30.12.2012 Views

Basem Ahmed Zoheir_Beida_JGE_2012.pdf

Basem Ahmed Zoheir_Beida_JGE_2012.pdf

Basem Ahmed Zoheir_Beida_JGE_2012.pdf

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Lode-gold mineralization in convergent wrench structures: Examples from South<br />

Eastern Desert, Egypt<br />

<strong>Basem</strong> A. <strong>Zoheir</strong><br />

Department of Geology, Faculty of Science, Benha University, 13518 Benha, Egypt<br />

article info<br />

Article history:<br />

Received 14 July 2011<br />

Accepted 29 December 2011<br />

Available online 5 January 2012<br />

Keywords:<br />

Lode-gold<br />

Convergent wrench structures<br />

Fluid mixing<br />

Wadi El <strong>Beida</strong>–Wadi Khashab<br />

Eastern Desert<br />

Egypt<br />

1. Introduction<br />

abstract<br />

It is believed that most significant gold deposits throughout the<br />

world are controlled by major and subsidiary shear zones (e.g.,<br />

Bonnemaison and Marcoux, 1990; Hodgson, 1989; Roberts, 1987).<br />

Ore fluidmigrationandgolddeposition processes in these deposits<br />

are attributed to a contrasting fluid-pressure/temperature regime<br />

between the first and second order structures (Eisenlohr et al.,<br />

E-mail address: basem.zoheir@gmail.com.<br />

0375-6742/$ – see front matter © 2012 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.gexplo.2011.12.005<br />

Journal of Geochemical Exploration 114 (2012) 82–97<br />

Contents lists available at SciVerse ScienceDirect<br />

Journal of Geochemical Exploration<br />

journal homepage: www.elsevier.com/locate/jgeoexp<br />

Data from regional- and mine-scale mapping, structural, geochemical and fluid inclusion studies reveal a synkinematic<br />

genesis of gold in convergent wrench structures that cut the Neoproterozoic greenstone belt in the<br />

south Eastern Desert of Egypt. Au-quartz lodes in the Wadi El <strong>Beida</strong>–Wadi Khashab area are associated with<br />

NNW-trending shear zones in pervasively silicified, ferruginated volcanic/volcaniclastic rocks, or along steeply<br />

dipping thrust segments bounding allochthonous ophiolitic blocks. Development of the mineralized shear zones<br />

is attributed to a wrench-dominated transpression assigned to D2 deformation throughout the Pan-African evolution<br />

of the South Eastern Desert (620–540 Ma?).<br />

Hydrothermal alteration associated with Au-quartz lodes comprises an inner quartz–sericite–pyrite assemblage<br />

progressed outwards into an outer quartz–chlorite–calcite assemblage. Mass balance calculations reveal a systematic<br />

volume- and mass-increase, addition of SiO2, K2O, Fe2O3 t , S and L.O.I. and slight depletion in Na2O in<br />

the altered wallrocks approaching the Au-quartz lodes. Erratic concentrations of S, MgO and CaO in the altered<br />

host rocks, however, suggest selective carbonatization, and sulfidation of Fe-rich host rocks proximal to the<br />

ore bodies.<br />

Ore bodies include quartz-only and quartz–carbonate lodes with disseminated pyrite, chalcopyrite, chalcocite,<br />

covellite, marcasite, subordinate pyrrhotite, sphalerite and gold. In addition to free gold inclusions in As-poor<br />

pyrite, microprobe and LA-ICP-MS spot analyses reveal the presence of traces of Au, Ag (10s to 100s ppmlevels),<br />

positively correlated with Cu contents (1000s ppm-levels). Analyses of pyrrhotite and marcasite indicate<br />

a comparable relationship. Chalcopyrite intimately associated with pyrite contains lower levels of refractory<br />

gold and silver (av. 12 ppm Au and 3 ppm Ag). Solid solution may have been responsible for invisible gold,<br />

whereas, free gold deposition is a function of remobilization, reconstitution and concentration of the earlier<br />

phase.<br />

Primary and pseudosecondary fluid inclusions in the auriferous quartz veins comprise low salinity aqueous–<br />

carbonic fluids (2–12 eq. wt.% NaCl). Homogenization temperatures of synchronous aqueous-dominant and<br />

carbonic-dominant (H2O–CO2–NaCl±CH4±N2) fluid inclusions (258–343 °C) correspond to 0.8–2.3 kbars and<br />

depths of 3 and 9 km (mesothermal conditions). Decreasing gold solubility and segregation from bi-sulfide complexes<br />

most probably resulted from interplay of dilution and mixing of an evolved carbonic-rich fluid with a<br />

more oxidized aqueous fluid, pressure fluctuation and wallrock-sulfidation.<br />

New geochemical data combined with available geophysical information indicate viable gold ore bodies in the<br />

study area, and suggest similar situation for zones with discernible signs of hydrothermal alteration along the<br />

major convergent wrench (shear) zones in the Eastern Desert of Egypt.<br />

© 2012 Elsevier B.V. All rights reserved.<br />

1989). According to Cox et al. (1991), fluid flow in shear-hosted<br />

gold deposits is seldom distributed uniformly along individual<br />

faults or shear zones, but is localized within distinct segments of<br />

these structures where permeability and/or gradients in hydraulic<br />

head are highest. Elements that commonly govern ore localization<br />

and deposit geometry are fault bends, jogs, fault intersections and<br />

splays (Cox et al., 1991).<br />

Transpression describes a style of deformation that involves collisional<br />

orogenesis accompanied by strike–slip shear across a zone<br />

(Dabo et al., 2008). Wrench-dominated transpression, a characteristic<br />

feature of obliquely convergent mobile belts, has been suggested to


explain the complex deformation kinematics in the Eastern Desert of<br />

Egypt (e.g., Abd El-Wahed, 2008; Abd El-wahed and Kamh, 2010;<br />

Abdeen et al., 2008; Fritz et al., 1996; Fritz et al., 2002; Greiling et<br />

al., 1994; Loizenbauer et al., 2001; Makroum, 2001; Shalaby, 2010;<br />

Shalaby et al., 2005; Wallbrecher et al., 1993). In south Eastern Desert<br />

of Egypt, discrete NW- to NNW-trending, kilometer-scale, shear zones<br />

cutting the ophioltic and island arc metavolcanic/volcaniclastic assemblages<br />

in the Wadi El <strong>Beida</strong>–Wadi Khashab district are likely related to<br />

the ca. 620–540 Ma, Najd wrench system (de Wall et al., 2001; Kröner<br />

et al., 1987; Sadek et al., 1996; Stern, 1985; Stern et al., 1989). Transpressional<br />

wrenching related to the Najd system was consequent to<br />

thrust-related structures associated with oblique convergence of the<br />

arc and back-arc assemblage onto the Saharan Metacraton (e.g., Abd<br />

El-Wahed, 2008 and references therein).<br />

Gold-bearing quartz veins are widespread in south Eastern Desert<br />

of Egypt, commonly showing spatial and temporal association with<br />

shear zones (e.g., Kusky and Ramadan, 2002; <strong>Zoheir</strong>, 2008, 2011).<br />

Abundant remnants of ancient gold mining on both flanks of the<br />

Wadi El <strong>Beida</strong> and Wadi Khashab, including stopped-out quartz<br />

veins and tailings, and stone huts likely date back to the Arab times<br />

(i.e. 11th century AD) or earlier. These mine working ruins could<br />

have been a part of the big gold rush in the Nubia desert at that<br />

time. Mine workings are confined to discrete NNW-trending shear<br />

zones that are associated with shear-band boudins of quartz and<br />

highly oxidized, ferruginated silicified metavolcanic/volcaniclastic<br />

rocks (gossan-like bodies). Previous studies (Kontny et al., 1999;<br />

Obeid et al., 2001; Ramadan, 1994) reported anomalous gold contents<br />

(10s ppm Au) in the silicified rocks. Significant Au concentrations<br />

have been reported in gossans, brecciated quartz and quartz carbonate<br />

veins rich in malachite (Kontny et al., 1999; Nano et al., 2002). Anomalous<br />

gold concentrations in the gossan-like bodies encouraged a preliminary<br />

self-potential magnetic and induced polarization surveying<br />

(Sultan et al., 2009), which revealed the presence of conductor bodies<br />

(55 to 175 m-across) at ~40–53 m depth in this area. Most magnetic<br />

and resistivity anomalies coincide with silicified, shear zones, while<br />

the low-resistivity zones coincide with the altered metavolcanic rocks<br />

(Sultan et al., 2009).<br />

Although evolution of this part of the Wadi El <strong>Beida</strong>–Wadi Khashab<br />

