, Central Iran micro-

continent (Lut block), Kopeh Dagh Mountain, Makran zone, and Eastern Iran zone

(Nezafati, 2015). Each unit has a comparatively distinct record of stratigraphy, magmatic

processes, metamorphism, orogenic events, tectonics, and general geological style. Part of

the range of mountains between the Himalayas and the Alps is the Zagros Orogen. The

Zagros Orogen is divided into four parallel tectonic zones that run from northeast to

southwest: (1) Urmia–Dokhtar Magmatic Assemblage (UDMA), (2) Sanandaj–Sirjan Zone

(SSZ), (3) Zagros Fold Thrust belt, and (4) Mesopotamian-Persian gulf foreland basin

(Stöcklin, 1968; Berberian and King, 1981; Alavi, 1994; Mohajjel and Fergusson, 2000).

The metamorphic and igneous zone of the Zagros Orogen, the SSZ, was named by

Stöcklin (1968). The SSZ extends ~1500 km from the northwest (Sanandaj) to southeast

(Sirjan) with a width of 150–200 km, parallel to the Zagros Fold Thrust belt. Compared to

the high Zagros Mountains, most of the SSZ has a relatively low relief, typically no more
21
than 1400 m. The Zagros Main Thrust bounds the southern margin of the SSZ and separates

it from Zagros. Central Iran is separated from the SSZ by a belt of steep and straight faults

including the Tabriz and Nain–Baft Faults (Şengör, 1979). Overall, the scarcity of Tertiary

volcanic rocks, high volumes of Mesozoic (and somewhat Tertiary) intrusions, the relatively

high abundance of Paleozoic volcanic rocks (Silurian, Devonian, and Permian), and

metamorphism due to Cimmerian movements are the main features of the SSZ (e.g.,

Aghanabati, 2004; Ghasemi and Talbot, 2006). The geodynamic evolution of the SSZ was

controlled by the opening and subsequent closure of the Neo-Tethys Ocean at the

northeastern margin of Gondwana (Alavi, 1994).

22
Figure 2.1. The position of the Qezel-Dagh barite deposit marked by a black box on the geological map of Iran from

Alavi (1994).

23
2.2 Geology of the study area

The Qezel-Dagh barite deposit is located on the SSZ's northwest end and on the southwest

slope of the Qezel-Dagh Mountain, both of which are listed in Berberian and King's 1981

geological categorization of Iran. (Figure 2.1). The Qezel-Dagh barite deposit is located 20

km northeast of Mahabad city, west Azarbaijan province, NW Iran (Figure 2.2).

The study area is a highly tectonized area that is associated with tectonic activities, such

as faults, fractures, and joints in rocks of the study area, especially in rhyolite rocks. The

faults in the study area have two general trends: north–south and east–west. Since the

predominant trend of barite veins in the western part of the Qezel-Dagh barite deposit is

east–west, it seems that east–west trending faults played a major role in barite mineralization

in the western part of the Qezel-Dagh barite deposit. In contrast, major veins and cavity

fillings in the eastern part of the Qezel-Dagh barite deposit have a north– south orientation.

In general, barite mineralization in the Qezel-Dagh area followed the main structural trend

of faults and fractures.

Following the Katangai orogeny and the evolution of extensional fractures on the Iranian

platform, especially in the northern part of the SSZ, volcanic lavas that mainly contain

alkaline rhyolites formed. The volcano-sedimentary rocks of the Mahabad Formation are

comparable to those that have been revealed in the Central Iran and Alborz tectonic zones,

which have been established as the Precambrian substructure of the Iranian plateau. Noori

Asli (2017) believes that the genesis of the Mahabad rhyolites was significantly influenced

by partial melting and crustal pollution. The Mahabad rhyolites, which are thought to have

formed as a result of the Proto-Tethys continental crust being subducted near the northern

active boundary of Gondwana, are thought to have formed in a continental volcanic arc

tectonic context. The Mahabad rhyolites with a north–south trend formed in response to

extension and rifting, followed by compressive tectonic phases causing barite mineralization

along N–S trending faults in the eastern part of the Qezel-Dagh barite deposit.

24
Figure 2.2. The position of the Qezel-Dagh barite deposit relative to the Mahabad city.

The Mahabad Formation (Neoproterozoic) is the oldest formation with the highest

expansion on the surface of the Qezel-Dagh barite deposit and also in the northeastern

regions of Mahabad. The Mahabad Formation is mainly composed of acidic volcanic rocks

(mainly rhyolite-rhyodacite) and, to a lesser extent, diorite without shale units. Shale unit of

the Mahabad Formation can be observed in parts of the trench of the Mahabad road towards

Bukan. Rhyolites of the Mahabad Formation are the main host rocks of barite mineralization

in the northeastern part of Mahabad. Moreover, there are outcrops of reef limestone of the

Qom Formation in the northwestern part of the Qezel-Dagh barite deposit. The youngest

sediments in the study area are alluviums and agricultural clay soils along with fragments

of rocks of the Mahabad Formation. Based on 1:1,000 geological map of the Qezel-Dagh

barite deposit and detailed fieldworks, the main lithologic units in the studied area are

rhyolites of the Mahabad Formation (late Proterozoic) and Quaternary units (Figure 2.3).

The main lithologic unit in the western part of the Qezel-Dagh barite deposit is the Mahabad

rhyolite and in the eastern part of the Qezel-Dagh barite deposit is thin-bedded tuffs. Least-

25
altered rhyolites of the Mahabad Formation are greenish gray to light brown in color (Figure

2.4) with morphology of rounded hills (Figure 2.5a).

26
Figure 2.3. Simplified geological map of the study area showing the local geological setting and lithological units for the Qezel-Dagh barite deposit.
27
Altered rhyolite samples are yellowish cream in color. This unit appears to be composed

of submarine lava flows. The upper part of this unit is severely brecciated (Figure 2.5b).

Rhyolite and ignimbrite were affected by metamorphic events and turned into meta-rhyolite

and meta-ignimbrite.

Figure 2.4. (a and b) Rhyolitic host rocks along with barite mineralization and Fe- and Mn-oxides in the western part of

the Qezel-Dagh barite deposit.

29
(a)

(b)

Figure 2.5. (a) View of the western part of the Qezel-Dagh barite deposit (looking to the east), (b) view of highly tectonized

rhyolite in the western part of the Qezel-Dagh barite deposit.

In terms of texture and color, rhyolite rocks in the eastern part of the Qezel-Dagh deposit

are different from those in its western part. Rhyolitic tuffs and rhyolites in the eastern part

of the deposit mainly have vitroclastic and glassy textures, respectively, whereas rhyolites

in the western part of the deposit mainly have a porphyry texture. Rhyolitic specimens in the

eastern part of the deposit are light green to gray in color, whereas they in the western part

30
of the deposit are mainly purple and light brownish gray in color. The color change in

rhyolitic host rock is mostly related to the alteration of feldspar minerals and the formation

of secondary mafic minerals, such as chlorite and epidote. Based on fieldworks, chloritic

alteration in the eastern part of the Qezel-Dagh deposit is more prevalent than its western

part. Moreover, foliation, lineation, and a slight deformation in some rhyolite specimens are

observed in the eastern part of the Qezel-Dagh deposit. The hanging wall host rock in the

eastern part of Qezel-Dagh deposit ranges from rhyolite, through tuff with a rhyolitic

composition, to rhyolite, distance from barite veins.

Rhyolite of the Mahabad Formation in the Qezel-Dagh area hosts barite and iron

(±copper) mineralization (Figures 2.6a–f). Mineralization at Qezel-Dagh follows a

predominant E–W trend that correlates well with the regional stress regime. Barite

mineralization occurs as veins in the western part of the Qezel-Dagh deposit and as lens and

layer along with iron (±copper) mineralization in the eastern part of the Qezel-Dagh deposit

(Figures 2.6a–d). The mineralization is hosted in open-space fillings and fractures. In other

words, mineralization in the Qezel-Dagh area is controlled by tectonic activities (Figure

2.6a).

Tuff rocks with a rhyolitic composition are host rock of barite and iron mineralization in

the eastern part of the Qezel-Dagh deposit. They are in contact with footwall and hanging-

wall alkaline rhyolite. View of the extraction workshop, lithological units, and ore minerals

are given in Figures 2.7–2.9.

31
( a) (b)

( c) (d)

( e) (f)

Figure 2.6. (a) Barite veins in the western part of the Qezel-Dagh barite deposit, (b) barite mineralization along with

Fe-oxides in the western part of the Qezel-Dagh barite deposit, (c) barite (± iron) mineralization in rhyolitic

host rock in the western part of the Qezel-Dagh barite deposit, (d) barite mineralization in open spaces and

fractures of rhyolitic host rock in the western part of the Qezel-Dagh barite deposit, (e) barite ore in contact

with rhyolitic host rock in the western part of the Qezel-Dagh barite deposit, and (f) barite with a cataclastic

texture in the western part of the Qezel-Dagh barite deposit and open spaces filled by Fe-oxides.

32
(a) (b)

(c) (d)

Figure 2.7. View of the eastern part of the Qezel-Dagh barite deposit (looking to the west south), (b) view of

agglomerate pyroclastic rocks in the eastern part of the Qezel-Dagh barite deposit (looking to the west north), (c) view of

the extraction workshop in the eastern part of the Qezel-Dagh barite deposit (looking to the west north), and (d) barite and

iron mineralization (hematite and specularite) in the eastern part of the Qezel-Dagh barite deposit.

33
(a) (b)

( c)

Figure 2.8. (a) Iron mineralization in rhyolitic host rock and rhyolitic tuff in the eastern part of the Qezel-Dagh deposit,
(b) rhyolitic tuff as host rock in the eastern part of the Qezel-Dagh deposit, and (c) lithic tuff in the eastern part of the

Qezel-Dagh deposit.

34
(a) (b)

(c) (d)

Figure 2.9. (a) Barite, iron, and copper ore in the eastern part of the Qezel-Dagh deposit, (b) barite and iron ore in the

eastern part of the Qezel-Dagh deposit, (c) iron and copper ore in the eastern part of the Qezel-Dagh deposit, and (d)

specularite mineralization in joints of host rocks in the eastern part of the Qezel-Dagh deposit.

35
CHAPTER 3: Petrology

36
3.1 Petrology

3.1.1 Petrology in the western part of the Qezel-Dagh barite

deposit
The oldest lithological unit in the study area is volcanic rocks of the Mahabad Formation

(Neoproterozoic) that are host rocks of mineralization. No geochemical and dating studies

have been performed on host rocks of ore bodies in the Qezel-Dagh area. Studied rhyolites

are semi-crystalline to semi-glassy, holocratic to leucocratic and phaneritic. These rocks

have, mainly, porphyry and glomeroporphyritic textures. Rhyolite of the Mahabad

Formation consists of a mineral assemblage of quartz, K-feldspar, plagioclase, and mafic

minerals, such as biotite and hornblende (Figures 3.1–3.6).

Quartz phenocrysts are well visible in rhyolitic to rhyodacitic specimens in some parts of

the study area, especially in the western part of the Qezel-Dagh deposit, because of their

igneous origin. Quartz occurs as primary phenocrysts and also as fine-grained crystals in the

matrix. Quartz crystals are the main rock-forming constituent of rhyolite rocks in thin

sections (35–55% volume of the rock), showing semi-shaped to amorphous phenocrysts

mainly with an embayed textural habit (Figures 3.3a and b). Such textures may be caused

by rapidly rising and sudden depressurized rhyolite melts (Shelley, 1993). In some samples,

this mineral occurs as open-space fillings in the form veins and veinlets in response to

tectonic stresses. Quartz also occur as a single mineral and polymineral, and, in rare cases,

as a spherulitic texture, indicating the role of hydrothermal processes. Cavities and open

spaces filled by quartz phenocrysts are similar to a glomeroporphyritic texture. But in fact

this texture is a recrystallized texture, not a glomeroporphyritic texture.

The alkaline feldspars of rhyolite are often sanidine and make up 40 to 60% volume of

the rock. This mineral occurs as phenocrysts and also as fine-grained crystals in the matrix,

and mainly partially altered to sericite. Sanidine phenocryst mineralogy suggests somewhat

higher liquidus temperatures than are perhaps normal for rhyolitic extrusive. Footwall rocks

in the western part of the Qezel-Dagh barite deposit are often green, due to the presence of
37
chlorite as a product of alteration. Fine-grained matrix of rhyolite rock is mainly composed

of sericite, muscovite, quartz, and, in rare cases, biotite (Figures 3.3 and 3.4). Opaque

minerals, hydrous iron oxides, and, to a lesser extent, rutile, and anatase are accessory

mineral phases of the rhyolite rock.

(a) ( b)

(c) ( d)

Figure 3.1. Photomicrographs of rhyolite of the Mahabad Formation as host rock of barite mineralization in the QezelDagh

deposit. Quartz, K-feldspar, and plagioclase are the main mineral of rhyolite. Photos of (a) and (c) are in crossed

polarized light and photos of (b) and (d) are in plane polarized light of (a) and (c). Abbreviations: Qz = quartz,

Afs = alkali feldspar, Opq = opaque mineral (Whitney and Evans, 2010).

38
(a) ( b)

(c) ( d)

Figure 3.2. Photomicrographs of rhyolite of the Mahabad Formation as host rock of barite mineralization in the QezelDagh

deposit. Quartz, K-feldspar, and plagioclase are the main mineral of rhyolite. Photos of (a) and (c) are in crossed

polarized light and photos of (b) and (d) are in plane polarized light of (a) and (c). Abbreviations: Qz = quartz,

Afs = alkali feldspar, Opq = opaque mineral (Whitney and Evans, 2010).

39
(a) (b)

(c) (d)

Figure 3.3. Photomicrographs of rhyolite of the Mahabad Formation with a porphyry texture as host rock of barite

mineralization in the Qezel-Dagh deposit. All photos are in crossed polarized light. Abbreviations: Qz = quartz,

Afs = alkali feldspar (Whitney and Evans, 2010).

(a) (b)

(c) (d)

40
Figure 3.4. Photomicrographs of rhyolite of the Mahabad Formation with a porphyry texture as host rock of barite

mineralization in the Qezel-Dagh deposit. All photos are in crossed polarized light. Abbreviations: Qz = quartz,

Afs = alkali feldspar (Whitney and Evans, 2010).

(a) (b)

(c) (d)

Figure 3.5. Photomicrographs of rhyolite of the Mahabad Formation with a porphyry texture as host rock of barite

mineralization in the Qezel-Dagh deposit. All photos are in crossed polarized light. Abbreviations: Qz = quartz,

Afs = alkali feldspar (Whitney and Evans, 2010).

41
(a) (b)

(c) (d)

Figure 3.6. Photomicrographs of rhyolite of the Mahabad Formation with a microgranular texture as host rock of barite

mineralization in the Qezel-Dagh deposit. All photos are in crossed polarized light. Abbreviations: Qz = quartz,

Afs = alkali feldspar, Opq = opaque mineral (Whitney and Evans, 2010).

3.1.1.1 Barite mineralization


Chalcopyrite, pyrite, covellite, malachite, azurite, and hematite are base-metal minerals

that are linked with barite. According to field observations, ore bodies (veins and veinlets)

originated as stratabound structures. Barite veins in the Qezel-Dagh deposit are affected by

tectonic activities and alteration processes. Barite mineralization mainly occurs in rhyolites

of the Mahabad Formation as open-space fillings in faults, fractures, and cavities through

the migration of ore fluids. The trend of barite mineralization is almost E– W, NW–SE, and

NE–SW in accordance with the trend of factures in host rocks.

Barite ores in the western part of the Qezel-Dagh deposit are white to light cream in color

and, sometimes, yellowish white to yellowish brown, due to the impregnation with Fe-

oxides. In some parts of the deposit, barite mineral ore has a high grade. This mineral occurs

as well-formed phenocrysts, which is fine-grained barite towards margins of the veins and

42
at the contact with surrounding rocks. Based on microscopic observations, three forms of

barite mineralization are identified as follows:

1- Barite mineralization occurs semi-shaped to elongated phenocrysts (Figures 3.7 and

3.8) (Brt. 1: barite I). Some of these phenocrysts affected by tectonic forces have a cataclastic

texture (Figure 3.7).

2- Barite mineralization occurs along cleavage surfaces of phenocrysts (barite I) (Figures

3.7 and 3.8) (Brt. 2: barite II).