district is very much affected by the Pan-African Najd wrench system<br />

(620–540 Ma), no adequate information is available to envisage the<br />

relationship between gold mineralization and deformation of the<br />

country rocks in this area. The present contribution is intended to<br />

provide detailed information on the geological–structural setting<br />

and genesis of the auriferous silicified shear zones and quartz lodes<br />

in relation to the regional wrench structures based on field work,<br />

petrographic and geochemical methods.<br />

2. Materials and methods<br />

High-resolution mapping of the lithological units, alteration halos<br />

and major structures in the study area was aided by using false color<br />

composite images and color composite ratio satellite images (ETM+<br />

and ASTER). Field work encompassed sampling of the mineralized<br />

shear zones, hydrothermally altered and fresh host rocks. Structural<br />

measurements, marker sketching and lithological identification have<br />

been foremost parts of the field work. Petrographic and ore microscopy<br />

studies are done on 80 polished thin sections of altered/mineralized and<br />

fresh/un-mineralized quartz lodes and host rocks. Out of these polished<br />

sections, 31 samples were chosen for electron microprobe studies<br />

(EMPA).<br />

Thirty-one representative samples of fresh host rocks, hydrothermally<br />

altered wallrocks and ore bodies were chosen for geochemical<br />

studies. Bulk rock analyses for major and trace elements including<br />

Au were done at the Activation Laboratories Ltd. (Canada) using<br />

two packages (codes: 4B and 1H) using the simultaneous multielement<br />

neutron activation INAA and inductive coupled plasma ICP<br />

B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

(total digestion) techniques. These two packages offer precise detection<br />

limits for most important elements, e.g. Au (2 ppb), Ag (0.3 ppm), Bi<br />

(2 ppm), Sb (0.1 ppm) and As (0.5 ppm).<br />

A Cameca SX100 Electron Microprobe at the Institute for Mineralogy<br />

and Mineral Resources, TU-Clausthal (Germany) was used to determine<br />

the composition of ore minerals. The operating conditions applied for<br />

most elements: 30 keV and 2 nA beam current. For analysis of Au, Ag,<br />

Ni, Co, Bi and Sb, a 100 nA beam current was applied, and count times<br />

ranged from 10 to 400 s. For high-resolution trace element geochemistry<br />

of sulfides, some samples were spot-analyzed by a quadrupole LA-<br />

ICP-MS (Elan 6000) at the US Geological Survey Laser Ablation ICP-MS<br />

Laboratory, Denver.<br />

Fluid inclusion microthermometric measurements were carried out<br />

on ~150 μm-thick doubly-polished wafers using a Linkam TMH-600<br />

heating/freezing stage at the fluid-inclusion laboratory of the Department<br />

of Geology and Geochemistry at Stockholm University. Calibrations<br />

were made against known melting points of pure substances.<br />

Heating rates of 0.5 and 1 °C/min were applied to record phase changes<br />

below 30 °C, whereas a heating rate of 5 °C/min was applied for phase<br />

changes above this temperature. The accuracy is ±0.2 °C for the measured<br />

melting temperatures and within ±2 °C for the homogenization<br />

temperatures. Parameters used in fluid inclusion microthermometry include:<br />

first melting temperatures of CO 2 (T mCO2), freezing depression<br />

(melting) temperatures of clathrate (Tm clath), ice melting temperatures<br />

(T mice), partial homogenization temperatures of CO 2 (T hCO2), and total<br />

homogenization temperatures (Th total).<br />

3. Geology and structural setting<br />

The Neoproterozoic greenstone belt exposed in the Wadi El<br />

<strong>Beida</strong>–Wadi Khashab area comprises variably deformed ophiolites<br />

and island-arc volcanic–plutonic rocks. This metamorphic pile is<br />

cut by extensive syn-orogenic granitoids and discrete intrusions of<br />

late/post-orogenic gabbros and granites. These Precambrian rocks<br />

are un-conformably overlain by Cretaceous sandstones (Nubian<br />

Sandstone) and intruded by Red Sea rift-related, Tertiary basalt<br />

(Ramadan and Kontny, 2004).<br />

Metasomatized serpentinite, metagabbro and less common<br />

pillow metabasalt are overthrusted on the island-arc metavolcanic/<br />

volcaniclastic rocks along steeply east-dipping thrust/strike–slip fault<br />

structures (e.g., Wadi Khashab and Wadi El <strong>Beida</strong>, Fig. 1). Serpentinite<br />

and associated talc carbonate rocks form large masses, elongated in a<br />

NW–SE direction (Gebel Sirsir and parts of Gebel El <strong>Beida</strong>), or<br />

occur as irregular slices tectonically incorporated within the volcaniclastic<br />

terrains. Ophiolitic metagabbro is intimately associated<br />

with serpentinite and metabasalt and locally exhibits compositional<br />

layering. Pillowed, amygdaloidal metabasalts associated with discrete,<br />

stretched chert lenses form a NW–SE belt alternating with<br />

the ultramafic rocks close to El <strong>Beida</strong> water well. The island-arc<br />

metavolcanic/volcaniclastic assemblages are mainly foliated metaandesite/basaltic<br />

andesite and intermediate to acidic tuffs (i.e.,<br />

mudstones, andesitic tuffs and pyroclastics), forming the NW–SE El<br />

<strong>Beida</strong>–Khashab mountainous belt. Arc-related gabbro–diorite intrusions<br />

cut the metavolcanic/volcaniclastic rocks, forming low-relief hills in the<br />

western part of the study area (Fig. 1). The syn-orogenic intrusions are<br />

characterized by a wide compositional range, from tonalite to monzogranite.<br />

They cut the gabbro–diorite and metavolcanic/volcaniclastic rocks,<br />

and show variable degrees of shearing and mineral foliation. The Postorogenic<br />

gabbro forms a small circular intrusion in the southwestern<br />

part of the mapped area (G. Homraii) cutting metavolcanic rocks and<br />

syn-orogenic granitoids. Isolated intrusions of post-orogenic granites<br />

(mainly monzogranite) cut serpentinite and syn-orogenic granitoids in<br />

the eastern and southwestern parts of the study area, respectively.<br />

The study area evolved throughout a multistage deformation history<br />

(e.g., Abdeen et al., 2008; El Amawy et al., 2000; Kontny et al.,<br />

1999; Nano et al., 2002; Obeid et al., 2001). Neoproterozoic terrane<br />

83


84 B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

Fig. 1. Geological map of the Wadi El <strong>Beida</strong>–Wadi Khashab region in south Eastern Desert of Egypt.<br />

Modified from Ashmawy (1987), EGSMA (1992), El Amawy et al. (2000) and Abdeen et al. (2008).<br />

accretion structures (~720 Ma?), best developed southwest of the study<br />

area, have been completely obliterated by post-accretionary shortening<br />

and transpression structures (Table 1). The post-collisional evolution includes<br />

an early NNE–SSW crustal shortening (D1), which led to the formation<br />

of WNW–ESE tight intrafolial or overturned folds (F 1) and axial<br />

planar foliations (S 1). The latter is parallel to the primary (lamination)<br />

bedding (So) in the island arc metavolcanic/volcaniclastic rocks.<br />

Southward-vergence is suggested based on the geometry of these<br />

asymmetrical folds. A subsequent sinistral transpression (D2) was<br />

likely due to NW-ward nappe stacking led to transpression along


Table 1<br />

Summary of the deformation events affected the study area based on the observed structures.<br />

Episode Deformation events Structural elements Kinematics Veins and alteration<br />

Terrane exhumation<br />

(e.g., Abdeen et al., 2008)<br />

Post-collisional transpression<br />

(Najd fault system; 575–520 Ma;<br />

Greiling et al., 1994; de Wall<br />

et al., 2001)<br />

Post-collisional shortening<br />

(e.g., Greiling, 1987; de Wall<br />

et al., 2001)<br />

Weak brittle deformation,<br />

which may have a relation<br />

to the Red Sea opening<br />

(e.g., Abdeen et al., 2008).<br />

(D2) Kinematic partitioning<br />

of ~E–W shortening and<br />

sinistral transpression<br />

(e.g., Abd El-Wahed and<br />

Kamh, 2010).<br />

(D1) NNE–SSW shortening,<br />

crustal thickening accompanied<br />

island-arc accretion and nappe<br />

stacking resulted in a partitioned<br />

displacement with northwest<br />

directed thrusting in the<br />

internal portions and west to<br />

southwest directed thrusting in<br />

the external portions of the<br />

orogeny (e.g., Stern, 1994).<br />

thrust segments at the base of the ophiolitic blocks and steeplydipping,<br />

left lateral NNW–SSE ductile shear zones and major faults<br />

deformed most of the exposed lithologies and overprinted the preexisting<br />

lithologic/tectonic contacts. Second-order, sinistral shear<br />

zones form a positive flower structure (e.g., Clendenin, 1993;<br />

Waldron et al., 2007). The steeply dipping strike–slip fault segments<br />

bifurcate into reverse faults and then flatten into moderately dipping<br />

NNW-trending thrust zones. Related kilometer-scale, NNW–SSE<br />

B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

—ENE-WSW dextral strike–slip<br />

faulting affected the syn-orogenic<br />

granodiorite and pre-existed rocks.<br />

— Normal faults and joints trending<br />

invariousdirectionsspreadoverthe<br />

study area.<br />

— Intrusion of maficdykeswarms<br />

in E–WorENE–WSW directions.<br />

— NNW–SSE left-stepping folds<br />

— Steeply-dipping, left lateral<br />

NNW–SSE ductile shear zones and<br />

major faults<br />

— Sub-horizontal stretching<br />

lineations on the steeply-dipping<br />

NW–SE thrust planes.<br />

— Conjugate sets of steeply-dipping,<br />

NW-trending sinistral and<br />

NE-trending dextral fault zones.<br />

— Crenulation cleavage (S2) superimposed on S1 in the<br />

metavolcaniclastic rocks.<br />

Sketch drawing explaining the<br />

geometric relationships between<br />

the different D2 fabrics.<br />

— WNW–ESE tight intrafolial or<br />

overturned folds (F1) commonly<br />

vergent to SW.<br />

— Axial planar foliations (S1), parallel<br />

to the primary (lamination) bedding<br />

So in the island<br />

arc metavolcanic/volcaniclastic rocks.<br />

— Less preserved ENE-plunging<br />

stretching lineation (L1)intheisland<br />

arc metavolcanic/volcaniclastic rocks<br />

(e.g., elongate pebbles and stretched<br />

porphyroblasts),<br />

developed on or close to the NW–SE<br />

thrust planes.<br />

Schematic sketch showing the geometry<br />

of WNW–ESE striking F1 folds in relation<br />

to the related foliation and lineation, and<br />

barren quartz pods.<br />

— The dextral faults<br />

may be the conjugate<br />

shear fractures to the<br />

NNW–SSE oriented<br />

sinistral wrench faults<br />

or are related to a late<br />

brittle deformation event<br />

(e.g., Abdeen et al., 2008).<br />

Crenulation cleavage (S2) imposed upon S1 in the<br />

metavolcaniclastics.<br />

Kinematics of the shear<br />

zones include:<br />

— S–C fabrics, foliation<br />

deflection in a NNW–SSE<br />

direction.<br />

— en echelon, slightly<br />

sigmoidal quartz lenses,<br />

related to progressive<br />

transpression of the<br />

NNW–SSE trending,<br />

sinistral wrench zone.<br />

SW-vergent folds (F1),<br />

and consistently<br />

ENE-plunging<br />

lineations indicating<br />

top-to-the WSW-tectonic<br />

transport<br />

folds (F2), showing en-echelon geometry, verge commonly to the<br />

WSW and are associated with pervasive axial planar crenulation<br />

cleavage (S2), which imposes upon S1 in the metavolcaniclastic<br />

rocks. Boudinaged, stretched gold-bearing, quartz lodes are commonly<br />

observed where S2 is superimposed on S1. The left-stepping<br />

folds are consistent with the sinistral sense of wrench motion, and<br />

have been rotated toward parallelism with the master fault as commonly<br />

seen in wrench systems (e.g., Harding, et al., 1985; Wilcox<br />

–<br />

— Stretched gold-bearing,<br />

quartz lodes common in<br />

domains where S 2 is<br />

imposed upon S 1.<br />

— Silicified, highly sheared<br />

metavolcanic and ophiolitic<br />

rocks along NW and NNW<br />

shear zones.<br />

— Highly stretched,<br />

boudingated quartz<br />

pockets, with no sulfides<br />

and barren in gold.<br />

— Albitization of the<br />

metavolcanic rocks is<br />

commonly associated with<br />

the highly sheared rocks<br />

alongside the quartz pods.<br />

85


86 B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

et al., 1973). Finally, weak brittle deformation led to ENE–WSW dextral<br />

strike–slip faulting that affected the syn-orogenic granodiorite<br />

and pre-existing rocks. This deformation event might be related to<br />

the Red Sea rift (e.g., Abdeen et al., 2008).<br />

Weathering of the mineralized structures is a striking feature in<br />

the study area, where zones of bleached wallrocks extend for hundreds<br />

of meters along the major fault/fracture zones and the splays<br />

(Fig. 2a). Granophyric dikes are ubiquitous in the altered shear<br />

zones, commonly associated with quartz veins and disseminated sulfides<br />

(Fig. 2b). Whereas, aplitic dikes cut less deformed metavolcanic/<br />

volcaniclastic rocks in Wadi Khashab (Fig. 2c). Quartz veins are less<br />

abundant in the shear zones compared to the granophyric dikes<br />

(Fig. 2d–f). However, networks of quartz veinlets cut the altered<br />

metavolcanic/volcaniclastic rocks in zones of dense faults/fractures<br />

intersections. Some quartz veins extend for 40 m and vary in thickness<br />

from a 2 to 60 cm. Alongside the shear zone, network of quartz<br />

veinlets cuts pervasively silicified–sericitized–sulfidized metavolcanic/volcaniclastic<br />

rocks (Fig. 2g). The main ore bodies are made up<br />

of quartz with subordinate carbonate and ferruginated wallrock<br />

selvages. Discrete barren quartz lenses are common in brecciated<br />

granodiorite along the western flank of Wadi Khashab.<br />

4. Mineralization style and lode characteristics<br />

Gold-sulfide mineralization is confined to discrete shear zones of<br />

highly silicified, ferruginated volcanic/volcaniclastic rocks commonly<br />

associated with sulfide-bearing granophyric dikes and quartz±<br />

carbonate veins (Tables 2a, 2b and 3). The mineralized shear zones<br />

form splays off the much larger NNW-trending shear zones, i.e.,<br />

the Wadi Khashab sinistral fault zone. Sub-parallel shear zones<br />

with disseminated pyrite in variably sheared, silicified metavolcanic/<br />

volcaniclastic rocks have been documented for several hundreds of meters<br />

in the study area (see Fig. 1). This shear zone set occurs within an<br />

approximately 1000 m-wide envelope of variably strained host rocks;<br />

individual shears have widths up to 5–10 m, whereas faults are more<br />

confined. The area between the major shear zones contains a complex<br />

pattern of anastomosing subvertical shear zones and fabrics wrapping<br />

around the rigid blocks of the metavolcanic/volcaniclastic rocks. The<br />

Fig. 2. Field photographs of altered/sheared rocks and quartz lodes in Wadi El <strong>Beida</strong>–Wadi Khashab area. (a) Extensive pervasively silicified–sericitized NNW-trending alteration<br />

zone in meta-andesites south of Wadi El <strong>Beida</strong>. (Photo looking to N). Notice the pervasive kaolinite alteration (1) bounding outer quartz–chlorite–carbonate zone (2) and the inner<br />

sulfide-bearing quartz–sericite (3) alteration zone, (b) Sulfide-bearing granophyric dike (G) along sheared/altered rock zone south of El <strong>Beida</strong> water well. (Photo looking to S), (c)<br />

Barren aplitic granophyric dike (A) cutting meta-agglomerate and breccias along Wadi Khashab. (Photo looking to SW), (d) Mineralized sheared quartz lenses cutting pervasively<br />

sericitized and furrigenated metavolcaniclastic rocks. (Photo looking N), (e) Bifurcated sheared quartz vein (Q) cutting and enclosing severely altered metavolcaniclastic rocks<br />

along Wadi Abu Hegleig south of El <strong>Beida</strong> water well. (Photo looking to NE), (f) Quartz veins cutting foliated gabbro–diorite complex west of Wadi Khashab. (Photo looking to<br />

NW), (g) Quartz network in highly sericitized metavolcaniclastic rocks next to mineralized quartz lodes in west of Wadi El <strong>Beida</strong>. (Photo looking to N), (h) Admixed milky and<br />

grayish quartz and wallrock material in worked quartz lodes from west of Wadi El <strong>Beida</strong>., (i) Brecciated quartz filling fractures in severely sericitized, sulfidized granophyric<br />

dikes east of Wadi Khashab. (Photo looking N).


average azimuth of the overall strike of the anastomosing shear network<br />

is 150°. The ophiolitic metabasalt and arc-related volcanic agglomerate<br />

and breccias flanking the shear zones are severely<br />

fragmented and intermixed. Crenulations on the WNW–ESE foliation<br />

(S1), pervasive NNW–SSE shears are common deformation patterns<br />

along the altered shear zones. The sense of displacement can be deduced<br />

unequivocally from several kinematic indicators (e.g. stretched<br />

objects and stretching lineations, asymptotic schistosity, rotated porphyroclasts,<br />

etc.), which consistently point toward left-lateral shearing.<br />

The auriferous quartz veins occur as closely-spaced, parallel or subparallel<br />

swarms of tabular and subordinate lensoidal veins (3–100 cmthick),<br />

extending laterally for a few tens of meters. Striations on the vein<br />

walls are almost invariably plunge steeply to north. Quartz is dominant,<br />

and carbonate minerals may constitute more than 20% in parts of<br />

these veins, commonly with disseminated opaque and clay minerals.<br />

Wallrock materials are present as strongly foliated slivers in quartz<br />

veins commonly associated with carbonate. Vugs and open-space<br />

filling textures are mostly absent. Vein quartz textures vary from<br />

buck (large porphyroblasts), comb (recrystallized grains) and microcrystalline<br />