3- Barite mineralization occurs amorphous fine-grained crystals in the margin of

recrystallized barite phenocrysts (Figures 3.7 and 3.8) (Brt. 3: barite III).

43
( a) (b)

(c) (d)

(e) (f)

Figure 3.7. Photomicrographs of different forms of barite mineralization in the western part of the Qezel-Dagh deposit.

Abbreviations: Brt. 1 = the first form of barite, Brt. 2 = the second form of barite, Brt. 3 = the third form of

barite (Whitney and Evans, 2010). All photos are in crossed polarized light.

44
(a) (b)

(c) (d)

Figure 3.8. Photomicrographs of different forms of barite mineralization in the western part of the Qezel-Dagh deposit.

Abbreviations: Brt. 1 = the first form of barite, Brt. 2 = the second form of barite, Brt. 3 = the third form of

barite (Whitney and Evans, 2010). All photos are in crossed polarized light.

3.1.1.2 Mineralogy in the western part of the Qezel-Dagh barite

deposit
Based on mineralogical studies, hematite and goethite are the main ores in rhyolitic host

rocks and barite ores (Figures 3.9 and 3.10). Moreover, inclusions of pyrite within goethite

are the main ores and sometimes gangues are identified.

45
( a) (b)

(c) (d)

Figure 3.9. Photomicrographs of hematite mineralization with massive and disseminated textures in the western part of

the Qezel-Dagh deposit. Abbreviations: Hem = hematite, Gth = goethite (Whitney and Evans, 2010). All photos

are in reflected light.

46
( a) ( b)

( c) ( d)

Figure 3.10. Photomicrographs of hematite mineralization with massive and open-space fillings textures in the western

part of the Qezel-Dagh deposit. Abbreviations: Hem = hematite, Gth = goethite (Whitney and Evans, 2010). All

photos are in reflected light.

3.1.2 Petrology in the eastern part of the Qezel-Dagh barite

deposit
Quartz and alkaline feldspars, often sanidine and, to a lesser extent, orthoses occur as

phenocrysts and also as fine-grained crystals in the matrix of host rock. Alkaline feldspars

are altered to clay minerals. Pyrite, chalcopyrite, hematite, goethite, and limonite are opaque

minerals of rhyolite rocks. Rhyolitic tuff has a glassy matrix consisting of finegrained

crystals of quartz and alkaline feldspar that contain 50% volume of the rock (Figures 3.11–

3.13).

47
( a) (b)

( c) (d)

( e) (f)

Figure 3.11. Photomicrographs of glassy tuff with a vitroclastic texture as host rock of barite and iron

mineralization in the Qezel-Dagh deposit. Fine-grained matrix consists of alkali feldspar and quartz crystals.

Photos of (a), (c), (e), and (f) are in crossed polarized light and photos of (b) and (d) are in plane polarized

light of (a) and (c). Abbreviations: Afs = alkali feldspar, Chl = chlorite, Qz = quartz, Opq = opaque mineral

(Whitney and Evans, 2010).

48
(a) (b)

(c) (d)

(e) (f)

Figure 3.12. Photomicrographs of glassy tuff with a vitroclastic to vitrophyric texture as host rock of barite and

iron mineralization in the Qezel-Dagh deposit. Photos of (a), (c), and (e) are in crossed polarized light and photos

of (b), (d), and (f) are in plane polarized light of (a), (c), and (e). Abbreviations: Afs = alkali feldspar, Brt =

barite, Chl = chlorite, Qz = quartz, Opq = opaque mineral (Whitney and Evans, 2010).

49
(a) (b )

(c) (d)

(e) (f)

Figure 3.13. Photomicrographs of glassy tuff as host rock of barite and iron mineralization in the Qezel-Dagh

deposit. Photos of (a), (c), and (e) are in crossed polarized light and photos of (b), (d), and (f) are in plane

polarized light of (a), (c), and (e). Abbreviations: Afs = alkali feldspar, Brt = barite, Qz = quartz, Opq =

opaque mineral (Whitney and Evans, 2010).

3.1.2.1 Barite mineralization


50
Based on microscopic observations, three forms of barite mineralization are identified as

follows:

1- Barite mineralization occurs well-shaped (Figures 3.14 and 3.15) and elongated crystals

often with a wavy extinction (Figure 3.14) (Brt. 1: barite I). The boundary among barite

crystals is sometimes simple and in some cases jagged. Barite crystals with recrystallized

margins often have a wavy extinction and sometimes a mylonitic texture, indicating the

effect of tectonic stresses (Figures 3.15 and 3.16).

2- Barite mineralization occurs along cleavage surfaces of phenocrysts (the first form:

barite I) (Brt. 2: barite II).

3- Barite mineralization occurs fine-grained crystals in the margin of barite phenocrysts

of the first form (Figures 3.14 and 3.15) (Brt. 3: barite III).

In some parts of the eastern part of the Qezel-Dagh barite deposit, Fe-rich fluid filled

open spaces among barite minerals, industrially known as ferrous barite (Figure 3.16).

51
(a) (b )

(c) (d )

(e) (f)

Figure 3.14. Photomicrographs of different forms of barite mineralization in the eastern part of the Qezel-Dagh deposit.

Abbreviations: Brt. 1 = the first form of barite, Brt. 2 = the second form of barite, Brt. 3 = the third form of

barite (Whitney and Evans, 2010). All photos are in crossed polarized light.

52
(a ) (b )

(c ) (d )

Figure 3.15. Photomicrographs of different forms of barite mineralization in the eastern part of the Qezel-Dagh deposit.

Abbreviations: Brt. 1 = the first form of barite, Brt. 2 = the second form of barite, Brt. 3 = the third form of

barite (Whitney and Evans, 2010). All photos are in crossed polarized light.

53
(a ) (b )

(c ) (d )

Figure 3.16. Photomicrographs of different forms of barite mineralization associated with Fe in the eastern part of the

Qezel-Dagh deposit. Photos of (a) and (c) are in crossed polarized light and photos of (b) and (d) are in plane

polarized light of (a) and (c). Abbreviations: Brt. 1 = the first form of barite, Brt. 2 = the second form of

barite, Brt. 3 = the third form of barite, Opq = opaque mineral (Whitney and Evans, 2010).

3.1.2.2 Mineralogy in the eastern part of the Qezel-Dagh barite

deposit
Opaque minerals in the eastern part of the Qezel-Dagh barite deposit, on the basis of a

decreasing order of their percentage by volume, are as follows:

3.1.2.2.1 Hematite
More than 80% of the volume of hematite in barite ores are the result of the oxidation of

pyrite through the reaction with alkaline solutions. Sometimes pyrite is completely converted

to hematite and no trace of primary pyrites in hematites is not obvious (Figure

54
3.17). However, in some cases, pyrite partially converted to hematite and negligible traces

of primary pyrites in hematite are visible (Figures 3.18 and 3.19). In addition, in some cases,

pyrite occurs as non-altered and intact crystals with different forms in relation to barite

mineralization and sulfide minerals (Figures 3.17 and 3.18).

Flaky hematite, called specularite, is the major constituent and most abundant mineral in

rhyolite rocks in the eastern part of the Qezel-Dagh deposit. Specularite is found as single

crystals and sometimes in association with barite in the eastern part of the QezelDagh deposit

(Figure 3.20a and b). In this case, barite ore is converted to Fe-bearing barite. The most

important characteristic of hematite in the Qezel-Dagh deposit is the presence of blady

texture (Figure 3.20a and b) and vein and veinlet (Figure 3.19). This mineral is formed in

association with barite by the penetration of Fe solutions in joints, fractures, and open spaces.

Disseminated and colloform textures are other textures in hematites in the eastern part of the

Qezel-Dagh deposit. Disseminated texture is the specified texture of pyrite transformed to

hematite (Figure 3.19). In some specimens, hematite with a colloform texture is transformed

to goethite through hydration.

3.1.2.2.2 Goethite
Goethite formed through secondary supergene processes and further dehydration of

secondary hematites derived from the alteration of pyrite. The association of goethite with

hematite and the formation of colloform texture can confirm this claim. Colloform texture

(Figure 3.19), bluish anisotropy, and abundant internal reflections along with porous kidney-

shaped bands with a stringy texture are the other characteristic features of goethite under the

microscope. The formation of goethite and limonite in barite ore as disseminated and often

irregular crystals along fractures and the formation of sericite derived from the alteration of

alkali feldspars and chlorite in host rock indicate the oxidation of deposition environment

and the effect of supergene processes.

55
(a) ( b)

(c) ( d)

(e) (f)

Figure 3.17. Photomicrographs of Fe mineralization with replacement and semi-shaped granular textures and pyrite as

inclusions in Fe oxides and gangues in the eastern part of the Qezel-Dagh deposit. Abbreviations: Gth = goethite, Hem

= hematite, Lm = limonite, Py = pyrite (Whitney and Evans, 2010). All photos are in reflected light.

3.1.2.2.3 Pyrite
Pyrite is present as an accessory mineral in rhyolitic host rock along with barite

mineralization. In polished sections, this mineral with a pale yellow color is present as

disseminated crystals in host rock. Disseminated texture is the predominant texture in pyrites

and is often observed as semi-shaped to circular forms (Figure 3.18) in relation to


56
chalcopyrite, goethite, limonite, sericite, and chlorite in paragenetic sequence (Figure 3.18).

Pyrite is converted to Fe oxides in large quantities in response to a change in

physicochemical conditions of the mineral formation environment. In other words, sulfide

minerals are replaced by oxide minerals. Inclusions of pyrite in secondary hematite confirm

this claim. Hematite is also transformed into secondary minerals of goethite and limonite

during subsequent supergene processes. Due to their high oxygen fugacities and relatively

alkaline pH, penetrating solutions have a high oxidizing power. Therefore, they have

potential to change the chemical composition of sulfide minerals in joints and fractures and

these minerals are finally converted to oxide minerals. These solutions can partially or

completely convert pyrite in rhyolitic host rock to hematite.

Microscopic studies show the conversion of pyrite to hematite with a border and marginal

replacement texture, in which pyrite from its margins and boundary of this mineral with

other minerals is often converted to hematite. As a whole, pyrite is converted to hematite

during supergene processes. So that, secondary hematite follows texture of primary pyrite

and occurs as disseminated crystals. In some polished sections, semiformed primary pyrites

with a disseminated texture are identified, a conversation from the center to margin from

hematite to limonite (Figure 3.17). It is caused by transport of fluids with oxygen fugacity

among joints and fractures in the rock or on a microscopic scale by the effect of oxidation

solutions on cleavage surfaces.

3.1.2.2.4 Chalcopyrite
Chalcopyrite and pyrite are present as accessory minerals in rhyolitic host rock (Figure

3.18). This mineral has a yellow to brassy yellow color with a high reflectivity. This mineral

can be identified by a blady twinning and poor anisotropy under a microscope. This mineral

is found as semi-shaped to amorphous crystals in microscopic sections. This mineral can be

distinguished from pyrite by crystalline shape, cleavage, blady twinning, and brassy yellow

color.

57
After deposition of hydrothermal fluids and the formation of chalcopyrite, this mineral is

partially converted to covellite and other Cu sulfide minerals in response to changes in pH

and Eh conditions of the formation environment. As environmental conditions change from

reduction to oxidation, iron in chalcopyrite structure causes the generation of Fe oxides and

hydroxides, and following releasing the copper ions, if carbonate is present, Cu carbonate

minerals, such as malachite can generate.

(a) (b)

(c) (d)

Figure 3.18. Photomicrographs of chalcopyrite mineralization with amorphous to semi-shaped granular texture, pyrite
with spherical and disseminated textures, and hematite with replacement texture in the eastern part of the Qezel-Dagh

deposit. Abbreviations: Ccp = chalcopyrite, Hem = hematite, Py = pyrite (Whitney and Evans, 2010). All photos

are in reflected light.

58
(a ) (b )

(c ) (d)

(e ) (f)

Figure 3.19. Photomicrographs of Fe mineralization as vein, open-space fillings, and colloform textures in the eastern

part of the Qezel-Dagh deposit. Abbreviations: Gth = goethite, Hem = hematite, Lm = limonite, Py = pyrite

(Whitney and Evans, 2010). All photos are in reflected light.

59
(a) (b )

Figure 3.20. Photomicrographs of (a) specularite with a blady texture and (b) specularite with a semi-shaped granular

texture in the eastern part of the Qezel-Dagh deposit. Abbreviations: Hem = hematite, Py = pyrite (Whitney and

Evans, 2010). All photos are in reflected light.

3.2 Paragenetic sequence of minerals


Paragenesis is the association of minerals in rocks and rock suites considered in relation

to their origin (Bowes, 1989). The paragenetic sequence in mineral formation is an important

concept in deciphering the detailed geologic history of ore deposits. The sequence is worked

out through detailed microscopic studies in polished ore mineral sections, petrologic thin

sections and fluid inclusion studies as well as macroscopic field relations. In its broadest

sense, this term can be applied to any mineral assemblage or sequence of minerals, including

those in igneous, sedimentary and metamorphic rocks, as well as ore deposits (Kamilli,

1998). Mineral paragenesis is closely related to the concept of mineral zoning, where the

variation occurs in space, rather than time.

Paragenetic sequence is the sequential order of mineral deposition during formation of an

ore deposit, or sequence of mineral paragenesis. In other words, the concept of mineral

paragenesis applies to characteristic associations of minerals in ore deposits that connote

contemporaneous formation, presumably in chemical equilibrium (Kamilli, 1998). Since the

generation of a deposit is related with a set of processes, each of these processes has a

specific effect on host rock and associated minerals, study of the chronological order of

mineral formation and its relationship with other minerals can describe the relationship of

60
different ores, gangues, and alteration minerals. In addition to ores and minerals, mineral

paragenesis can explain not only the deposition time of gangue minerals and time period of

their deposition, but also the peak of formation of minerals that usually shown as thickness

of each bar relative to the main events spent in the region. In general, mineral paragenesis

describes sequences or assemblages of minerals related by common formation conditions.

Considering that mineralogical studies and mineral paragenesis provide useful information

about depositional environment of ore minerals, in this study, paragenetic sequence of

minerals is given in Figure 3.21.

Based on petrographic and mineralogical studies, in order of decreasing frequency,

mineral paragenesis in the Qezel-Dagh barite deposit is barite, specularite, goethite,

limonite, sericite, pyrite, chalcopyrite, and malachite. Barite along with silica and pyrite and

chalcopyrite are precipitated from hydrothermal fluids as primary minerals. However,

secondary minerals are the product of these primary minerals during supergene processes.

61
Figure 3.21. Paragenetic sequence of minerals in the Qezel-Dagh barite deposit.

62
CHAPTER 4:

Geochemistry and isotopic

analyses

63
4.1 Geochemistry of rhyolite of the Mahabad Formation

Table 4.1 lists trace element compositions of the least-altered rhyolite of the Mahabad

Formation at Qezel-Dagh. Contents of large ion lithophile elements (LILE) Rb, Ba, Cs, and

Sr in rhyolite of the Mahabad Formation are in the range of 7–190 ppm Rb, 154–6716 ppm

Ba, 0.7–18.7 ppm Cs, and 11.7–504 ppm Sr. The rhyolites of the Mahabad Formation exhibit

low concentrations of high field strength elements (HFSE), such as 0.76–8.19 ppm Hf, 10–

174 ppm Zr, 1.1–7.8 ppm Nb, 0.27–1.27 Ta, 44–1721 ppm Ti, and 0.89–18.99 ppm Th.

Similarly, the rhyolites are characterized by low concentrations of high field strength

elements (HFSE), such as 1.7–7.3 ppm Sc, 4–20 ppm V, 5–15 ppm Cr, 1–3 ppm Ni, 8– 744

ppm Cu, expect for one sample has 34589 ppm Cu, and 35–438 ppm Zn. Among HFSE, Ta

has the lowest concentration and among LILE, Cs has the lowest concentrations.

Results of rare earth element (REE) analysis are given in Table 4.2, and in

chondritenormalized REE patterns are presented in Figure 4.1. Contents of total RRR

(ΣREE: La– Lu) in rhyolite of the Mahabad Formation are in the range of 72.45–433.88 ppm

with an average of 189.73 ppm. The concentration of light REE (LREE: La–Eu = 59.67–

409 ppm) is much higher than that of heavy REE (HREE: Gd–Lu except Ho = 10.98–30.23

ppm). The chondrite-normalized REE patterns display a LREE enrichment, HREE

depletion, mainly, pronounced negative to slightly positive Eu anomaly (Eu/Eu* = 0.34–

1.19; average = 0.61), and pronounced Yb positive anomaly (Figure 4.1). There is not a

significant difference between values of chondrite-normalized patterns of the barite samples,

likely evolved from a common primary magma.

64
Table 4.1. Chemical analyses of major and trace elements, and elemental ratios from whole rock of rhyolite
samples from the Mahabad Formation.
Samples P2-S1 P2-S2 P2-S4 P2-S5 P2-S6 P2-S7 P2-S9 P2-S10 P2-S15 P2-S16 P2-S24
Ta (ppm) 0.64 0.94 0.85 0.6 1.27 0.35 0.98 0.33 0.55 0.98 0.27
Te <0.1 0.6 <0.1 0.52 <0.1 1.38 <0.1 <0.1 0.64 <0.1 <0.1
Th 7.58 14.43 10.7 9.93 16.19 0.89 12.55 2.41 9.37 18.99 <0.1
Ti 411 830 769 335 1721 138 1109 44 521 992 <1
Tl 0.19 2.43 2.16 8.11 5.59 1.44 4.65 3.2 3.28 6.58 1.54
U 3.4 5.4 4 2.6 9.3 4.9 5.6 3.4 4.9 4.3 1
V 7 4 5 10 16 10 11 7 20 17 5
W 1.7 1.4 <1 1.1 6.1 16.9 7.4 2.3 3.9 4 <1
Y 18.5 48.1 17.3 7.1 33 11.7 13.9 13.9 11.9 12.9 22.7
Zn 100 72 50 148 54 87 53 352 35 71 438
Zr 77 122 98 47 174 30 79 17 65 101 10
Ag 0.2 0.2 0.6 5.9 8.7 8 1.2 2.1 4.4 2.6 12.9
Al 31930 60410 43481 40457 76826 14245 51526 9446 37039 66747 2955
As 12.5 4.9 7 >100 >100 >100 >100 >100 >100 >100 >100
Ba 462 6038 374 154 266 1820 4849 >1 263 434 6716
Be 0.3 2.1 1.5 1.2 2.2 0.5 1.3 <0.2 0.9 1.5 <0.2
Bi 6.1 1.1 0.5 20.4 154 187 36.1 13.1 286 83.1 55.2
Ca 888 780 660 8744 45238 2564 32979 3209 1106 1212 42931
Cd <0.1 <0.1 <0.1 <0.1 0.2 1.6 <0.1 0.5 0.4 0.6 6.1
Co 4.6 5.6 6.1 9.4 3.7 15.6 4.3 24.6 8.6 9 34
Cr 6 5 7 7 13 10 11 8 15 10 6
Cs 0.7 6.1 10.1 7.1 18.7 1.9 7 0.8 5.2 5.9 <0.5
Cu 18 8 24 93 21 164 156 744 335 183 34589
Fe 69776 51091 40859 >10 52317 >10 81866 >10 >10 >10 >10
Hf 3.59 60.3 4.7 2.61 8.19 1.07 4.51 0.76 2.97 5.07 <0.5
In <0.5 <0.5 <0.5 <0.5 <0.5 2.11 0.7 3.08 0.55 <0.5 6.06
K 1965 43776 18032 11407 23850 4763 18659 2896 13028 34610 1257
Li 39 19 17 14 22 5 26 5 16 18 5
Mg 19292 4186 1660 2556 5073 1392 4358 1546 3038 4420 782
Mn 321 816 1286 838 98 241 168 6160 161 195 3803
Mo <0.1 <0.1 1 <0.1 6 32 7 1 10 5 6
Na 171 421 306 154 324 <100 178 <100 265 379 <100
Nb 3.4 6 5.1 2.9 7.8 3 7.6 3.1 2.7 4.8 1.1
Ni 3 2 2 1 2 2 1 3 1 1 <1
P 38 54 35 291 143 318 181 216 256 414 422
Pb 3 14 5 29 109 161 15 12 94 78 64
Rb 13 190 97 85 147 45 151 30 96 166 7
S 225 1996 151 437 1531 1809 1519 2976 592 2878 2497
Sb 6.7 8 12.2 24.5 21.3 36.9 13.7 21.2 28.6 23 >0.01
Sc 4.4 6.4 7.3 5.7 4.1 2.7 2 1.7 2.3 4.8 2
Se <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5
Sn 1.4 1.4 0.9 1.6 5.1 3.9 2.7 0.8 3.4 1.4 0.8
Sr 12.1 504 11.7 14.7 29.5 39.8 67.9 133 20 32.5 131
Zr/Ti 0.19 0.15 0.13 0.14 0.10 0.22 0.07 0.39 0.12 0.10 –
Nb/Y 0.18 0.12 0.29 0.41 0.24 0.26 0.55 0.22 0.23 0.37 0.05

65
Table 4.2. Chemical analyses of REE and some elemental indices from whole rock of rhyolite samples from
the Mahabad Formation.
Samples P2-S1 P2-S2 P2-S4 P2-S5 P2-S6 P2-S7 P2-S9 P2-S10 P2-S15 P2-S16 P2-S24
La (ppm) 39 32 11 27 39 23 45 119 15 31 9
Ce 85 71 29 60 92 45 84 165 30 63 17
Pr 13.72 10.74 3.16 10.7 12.97 8 13.22 22.46 4.43 8.94 4.04
Nd 56.4 45.8 12.6 46.4 53.8 36.6 61 77.9 20.6 35.6 26.2
Sm 10.81 11.82 3.44 12.02 13.64 11.57 16.95 18.86 3.31 9.16 16.05
Eu 1.42 3.17 0.47 1.26 1.59 1.77 2.71 5.78 0.33 1.1 3.81
Gd 8.29 8.98 3.49 7.69 10.04 8.16 5.9 11.62 2.61 6.42 14.06
Tb 0.89 1.36 0.56 0.67 1.12 0.86 0.67 1.2 0.47 0.65 1.63
Dy 4.96 9.36 3.74 2.55 6.76 3.9 3.29 5.35 2.55 3.13 7.72
Er 2.91 5.87 2.58 1.42 4.32 1.42 1.9 1.79 1.7 1.87 2.03
Tm 0.45 0.85 0.43 0.22 0.59 0.19 0.29 0.22 0.28 0.29 0.26
Yb 2.2 3.1 1.6 3 2.7 4.8 1.9 4.5 3.1 2.8 3.8
Lu 0.48 0.71 0.38 0.18 0.58 0.17 0.27 0.2 0.27 0.25 0.22

LREE (La–Eu) 206.35 174.53 59.67 157.38 213 125.94 222.88 409 73.67 148.8 76.1
HREE (Gd–Lu) 20.18 30.23 12.78 15.73 26.11 19.5 14.22 24.88 10.98 15.41 29.72
REE (La–Lu) 226.53 204.76 72.45 173.11 239.11 145.44 237.1 433.88 84.65 164.21 105.82
0.46 0.94 0.41 0.41 0.83 1.19 0.34 0.44 0.77
0.89 0.92 1.18 0.99 0.82 0.72 0.88 0.90 0.68

Figure 4.1. Chondrite-normalized REE patterns for rhyolite samples of the Mahabad Formation.

Values of the chondrite are from McDonough and Sun (1995).

On the diagram of Winchester and Floyd (1977), the samples show a mainly rhyodacitic and
dacitic composition (Figure 4.2a). The volcanic rocks are located mainly in the rhyolite,
rhyodacite, and dacite fields on the Zr/Ti–Nb/Y diagram of Pearce (1996) (Figure 4.2b).

66
67
Figure 4.2. Plot of rhyolite of the Mahabad Formation on Zr/Ti vs. Nb/Y diagram of (a) Winchester

and Floyd and (b) of Pearce (1996).

On the Th–Co diagram of Hastie et al. (2007), rhyolites of the Mahabad Formation mainly

plot within the field of high-K calc-alkaline and shoshonite series (Figure 4.3). According

to investigations by Fathi Sadi (2012), calc-alkaline and shoshonite magmatism, which

developed in a post-collisional tectonic environment, are typical of rhyolitic rocks. Rhyolitic

rock development in the Kahar development hence points to the prevalence of extensional

regimes following continental collisions in the Qezel-Dagh region (approximately 550 Ma

ago).

Figure 4.3. Plot of rhyolite of the Mahabad Formation on Th-Co diagram of Hastie et al. (2007).

Rhyolites of the Mahabad Formation in primitive mantle-normalized multi-element

diagrams are characterized by an enrichment of Pb and Nd and depletion of Sr, Y, and some

HFSE (e.g., Ti and Nb) (Figure). Pronounced depletions of Ti and Nb in the studied samples

reflect the involvement of mineral phases carrying these elements (e.g., amphibole, titanite,

ilmenite, and rutile) in the source. According to Nicholson et al. (2004), in chondrite-

normalized REE and primitive mantle-normalized multi-element diagrams, depletions in Ti

68
and Nb along with enrichments in LREE and Pb indicate a subduction-related continental

arc setting for rhyolites of the Mahabad Formation. Moreover, enrichments in U and Pb, and

negative Nb, P, Sr, and Ti anomalies reflect that rhyolites of the Mahabad Formation at

Qezel-Dagh are genetically fractionated granitoids. When the data were plotted on a

primitive mantle-normalized diagram, rhyolites samples of the Mahabad Formation

displayed strong enrichment in some large LILE (e.g., Cs and Ba) and LREE, as well as

strong negative anomalies for P, Zr, Nb, and Ti, and strong positive anomalies for Pb.

Figure 4.4. Primitive mantle-normalized multi-element spider diagrams for rhyolite of the

Mahabad Formation. Values of the primitive mantle are from Sun and McDonough (1989).

69
4.2 Geochemistry of barite

4.2.1 Introduction
Barite's elemental and isotopic composition may be geochemically analyzed to provide

important details about the sample's origin, including whether it was hydrothermal,

diagenetic, cold seep, or marine. Concentrations of REE in hydrothermal fluids may yield

useful data regarding the origin of ore-forming elements (Kato, 1999). As REE can be

fractionated during geochemical processes, REE patterns normalized to standard material

have been widely used as a tool to reveal various geochemical processes (Kato, 1999).

The Ba source for the barite can be determined based on the geochemical study, using

REE to place it in the environment of its deposition, i.e., continental or marine (Guichard et

al., 1979; Baioumy, 2015). According to Guichard. (1979), barite, which occurs in both

terrestrial and marine environments, could offer a rewarding system to study REE

geochemistry because: (1) the REE pattern for seawater has a characteristic “V” pattern and

negative Ce anomaly (Hogdahl et al., 1968). (2) Marine sediments are distinct from basic

rock or hydrothermal patterns (Goldberget 1963; Haskinal., 1968) and may distinguish the

marine versus magmatic modes of barite formation. (3) The enrichment of Eu(II) at the

expense of Eu(III) in barite could reveal redox conditions at the loci of formation. (4) The

natural fractionation of REE by low temperature aqueous minerals such as barites might be

useful in interpreting crystallization chemistry, rare-earth solution complexing and the

utilization of REE for interpreting both diagenetic and late stage magmatic events.

The distinction between marine and non-marine barites has been made using both REE

concentrations and patterns. High REE concentrations are typical in deep-marine barites,

while low REE concentrations are seen in terrestrial and hydrothermal barites. According to

Guichard (1979), deep sea barites often contain 10-100 times more REE than those

discovered on continents. The Southern Californian rifted continental margin barite, which

was produced by low-temperature hydrothermal fluids flowing along faults, likewise had

extremely low quantities of REE, according to Hein's 2007 research.

70
Regarding the REE pattern, deep-marine barites exhibit chondrite-normalized Eu minima

but not negative Ce anomalies of seawater, whereas those types of reducing sedimentary and

metamorphic origin exhibit chondrite-normalized positive Eu anomalies (e.g. Guichard

1979). Hein (2007) exploited LREE enrichment in chondrite-normalized REE patterns of

the Southern California rifted continental margin barite to show that this barite has a typical

marine hydrothermal origin.

4.2.1.1 Major oxides and trace elements


Other than the BaSO4 contents, which range between 79.9 and 88.01 weight percent,

barite samples show modest quantities of significant oxides as Al2O3 (0.32-1.93 wt.%),

Fe2O3 (0.15–3.76 wt.%), TiO2 (0.18–0.22 wt.%), K2O (0.13–0.42 wt.%), and Na2O (1.03–

1.32 wt.%). MgO and P2O5 exist at extremely low amounts, less than 0.1 wt.% (Table

4.3).

The concentrations of trace elements Rb, Cs, Co, Nb, Ta, Th, U, Y, and Hf are below 5

ppm in the analyzed samples. The concentration of Sr falls in the range of 1110.2–2168.3

ppm, corresponding to 1.72–2.45 wt.% SrO in barite. The low concentration of Sr suggests

that barite belongs to the end-member barite in the barite-celestite solid solution series. The

high Sr concentrations in barite deposits is related to the similarities in the atomic radii of Sr

and Ba, which leads to isomorphic replacement of barium found in barite crystals by

strontium. No matter where the barite comes from, it appears to have the same amount of Sr

(Safina, 2015). The range of Sr in the examined barite samples, however, is within the range

reported for barite in both historic and contemporary seafloor hydrothermal sites (400-

100,000 ppm; Safina- 2015), showing an origin of the barite deposits from a lowtemperature

hydrothermal solution in the Qezel-Dagh deposit, as suggested by Kato and Nakamura

(2003) and Jurković et al. (2010) for barite deposits from western Australia and southeastern

Bosnia.

71
Table 41.3. Chemical analyses of major and trace elements of barite samples from the Qezel-Dagh

deposit.

Samples Sher-S2 Sher-S3 Sher-S4 Sher-S7 Sher-S8 Sher-S11 Sher-S13 Sher-S15 Sher-S16 Sher-S17
SiO2 5.26 7.15 8.69 5.85 8.26 9.81 5.58 9.05 5.21 7.82
Al2O3 1.28 1.47 1.79 1.68 0.99 1.93 0.51 0.32 0.43 1.82
Fe2O3 1.35 3.76 2.22 3.54 1.14 2.36 0.32 0.18 0.15 0.36
CaO 1.32 0.03 0.92 0.04 0.61 1.45 1.18 0.81 0.52 0.74
MgO 0.07 0.05 0.06 0.04 0.5 0.06 0.06 0.05 0.05 0.06
Na2O 1.32 1.03 1.23 1.25 1.23 1.26 1.05 1.24 1.31 1.25
K2O 0.26 0.22 0.35 0.33 0.23 0.42 0.15 0.13 0.13 0.37
SO3 20.6 19.44 19.56 19.72 19.96 19.36 20.61 20.13 20.91 20.21
TiO2 0.21 0.22 0.22 0.19 0.19 0.18 0.22 0.19 0.21 0.22
MnO 0.03 0.01 0.02 0.08 0.06 0.02 0.08 0.08 0.06 0.05
P2O5 0.02 0.03 0.01 0.02 0.02 0.02 0.03 0.03 0.03 0.02
SrO 2.11 2.18 2.09 2.45 1.94 2.02 2.15 1.72 2.12 2.23
BaO 65.07 62.33 61.29 62.88 63.05 60.54 66.03 64.02 67.1 61.71
LOI 1.06 2.05 1.28 1.67 1.71 0.51 1.81 2.03 1.75 3.09
Total 99.96 99.97 99.73 99.74 99.89 99.94 99.78 99.98 99.98 99.95
Co 1.3 1.6 1.5 1.5 1.6 1.2 1.2 1.4 1.3 1.6
Cr 26 14 28 15 44 11 41 42 63 74
Cs 0.24 0.22 0.09 0.18 0.08 0.06 0.14 0.05 0.14 0.04
Cu 6 6 7 5 8 7 7 8 9 9
Ga 4.9 6.6 4.5 5.4 5.6 3.6 5.2 5.6 4.5 4.4
Hf 0.6 0.5 0.5 0.3 0.4 0.5 0.3 0.4 0.5 0.5
Mo 6 17 12 4 14 9 11 12 7 11
Nb 3 0.4 0.4 0.5 0.4 0.5 0.3 0.4 0.4 0.3
Ni 25 20 22 25 22 22 21 22 23 21
Pb 9 13 13 14 15 12 13 14 13 16
Rb 0.9 0.8 0.8 0.4 0.7 0.9 0.8 0.7 0.7 0.8
Sn 5 3 4 4 5 3 6 5 4 7
Sr 1550.3 1210.3 1835.1 1110.2 1584.5 1983.2 2168.3 1487.2 1470.2 2155.3
Ta 0.5 0.6 0.4 0.3 0.4 0.4 0.3 0.3 0.3 0.3
Th 1.77 0.52 0.88 0.76 1.62 0.32 2.33 1.81 2.51 2.33
Tl <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5
U 0.66 0.92 0.32 0.65 0.38 0.12 0.51 0.26 0.61 0.13
V 7 18 10 12 11 6 9 7 11 8
W 4 2 3 3 3 3 4 4 3 3
Y 0.9 0.5 0.6 0.6 0.8 0.7 1.9 0.8 0.8 0.9
Zn 5 8 6 6 7 5 6 6 5 7
Zr 3 3 4 2 3 4 4 3 4 4

1
4.2.1.2 Rare earth substances
Table 4.4 provides a list of the REE (La-Lu) proportions found in samples of barite. The

total concentration of REE ranges from 5.95 to 54.09 ppm with an average value of 24.23

ppm. The low REE concentrations of barites from the Qezel-Dagh deposit likely suggest a

hydrothermal origin for the barites. The values of light REE (LREE: La–Eu) and heavy

72
REE (HREE: Gd–Lu) range from 4.1 to 24.0 ppm and 1.85 to 37.79 ppm, respectively. The

LREE/HREE ratio in the barite samples is in the range of 2.22 to 10.8, except for one sample

that is 0.43, indicating that barite is enriched in LREE relative to HREE. The (Nd/Yb) N and

(La/Yb)N ratios are used to better understand the enrichment trends in terms of HREE and

LREE contents in barites. These ratios for the studied barites range from 1.37 to 22.14 with

an average of 11.56 for (Nd/Yb)N ratio and from 1.61 to 24.26 with an average of 15.05 for

(La/Yb)N ratio, indicating that in the studied samples, LREE is enriched relative to HREE.