(minute subgrains) quartz, and alternating or crosscutting<br />

bands of different grain size are typical in many veins. Microstructures<br />

indicate variable degrees of dynamic recrystallization by grain boundary<br />

migration, subgrain rotation and bulging recrystallization in different<br />

domains of the individual shear zone. Ribboned, buck quartz is<br />

dominant in most quartz veins (Fig. 3a), however, the comb quartz variety<br />

(b20% vol.)±carbonate minerals occupy the interspaces between<br />

the large quartz ribbons in sulfide-bearing veins (Fig. 3b). A positive<br />

correlation between well-developed comb textures and disseminated<br />

sulfide minerals is observed in many samples, given that interstitial<br />

spaces between comb quartz crystals have everywhere been sealed by<br />

sulfide minerals and sericite–carbonate±Fe–hydroxide. Laminated<br />

quartz veins are made up of alternating bands of severely recrystallized<br />

quartz+wallrock material+sulfide and/or carbonate minerals and<br />

bands of large less-deformed quartz grains (Fig. 3c). Variations in the<br />

degree of plastic deformation and in the abundance of microfractures<br />

are observed between adjacent ribbons and laminae. Some of the mineralized<br />

lodes are composed mainly of sub-rounded quartz grains with<br />

highly serrate grain boundaries reflecting significant degrees of bulk recrystallization<br />

(Fig. 3d). Features of conjugate and progressive shearing<br />

are common in many of the highly deformed quartz veins (i.e., S–C<br />

structures; Fig. 3e). In the fragmented parts of the mineralized lodes,<br />

veinlets of fine-grained, subhedral blocky quartz show drusy growth<br />

textures normal to the vein wall. Lack of displaced markers and dominance<br />

of the comb quartz textures (Fig. 3f), clearly imply that these<br />

late veins occupy true extension fractures (e.g., Simpson and Schmidt,<br />

1983).<br />

The analyzed samples of mineralized quartz veins give 2.1–<br />

6.6 ppm Au, and sulfidized granophyric dikes are also gold-bearing<br />

B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

(0.2–2.1 ppm Au; Table 2a). Significant correlation coefficient (r)<br />

values between the analyzed granophyric dikes and quartz lodes<br />

(Table 2b) indicate a positive correlation between Au, Ag and Cu.<br />

No discernible correlation occurs between As and either Au or Ag.<br />

5. Geochemistry of the mineralized shear zones<br />

Table 2a<br />

Concentrations of some trace elements in sulfidized granophyre dikes and quartz veins from the Wadi El <strong>Beida</strong>–Wadi Khashab area.<br />

Partial to complete destruction of the wallrock textures and<br />

primary minerals is distinct in the mineralized shear zones. Iron<br />

oxides/hydroxides fill fractures in the host rocks and replace disseminated<br />

sulfide grains in quartz veins and in the silicified parts of the<br />

shear zones. Malachite stains the sheared rocks, vugs and cleavage or<br />

fracture walls. Based on the dominant hydrothermal alteration phases,<br />

alteration progressed from outer silicified–chloritized–carbonatized to<br />

inner silicified–sericitized–pyritized rocks approaching the Au-quartz<br />

lodes and sulfidized granophyric dikes. In places, alteration is overprinted<br />

by kaolinite and grayish clay mixtures.<br />

Direct observations of the analyses (Table 3) indicate higher SiO 2,<br />

K2O, SO2 and Fe2O3* and lower MgO and CaO contents in the goldbearing,<br />

hydrothermally altered wallrocks compared to the least altered<br />

rocks. Loss on ignition (L.O.I.) is considerably high in samples<br />

displaying intense alteration. Au, Ag, and Cu increase systematically<br />

from the silicified–chloritized–carbonatized rocks to the inner silicified–sericitized–pyritized<br />

wallrocks (up to ~3 ppm Au, 2.3 ppm Ag,<br />

and 236 ppm Cu; Table 3).<br />

The chemical compositions of fresh and hydrothermally altered<br />

metavolcanic/volcaniclastic host rocks have been used to infer the<br />

mass gains and losses of components as a function of hydrothermal<br />

alteration. In this study, the isocon method of Grant (1986) is used<br />

to evaluate whether the compositional changes involved significant<br />

changes in concentrations of components as well as in mass and volume.<br />

The average values and 1σ errors of samples were used to minimize<br />

composition heterogeneity of the volcaniclastic rocks.<br />

Rock forming elements are plotted as oxides in wt.% and trace elements<br />

in ppm in double-logarithmic plots (Fig. 4). The isocon is defined<br />

as best fit line through relative immobile elements including Ti,<br />

P and Y (e.g., Leitch and Lentz, 1994; Selverstone et al., 1991). Elements<br />

that plot above the isocon have been enriched during alteration,<br />

whereas elements that plot below the reference isocon are<br />

depleted with respect to the chosen reference frame (cf. Grant,<br />

1986). The isocon diagram of the outer quartz–chlorite–calcite alteration<br />

zone vs. fresh host rocks shows enrichments in most of the elements,<br />

except K, Na, Co, Zn and Ba, corresponding to chlorite–calcite<br />

alteration. Gains in S, Cu, Pb, Ag and Au can be attributed to sulfide<br />

mineralization in the outer parts of the shear zones. The inner<br />

quartz–sericite–pyrite alteration zone is characterized by gains in K,<br />

S, Cu, Sb, As Ag and Au vs. relative to the fresh host rocks (Fig. 4).<br />

ppm Sulfidized granophyric dikes Au-quartz lodes<br />

be-xx be37 be38 be40 be41 be-18 be-12 kh-21 kh-33 kh-7 kh-16 be-51 be-35 be-34<br />

Au 0.2 0.9 1.7 2.1 1.5 3.2 6.2 5.5 4.3 2.1 4.9 6.6 5.9 4.8<br />

Ag b0.3 b0.3 0.7 0.4 0.6 1.8 2.6 2.3 1.7 0.7 3.9 4.5 3.2 2.6<br />

Cu 167 264 529 333 308 557 711 591 521 272 478 824 630 551<br />

Pb 13 45 56 32 15 16 11 5 9 8 14 12 5 8<br />

Zn 104 276 112 38 65 113 133 89 228 112 203 211 109 75<br />

Ba 630 370 277 300 207 207 110 58 b50 112 b50 b50 b50 122<br />

Sr 165 208 89 232 93 51 33 26 14 63 12 11 8 27<br />

As 2 b0.5 8 3 b0.5 b0.5 8 6 b0.5 9 7 5 4 8<br />

Ni 16 23 28 6 4 b1 8 2 9 11 b1 3 10 12<br />

Co 10 11 21 b1 b1 2.0 b1 3.0 b1 2.0 b1 b1 2.0 2.3<br />

V 51 38 31 35 24 3 b2 b2 b2 b2 2 3 6 4<br />

Cr 11 9 18 23 6 b2 b2 b2 b2 b2 b2 b2 b2 b2<br />

Sb 0.7 1 2 b0.1 1.2 0.9 1.8 2.3 b0.1 1.6 b0.1 2.4 1.8 1.8<br />

b0.3 = below detection limit (0.3).<br />

87


88 B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

Table 2b<br />

Correlation coefficients (r) of Au and most trace elements in the ore lodes and sulfidized granophyric dikes.<br />

Au Ag Cu Pb As Ni Co V Zn Ba Sr<br />

Au 1.000<br />

Ag 0.920 1.000<br />

Cu 0.899 0.834 1.000<br />

Pb −0.558 −0.506 −0.268 1.000<br />

As 0.387 0.367 0.328 −0.094 1.000<br />

Ni −0.524 −0.548 −0.334 0.694 0.092 1.000<br />

Co −0.543 −0.486 −0.268 0.767 0.080 0.869 1.000<br />

V −0.811 −0.730 −0.683 0.630 −0.381 0.584 0.606 1.000<br />

Zn 0.135 0.231 0.140 0.157 −0.216 0.157 0.077 −0.113 1.000<br />

Ba −0.833 −0.779 −0.711 0.474 −0.331 0.520 0.572 0.907 −0.191 1.000<br />

Sr −0.791 −0.793 −0.728 0.620 −0.386 0.431 0.376 0.874 −0.117 0.815 1.000<br />

Bold data indicates significant positive correlation.<br />

In order to quantify the volume (and mass) changes by hydrothermal<br />

alteration, Grant's (1986) equations:<br />

ΔV ¼ ð1=SÞ ρ a =ρ o –1 100; ð1Þ<br />

ΔM ¼ ðð1=SÞ–1Þ 100; ð2Þ<br />

where S is the gradient of the isocon, ΔV and ΔM are losses or gains in<br />

percent and ρ a /ρ o is the ratio for the specific density of altered rock<br />

and unaltered protolith, respectively. Two isocons were defined for<br />

each plot using the 1σ error bars of the immobile elements yielding<br />

a minimum and maximum gradient, defining the minimum and maximum<br />

values for mass and volume changes. From these, the mean<br />

values and error bars were calculated.<br />

Based on the mass balance calculations (Table 3), the inner, silicified–sericitized–pyritized<br />

wallrocks have experienced considerable<br />

volume and mass changes, compared to the outer, silicified–chloritized–carbonatized<br />

wallrocks (Fig. 4a,b). Addition of SiO2, K2O, L.O.I.,<br />

Table 3<br />

Geochemical data and results of mass balance calculations of hydrothermally altered sheared metavolcanic/volcaniclastic rocks from Wadi El <strong>Beida</strong>−Wadi Khashab area.<br />