Such LREE contents enrichments in the barite may support epithermal process originated at

the upper continental crust. Similar enrichment values are observed for Southern California

barite ores studied by Hein et al. (2007). The Eu and Ce anomalies are calculated as Eu/Eu*

= EuN/(√SmN × GdN) and Ce/Ce* = 2CeN/(LaN + PrN), respectively. The suffix “N” refers to

the normalization of concentrations against the chondrite of McDonough and Sun (1995).

The studied barite samples show positive Eu and Ce anomalies (Eu/Eu* = 2.56–8.92 and

Ce/Ce* = 1.61–3.55).

Table 4.4. Chemical analyses of REE and elemental indices of barite samples from the Qezel-Dagh
deposit.

Samples Sher-S2 Sher-S3 Sher-S4 Sher-S7 Sher-S8 Sher-S11 Sher-S13 Sher-S15 Sher-S16 Sher-S17
La 2.4 0.8 2.4 0.4 1.9 3.1 3.3 2.3 1.7 2.5
Ce 11.5 5.2 10.1 2.6 8.6 12.4 11.9 9.3 7.8 10.3
Pr 0.79 0.13 0.67 0.07 0.61 0.91 0.94 0.71 0.56 0.81
Nd 3.9 0.7 3.5 0.3 3.1 4.7 4.8 3.5 2.9 4.4
Sm 1.78 0.87 1.51 0.42 1.28 1.78 1.58 1.33 1.31 1.56
Eu 1.11 0.61 0.95 0.31 0.81 1.11 0.96 0.83 0.82 0.96
Gd 0.63 0.05 0.57 0.03 0.49 0.77 0.83 0.59 0.49 0.73
Tb 0.11 0.06 0.09 0.03 0.08 0.11 0.11 0.08 0.07 0.09
Dy 0.41 0.05 0.38 0.03 32 0.51 0.53 0.36 0.31 0.47
Ho 0.16 0.06 0.17 0.03 0.16 0.22 0.23 0.17 0.19 0.25
Er 1.06 2.98 1.25 1.62 4.12 0.32 3.46 6.03 4.01 1.36
Tm 0.11 0.08 0.08 0.04 0.08 0.09 0.08 0.07 0.08 0.08
Yb 0.11 0.05 0.09 0.03 0.8 0.12 0.11 0.09 0.07 0.07
Lu 0.08 0.08 0.08 0.04 0.06 0.08 0.06 0.06 0.07 0.08
LREE (La–Eu) 21.48 8.31 19.13 4.1 16.3 24 23.48 17.97 15.09 20.53
HREE (Gd–Lu) 2.67 3.41 2.71 1.85 37.79 2.22 5.41 7.45 5.29 3.13
REE (La–Lu) 24.15 11.72 21.84 5.95 54.09 26.22 28.89 25.42 20.38 23.66
LREE/HREE 8.04 2.44 7.06 2.22 0.43 10.81 4.34 2.41 2.85 6.56
(La/Yb)N 14.82 10.87 18.12 9.06 1.61 17.55 20.38 17.36 16.50 24.26
(Nd/Yb)N 12.49 4.93 13.70 3.52 1.37 13.80 15.37 13.70 14.60 22.14
Eu/Eu* 3.20 8.92 3.12 8.42 3.12 2.89 2.56 2.86 3.12 2.74
Ce/Ce* 2.01 3.55 1.90 3.47 1.92 1.77 1.61 1.75 1.93 1.74
Ce/La 4.79 6.50 4.21 6.50 4.53 4.00 3.61 4.04 4.59 4.12

73
Chondrite-normalized REE patterns for barite samples from the Qezel-Dagh deposit are

shown in Figure 4.5. These patterns of the barite sample indicate that LREE are enriched

compared to HREE, and that REE contents of the Qezel-Dagh barite deposit are generally

low. The enrichment of LREE compared to HREE is a typical feature of barite, because the

greater similarity of LREE ion sizes (compared to HREE) with those of Ba2+ results in LREE

being concentrated in barite when REE are substituted in the barite structure Guichard

(1979). The LREE-enriched patterns in a few marine precipitates are explained as having

formed as a result of deposition in those marine settings that had significant inputs of fluid

from submerged hydrothermal vents (Chenal, 2006). However, in the case of barite, this

conclusion cannot be reached due to crystallographic challenges with HREE replacement in

crystal structure. As a result, barite samples from the Qezel-Dagh deposit have LREE-

enriched patterns that likely indicate the presence of active hydrothermal vents close to the

site of barite deposition.

Figure 4.5. Chondrite-normalized REE patterns for barites from the Qezel-Dagh deposit. Values of

the chondrite are from McDonough and Sun (1995).

74
Ce/La ratio and Ce anomaly are the final two REE geochemical parameters that may be

used to determine the circumstances in which barite developed. According to Guichard et

(1979), the Ce/La ratio in barite can be used as a distinguishing feature between marine and

terrestrial barite. This ratio is <1 and resembles that of deep-marine barite, while it is >1 and

resembles that of basic rocks and clays in terrestrial (vein) barite (Gui-chard al., 1979). The

studied barites have Ce/La ratios ranging from 3.61 to 6.50, with an average value of 4.69,

similar to terrestrial barite.

REE, especially the Eu anomaly, have been used as a tracer to evaluate possible

submarine hydrothermal inputs. Positive Eu anomalies are commonly found in ores and

chemical precipitates associated with massive sulfide deposits of submarine hydrothermal

exhalative origin (Jiang, 2006). It is believed that circulating hydrothermal fluids react with

Eu-enriched feldspar-bearing lithologies in the deep marine basin, which would mobilize Eu

and, thus, cause a pronounced increase of Eu in the ore-forming solutions and their

precipitates (Graf, 1977). The positive Eu anomalies of the barite samples from the Qezel-

Dagh deposit in chondrite-normalized REE patterns suggest a hydrothermal source for barite

mineralization at Qezel-Dagh.

At Qezel-Dagh deposit, Ce/La and Ce/Ce* ratios are >1 in all samples. It is well known

that Ce4+ is effectively scavenged by Mn- and Fe-oxides, resulting in a positive Ce anomaly

in deposits containing them (Murray, 1990). It is fairly probable that the anomalous high

Ce/La and Ce/Ce* ratios are resulted from impurities of Mn- and Feoxides in the analyzed

barites samples. Moreover, positive Ce anomaly is reported by Gui chard, (1979) in a barite

sample formed within a manganese crust from the Indian Ocean.

4.2.2 Comparing the Qezel-Dagh barite deposit with other places

where the mineral occurs

75
The Qezel-Dagh barite sample's chondrite-normalized REE pattern has a zigzag pattern

and is near to 10 as illustrated in Figure 4.6. Reviewing the variations in the REE values of

the barite samples reveals substantial positive anomalies for Ce, Eu, Er, and Lu, as well as

negative anomalies for La, Pr, Dy, and Yb. contrasting the chondrite-normalized REE

pattern of samples of barite to that of barite with an oceanic origin (Gui chard al., 1979;

Murray, 1990; Alexander, 2008) Several similarities between the Qezel-Dagh barite deposit

and low temperature hydrothermal barite deposits have been noted (Gui-chard, 1979; Hein,

2007).

76
Figure 4.6. Comparison of chondrite-normalized REE patterns of the Qezel-Dagh barite deposit

and other barite deposits. Data are from Griffith and Paytan (2012), Guichard et al. (1979), and

Hein et al. (2007).

4.3 Isotopic studies

77
4.3.1 Introduction
The source(s) of the fluids from which barite precipitated can be determined by the

isotopic and chemical fingerprints of major (such as S, O, Ba, and Sr) and trace elements

(such as Ca, REE, and Pb) in the mineral. Because of this, isotopic analysis of a barite

sample's main elemental components may frequently restrict the sample's provenance. The

chemical and isotopic composition recorded in barite may, in turn, provide information on

the chemistry of formation fluids and specifically could be used to reconstruct changes in

seawater chemistry over time (Griffith and Paytan, 2012). Variations in sulfur isotope

compositions in natural systems result from different chemical exchange reactions and can

be enhanced by bacterial reduction of sulfate (Hoefs and Sywall, 1997; Sealet, 2000). Sulfur

isotopic signatures of marine and evaporated sulfate minerals have changed throughout

geological time, providing a record of secular variations in seawater sulfate (Claypoolal,

1980).

The sulfur isotopic composition (δ34S) of barite is quite similar to the SO4 from which it

precipitated (<0.4 per mil difference) (Kusakabe and Robinson, 1977). Accordingly, barite

δ34S will record the sulfur isotopic composition of the formation fluids. Depending on the

mode of barite formation, several potential sources of SO4 may be available for barite

precipitation, including seawater SO4 (δ34S = +21‰, in the modern ocean), magmatic sulfate

(δ34S = +1 to +2‰), pore water SO4 modified by microbial reduction, SO4 from calcium

sulfate minerals and SO4 produced by the oxidation of reduced sulfur species (Griffith and

Paytan, 2012).

The sulfur isotopic signature in barite can be used to distinguish the mode of barite

formation (Paytan et al., 2002). Marine barite records contemporaneous seawater sulfur

isotope values. Marine barite which precipitates in the water column records changes in

oceanic sulfate δ34S through time (Paytan , 1998, 2004). Because sulfur has a long residence

time in the ocean (>10 Myr; Walker, 1986), the isotopic composition of seawater sulfate is

uniform throughout the oceans at any given time (Reeset, 1978). The oceanic sulfate δ34S at

78
any given time is controlled by the relative proportion of sulfide and sulfate input and

removal from the oceans and their isotopic compositions and this, in turn, influences the

oxygen content of the atmosphere (Holland, 1973; Berner and Canfield, 1989; Berner, 1999;

Paytan and Arrigo, 2000). In addition, knowledge of the δ 34S of seawater can shed light on

potential changes in the S sources to the ocean such as river runoff, volcanism and

hydrothermal activity (Paytan et al., 2004), and the size of the oceanic sulfur reservoir

(Wortmann and Chernyavsky, 2007). The use of barite for the reconstruction of seawater

δ34S increased the temporal resolution and uncertainty associated with records based on

evaporites (Claypool et al., 1980) and revealed previously unrecognized fine structure in the

record.

Diagenetic barite typically precipitates from fluids that have undergone some degree of

SO4 loss due to BSR, the precipitated barite is expected to have high sulfur isotope ratios

similar to those in the residual SO4 of these fluids (Griffith and Paytan, 2012). In the Peru

margin and in the Japan Sea, Torres et al. (1996b) measured highly enriched sulfur isotope

ratios in barite (δ34S up to +84‰). Torres et al. (1996b) suggested that these barites formed

as a result of sulfate reduction of marine barite, its remobilization and subsequent re-

precipitation as a diagenetic barite front.

The sulfur isotope ratio in hydrothermal barites is equivalent to or slightly less than that

of contemporaneous seawater due to the relative quantity of sulfur integrated from seawater

SO4 (+21 at present) or from the oxidation of hydrothermal H2S with an isotopic signature

of around +1 to 2. (Griffith and Paytan, 2012), depending on the relative contribution of

hydrothermal sulfate derived from the oxidation of H2S (Ehya, 2012; Griffith, 2018). The

BSR of seawater may also take place in hydrothermal environments, resulting in a high

sulfur isotope ratio in some hydrothermal barite deposits (Griffith and Paytan, 2012).

However, barite samples from the Guaymas Basin hydrothermal vent area in the Gulf of

California with δ34S values higher than seawater and hydrothermal sulfide are an exception.

This is likely due to sulfate from fluids that were modified as they moved through the

sediment-covered ridge system where BSR was active (Griffith et al., 2018). Lüders et al.
79
(2001) explained the range of sulfur isotope values in barites found in stockwork, smokers

and mounds on the sea floor in the Central Okinawa Trough of Japan to suggest mixing of

hydrothermal solutions with seawater to form the barites. Locally heavy δ 34S of barite was

related to partial SO4 reduction as well (Lüders et al., 2001).

Marine cold seep barites are distinguished by δ34S values that range significantly from

values similar to seawater to levels that are four times the sulfate amount found in seawater.

Cold seep barite sulfur isotope values represent a mixture of seawater and pore fluid

signatures. Stevens et al. (2015) reported cold seep barite from the Gulf of Mexico that

exhibits low variability in δ34S and is slightly enriched in 34


S compared to bottom water

sulfate. Extremely high δ34S values are found in barites precipitated from fluids that have

undergone varying degrees of microbial sulfate reduction, either by dissimilatory sulfate

reduction or anaerobic oxidation of methane associated with sulfate reduction (Griffith et

al., 2018, and references therein). In addition, the slow replenishment from the overlying

seawater causes sulfate in the pore water to gradually becomes depleted in 32S isotope due

to progressive BSR consumption (Zhou et al., 2015). Similarly, in the California continental

margin, Naehr et al. (2000) interpreted systematic changes in S isotope ratios of barite

samples precipitated on the sea floor to reflect changes in the relative contribution of SO 4

from pore fluids and seawater resulting from changes in pore water flow rates.

Detailed δ34S analyses suggest that most massive bedded (stratiform) barite deposits have

a seawater sulfate source and thus probably formed from rapid Ba-rich fluid flow to the sea

floor (Jewell, 2000). Results of sulfur isotope analyses for numerous SEDEX barite deposits

have been interpreted to reflect barite formation from either pristine seawater sulfate

(Goodfellow and Jonasson, 1984; Jewell, 2000) or seawater sulfate isotopically modified by

BSR (Rye et al., 1978; Paytan et al., 2002; Torres et al., 2003). Pelagic marine barites record

the δ34S value of seawater sulfate with a small isotopic deviation (<0.4‰; Paytan et al.,

1998), as widely used in the reconstruction of the sulfur isotope curve for seawater sulfate

through geologic time (e.g., Paytan et al., 1998, 2004).

80
In the present day ocean, the δ18OSO4 is about +9.3‰ Vienna Standard Mean Ocean Water

(VSMOW), which is significantly out of equilibrium with the ocean's 18O, or +0 VSMOW

(Lloyd, 1968). According to Chiba and Sakai (1985), this discrepancy can be attributed to

isotopic non-equilibrium, which persists under most oceanic circumstances of temperature

and pH due to the sluggish kinetics of isotopic exchange between oxygen in SO4 and

seawater. "VSMOW" (Lloyd, 1968). Equilibrium is not expected to occur within the

estimated oceanic residence time of seawater SO4 (>10 Myr). Therefore, the seawater

δ18OSO4 will reflect the source of SO4 to the ocean (weathered evaporites +11 to +13‰ and

oxidative pyrite weathering (–4 to +2‰), the biogeochemical cycling of SO4 in the ocean

(for example, SO4 reduction and sulfide oxidation) and the sinks of SO4 from the ocean (for

example, gypsum, anhydrite and pyrite burial) (Claypool et al., 1980). Marine barite should

reflect the contemporaneous marine SO4 isotopic composition and thus will also not record

equilibrium values (Griffith and Paytan, 2012).