Least-altered host rocks Quartz (Qtz-chl-carb) alteration Inner (Qtz-Ser-Py) alteration<br />

n 4 6 7<br />

average 1σ average 1σ Gain/loss 1σ average 1σ Gain/loss 1σ<br />

g/100 g<br />

SiO2 wt.% 62.78 0.67 63.01 0.90 2.95 2.18 69.14 1.24 14.34 3.01<br />

a<br />

TiO2 wt.% 0.48 0.06 0.46 0.03 0.00 0.01 0.42 0.07 −0.02 0.01<br />

a<br />

Al2O3 wt.% 13.54 0.55 12.79 0.49 −0.20 0.33 11.87 0.53 −0.30 0.49<br />

t<br />

Fe2O3 wt.% 5.82 0.44 6.77 0.73 1.24 0.23 6.92 0.55 1.92 0.30<br />

MnO wt.% 0.07 0.05 0.23 0.06 0.17 0.01 0.19 0.08 0.14 0.01<br />

MgO wt.% 4.25 0.78 4.97 0.52 0.93 0.17 0.93 0.49 −3.21 0.04<br />

CaO wt.% 4.28 0.76 6.46 0.41 2.46 0.21 2.01 0.53 −2.04 0.09<br />

Na2O wt.% 2.73 0.37 2.29 0.12 0.39 0.01 1.65 0.42 −0.89 0.03<br />

K2O wt.% 3.18 0.58 2.16 0.45 −0.93 0.05 5.14 1.19 2.55 0.22<br />

a<br />

P2O5 wt.% 0.29 0.03 0.28 0.05 0.00 0.01 0.26 0.07 0.00 0.01<br />

L.O.I. wt.% 1.85 0.47 2.66 0.28 0.92 0.10 2.25 0.62 0.66 0.10<br />

S wt.% 0.08 0.01 0.87 0.43 0.83 0.02 1.03 0.26 1.07 0.04<br />

g/1000 kg<br />

Au ppm – – 0.51 0.17 0.53 0.13 2.23 0.96 2.37 0.55<br />

Ag ppm – – 0.2 0.3 0.21 0.01 1.1 1.2 1.23 0.05<br />

Cu ppm 29.8 7.2 94.7 82 66.4 31.5 143.3 85.3 108.5 40.89<br />

Cd ppm 4.0 – – – −4.0 – – – −4.0 –<br />

Mo ppm – – 12.3 20 12.83 0.41 2.5 1.9 2.79 0.11<br />

Pb ppm 42.3 5.1 20.2 30 −21.23 0.67 58.6 33.7 23.06 2.56<br />

Ni ppm 25.8 4.8 22 18 −2.85 0.73 26.3 18.2 3.53 1.15<br />

Zn ppm 94 18.6 61.3 112 −30.1 2.0 83.9 58.4 −0.42 3.66<br />

As ppm – – 12.1 9.5 12.62 0.40 5.6 3.81 6.25 0.24<br />

Ba ppm 591 96.7 627 134 158.2 4.23 61.2 13.1 −522.7 2.67<br />

Co ppm 23 3.67 21.3 32 −0.78 0.71 14.8 5.80 −6.49 0.65<br />

Cr ppm 40.8 9.4 34 193 0.1 4.1 46.0 173.4 10.84 10.7<br />

Eu ppm 0.7 0.00 1.1 1.3 0.45 0.04 0.7 0.60 0.08 0.03<br />

Sb ppm 0.5 0.00 1.3 2.2 1.29 0.04 1.4 0.7 1.06 0.06<br />

Sc ppm 17.8 1.48 21.1 2.3 5.25 0.07 9.1 8.08 −7.65 0.40<br />

Sr ppm 170.3 41.8 310.8 258 208.3 7.01 81.6 50.8 79.28 3.56<br />

Th ppm 0.63 0.21 0.80 0.2 0.20 0.03 0.53 0.60 −0.04 0.02<br />

V ppm 129 19.7 107.6 41 −16.7 21.25 92.6 69.2 −25.73 4.04<br />

W ppm – – 4 4.17 0.13 5 – 5.58 0.22<br />

Y a<br />

ppm 26 4.5 24.8 3.7 −0.13 0.00 23.1 3.3 −0.22 0.00<br />

La ppm 6.2 2.1 8.6 16 2.77 0.27 7.3 2.73 1.94 0.16<br />

Density gr/cm3 2.80 2.83 2.86<br />

n = number of analyzed samples.<br />

– Below detection limit.<br />

Gains/losses are given as g/100 g for major elements, as g/10 3 kg for trace elements, and as g/10 6 kg for Au.<br />

a<br />

Immobile elements chosen for the definition of the isocon.


Fe2O3* and S, and loss of CaO, MgO, and Na2O characterize the major<br />

chemical changes in the inner wallrocks. On the other hand, the silicified–chloritized–carbonatized<br />

rocks showed addition of most major<br />

and trace elements and depletion in K2O, V and Co (Table 3). Addition<br />

of SiO2, K2O, and S in the altered wallrocks that are relative to the<br />

fresh (unaltered) host rocks coincides with the pervasive silicification,<br />

sericitization and pyritization (Table 3). As chlorite and carbonate<br />

replaced most of the protolith mineralogy, addition of CaO, Fe 2O 3*<br />

and MgO occurred at the expense of K2O and Na2O in the carbonatized–chloritized<br />

wallrocks. The uniformity of element mobility and<br />

the relatively low error of the average values indicate that geochemical<br />

changes due to wall rock alteration were relatively homogenous<br />

on the deposit scale, with mass and volume increase towards the<br />

ore bodies.<br />

B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

Fig. 3. Photomicrographs of microstructures in quartz veins from Wadi El <strong>Beida</strong>–Wadi Khashab area (crossed polarized light). (a) Ribbon-dominant vein specimen of preferably<br />

oriented large quartz crystals (qtz) and interstitial carbonate (cc), (b) Sheared large quartz ribbons. Fine-grained recrystallized grains developed along grain boundaries and<br />

along the shear planes indicative of grain-boundary migration/recrystallization in quartz veins, (c) Large quartz ribbons with serrate boundaries filled with sulfide minerals<br />

(dark) and cut by conjugate fracture sets locally with fine subgrains, (d) Bulging recrystallization of large quartz crystals, where grain boundaries are filled with sulfide minerals<br />

and wallrock material. Primary isolated and clustered fluid inclusions are abundant in the large quartz crystals, (e) Intensely re-crystallized quartz ribbons with fine-grained polygonal<br />

subgrains defining the C and S-planes. The gray quartz along S–C planes was likely derived by progressive resorption and recrystallization of the comb and buck quartz.<br />

Microscopic sulfide inclusions are ubiquitous, possibly contributing to the gray color of this quartz type, (f) texturally zoned extensional (syntaxial) veinlet with internal elongated<br />

blocky microstructures cutting the sheared mineralized quartz veins.<br />

6. Ore minerals<br />

Disseminated sulfides are common in the quartz veins, granophyric<br />

dikes and altered host rocks in shear zones. Principal ore minerals include<br />

pyrite, chalcopyrite, chalcocite, covellite, marcasite, and gold. Pyrite<br />

is ubiquitous, commonly as discrete and aggregated grains of<br />

variable size (from specks up to 3 mm-across). Inclusions of chalcopyrite,<br />

sphalerite, pyrrhotite and gold, less than 50 μm-across are common<br />

in many of the large pyrite crystals in the mineralized quartz veins and<br />

granophyric dikes (Fig. 5a,b). Pyrite is partly overgrown by chalcopyrite<br />

and/or chalcocite (Fig. 5c,d). Alteration of pyrite to goethite is common.<br />

Electron microprobe analyses (Table 4) reveal the presence of traces of<br />

Cu and Ni in pyrite and marcasite (up to 0.12 and 0.16, respectively).<br />

The analyzed pyrite is As-poor or barren and oscillatory zoning of As<br />

89


90 B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

Fig. 4. Isocon diagram after Grant (1986) for: (a) the outer quartz–chlorite–carbonate<br />

alteration, and (b) the inner quartz–sericite–pyrite alteration, plotted against the unaltered<br />

host rock metavolcanics.<br />

in pyrite in the back-scattered electron (BSE) images is totally absent.<br />

Chalcopyrite forms patchy intergrowths with sphalerite and pyrite. Locally,<br />

chalcopyrite is variably replaced by chalcocite, covellite and<br />

digenite (Fig. 5d). A few grains of chalcocite were observed associated<br />

with chalcopyrite. Also, chalcocite intergrows with pyrite and marcasite<br />

in many samples. Sphalerite occurs as dark gray (Fe-rich), irregular and<br />

anhedral masses ranging in size from tiny specks to crystals more than<br />

1 mm across. The mineral is intimately intergrown with pyrite, chalcopyrite,<br />

and pyrrhotite. Large sphalerite patches show abundant chalcopyrite<br />

blebs ‘chalcopyrite disease’. Marcasite is locally associated with<br />

pyrite as irregularly-formed discrete grains or fine-grained aggregates.<br />

Marcasite, locally colloform, replaces pyrite and chalcopyrite in many<br />

samples (Fig. 5e,f). Gold forms small inclusions in pyrite, or disseminated<br />

in fine-grained quartz, sericite and calcite matrix in the intensively<br />

altered wallrocks. Gold occurs also as thin wires rimming pyrite crystals<br />

(Fig. 5g). The analyzed gold grains show, generally b10 wt.% Ag, and<br />

consistently high fineness levels (1000 *Au/Au+Ag>900). Hematite<br />

forms veinlets commonly associated with quartz (quartz–hematite<br />

stockwork) or occurs as foliation-parallel laminations of acicular laths<br />

in the basaltic andesite host rocks. The hematite laminations contain<br />

abundant relict magnetite inclusions, likely of pre-mineralization/alteration<br />

origin. Hematite associated with micro-crystalline quartz in<br />

quartz veins, however, contains no magnetite inclusions. Discrete gold<br />

inclusions can be found in the hematite veins where carbonate and supergene<br />