In hydrothermal barite, formed at temperatures above 150 °C and near-neutral pH, the

oxygen isotopic composition of SO4 is in isotopic equilibrium with environmental water

(Chiba and Sakai, 1985). The oxygen isotopic fractionation factor of the SO4–H2O system

is temperature dependent and can therefore be used as a geothermometer to infer the

conditions under which SO4 formed (Ohmoto and Lasaga, 1982). Some marine

hydrothermal barites have δ18O values similar to that of modern seawater, while others are

slightly lower than that of seawater. They may also have values higher than seawater sulfate,

suggesting the possible importance of BSR in this environment (Ehya, 2012; Griffith et al.,

2018).

Diagenetic barite formed at an oxic/anoxic interface is thought to have anomalously high

δ18OSO4 due to barite dissolution associated with BSR, oxygen isotope exchange reactions

associated enzymatically activated sulfate reactions and subsequent reprecipitation within

the pore fluids (Turchyn and Schrag, 2006; Wortmann et al., 2007). The BSR will increase

the δ18OSO4 by preferentially reducing the SO4 of light oxygen isotopes leaving a heavy
18
residual SO4 pool, although the exact parameters affecting this O fractionation are not
81
known (Knöller et al., 2006). Pore water profile measurements suggest that the fractionation

during BSR is typically between 2‰ and 10‰ (Aharon and Fu, 2000). Oxygen isotope

exchange reactions associated with enzymatically-activated sulfate seem to be associated

with an oxygen isotope equilibrium fractionation factor of about 30‰ also resulting in pore

water sulfate oxygen isotope enrichment (Wortmann et al., 2007). The combination of the

δ34S and δ18O of SO4 can help differentiate between sulfur transformation activities that have

taken place in the environment (Figure 4.7).

Figure 4.7. Plot of the sulfur and oxygen isotopic composition of barite. Juan de Fuca O and S isotope

data are from Goodfellow et al. (1993) and Paytan et al. (2002), respectively. Sea of Okhotsk data are from
Greinert et al. (2002) and Sea of Japan data are from Sakai (1971). Marine barite S isotope data are from Paytan

et al. (1998) and O isotope data are from Turchyn and Schrag (2004). Average marine barite

O isotope data were based on measurements of recent marine barite from five sites (Turchyn and Schrag, 2004).

The reason for the apparent offset between marine barite and seawater δ18OSO4 compels further investigation.

Results of oxygen isotope analyses for some SEDEX barite deposits matched seawater

sulfate, but δ18O was also found to covary with δ34S. This covariation was variously

attributed to BSR within sediment pore waters (Cecile et al., 1983; Johnson et al., 2004) or

euxinic seawater (Cecile et al., 1983), mixing of marine waters that had undergone different
82
degrees of BSR during subseafloor convection (Rao, 1997), and oxidation of isotopically

light sulfide minerals (Cortecci et al., 1989). The δ 18O values of marine pelagic barite exhibit

an offset of 2.0–2.5‰ to lower values compared to that of seawater sulfate, which can be

attributed to the incorporation of sulfate from oxidized organic S compounds (Markovic et

al., 2016). Marine cold seep barites have δ18O values higher than that of seawater sulfate, as

the widest range of δ18O values has been recorded in barite of this species. This extremely

wide range indicates the importance of BSR in host sediments, resulting in higher δ 18O

values in the residual sulfate (Seal et al., 2000). However, cold seep barites from the Gulf of

Mexico exhibits low δ18O variability and are slightly enriched in 18


O compared to bottom

water sulfate (Stevens et al., 2015). The lowest δ 18O value (+4.66‰) was measured in

diagenetic barite from Ocean Drilling Program Site 765, suggesting that the pore fluids from

which the barite precipitated have different chemistry than present-day seawater (Paytan et

al., 2002).

Analytical method
Under a binocular microscope, barite samples were meticulously divided using a micro

drill. Numerous measurements were taken on the samples and on internal standards to

confirm the entire process. In accordance with the instrument's declared specs, the accepted

standard deviation (1σ) for δ18O is ≤ 0.3‰, while that for δ34S is

0.2‰. Isotopic ratios are given in δ notation as per mil (‰) relative to VSMOW for oxygen

and to VCDT for sulfur.

Sulfur and oxygen isotopes and their sources


The δ34S and δ18O values for barite samples (n = 8) from the Qezel-Dagh deposit are

summarized in Table. Sulfur isotope signatures for barite mineralization from the Qezel-

Dagh deposit have general values of δ 34S ranging from +15.41 to +23.14‰. These δ34S

values are significantly higher than the expected value for sulfur of a mantle-derived

83
magmatic source (~0‰) and thus oppose a magmatic origin of the mineralization. These

values correlate well with the typical composition of marine sulfate in Neoproterozoic and

Paleozoic time (12–38‰, Claypool, 1980; Strauss, 2001; Kampschulte and Strauss, 2004).

The isotopic composition of oxygen and sulfur in barite samples can indicate the origin of

sulfate (SO42−) in barite (Claypoolal., 1980; Cecileet, 1983; Clark., 2004; Johnson, 2004).

The δ18O values fall within a range of +4.12 to

+17.88‰. These values are within the δ18O range for Neoproterozoic and Paleozoic marine

sulfate (10–20‰, Claypool, 1980).

Table 4.5. Oxygen and sulfur isotopic compositions for barites from the Qezel-Dagh deposit.
Sample numbers δ 34S (‰) δ18O (‰)

Sher-S2 +16.99 +4.12


Sher-S7 +18.22 +6.41
Sher-S8 +23.14 +7.14
Sher-S11 +21.42 +14.55
Sher-S13 +16.25 +16.18
Sher-S15 +22.10 +17.88
Sher-S16 +19.12 +10.10
Sher-S17 +15.41 +8.23

Seawater is an important sulfur reservoir with its high SO 4–2 concentrations. The sulfur

isotopic composition of seawater has varied in geological time with δ 34S between +10 and

+30‰, and in modern seawater it has a value between +17 and +20‰ (Claypool et al., 1980;

Ohmoto, 1986) (Figure 4.8). During the Paleozoic Era, the δ 34S values of seawater were

between +15 and +35‰ (Faure, 1986; Hanor, 2000), whereas the maximum δ 34S values for

seawater sulfate during the Cambrian–Ordovician period were between +25 and +35‰. The

δ34S values of barite samples are high, strongly positive values and show a wide distribution

of δ34S values. The isotope data in Figures 4.8 indicate that the mineralization has formed

from seawater sulfate, which had been isotopically modified. As reported by previous

studies (e.g., Goodfellow and Jonasson, 1984; Large et al., 2005), the distribution of δ 34S

84
values was related to two different sources of sulfur (bacterial and thermochemical sulfide

reduction) and/or the formation under closed system conditions

(and related Rayleigh Type effects) during the evolution of the deposit. Positive and highly

homogeneous δ34S values are not consistent with Rayleigh Type fractionation in a closed

system.

Figure 4.8. Distribution of δ34S values for barite from the Qezel-Dagh deposit, which plotted on a

chronologic diagram of the average δ34S content of marine, evaporates (Horita et al., 2002) from a

85
modified version of Claypool et al. (1980). Horizontal and vertical lines through the markers are the ranges of

values and the average value, respectively.

Plots of δ34S versus δ18O values of barite have been used to discriminate between barite

of different origins (Johnson et al., 2004, 2009; Griffith et al., 2018). As shown in Figure

4.9, diagenetic and marine cold seep barites exhibit a wide range of δ34S and δ18O values,

ranging from seawater-like values to values several times higher than that of modern

seawater sulfate. Therefore, diagenetic and marine cold seep barite are poor analogous to

barite from the Qezel-Dagh deposit. Hydrothermal barites have δ34S and δ18O values similar

to or slightly lower than the present-day seawater, and the Qezel-Dagh barites also have δ34S

values similar to contemporaneous seawater. The Qezel-Dagh deposit differs from marine

(pelagic) barites when their δ34S values are compared, because the δ34S values in the later

are close to that of coeval seawater, but they have a narrow range in δ 34S values, while in

the Qezel-Dagh deposit in spite of δ34S values having close to that of coeval seawater they

have a wide range in δ34S values. Therefore, marine hydrothermal barite can be considered

as a strong analog to the Qezel-Dagh barite deposit.

86
Figure 4.9. Plot of δ34S versus δ18O values of barite from the Qezel-Dagh deposit compared to

isotopic composition of present seawater, and to barites from main modern settings in which

barites are formed (data from Griffith and Paytan, 2012).

According to Turchyn et al. (2006), the reversal in δ 18O at high δ34S is useful as an

indicator of anaerobic oxidation of methane in the formation of stratiform barites. The data

are plotted on a δ18O versus δ34S diagram that span a broad range and define a linear trend

even at high δ34S values. However linear or concave-upward trends are common in

laminated, massive, and breccia barite (Johnson et al., 2009), the number of analyses is too

few to either verify trend or rule it out. The Qezel-Dagh barite deposit is compared with

other barite deposits that formed in different age and tectonic setting (Figure 4.10).

Figure 4.10. Comparison of δ34S values for barite from the Qezel-Dagh deposit and other barite

deposits.

87
CHAPTER 5: Conclusions

and suggestions

88
5.1 Conclusions
The Sanandaj-Sirjan metamorphic zone's northwest region contains the Qezel-Dagh barite

deposit, which is 20 kilometers northeast of Mahabad City. The main lithologic unit in the

western part of the Qezel-Dagh barite deposit is the Mahabad rhyolite and in the eastern part

of the Qezel-Dagh barite deposit is thin-bedded tuffs. Rhyolite of the Mahabad Formation

hosts barite and iron (±copper) mineralization. Barite mineralization occurs as veins and

open-space fillings in the western part of the Qezel-Dagh deposit and as lens and layer along

with iron (±copper) mineralization in the eastern part of the Qezel-Dagh deposit. Rhyolite

of the Mahabad Formation consists of a mineral assemblage of quartz, K-feldspar,

plagioclase, and mafic minerals, such as biotite and hornblende. As evidenced by high

(La/Yb)N ratios ranging from 1.61 to 24.26 and pronounced positive Ce anomalies ranging

from 1.61 to 3.55, chondritenormalized REE patterns for barite from the Qezel-Dagh deposit

show an enrichment of light rare earth elements (LREE) relative to heavy rare earth elements

(HREE). The δ34S values in the analyzed barites from the Qezel-Dagh deposit range from

+15.41 to +23.14‰, consistent with the typical composition of marine sulfate in

Neoproterozoic and Paleozoic time (12–38‰). The δ18O values of barite samples vary from

+4.12‰ to +17.88‰. Supporting features for a submarine hydrothermal origin of barite in

Qezel-Dagh deposit are the following: (1) open-space filling textures of barite reveal the

deposition of this mineral by hydrothermal activities; (2) high Sr concentrations (1110.2–

2168.3 ppm), low ΣREE concentrations (5.95–54.09 ppm), LREE enriched chondrite-

normalized REE patterns of barite, and pronounced positive Eu anomalies (2.56–8.92) are

commonly considered as particular features of marine precipitates (including barite)

deposited from hydrothermally influenced seawater; and (3) sulfur isotope data are also

consistent with isotope values of contemporaneous seawater. Based on these features, it is

suggested that there were active submarine hydrothermal vents in Sanandaj–Sirjan zone

during late Neoproterozoic and Paleozoic. The hydrothermal fluids escaping from the vents

carried sufficient barium to precipitate barite locally. As in the case for barite deposits

associated with volcanic rocks, hydrothermal fluids acquired barium during the circulation

89
of seawater in underlying volcanic source rocks. Barite deposition occurred on the seafloor

where ascending hydrothermal barium-bearing fluids encountered sulfate bearing marine

waters in a manner similar to that found in modern analogs on the ocean floor.

Suggestions
1- Sulfur isotopic analysis of sulfide minerals (e.g., pyrite and chalcopyrite) associated

with barite mineralization for determining the source of sulfur;

2- Microthermometry of fluid inclusions with the aim of determining salinity and

temperature of ore-forming fluid and physicochemical nature of the ore-forming fluid; 3-


87
Sr/86Sr isotopic analysis of barite samples with the aim of determining the source of barite

and also providing an indicator of the depositional environment of barite.

90
References

Aghanabati, A., 2004. Geology of Iran: Geological Survey of Iran, 600 pp.

Aharon, P., Fu, B., 2000. Microbial sulfate reduction rates and sulfur and oxygen isotope

fractionations at oil and gas seeps in deep water Gulf of Mexico. Geochimica et

Cosmochimica Acta, 64, 233–246.

Alavi, M., 1991. Sedimentary and structural characteristics of the Paleo-Tethys remnants in

Northeastern Iran. The Geological Society of America Bulletin 103, 983–992.

Alavi, M., 1994. Tectonics of the Zagros Orogenic belt of Iran; new data and interpretations.

Tectonophysics 299, 211–238.

Alexander, B.W., Bau, M., Andersson, P., Dulski, P., 2008. Continentally-derived solutes in

shallow Archean seawater: rare earth elements and Nd isotope evidence in iron formation

from the 2.9 Ga Pongola supergroup, South Africa. Geochimica et Cosmochimica Acta,

72, 378–394.

Aliyari, F., Rastad, E., Zengqian, H., 2007. Orogenic gold mineralization in the Qolqoleh

Deposit, Northwestern Iran. Resource Geology, 57(3), 269–282.

Almasi, A., Jafarirad, A., Kheyrollahi, H., Rahimi, M., Afzal, P., 2014. Evaluation of

structural and geological factors in orogenic gold type mineralisation in the Kervian area,

north-west Iran, using airborne geophysical data. Exploration Geophysics, 45(4), 261–

270.

Aloisi, G., Wallmann, K., Bollwerk, S.M., Derkachev, A., Bohrmann, G., Suess, E., 2004.

The effect of dissolved barium on biogeochemical processes at cold seep. Geochimica et

Cosmochimica Acta, 68(8), 1735–1748.

Aquilina, L., Dia, A.N., Boulègue, J., Bourgois, J., Fouillac, A.M., 1997. Massive barite

deposits in the convergent margin off Peru: Implications for fluid circulation within

subduction zones. Geochimica et Cosmochimica Acta, 61(6), 1233–1245.

91
Asghari, G., Alipour, S., Azizi, H., Mirnejad, H., 2018. Geology and mineral chemistry of

gold mineralization in Mirge-Naqshineh occurrence (Saqez, NW Iran): implications for

transportation and precipitation of gold. Acta Geologica Sinica-English, 92, 210–224.

Badrzadeh, Z., Barrett, T.J., Peter, J.M., Gimeno, D., Sabzehei, M., Aghazadeh, M., 2011.

Geology, mineralogy, and sulfur isotope geochemistry of the Sargaz Cu–Zn volcanogenic

massive sulfide deposit, Sanandaj-Sirjan Zone. Iran. Mineralium Deposita, 46(8), 905–

923.

Baioumy, H.M., 2015. Rare earth elements, S and Sr isotopes and origin of barite from

Bahariya Oasis, Egypt: Implication for the origin of host iron ores. Journal of African

Earth Sciences, 106, 99–107.

Barry, J.P., Greene, H.G., Orange, D.L., Baxter, C.H., Robinson, B.H., Kochevar, R.E.,

Nybakken, J.W., Reed, D.L., McHugh, C.M., 1996. Biologic and geologic characteristics

of cold seep in Monterey Bay, California. Deep Sea Research Part I: Oceanographic

Research Papers 43(11–12), 1739–1762.

Behyari, M., Rahimsouri, Y., Hoseinzadeh, E., Kurd, N., 2019. Evaluating of lithological

and structural controls on the barite mineralization by using the remote sensing, Fry and

fractal methods, Northwest Iran. Arabian Journal of Geosciences, 12(5), 1–11.

Berberian, M., King, G.C.P., 1981. Towards a paleogeography and tectonic evolution of

Iran. Canadian journal of earth sciences, 18(2), 210–265.