Cu-hydroxides are common (Fig. 5h).<br />

Goethite and malachite are supergene phases associated with sulfide<br />

minerals in voids or along micro-fractures. Ramadan and Kontny (2004)<br />

described a complex alteration of chalcopyrite to anilite (Cu 6.7Fe 0.3S 4)<br />

and covellite (CuS), and finally to delafossite (CuFeO2) basedonSEM<br />

and electron microprobe studies of the samples collected from El<br />

<strong>Beida</strong> mineralization.<br />

Analysis by LA-ICP-MS of several pyrite grains (Table 5a) revealed<br />

the presence of trace amounts of Au (8.5–120.1 ppm), Ag (2.2–<br />

57.8 ppm), As (42–468 ppm), Sb (up to 4.7 ppm), Ni (66–415 ppm),<br />

Co (218–701 ppm), Cu (120–6442 ppm), and Pb (26–121 ppm). Au<br />

values positively correlate with Ag and Cu, whereas, no correlation was<br />

observed between Au and As. Pyrrhotite contains traces of Au (26–<br />

35 ppm), Ag (2–27 ppm), As (23–45 ppm), Ni (29–1009 ppm), Co (54–<br />

339 ppm), Cu (47–129 ppm), Zn (24–31 ppm). Au positively correlates<br />

with Ag, and no distinct correlation has been observed between Au and<br />

the other elements. LA-ICP-MS measurements on marcasite reveal the<br />

presence of Au (up to 36 ppm), Ag (up to 3.3 ppm), As (up to 22 ppm),<br />

Ni (up to 34 ppm), Co (11–79 ppm), Cu (up to 263 ppm). Chalcopyrite<br />

contains traces of Au (9.3–18.4 ppm) and Ag (1.5–4.2 ppm) that do not<br />

show correlation with any other trace element (i.e., 357–611 ppm Pb,<br />

1345–899 ppm Zn; Table 5b).<br />

7. Fluid inclusions<br />

Upon a petrographic study, thirteen double-polished thick sections<br />

(~150 μm thick) that represent all types of the mineralized<br />

quartz veins were selected for microthermometric measurements.<br />

Selection of these samples was based on the availability of workable<br />

inclusions in synchronous groups (cf. Touret, 2001). Solitary inclusions<br />

and those that occur in clusters, or along intra-granular planar<br />

arrays (pseudosecondary inclusions) in less deformed quartz grains<br />

were considered for detailed petrography and microthermometric<br />

studies. Fluid inclusions on trails crossing grain boundaries are interpreted<br />

as secondary inclusions, and are discarded from further investigation.<br />

From each fluid inclusion assemblage (FIA: Goldstein<br />

and Reynolds, 1994), more than five representative inclusions with<br />

substantial size are measured for better reliability of the microthermometric<br />

results.<br />

Fluid inclusions in the mineralized quartz veins belong to the H 2O–<br />

CO2–NaCl±CH4(±N2) system. On the basis of their phase contents at<br />

ambient laboratory temperature, the observed fluid inclusions are<br />

grouped into three types, namely: type-I: bi-phase aqueous inclusions<br />

(liquid+vapor aqueous), type-II: bi-phase aqueous–carbonic (liquid<br />

CO2 bubble surrounded by liquid H2O), and type-III tri-phase, carbonicrich<br />

aqueous–carbonic inclusions (liquid H 2O–liquid CO 2–vapor CO 2),<br />

generally with no daughter minerals (Fig. 6a–d). The liquid CO2 bubbles<br />

in type-II inclusions have typically dark boundaries (meniscus) against<br />

the liquid H2O fraction, and show a characteristic phase change on<br />

both cooling and warming runs. Most of the observed fluid inclusions<br />

are 2–10 μm across, but a few ones measure up to 20 μm-across. The<br />

type-I inclusions are relatively large in size compared to other inclusion<br />

types in a single trail. Most of the type-I and type-II inclusions are elliptical<br />

or oval in shape, whereas inclusions of type-III commonly exhibit a<br />

negative crystal habit. The liquid to vapor ratios are used to discriminate<br />

between type-I and type-II inclusions. Whereas, proportions of the<br />

aqueous and carbonic phases are used to classify inclusions of type-II<br />

and type-III. The type-III inclusions show consistently small aqueous<br />

fraction (20–30%).<br />

7.1. Microthermometric results<br />

Microthermometric data of measured inclusions are summarized in<br />

Table 6 and Fig. 7. Composition and density of H2O–CO2 phases were estimated<br />

using the H 2O–CO 2–CH 4 ternary system (Jacobs and Kerrick,<br />

1981). Salinities of the aqueous inclusions (type-I) were calculated<br />

from the final ice melting temperatures (T mice) using the equation of<br />

Bodnar (1993). Whereas, salinities of the aqueous–carbonic inclusions<br />

(type-II and type-III) were based on the final melting temperatures of<br />

clathrate (in the presence of liquid CO2) using the computer program


package “clathrates” of Bakker (1997) and Bakker and Brown (2003).<br />

The equations of state of Brown and Lamb (1989) have been used to calculate<br />

the isochores for selected aqueous–carbonic inclusions.<br />

Total freezing was achieved by cooling down to −110 °C. Because of<br />

the difficulty of viewing phase changes in small inclusions, clathrate and<br />

ice melting measurements could not be performed on all inclusions for<br />

which T htotwas measured. Melting of CO 2 (T mCO2)intheaqueous–<br />

carbonic inclusions (type-II and type-III) occurred at −56.6 to<br />

−61.9 °C, but commonly between −57 and −59 °C (Fig. 7a). Melting<br />

of ice (Tm ice) in type-I inclusions occurred between −1.5 and −8.9 °C,<br />

but mostly clustered between −4 and−7 °C. In the aqueous-carbonic<br />

B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

Fig. 5. Photomicrographs of the ore minerals of Wadi El <strong>Beida</strong>–Wadi Khashab mineralization (a) irregular relict patches of pyrrhotite (po) and sphalerite (sph) in a large pyrite (py)<br />

grain, (b) Disseminated gold blebs in carbonate horizon and associated with fine grains of pyrite (py), (C) Chalcopyrite (cpy) intergrown with pyrite and replaced by rhythmic covellite<br />

(cov) and chalcocite (cc), (D) Chalcopyrite intergrown with sphalerite (sph) and replaced by chalcocite (cc) and digenite (dig). Note development of malachite (mal) into a<br />

void, (e) collorform marcasite pseudomorphs pyrite (Backscattered Electron image, BSE), (f) Marcasite replacing pyrite–chalcopyrite intergrowth. (BSE), (g) Gold grain associated<br />

with a subhedral pyrite crystal. (BSE), (h) Malachite (mal) and calcite grains and veins cutting hematite lamination in the host metavolcanic rocks. Notice the dispersed gold speck<br />

(Au) in hematite (close up view given). (BSE).<br />

inclusions (type-II and type-III), clathrate melting varied from 8.2 to<br />

2.7 °C, but mostly between 7 and 5.5 °C (Fig. 7b). Inclusions of type-II<br />

showed clathrate melting at a wider temperature range (3.6–8.2 °C)<br />

relative to those of type-III (6.2–8.4 °C). Partial homogenization of CO2<br />

in type II and type III occurred commonly to liquid, but a few inclusions<br />

of type III displayed vapor bubble expansion and homogenization into<br />

the vapor phase (at 9.8–17.2 °C). A few two-phase aqueous inclusions<br />

exhibited H2O–CO2 clathrate melting upon warning, although no separate<br />

CO 2 phase is present. The presence of clathrate indicates that these<br />

inclusions contain a small amount of CO2, though melting points are<br />

barely observed (measured at ~7 °C in two inclusions). In the type II<br />

91


92 B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

Table 4<br />

Representative EMPA data of sulfide minerals disseminated quartz veins from Wadi El <strong>Beida</strong>–Wadi Khashab area.<br />

Wt.% Pyrite Marcasite<br />

Fe 46.01 46.25 45.86 44.90 46.53 46.20 46.42 45.86 45.60 46.17 46.10 46.09 44.49 44.85 44.91<br />