Berner, R.A., 1999. Atmospheric oxygen over Phanerozoic time. Proceedings of the

National Academy of Sciences, 96(20), 10955–10957.

Berner, R.A., Canfield, D.E., 1989. A new model for atmospheric oxygen over Phanerozoic

time. American Journal of Science, 289(4), 333–361.

Bernstein, R.E., Byrne, R.H., Betzer, P.R., Greco, A.M., 1992. Morphologies and

transformations of celestite in seawater: The role of acantharians in strontium and barium

geochemistry. Geochimica et Cosmochimica Acta, 56(8), 3273–3279.

92
Bernstein, R.E., Byrne, R.H., Schijf, J., 1998. Acantharians: a missing link in the oceanic

biogeochemistry of barium. Deep Sea Research Part I: Oceanographic Research Papers,

45, 491–505.

Bertram, M.A., Cowen, J.P., 1997. Morphological and compositional evidence for biotic

precipitation of marine barite. Journal of Marine Research, 55, 577–593.

Bishop, J.K.B., 1988. The barite-opal-organic carbon association in oceanic particulate

matter. Nature, 332(6162), 341–343.

Bolze, C.E., Malone, P.G., Smith, M.J., 1974. Microbial mobilization of barite. Chemical

Geology, 13(2), 141–143.

Bonel, K.A., 2005. Barytes. British Geological Survey, Natural Environment Research

Council. Mineral Profile Report, pp. 28.

Boni, M., Gilg, H.A., Balassone, G., Schneider, J., Allen, C.R., Moore, F., 2007. Hypogene

Zn carbonate ores in the Angouran deposit, NW Iran. Mineralium Deposita, 42(8), 799–

820.

Boveiri, M., Rastad, E., 2016. Nature and origin of dolomitization associated with sulphide

mineralization: new insights from the Tappehsorkh Zn-Pb (-Ag-Ba) deposit, Irankuh

Mining District, Iran. Geological Journal, 53(1), 1–21.

Bowes, D.R., 1989. The Encylopedia of Igneous and Metamorphic Petrology. New York:

Van Nostrand Reinhold, pp. 666.

Bréhéret, J.G., Brumsack, H.J., 2000. Barite concretions as evidence of pauses in

sedimentation in the Marnes Bleues Formation of the Vocontian Basin (SE France).

Sedimentary Geology, 130, 205–228.

Brobst, D.A., 1958. Barite resources of the United States. US Government Printing

Office.1072-B, 130–677.

Brobst, D.A., 1970. The geological framework of barite resources. Transactions of the

Institution of Mining and Metallurgy. Section A: Mining Industry 93, A123–A130.

93
Brumsack, H.J., Gieskes, J.M., 1983. Interstitial Water Trace-Metal Chemistry of Laminated

Sediments from the Gulf of California, Mexico. Marine Chemistry, 14, 89– 106.

Canet, C., Prol-Ledesma, R.M., Torres-Alvarado, I., Gilg, H.A., Villanueva, R.E.,

LozanoSanta Cruz, R., 2005. Silica-carbonate stromatolites related to coastal

hydrothermal venting in Bahia Concepcion, Baja California Sur, Mexico. Sedimentary

Geology, 174, 97–113.

Cecile, M.P., Shakur, M.A., Krouse, H.R., 1983. The isotopic composition of western

Canadian barites and the possible derivation of oceanic sulphate δ34S and δ18O age

curves. Canadian Journal of Earth Sciences, 20, 1528–1535.

Chen, D., Qing, H., Yan, X., Li, H., 2006. Hydrothermal venting and basin evolution

(Devonian, South China): constraints from rare earth element geochemistry of chert.

Sedimentary Geology, 183, 203–216.

Chiba, H., Sakai, H., 1985. Oxygen isotope exchange-rate between dissolved sulfate and

water at hydrothermal temperatures. Geochimica et Cosmochimica Acta, 49, 993–1000.

Chow, T.J., Goldberg, E.D., 1960. On the marine geo-chemistry of barium. Geochimica et

Cosmochimica Acta, 20, 192–198.

Church, T.M., Wolgemuth, K., 1972. Marine barite saturation. Earth and Planetary Science

Letters, 15(1), 35–44.

Clark, S.H., Poole, F.G., Wang, Z., 2004. Comparison of some sediment-hosted, stratiform

barite deposits in China, the United States, and India. Ore Geology Reviews, 24, 85– 101.

Clark, S.H.B., 1989. Barite - a model for deposition from stratified seawater based on barite

nodules in Paleozoic shale and mudstone of the Appalachian Basin US Geological Survey

Bulletin 1877, A1–A6.

Clark, S.H.B., Gallagher, M.J., Poole, F.G., 1990. World barite resources: a review of recent

production patterns and a genetic classification. In Congress of the council of mining and

metallurgical institutions 99, B125–B132.

94
Clark, S.H.B., Orris, G.J., 1991. Descriptive model of exhalative barite. In: Orris, G.J., Bliss,

J.D. (Eds.), Some Industrial Mineral Deposit Models: Descriptive Deposit Models. U.S.

Geological Survey Open-File Report 91-11A, 21–22.

Clark, S.H.B., Poole, F.G., Wang, Z., 2004. Comparison of some sediment hosted, stratiform

barite deposits in China, the United States, and India. Ore Geology Reviews, 24, 85–101.

Claypool, G.E., Holser, W.T., Kaplan, I.R., Sakai, H., Zak, I., 1980. The age curves of sulfur

and oxygen isotopes in marine sulfate and their mutual interpretation. Chemical geology,

28, 199–260.

Cortecci, G., Fontes, J.C., Maiorani, A., Perna, G., Pintus, E., Turi, B., 1989. Oxygen, sulfur,

and strontium isotope and fluid inclusion studies of barite deposits from the Iglesiente-

Sulcis mining district, southwestern Sardinia, Italy. Mineralium Deposita, 24, 34–42.

Crockford, P.W., Wing, B.A., Paytan, A., Hodgskiss, M.S.W., Mayfield, K.K., Hayles,

J.A., Middleton, J.E., Ahm, A.S.C., Johnston, D.T., Caxito, F., Uhlein, G., Halverson,

G.P., Eickmann, B., Torres, M., Horner, T.J., 2019. Barium-isotopic constraints on the

origin of post-Marinoan barites. Earth and Planetary Science Letters, 519, 234–244.

Daliran, F., Pride, K., Walther, J., Berner, Z.A., Bakker, R.J., 2013. The Angouran Zn (Pb)

deposit, NW Iran: evidence for a two stage, hypogene zinc sulfide–zinc carbonate

mineralization. Ore Geology Reviews, 53, 373–402.

de Brodtkorb, M.K., 1989. Nonmetalliferous Stratabound Ore Fields. Van Nostrand

Reinhold, New York.

Dean, W.E., Schreiber, B.C., 1977. Authigenic barite. In: Proceedings of the Deep Sea

Drilling Project, Initial Reports, (Eds J. Gardner and J. Herring), 41, 915–931.

Dehairs, F., Chesselet, R., Jedwab, J., 1980. Discrete suspended particles of barite and the

barium cycle in the open ocean. Earth and Planetary Science Letters, 49(2), 528–550.

Dia, A.N., Aquilina, L., Boulègue, J., Bourgois, J., Suess, E., Torres, M., 1993. Origin of

fluids and related barite deposits at vent sites along the Peru convergent margin.

Geology, 21, 1099–1102.

95
Dickens, G.R., 2001. Sulfate profiles and barium fronts in sediment on the Blake Ridge:

present and past methane fluxes through a large gas hydrate reservoir. Geochimica et

cosmochimica Acta, 65, 529–543.

Dube, T.E., 1988. Tectonic significance of Upper Devonian igneous rocks and bedded

barite, Roberts Mountains Allochthon, Nevada, USA. Canadian Society of Petroleum

Geologists, 14, 235–249.

Dymond, J., Suess, E., Lyle, M., 1992. Barium in deep-sea sediment: A geochemical proxy

for paleoproductivity. Paleoceanography, 7, 163–181.

Ehya, F., 2012. Rare earth element and stable isotope (O, S) geochemistry of barite from the

Bijgan deposit, Markazi Province, Iran. Mineralogy and Petrology, 104(1), 81–93.

Ehya, F., Lotfi, M., Rasa, I., 2010. Emarat carbonate-hosted Zn–Pb deposit, Markazi

Province, Iran: A geological, mineralogical and isotopic (S, Pb) study. Journal of Asian

Earth Sciences, 37(2), 186–194.

Fathi Sadi, M., 2012. Geochemistry and petrogenesis of rhyolitic rocks from East

Azarbaijan-NW Iran. MSC Thesis. University of Tabriz, Iran.

Faure, G., 1986. Principles of Isotope Geology, second ed. John Wiley and Sons Inc., New

York.

Faure, G., 1998. Principles and applications of geochemistry. Prentice Hall, New Jersey.

Fisher, N.S., Guillard, R.R.L., Bankston, D.C., 1991. The accumulation of barium by marine

phytoplankton grown in culture. Journal of marine research, 49, 339–354.

Fu, B., Aharon, P., Byerly, G.R., Roberts, H.H., 1994. Barite chimneys on the Gulf of

Mexico slope: Initial report on their petrography and geochemistry. Geo-Marine Letters,

14(2–3), 81–87.

Ganeshram, R.S., François, R., Commeau, J., Brown-Leger, S.L., 2003. An experimental

investigation of barite formation in seawater. Geochimica et Cosmochimica Acta, 67,

2599–2605.

Ganji, A., 2015. Barite mineralization in Iran. VI. BALKANMINE, Petrosani 2015, 7.

96
Ghasemi, A., Talbot, C.J., 2006. A new tectonic scenario for the Sanandaj–Sirjan Zone

(Iran). Journal of Asian Earth Sciences, 26(6), 683–693.

Ghazban, F., McNutt, R.H., Schwarcz, H.P., 1994. Genesis of sediment-hosted Zn-Pb-Ba

deposits in the Iran Kouh district, Esfahan area, west-Central Iran. Economic Geology,

89(6), 1262–1278.

Ghorbani, M., 2002. An introduction to economic geology of Iran. Tehran: Geological

Survey and Mineral Explorations of Iran (GSI).

Ghorbani, M., 2013. The economic Geology of Iran, mineral deposits and natural resources.

Springer Geology, Dordrecht.

Gilg, H.A., Boni, M., Balassone, G., Allen, C.R., Banks, D., Moore, F., 2006. Marble hosted

sulphide ores in the Angouran Zn-(Pb-Ag) deposit, NW Iran: interaction of sedimentary

brines with a metamorphic core complex. Mineralium Deposita, 41(1), 1–

16.

Glasby, G.P., Cherkashov, G.A., Gavrilenko, G.M., Rashidov, V.A., Slovtsov, I.B., 2006.

Submarine hydrothermal activity and mineralization on the Kurile and western Aleutian

island arcs, NW Pacific. Marine geology, 231(1-4), 163–180.

Goldberg, E.D., Arrhenius, G., 1958. Chemistry of pelagic sediments. Geochimica et

Cosmochimica Acta, 13, 153–212.

Goldberg, E.D., Koide, M., Schmitter, R.A., Smith, R.H., 1963. Rare-earth distributions in

the marine environment. Journal of Geophysical Research, 68, 4209–4217.

Gontharet, S., Pierre, C., Blanc-Valleron, M.M., Rouchy, J.M., Fouquet, Y., Bayon, G.,

Foucher, J.P., Woodside, J., Mascle, J., Party, T.N.S., 2007. Nature and origin of

diagenetic carbonate crusts and concretions from mud volcanoes and pockmarks of the

Nile deep-sea fan (eastern Mediterranean Sea). Deep Sea Research Part II: Topical

Studies in Oceanography, 54, 1292–1311.

Goodfellow, W.D., Grapes, K., Cameron, B., Franklin, J.M., 1993. Hydrothermal alteration

associated with massive sulfide deposits, Middle Valley, Northern Juan De Fuca Ridge.

Canadian Mineralogist, 31, 1025–1060.


97
Goodfellow, W.D., Jonasson, I.R., 1984. Ocean stagnation and ventilation defined by δ 34S

secular trends in pyrite and barite, Selwyn Basin, Yukon. Geology, 12, 583–586.

Graf, J.L., 1977. Rare earth elements as hydrothermal tracers during the formation of

massive sulfide deposits in volcanic rocks. Economic geology, 72(4), 527–548.

Greinert, J., Bollwerk, S.M., Derkachev, A., Bohrmann, G., Suess, E., 2002. Massive barite

deposits and carbonate mineralization in the Derugin Basin, Sea of Okhotsk: precipitation

processes at cold seep sites. Earth and Planetary Science Letters, 203(1), 165–180.

Griffith, E.M., Paytan, A., 2012. Barite in the ocean-occurrence, geochemistry and

palaeoceanographic applications. Sedimentology 59, 1817–1835.

Griffith, E.M., Paytan, A., Wortmann, U.G., Eisenhauer, A., Scher, H.D., 2018. Combining

metal and nonmetal isotopic measurements in barite to identify mode of formation.

Chemical Geology, 500, 148–158.

Guichard, F., Church, T.M., Treuil, M., Jaffrezic, H., 1979. Rare earths in barites:

distribution and effects on aqueous partitioning. Geochimica et Cosmochimica Acta,

43(7), 983–997.

Hanor, J.S., 2000. Barite-celestite geochemistry and environments of formation. Reviews in

Mineralogy and Geochemistry 40, 193–263.

Harben P.W., 1999. The industrial minerals handy book, 3 rd edition. Industrial Minerals

Information Ltd, London.

Hashemi, F., Mousivand, F., Rezaei-kahkhaei, M., 2014. Varandan Deposit: Kuroko-type

Ba-Pb-Cu volcanogenic massive sulfide mineralization in the Urumieh-Dokhtar

magmatic arc. In: 32nd meeting of the first Congress of the International Earth Science,

Geological Survey of Iran, Tehran, Iran, (in Persian with English abstract).

Haskin, L.A., Freyf, F.A., Schmitter, A., Smithr, H., 1968. Meteoritic, solar and terrestrial

rare earth distributions. In: Physics and Chemistry of the Earth. Pergamon Press.

Hastie, A.R., Kerr, A.C., Pearce, J.A., Mitchell, S.F., 2007. Classification of altered volcanic

island arc rocks using immobile trace elements: development of the Th–Co

discrimination diagram. Journal of petrology, 48(12), 2341–2357.


98
Haymon, R.M., Kastner, M., 1981. Hot-spring deposits on the East Pacific Rise at 21ºN -

Preliminary description of mineralogy and genesis. Earth and Planetary Science Letters,

53, 363–381.

Hein, J.R., Zierenberg, R.A., Maynard, J.B., Hannington, M.D., 2007. Barite-forming

environments along a rifted continental margin, Southern California Borderland. Deep

Sea Research Part II: Topical Studies in Oceanography, 54(11–13), 1327–1349.

Hoefs, J., Sywall, M., 1997. Lithium isotope composition of quaternary and tertiary biogene

carbonates and a global lithium isotope balance. Geochimica et Cosmochimica Acta, 61,

2679–2690.

Hogdahl, T., Melsom, S., Bowenv, T., 1968. Neutron activation analysis of lanthanide

elements in sea water. In: Advances in Chemistry, Series No. 73, American Chemical

Society, pp. 308.

Holland, H.D., 1973. Systematics of the isotopic composition of sulfur in the oceans during

the Phanerozoic and its implications for atmospheric oxygen. Geochimica et

Cosmochimica Acta, 37, 2605–2616.

Holser, W.T., Kaplan, I.R., 1966. Isotope geochemistry of sedimentary sulfates. Chemical

Geology, 1, 93–135.

Horita, J., Zimmermann, H., Holland, H.D., 2002. Chemical evolution of seawater during

the Phanerozoic; implications from the record of marine evaporites. Geochimica et

Cosmochimica Acta, 66(21), 3733–3756.

Huheey, J.E., Keiter, E.A., Keiter, R.L., 1993. Inorganic chemistry: principles of structure

and reactivity. Harper-Collins, New York, pp. 400.

Jewell, P.W., 1994. Paleoredox conditions and the origin of bedded barites along the late

Devonian North American continental margin. The Journal of Geology, 102(2), 151–

164.