S 52.66 52.92 52.66 52.83 52.23 52.80 52.71 52.73 52.39 52.71 52.62 52.52 53.75 53.54 53.81<br />

As – – – – – – – – – – – – – – –<br />

Cu – 0.07 0.09 0.04 – – 0.12 0.05 – – 0.09 – 0.09 – 0.11<br />

Ni 0.04 0.04 – 0.07 – 0.07 0.06 – 0.04 0.16 0.07 0.08 0.12 – 0.09<br />

Sum 98.71 99.21 98.52 97.94 98.76 99.07 99.31 98.64 98.03 99.04 98.88 98.69 98.41 98.39 98.92<br />

Chalcopyrite Sphalerite<br />

Fe 30.83 29.72 30.26 30.92 30.60 30.77 29.93 29.98 30.74 7.39 6.95 6.08 6.81 6.61 6.50<br />

S 33.86 34.33 34.02 35.13 34.25 34.79 35.27 34.29 34.84 34.04 33.88 34.54 34.53 34.19 34.29<br />

Cu 34.30 34.94 34.17 33.52 33.90 33.07 33.81 33.95 33.55<br />

Zn 0.03 0.05 0.15 0.11 0.15 0.03 0.23 0.11 0.03 57.64 58.29 58.79 57.64 58.44 58.49<br />

Sum 99.02 99.04 98.6 99.68 99.25 98.90 99.24 98.33 99.16 99.1 99.12 99.41 98.98 99.24 99.32<br />

Chalcocite Covellite Digenite Pyrrhotite Gold<br />

Fe 0.32 0.26 0.41 0.81 0.43 0.51 0.29 0.12 0.15 60.89 60.32 60.38<br />

S 21.32 22.09 22.19 32.17 31.98 32.24 22.15 23.23 22.87 38.65 38.42 37.96<br />

As 0.13 0.21 0.25<br />

Ni 0.16 0.12 0.09<br />

Cu 78.45 77.08 77.14 66.71 66.23 67.01 76.46 76.48 76.55 – – –<br />

Ag 0.12 – – 5.69 7.27 3.11<br />

Au 0.09 0.16 0.12 93.89 92.31 96.65<br />

Sum 100.16 99.43 99.74 99.73 98.64 99.79 98.9 99.83 99.57 100. 4 99.23 98.8 99.58 99.58 99.76<br />

– Below detection limit.<br />

Blank cells refer to unmeasured elements.<br />

aqueous–carbonic inclusions, Th CO2ranges between 29.2 and 8.7 °C,<br />

whereas type III inclusions showed T h CO2 between 3.4 and 20.2 °C<br />

(Fig. 7c). The total homogenization of all inclusion types (I, II and II) occurred<br />

commonly into liquid. In a few inclusions of type-III, homogenization<br />

does not involve the disappearance of the vapor bubble into the<br />

liquid phase, but the meniscus disappeared at higher temperatures and<br />

decrepitation is common. Decrepitation of the large H2O–CO2 fluid inclusions<br />

(>15um-long) at critical temperatures during measurement<br />

of total homogenization temperatures (Th total) was sporadic. The decrepitation<br />

temperatures varied from sample to sample, but mostly initiated<br />

around 330 °C. The undecrepitated aqueous–carbonic (type-II<br />

and type-III) and aqueous (type-I) inclusions showed comparable<br />

Table 5a<br />

Representative LA-ICP-MS measurements on some sulfides disseminated in the auriferous<br />

lodes in W. El-<strong>Beida</strong>-Khashab area.<br />

Spot Au Ag Sb Te Pb Bi As Ni Co Cu Zn<br />

be-Py01 – – – – 13 – 95 – 103 – –<br />

be-Py03 22.1 8.5 2.9 – 50 – 206 66 240 1022 –<br />

be-Py01 34.9 22.6 4.7 – – – – – 701 1225 48.1<br />

be-Py07 – – – – 26 – 468 103 – 335 –<br />

be-Py01 9.8 2.2 2.7 – 42 – – 190 308 120 –<br />

kh-Py05 120.1 57.8 – – 40 – 295 415 322 6442 43.5<br />

kh-Py06 17.2 36.4 2.6 – 75 – 42 – – 3228 –<br />

kh-Py09 8.5 11.7 – – 121 – 55 762 218 1119 –<br />

kh-Py11 24.6 13.6 – – – – – 297 – 139 23.9<br />

be-Po01 34.8 26.7 – – – – 45 1009 339 47 –<br />

be-Po03 25.7 5.9 – – 5 – 29 29 184 – 24.4<br />

be-Po15 30.5 24.0 – – – – – 111 254 – –<br />

be-Po07 – 7.9 – – – – 23 – 54 – 31.1<br />

be-Po01 28.0 16.3 – – – – – 37 95 129 –<br />

kh-Po05 – 1.8 – – – – – 109 58 – –<br />

kh-Mc6 – – – – 10 1.1 22 – 79 – –<br />

be-Mc9 20.6 2.4 – – 7 1.0 – 34 – 51 –<br />

kh-Mc2 12.9 1.6 2.7 – – – 16 – 11 – –<br />

be-Mc8 35.8 3.3 – – – – – 7 19 263 –<br />

kh-Cpy02 18.4 3.2 – – 602 – – 112 12 >>>> 1212<br />

kh-Cpy04 11.2 2.9 – – 441 – 24 24 – >>>> 903<br />

be-Cpy8 9.3 4.2 2.1 – 611 – – 53 7 >>>> 1345<br />

be-Cpy11 9.7 1.5 – – 357 – – 38 – >>>> 899<br />

Det. limit 0.1 0.6 1.8 16 2 0.2 10 6 1.6 17 20.8<br />

Py = pyrite, Po = pyrrhotite, Mc = marcasite, Cpy = chalcopyrite.<br />

total homogenization temperatures (Th total), broadly established on a<br />

considerable number of inclusions between 258 and 343 °C (Fig. 7d).<br />

However, the median homogenization temperatures from each cluster<br />

occur in relatively small ranges (mostly less than 30 °C).<br />

7.2. Fluid inclusions composition<br />

Data of fluid inclusion microthermometry indicate that chief components<br />

of the mineralizing fluids are H 2O and CO 2. Based on the ice<br />

melting temperatures, the aqueous inclusions (type-I) have salinities<br />

of 2.6–12.7 wt.% eq. NaCl. Based on the clathrate melting temperatures,<br />

type-II and type-III fluid inclusions have salinities between 2<br />

and 8.5 wt.% eq. NaCl, (Fig. 7e). Typically, the vapor-dominant inclusions<br />

(type-III) have lower salinities compared to the type-II inclusions.<br />

The aqueous inclusions (type-I) have a typical composition of 0.96–<br />

0.99 mol% H2O+0.04–0.01 mol% NaCl. Based on the Tm CO 2 and Th CO 2<br />

data of type-II and type-III inclusions, CO 2 density ranges between<br />

0.28 and 0.87 g/cm 3 . The bulk composition of measured fluid inclusions<br />

was estimated on the basis of volume percent of CO 2 and H 2O phases<br />

“degree of fill” at room temperature (Roedder and Bodnar, 1980). The<br />

volume percent of CO 2 in type-II inclusions ranges from 30 to 60%,<br />

which corresponds to XCO 2 (mole %) between 10 and 30. Whereas,<br />

most type-III inclusions have 70–90 vol.% CO 2, corresponding to 19–<br />

38 mol% of CO2. Melting below the typical melting temperature of<br />

pure CO 2 (−56.6 °C) implies the presence of dissolved gasses other<br />

than CO2. IfCH4 is assumed, approximately 6–14 mol% CH4 in the<br />

CO 2-bearing inclusions is estimated by applying the graphical method<br />

of mole composition of CO2–CH4 mixtures (Shepherd et al., 1985).<br />

7.3. Fluid mixing versus immiscibility<br />

The highly variable CO 2/H 2O ratios have been used as evidence of<br />

phase separation and heterogeneous trapping (Ramboz et al., 1982),<br />

fluid mixing (Anderson et al., 1992; Cassidy and Bennett, 1993;<br />

Pichavant et al., 1982), or post-entrapment modifications with variable<br />

loss of H 2O(Crawford and Hollister, 1986). The regular morphology<br />

(i.e., absence of necking-down) of the measured fluid inclusions<br />

and absence of vertical tendencies on the salinity–T h total diagram<br />

(Fig. 7f) and/or density variation from core to rim in a single quartz


grain (e.g., Huizenga and Touret, 1999) discards significant postentrapment<br />

modifications in the measured inclusions. Variable CO 2/<br />

H2O ratios in the measured inclusions may, therefore, be attributed<br />

to immiscibility or heterogeneous entrapment, likely attained by mixing<br />

or unmixing (phase separation).<br />

Slight changes in ambient physical conditions (e.g., pressure) during<br />

the development of a shear zone can result in separation of CO2<br />

and H 2O phases from a parent, homogeneous aqueous–carbonic<br />

fluid. Trapping of immiscible fluids will result in coexisting vaporrich<br />

inclusions, and liquid-rich inclusions that show opposite modes<br />

of total homogenization over the same temperature range (e.g.,<br />

Ramboz et al., 1982). This criterion is not met by the measured inclusions,<br />

where both liquid-rich and vapor-rich CO2-bearing fluid inclusions<br />

homogenize commonly to the liquid phase over a wide range of<br />

temperature, and without an obvious correlation between L:V ratio.<br />

Therefore, type-II and type-III inclusions cannot represent separated<br />

phases by immiscibility. Instead, heterogeneous trapping of two<br />

fluids is another possible mechanism that can result in preservation<br />

of CO2-rich fluid inclusions. Ramboz et al. (1982) outlined four criteria<br />

for determination of heterogeneous trapping of two fluids in<br />

the same inclusion: (1) simultaneous trapping of inclusions, (2) no<br />

evidence of leakage or necking, (3) scattered degree of fill, T h tot and<br />

bulk compositions, and (4) Th tot histogram that is non-symmetrical<br />

B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

Table 5b<br />

Correlation coefficients (r) between gold and other trace elements in sulfide minerals disseminated in quartz veins.<br />