Jewell, P.W., 2000. Bedded barite in the geologic record, in Glenn, C.R., Prévôt-Lucas,

Liliane, and Lucas, Jacques, Marine authigenesis—from global to microbial: Society of

Economic Paleontologists and Mineralogists Special Publication 66, 147–161.


99
Jewell, P.W., Stallard, R.F., 1991. Geochemistry and paleoceanographic setting of central

Nevada bedded barites. Journal of Geology, 99(2), 151–170.

Jiang, S.Y., Chen, Y.Q., Ling, H.F., Yang, J.H., Feng, H.Z., Ni, P., 2006. Trace-and rareearth

element geochemistry and Pb–Pb dating of black shales and intercalated Ni–Mo– PGE–

Au sulfide ores in Lower Cambrian strata, Yangtze Platform, South China. Mineralium

Deposita, 41(5), 453-467.

Johnson, C.A., Emsbo, P., Poole, F.G., Rye, R.O., 2009. Sulfur- and oxygen-isotopes in

sediment-hosted stratiform barite deposits. Geochimica et Cosmochimica Acta, 73, 133–

147.

Johnson, C.A., Kelley, K.D., Leach, D.L., 2004. Sulfur and oxygen isotopes in barite

deposits of the western Brooks Range, Alaska, and implications for the origin of the Red

Dog massive sulfide deposits. Economic Geology, 99(7), 1435–1448.

Jorgensen, B.B., Kasten, S., 2006. Sulfur cycling and methane oxidation. Marine

geochemistry. Springer, Berlin, Heidelberg, 271–309.

Jurković, I., Garašić, V., Hrvatović, H., 2010. Geochemical characteristics of the barite

occurrences in the Paleozoic complex of the Southeastern Bosnia and their relationship

to the barite deposits of the Mid-Bosnian Schist Mountains. Geologia croatica, 63(2),

241–258.

Kabata-Pendias, A., 2011. Trace Elements in Soils and Plants. CRC Press, Taylor and

Francis Group.

Kamilli R.J., 1998. Paragenesis. In: Geochemistry. Encyclopedia of Earth Science. Springer,

Dordrecht.

Kampschulte, A., Strauss, H., 2004. The sulfur isotopic evolution of Phanerozoic seawater

based on the analysis of structurally substituted sulfate in carbonates. Chemical Geology,

204(3-4), 255–286.

Karunakaran, C., 1973. Barytes, Indo-Soviet Symposium on recent trends in Exploration of


Minerals, Oil and Groundwater India. Nat. Sci. Acad., New Delhi.

100
Kato, Y., 1999. Rare earth elements as an indicator to origins of skarn deposits: examples of

the Kamioka Zn-Pb and Yoshiwara-Sannotake Cu (–Fe) deposits in Japan. Resource

Geology, 49(4), 183–198.

Kato, Y., Nakamura, K., 2003. Origin and global tectonic significance of Early Archean

cherts from the Marble Bar greenstone belt, Pilbara Craton, Western Australia.

Precambrian Research, 125(3–4), 191–243.

Kharaka, Y.K., Meast, A.S., Carothers, W.W., Law, L.M., Lamothe, P.J., 1987.

Geochemistry of metal-rich brines from central Mississippi Salt Dome Basin, USA.

Applied Geochemistry, 2, 543–561.

Knöller, K., Vogt, C., Richnow, H.H., Weise, S.M., 2006. Sulfur and oxygen isotope

fractionation during benzene, toluene, ethyl benzene, and xylene degradation by

sulfatereducing bacteria. Environmental science and technology, 40, 3879–3885.

Kontak, D.J., Kyser, K., Gize, A., Marshall, D., 2006. Structurally controlled vein barite

mineralization in the Maritimes Basin of eastern Canada: Geologic setting, stable

isotopes, and fluid inclusions. Geology, 101(2), 407-430.

Koski, R.A., Hein, J.R., 2004. Stratiform barite deposits in the Roberts Mountains

allochthon, Nevada: A review of potential analogs in modern sea-floor environments.

Koski, R.A., Lonsdale, P.F., Shanks, W.C., Berndt, M.E., Howe, S.S., 1985. Mineralogy and

geochemistry of a sediment-hosted hydrothermal sulfide deposit from the Southern

Trough of Guaymas Basin, Gulf of California. Journal of Geophysical Research 90(B8):

6695–6707.

Koski, R.A., Shanks III, W.C., Bohrson, W.A., Oscarson, R.L., 1988. The composition of

massive sulfide deposits from the sediment-covered floor of Escanaba Trough, Gorda

Ridge: Implications for depositional processes. The Canadian Mineralogist 26, 655– 673.

Krebs, R.E., 2006. The history and use of our earth’s chemical elements: a reference Guide.

Greenwood Publishing Group, London, pp. 422.

101
Large, R.R., Bull, S.W., McGoldrick, P.J., Walters, S.G., 2005. Stratiform and stratabound

Zn-Pb-Ag deposits in Proterozoic sedimentary basins, northern Australia. Economic

Geology, 100, 931–963.

Lloyd, R.M., 1968. Oxygen isotope behavior in the sulfate‐water system. Journal of

Geophysical Research, 73, 6099–6110.

Lotfi, M., Baharvandi, A., Ghaderi, M., Jafari, M.R., Tajeddin, H.A., 2017. Ore

mineralization and fluid inclusion and sulfur isotope studies on the Shakarbeig deposit,

Southwest Mahabad, Sanandaj-Sirjan zone. Geosciences 26, 103.

Lüders, V., Pracejus, B., Halbach, P., 2001. Fluid inclusion and sulfur isotope studies in

probable modern analogue Kuroko-type ores from the JADE hydrothermal field (Central

Okinawa Trough, Japan). Chemical Geology, 173(1–3), 45–58.

Lydon, J.W., Goodfellow, W.D., Jonasson, I.R., 1985. A general model for stratiform baritic

deposits of the Selwyn basin, Yukon Territory and District of Mackenzie. Geological

Survey of Canada Paper 85–1A, 651–660.

Lyons, T.W., Gellatly, A.M., McGoldrick, P.J., Kah, L.C., 2006. Proterozoic sedimentary

exhalative (SEDEX) deposits and links to evolving global ocean chemistry.

MemoirsGeological Society of America, 198, 169–184.

Macpherson, G.L., 1989. Lithium, boron and barium in formation waters and sediments,

northwestern Gulf of Mexico sedimentary basin. Ph.D., University of Texas, Austin, TX,

286 pp.

Maghfouri, S., 2017. Geology, geochemistry, ore controlling parameters and genesis of early

Cretaceous carbonate-clastic hosted Zn-Pb deposits in Southern Yazd basin, with

emphasis on Mehdiabad deposit. (Ph.D) Thesis, University of Tabriz, Iran.

Mahmoodi, P., Rastad, E., Rajabi, A., Peter, J.M., 2018. Ore facies, mineral chemical and

fluid inclusion characteristics of the Hossein-Abad and Western Haft-Savaran

sediment-hosted Zn-Pb deposits, Arak Mining District. Iran. Ore Geology Reviews, 95,

342–365.

102
Manheim, F.T., Bischoff, J.L., 1969. Geochemistry of pore waters from Shell Oil Company

drill holes on the continental slope of the northern Gulf of Mexico. Chemical Geology,

4, 63–82.

Marchev, P., Downes, H., Thirlwall, M.F., Moritz, R., 2002. Small-scale variations of
87
Sr/86Sr isotope composition of barite in the Madjarovo low-sulphidation epithermal

system, SE Bulgaria: implications for sources of Sr, fluid fluxes and pathways of ore

forming fluids. Mineralium Deposita, 37(6), 669–677.

Markovic, S., Paytan, A., Li, H., Wortmann, U.G., 2016. A revised seawater sulfate oxygen

isotope record for the last 4 Myr. Geochimica et Cosmochimica Acta, 175, 239–251.

Maynard, J.B., Okita, P.M., 1991. Bedded barite deposits in the United States, Canada,

Germany, and China; two major types based on tectonic setting. Economic Geology,

86(2), 364–376.

McDonough, W.F., Sun, S.S., 1995. The composition of the Earth. Chemical geology,

120(3–4), 223–253.

McManus, J., Berelson, W.M., Klinkhammer, G.P., Johnson, K.S., Coale, K.H., Anderson,

R.F., Kumar, N., Burdige, D.J., Hammond, D.E., Brumsack, H.J., McCorkle, D.C., 1998.

Geochemistry of barium in marine sediments: Implications for its use as a paleoproxy.

Geochimica et Cosmochimica Acta, 62, 3453–3473.

Mirmohammadi, M., Hajsadeghi, S., Asghari, O., Ahmad Meshkani, S., 2017. Geology and

mineralization at the copper-rich volcanogenic massive sulfide deposit in Nohkouhi,

Posht-e-Badam block, Central Iran. Ore Geology Reviews, 92, 379–396.

Mohajjel, M., Fergusson, C.L., 2000. Dextral transpression in Late Cretaceous continental

collision, Sanandaj–Sirjan zone, western Iran. Journal of Structural geology, 22(8), 1125–

1139.

Mohajjel, M., Fergusson, C.L., Sahandi, M.R., 2003. Cretaceous-Tertiary convergence and

continental collision, Sanandaj-Sirjan zone, western Iran. Journal of Asian Earth

Sciences, 21(4), 397–412.

103
Momenzadeh, M., 1976. Stratabound lead–zinc ores in the Lower Cretaceous and Jurassic

sediments in the Malayer-Esfahan district (west central Iran), lithology, metal content,

zonation and genesis. University of Heidelberg, Heidelberg, pp. 300.

Monnin, C., Cividini, D., 2006. The saturation state of the world’s ocean with respect to (Ba,

Sr)SO4 solid solution. Geochimica et Cosmochimica Acta, 70(13), 3290–3298.

Monnin, C., Jeandel, C., Cattaldo, T., Dehairs, F., 1999. The marine barite saturation state

of the world's oceans. Marine Chemistry, 65(3–4), 253–261.

Mousivand, F., Rastad, E., Hoshino, K., Watanabe, M., 2007. The Bavanat Cu-Zn-Ag

orebody: first recognition of a Besshi-type VMS deposit in Iran. Neues Jahrbuch für

Mineralogie-Abhandlungen. Neues Jahrbuch fur Mineralogie-Abhandlungen, 183(3),

297–315.

Mousivand, F., Rastad, E., Meffre, S., Peter, J.M., Solomon, M., Zaw, K., 2011. U–Pb

geochronology and Pb isotope characteristics of the Chahgaz volcanogenic massive

sulphide deposit, southern Iran. International Geology Review, 53(10), 1239–1262.

Murchey, B.L., Madrid, R.J., Poole, F.G., 1987. Paleozoic bedded barite associated with

chert in western North America. In: Hein JR (ed) Siliceous sedimentary rock-hosted ores

and petroleum. Van Nostrand Reinhold, New York.

Murray, R.W., Buchholtz ten Brink, M.R., Jones, D.L., Gerlach, D.C., Russ III, G.P., 1990.

Rare earth elements as indicators of different marine depositional environments in chert

and shale. Geology, 18(3), 268–271.

Naehr, T.H., Stakes, D.S., Moore, W.S., 2000. Mass wasting, ephemeral fluid flow, and

barite deposition on the California continental margin. Geology, 28, 315–318.

Natalin, B.A., Şengör, A.C., 2005. Late Palaeozoic to Triassic evolution of the Turan and

Scythian platforms: the pre-history of the Palaeo-Tethyan closure. Tectonophysics, 404,

175–202.

Nazemi, M., 2005. Investigate the economic geology of Barite ore in the Misho Mountains

and determine its genesis. MSC thesis. University of Tabriz, Iran.

104
Nezafati, N., 2015. Mineral resources of Iran (an overview), Internationals
AlumniSymposium.
Nicholson, K.N., Black, P.M., Hoskin, P.W.O., Smith, I.E.M., 2004. Silicic volcanism and

back-arc extension related to migration of the Late Cainozoic Australian–Pacific plate

boundary. Journal of Volcanology and Geothermal Research, 131, 295–306.

Niroomand, S., Goldfarb, R.J., Moore, F., Mohajjel, M., Marsh, E.E., 2011. The Kharapeh

orogenic gold deposit: geological, structural, and geochemical controls on epizonal

ore formation in West Azerbaijan Province, Northwestern Iran. Mineralium Deposita,

46(4), 409–428.

Noguchi, T., Shinjo, R., Ito, M., Takada, J., Oomori, T., 2011. Barite geochemistry from

hydrothermal chimneys of the Okinawa Trough: insight into chimney formation and

fluid/sediment interaction. Journal of Mineralogical and Petrological Sciences, 106, 26–

35.

Noori Asl, R., 2017. Petrological studies of Mahabad rhyolite. MSc thesis. University of

Tabriz (In Persian).

Nosratpoor, H., 2008. A study of gold mineralization in Ghabaghlojeh shear zone (southwest

Saqqez, Kurdestan province) MS thesis Tehran University.

Ohmoto, H., 1986. Stable isotope geochemistry of ore deposits, in: Valley et al. (Eds.),

Stable Isotopes in High Temperature Geological Processes. Reviews in Mineralogy, 16,

491–559.

Ohmoto, H., Lasaga, A.C., 1982. Kinetics of reactions between aqueous sulfates and sulfides

in hydrothermal systems. Geochimica et Cosmochimica Acta, 46, 1727–1745.

Ohmoto, H., Rye, R.O., 1979. Isotopes of sulfur and carbon. Geochemistry of Hydrothermal

Ore Deposits (Barnes, HL, ed.), 509–567.

Paytan, A., Arrigo, K.R., 2000. The sulfur-isotopic composition of Cenozoic seawater

sulfate: implications for pyrite burial and atmospheric oxygen. International Geology

Review, 42, 491–498.

105
Paytan, A., Kastner, M., Campbell, D., Thiemens, M.H., 1998. Sulfur isotopic composition

of Cenozoic seawater sulfate. Science, 282, 1459–1462.

Paytan, A., Kastner, M., Campbell, D., Thiemens, M.H., 2004. Seawater sulfur isotope

fluctuations in the Cretaceous. Science, 304, 1663–1665.

Paytan, A., Mearon, S., Cobb, K., Kastner, M., 2002. Origin of marine barite deposits: Sr

and S isotope characterization. Geology, 30, 747–750.

Pearce, J.A. 1996. A user’s guide to basalt discrimination diagrams. In: Wyman D.A. (Ed.):

Trace element geochemistry of volcanic rocks: applications for massive sulphide

exploration. Geological Association of Canada, Short Course Notes, 12(79), 79–113.

Pfaff, K., Hildebrandt, L.H., Leach, D.L., Jacob, D.E., Markl, G., 2010. Formation of the

Mississippi Valley-type Zn-Pb-Ag deposit in the extensional setting of the upper

Rhinegraben in the Wiesloch area, SW Germany. Mineralium Deposita, 45, 647–666.

Poole, F.G., 1988. Stratiform barite in Paleozoic rocks of the western United States. In

Proceedings of the Seventh Quadren-nial IAGOD Symposium (ed. E. Zachrisson). E.

Schweizerbar’sche Verlagsbuchhandlung, Stuttgart, 309–319.

Rahimpour-Bonab, H., Shekarifard, A., 2002. Barite Ore Deposits in the Central

IranArdakan Province: Genesis and Alterations. Iranian Journal of Science and

Technology, 3(1), 69–91.

Rajabi, A., Canet, C., Rastad, E., Alfonso, P., 2014. Basin evolution and stratigraphic

correlation of sedimentary-exhalative Zn–Pb deposits of the Early Cambrian

ZariganChahmir basin, Central Iran. Ore Geology Reviews, 64, 328–353.

Ramos, V.A., de Brodtkorb, M.K., 1989. Celestite, barite, magnesite and fluorspar:

stratabound through time and space. Nonmetalliferous stratabound ore fields, 297–321.

Rao, X., 1997. Genesis of stata-bound ore deposits in central Guangxi, South China: fluid

inclusion and isotope constraints. Ph.D. dissertation, Yale University.

Rastad, E., 1981. Geological, mineralogical, and facies investigations on the Lower

Cretaceous stratabound Zn–Pb–(Ba–Cu) deposits of the Irankuh Mountain Range,

106
Esfahan, west Central Iran (Unpublished Ph.D. thesis). University of Heidelberg,

Heidelberg, p. 334.

Rees, C.E., Jenkins, W.J., Monster, J., 1978 Sulfur isotopic composition of ocean water

sulfate. Geochimica et Cosmochimica Acta, 42, 377–381.

Richards, J.P., Wilkinson, D., Ullrich, T., 2006. Geology of the Sari Gunay epithermal gold

deposit, northwest Iran. Economic geology, 101(8), 1455–1496.

Riedinger, N., Kasten, S., Groger, J., Franke, C., Pfeifer, K., 2006. Active and buried

authigenic barite fronts in sediments from the Eastern Cape Basin. Earth and Planetary

Science Letters, 241, 876–887.

Rudnick, R.L., Gao, S., Holland, H.D., Turekian, K.K., 2003. Composition of the continental

crust. The crust, 3, 1–64.

Rushdi, A.I., McManus, J., Collier, R.W., 2000. Marine barite and celestite saturation in

seawater. Marine Chemistry 69, 19–31.

Rye, R.O., Shawe, D.R., Poole, F.G., 1978. Stable isotope studies of bedded barite at East

Northumberland Canyon in Toquima Range, central Nevada. US Geological Survey

Journal of Research, 6, 221–229.

Safina, N.P., Melekestseva, I.Y., Nimis, P., Ankusheva, N.N., Yuminov, A.M., Kotlyarov,

V.A., Sadykov, S.A., 2016. Barite from the Saf’yanovka VMS deposit (Central Urals) and

Semenov-1 and Semenov-3 hydrothermal sulfide fields (Mid-Atlantic Ridge): a comparative

analysis of formation conditions. Mineralium Deposita, 51(4), 491–507.

Sakai, H., 1971. Sulfur and oxygen isotopic study of barite concretions from banks in the

Japan Sea of the Northeast Honshu, Japan. Geochemical Journal, 5, 79–83.

Sassen, R., Sweet, S.T., DeFreitas, D.A., Eaker, N.L., Roberts, H.H., Zhang, C., 2004. Brine

vents on the Gulf of Mexico slope: Hydrocarbons, carbonate-barite-uranium

mineralization, red beds, and life in an extreme environment. Society of Economic

Paleontologists and Mineralogists, 24, 258–70.

107
Schenau, S.J., Prins, M.A., De Lange, G.J., Monnin, C., 2001. Barium accumulation in the

Arabian Sea: Controls on barite preservation in marine sediments. Geochimica et

Cosmochimica Acta, 65, 1545–1556.

Schulz, K.J., DeYoung, J.H., Seal, R.R., Bradley, D.C. eds., 2017. Critical mineral resources

of the United States: economic and environmental geology and prospects for future

supply. Geological Survey.

Seal, R.R., Alpers, C.N., Rye, R.O., 2000. Stable isotope systematics of sulfate minerals.

Reviews in Mineralogy and Geochemistry. In: Alpers, C.N., Jambor, J.L., Nordstrom,

D.K. (Eds.), Sulfate Minerals: Crystallography, Geochemistry, and Environmental

significance. Mineralogical Society of America, Washington D.C, 541–602.

Şengör, A.M.C., 1979. Mid-Mesozoic closure of Permo-Triassic Tethys and its implications.

Nature 279, 590–593.

Shawe, D.R., Poole, F.G., Brobst, D.A., 1969. Newly discovered bedded barite deposits in

East Northumberland canyon, Nye County, Nevada. Economic Geology, 64, 245–254.

Shelley, D., 1993. Igneous and metamorphic rocks under the microscope: classification,

textures, microstructures and mineral preferred orientation. Chapman and Hall, United

Kingdom, 1–445.

Stampfli, G.M., 2000. The Tethyan Ocean. In: Bozkurt E, Winchester JA, Piper JDA (eds)

Tectonic and magmatism in Turkey and surrounding area, Vol. 173. Geological Society,

London, Special Publication, 1–23.

Stevens, E.W.N., Bailey, J.V., Flood, B.E., Jones, D.S., Gilhooly, W.P., Joye, S.B., Teske,

A., Mason, O.U., 2015. Barite encrustation of benthic sulfur-oxidizing bacteria at a

marine cold seep. Geobiology, 13, 588–603.

Stöcklin, J., 1968. Structural history and tectonics of Iran: a review. AAPG bulletin 52,

1229–1258.

Stöcklin, J., 1977. Structural correlation of the Alpine ranges between Iran and central Asia.

Mém Hors Sér Soc Géol Fr 8, 333–353.

108
Stoffers, P., Worthington, T.J., Schwarz-Schampera, U., Hannington, M.D., Massoth, G.J.,

Hekinian, R., Schmidt, M., Lundsten, L.J., Evans, L.J., Vaiomo'Unga, R., Kerby, T.,

2006. Submarine volcanoes and high-temperature hydrothermal venting on the Tonga

arc, southwest Pacific. Geology, 34(6), 453–456.

Strauss, H., Banerjee, D.M., Kumar, V., 2001. The sulfur isotopic composition of

Neoproterozoic to early Cambrian seawater—evidence from the cyclic Hanseran

evaporites, NW India. Chemical Geology, 175(1–2), 17–28.

Tajeddin, H.A., Rastad, E., Yaghoubpour, A., Maghfouri, S., Peter, J.M., Goldfarb, R.,

Mohajjel, M., 2019. The Barika gold-bearing Kuroko-type volcanogenic massive sulfide

(VMS) deposit, Sanandaj-Sirjan zone. Iran. Ore Geology Reviews, 113, 103081. Torres,

M.E., Bohrmann, G., Dubé, T.E., Poole, F.G., 2003. Formation of modern and Paleozoic

stratiform barite at cold methane seeps on continental margins. Geology, 31(10), 897–900.

Torres, M.E., Bohrmann, G., Suess, E., 1996a. Authigenic barites and fluxes of barium

associated with fluid seeps in the Peru subduction zone. Earth and Planetary Science

Letters, 144(3–4), 469–481.

Torres, M.E., Brumsack, H.J., Bohrmann, G., Emeis, K.C., 1996b. Barite fronts in

continental margin sediments: a new look at barium remobilization in the zone of sulfate

reduction and formation of heavy barites in diagenetic fronts. Chemical Geology, 127(1–

3), 125–139.

Torres, M.E., McManus, J., Huh, C.A., 2002. Fluid seepage along the San Clemente Fault

scarp: basin-wide impact on barium cycling. Earth and Planetary Science Letters, 203,

181–194.

Turchyn, A.V., Schrag, D.P., 2004. Oxygen isotope constraints on the sulfur cycle over the

past 10 million years. Science, 303, 2004–2007.

Turchyn, A.V., Schrag, D.P., 2006. Cenozoic evolution of the sulfur cycle: Insight from

oxygen isotopes in marine sulfate. Earth and Planetary Science Letters, 241(3–4), 763–

779.

109
Turchyn, A.V., Sivan, O., Schrag, D.P., 2006. Oxygen isotopic composition of sulfate in

deep sea pore fluid: evidence for rapid sulfur cycling. Geobiology, 4(3), 191–201.

Valenza, K., Moritz, R., Mouttaqi, A., Fontignie, D., Sharp, Z., 2000. Vein and karst barite

deposits in the western Jebilet of Morocco: fluid inclusion and isotope (S, O, Sr) evidence

for regional fluid mixing related to central Atlantic rifting. Economic Geology, 95, 587–

606.

Vazifeh, R., 2013. The study of polymetallic mineralization related to Barite veins in

North Daryan (Mishow Mountain-East Azerbaijan), MSC Thesis, University of Tabriz,

Iran.

Walker, J.C., 1986. Global geochemical cycles of carbon, sulfur and oxygen. Marine

Geology, 70, 159–174.

Whitney, D.L., Evans, B.W., 2010. Abbreviations for names of rock-forming minerals.

American mineralogist 95, 185–187.

Williams-Jones, A.E., Samson, I.M., Olivo, G.R., 2000. The genesis of hydrothermal

fluorite-REE deposits in the Gallinas Mountains, New Mexico. Economic Geology,

95(2), 327–341.

Wortmann, U.G., Chernyavsky, B., Bernasconi, S.M., Brunner, B., Böttcher, M.E., Swart,

P.K., 2007. Oxygen isotope biogeochemistry of pore water sulfate in the deep biosphere:

dominance of isotope exchange reactions with ambient water during microbial sulfate

reduction (ODP Site 1130). Geochimica et Cosmochimica Acta, 71(17), 4221–4232.

Wortmann, U.G., Chernyavsky, B.M., 2007. Effect of evaporite deposition on early

Cretaceous carbon and Sulphur cycling. Nature, 446, 654–656.

Zarasvandi, A., Zaheri, N., Pourkaseb, H., Chrachi, A., Bagheri, H., 2014. Geochemistry

and fluid-inclusion microthermometry of the Farsesh barite deposit. Iran. Geologos, 20,

201–214.

110
‫‪Zhang, H.F., Li, S.R., Santosh, M., Liu, J.J., DiWu, C.R., Zhang, H., 2013. Magmatism and‬‬

‫‪metallogeny associated with mantle upwelling: zircon U–Pb and Lu–Hf constraints from‬‬

‫‪the gold-mineralized Jinchang granite, NE China. Ore Geology Reviews, 54138– 156.‬‬

‫‪Zhou, X., Chen, D., Dong, S., Zhang, Y., Guo, Z., Wei, H., Yu, H., 2015. Diagenetic barite‬‬

‫‪deposits in the Yurtus Formation in Tarim Basin, NW China: Implications for barium‬‬

‫‪and sulfur cycling in the earliest Cambrian. Precambrian Research, 79–87.‬‬

‫‪Zolfi, L., Simmonds, V., 2015. Mineralogy and genesis of Barite mineralization in Shanjan,‬‬

‫‪North of Shabestar, NW Iran. Geochemistry, 4(1), 1–14.‬‬

‫چکید ه‬

‫ر‬
‫کیلومتی شمال رشق شهرستان مهاباد در قسمت‬ ‫کان ی ز یای باریت – آهن – مس مورد مطالعه در ‪20‬‬

‫ستجان قرار دارد و از رگه ها و توده های معد یی پرکننده فضای باز تشکیل‬ ‫ی‬
‫غری زوندگرگوی سنندج – ر‬
‫شمال ی‬
‫ی‬
‫معدی کوارتز‪،‬‬ ‫شده است‪ .‬باریت در ریولیت ها یسازند مهاباد ر ی‬
‫متبان ی یم شود‪ .‬ریولیت سازند مهاباد از مجموعه‬

‫کای های مافیک مانند ر‬


‫بیوی ت و هورنبلند تشیک ل شده است‪ .‬ریویل تهای سازند‬ ‫‪K-‬فلدسپات‪ ،‬پالژیوکالز و ی‬

‫ً‬ ‫ر‬ ‫ً‬


‫عمدتا ر ر‬
‫داسیب را نشان م یدهند و عمدتا در میدان کالک آلکال ین و‬ ‫یولیب‪ ،‬ریوداسیت ی و‬ ‫کیب‬
‫مهاباد تر ی‬

‫شوشونیت با پتاسیم باال رسم میکنند‪ .‬ریولیتهای سازند مهاباد در نمودارهای چند عنرص ییتمالشده با گوشته اولیه‬

‫با غن یسازی شب و ‪ Nd‬و کاهش ‪ Sr،Y‬و مقداری( ‪ HFSE‬مانند ‪ Ti‬و )‪ Nb‬مشخص میشون دکه نشاندهنده‬

‫‪111‬‬
‫تنظیم قوس قارهای مربوط به فرورانش است‪ .‬ریولیت های سازند مهاباد‪ .‬آنها با غنیسازی عنارص خایک کمیاب‬
‫ً‬ ‫ری‬
‫سنگی)‪ ، (HREE‬عمدتا ناهنجاری ‪ Eu‬منف ی تا کیم‬ ‫سبک)‪ ، (LREE‬کاهش عنارص خایک کمیاب‬

‫مثبت = *‪34.0-(19.1 (Eu/Eu‬و ناهنجار ی مثبت ‪ Yb‬مشخص میشوند‪ .‬باریت با محتویات کل عنارص‬

‫خایک کمیاب‪(REE) (5.95-‬‬

‫)‪54.09 ppm‬مشخص یم شود‪ .‬الگوهای ‪ REE‬نرمال شده با کندریت نشان یم دهد که باریت ها در عنارص‬

‫سنگی )‪ (HREE‬ی‬
‫غب شده اند همانطور که با‬ ‫ری‬ ‫خایک کمیابسبک )‪ (LREE‬در مقایسه با عنارص خایک کم یاب‬

‫ری‬
‫همچنی مثبت تلفظ م ی شود نشان داده‬ ‫متغت است و‬
‫ر‬ ‫نسبت های باال ‪(La/Yb)N‬که از ‪ 61.1‬تا ‪26.24‬‬

‫شده است‪ .‬ناهنجاری های ‪ Ce‬از‬

‫‪ 61.1‬تا ‪ .55.3‬مقادیر ‪ δ34S‬در باریت های تحلیل شده از کانسار قزل داغ از‪ 41.15 +‬تا‪ ،‰ 14.23 +‬سازگار با‬

‫یای در زمان نئوپروتروزوییک و پالئوزوئیک )‪ (‰ 12-38‬است‪ .‬مقادیر ‪δ34S‬‬


‫ترکیب معمویل سولفات در ی‬
‫گستدهای از مقادیر ‪ δ34S‬را نشان م یدهند‪ .‬ر‬
‫گستش‬ ‫نمونههای باریت باال‪ ،‬مقادیر بسیار مثبت هستند و توزی ع ر‬

‫ر‬ ‫ً‬
‫استشای ط محییط یکنواخت را در شاش میدا ن کانیسازی نشان دهد‪.‬‬ ‫نسبتا باریک در مقادیر ‪ δ34S‬ممکن‬

‫ایزوتوی‬ ‫داده های ایزوتوپ گوگرد نشان یم دهد که ی‬


‫کای سازی ازسولفات آب دریا تشکیل شده است که به صورت‬
‫ی‬

‫اصالح شده است‪ .‬مقادیر ‪ δ18O‬برای نمون ههای باریت از محدوده کانسار قزلداغ در محدوده ‪ +12.4‬تا‬

‫یای نئوپروتروزوییک وپالئوزوئیک )‪ (‰ 10-20‬قرار‬


‫میگتد که در محدوده ‪ δ18O‬برای سولفات در ی‬
‫‪ +88.17‬قرار ر‬

‫میگتد‪ .‬بافتهای پرکننده فضای باز باریت‪ ،‬غلظ تهای ‪ Sr‬باال)‪ ، )0111.2-2168.3 ppm‬غلظت کم ‪ΣREE‬‬
‫ر‬

‫)‪ ، (5.95-54.09 ppm‬الگوهای ‪ REE‬نرمالشده با کندریت غنیشده با‪ ، LREE‬و ناهنجاری هایمثبت‪Eu -(8‬‬

‫یردریای برا ی‬ ‫ری‬


‫همچنی مقادیر ‪ δ34S‬شواه دی را برای حمایت از منشا هیدروترمال ز‬ ‫‪ ، (2.56‬ب هعنوان ‪.2/8‬‬
‫ی‬

‫باریت در کانسار قزل داغ فراهم میکند‪.‬‬

‫واژگان کلیدی‪ :‬بار یت‪ ،‬سازند مهاباد ‪ ،‬منطقه سنندج‪ -‬ر‬


‫ستجان‪ ،‬شمال غرب ایرا‬
‫ن‬

‫‪112‬‬
‫دانشکده علوم‬

‫گروه زمینشناس ی‬

‫پایاننامه‬

‫برای دریافت درجه کارشنایس ارش د در رشته‬


‫زمینشنایس گرایش اقتصاد ی‬

‫عنوان‬

‫هایغرن و‬
‫ی‬ ‫ژئوشیمیان و دگرسان در بخش‬
‫ی‬ ‫مقایسه ویژگ های کان سازی‪،‬‬
‫ر ی‬
‫غرن‪ ،‬شمال‬
‫شق کانسار باریت قزل داغ مهاباد‪ ،‬استان آذربایجا ن ی‬
‫غرب ایرا ن‬

‫استاتید راهنما‬

‫دکت ر یوسف رحیم سور ی‬


‫ی‬
‫دکت عیل عابدین ی‬

‫پژوهشگر‬

‫تحسی حیم د‬
‫ر‬ ‫شتوان‬
‫ر‬
‫‪1401‬‬
‫مرداد‬

‫‪113‬‬

You might also like