Au Ag Sb Pb As Ni Co Cu Zn<br />

Au 1.000<br />

Ag 0.800 1.000<br />

Sb −0.078 0.112 1.000<br />

Pb −0.208 −0.217 −0.019 1.000<br />

As 0.319 0.222 0.075 −0.165 1.000<br />

Ni 0.253 0.394 −0.228 −0.097 0.120 1.000<br />

Co 0.391 0.463 0.486 −0.319 0.040 0.327 1.000<br />

Cu 0.788 0.847 0.146 −0.123 0.446 0.201 0.298 1.000<br />

Zn −0.190 −0.241 −0.033 0.983 −0.207 −0.157 −0.318 −0.177 1.000<br />

and flattened. In the present study, type-II and type-III inclusions<br />

meet most of these four criteria. The type-I inclusions are mostly<br />

aqueous-only, but a few of them contain non-virtual fraction of CO2.<br />

This may imply incomplete mixing of originally aqueous and carbonic<br />

fluids. Linear “mixing trends” on Th total versus eq. wt.% NaCl plot are<br />

commonly used to test mixing of two fluids (Fig. 7f; Shepherd et al.,<br />

1985). Heterogeneous mechanical trapping of varying proportions<br />

of the two end-member fluids (i.e., H 2O and CO 2±CH 4) is likely to result<br />

in poorly defined mixing trends (Shepherd et al., 1985), so this<br />

test may constrain the measured type-II and type-III inclusions. It is<br />

therefore, concluded that the investigated fluid inclusions are best<br />

explained by heterogeneous trapping of fluids during gold mineralization<br />

and vein growth. While type-II represent mixed aqueous and<br />

carbonic fluids, fluid mixing (miscibility) in the mineralized quartz<br />

veins might have been incomplete or succeeded by partial phase separation<br />

to generate the aqueous-only (type-I) and carbonic-dominant<br />

(type-III) fluid inclusions, due to pressure fluctuation throughout<br />

evolution of the shear zones.<br />

7.4. Conditions of gold deposition<br />

The measured fluid inclusions are considered primary or pseudosecondary<br />

in the sense of Roedder (1984). Combination of some features<br />

Fig. 6. Photomicrographs showing different types, shapes, and compositions of fluid inclusions in the mineralized quartz lodes at room temperature. (a) and (b) Mixture of H 2O-rich<br />

(type I) and CO 2-bearing (type-II, type-III) fluid inclusions, (c) and (d) variable degree of filling in three-phase CO 2-rich inclusions (type-III).<br />

93


94 B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

Table 6<br />

Characteristics and microthermometric data of the different fluid inclusion types in mineralized quartz veins.<br />

Type I: H2O–NaCl inclusions (aqueous) Type II: H2O–NaCl–CO2±CH4±N2 inclusions<br />

(aqueous-rich)<br />

Abundance: ~25% of the bulk fluid inclusion population<br />

Mode of occurrence: isolated, clustered or trail-bound<br />

in association other incl. types<br />

Shape: sub-rounded, oval, elongate<br />

Size: ~4–20 μm<br />

Occurrence: isolated, in clusters, or on trails<br />

Vol.% H 2O vapor=5–20<br />

T m ice =−1.5 to −8.9 °C<br />

T h total=262 to 337 °C<br />

Salinity =2.7–12.7 wt.% NaCl eq.<br />

dH 2 O (g/cm 3 )=1.01–1.08 g/cm 3<br />

The used equations of state are given in the text.<br />

including: (i) the simple ore mineralogy, (ii) occurrence of gold specks<br />

and sulfide grains in cores of the less-deformed quartz crystals, and<br />

(iii) absence of severe post-entrapment modifications, suggests that<br />

gold introduction was coeval with formation of the quartz veins, typical<br />

for shear-related orogenic gold quartz vein systems (e.g., Dubé and<br />

Gosselin, 2007 and references therein). Association of the three inclusion<br />

types in primary/pseudosecondary sites in the mineralized quartz<br />

veins suggests gold deposition through fluid mixing (miscibility). The<br />

aqueous-rich and carbonic-rich fluid inclusions (type-II and type-III)<br />

have, therefore, inherently variable bulk compositions and densities.<br />

The entire range of Th totalof the aqueous (type-I) and aqueous–<br />

carbonic inclusions (type-II and type-III) is comparable with temperature<br />

conditions inferred from quartz textures of the mineralized lodes<br />

(e.g., Stipp et al., 2002). The total homogenization temperatures of the<br />

measured inclusions (258–343 °C), correspond to 0.8 to 2.3 kbars<br />

using isochores for the aqueous–carbonic inclusions. The calculations<br />

of pressure assuming that entrapment of fluids occurred under lithostatic<br />

conditions indicate depths of 3 and 9 km. Variations in the estimated<br />

P–T conditions could have been the result of exhumation and<br />

pressure fluctuation when the fluid straddled the brittle–ductile transition.<br />

The brittle–ductile deformation is consistent with vein-quartz textures<br />

that vary from brecciated to ribboned.<br />

8. Discussion and conclusions<br />

It has been established that numerous en-echelon sinistral NNWtrending<br />

Riedel shear zones in the Eastern Desert of Egypt are related<br />

to positive flower structures associated with the (620–540 Ma) Najd<br />

transcurrent shear system (e.g., Abdeen and Greiling, 2005; Fritz et al.,<br />

1996). Formation of the NNW-trending, sinistral shear zones in the<br />

Wadi El <strong>Beida</strong>–Wadi Khashab area is attributed to a regional NNW–<br />

SSE folding and transpression (D2), during which extensive shear<br />

zones and conjugate fault systems were developed. The mineralized<br />

shear zones in the study area are part of the high-angle convergent<br />

wrench structure, namely Wadi Kharit–Wadi Hodein zone (e.g.,<br />

Abdeen et al., 2008; de Wall et al., 2001; Greiling et al., 1994; <strong>Zoheir</strong>,<br />

2011). Spatial association of the investigated auriferous shear zones<br />

and regional convergent structures is typical for the orogenic, lodegold<br />

setting (e.g., Goldfarb et al., 2001; Huston et al., 2007).<br />

Within the shear zones, quartz veins, quartz network and granophyric<br />

dikes with disseminated sulfides are gold-bearing (~0.2–7ppm<br />

Abundance: ~45% of the bulk inclusions population<br />

Shape: equant, elongate, irregular<br />

Size: b2–15 μm<br />

Occurrence: isolated or on intra-granular trails<br />

DF=0.4–0.7<br />

T mCO2 =−57.3 to −62.4 °C<br />

T hCO2 =8.7 to 29.2 °C (to liquid)<br />

T m Clath=3.6 to 8.2 °C<br />

T h total=287 to 340 °C<br />

d CO2 (g/cm 3 )=0.62–0.87 g/cm 3<br />

Salinity =3.6–12.3 wt.% NaCl eq.<br />

XCO 2 (mol%)=10–30<br />

Bulk density =0.83–0.99 g/cm 3<br />

Type III: H2O–NaCl–CO2±CH4±N2 inclusions<br />

(carbonic-rich)<br />

Abundance: ~30% of the bulk inclusions population<br />

Shape: oval, oblate, negative crystal<br />

Size: b2–10 μm<br />

Occurrence: isolated, along intra-granular trails<br />

DF= b0.2–0.3<br />

T mCO2 =−56.7 to −59.6 °C<br />

T hCO2 =3.4 to 20.2 °C (into liquid)<br />

T hCO2 =9.8 to 17.2 °C (into vapor)<br />

T m Clath=5.0 to 8.9 °C<br />

T h total=301 to 342 °C<br />

dCO 2 (g/cm 3 )=0.12–0.91 g/cm 3<br />

Salinity =~2–9 wt.% NaCl eq.<br />

X CO2 (mol%)=12–77<br />

Bulk density =0.27–0.94 g/cm 3<br />

Au). The bulk gold-sulfide mineralization in the quartz lodes is confined<br />

to comb quartz mixed with ferruginated domains of sericite and carbonate.<br />

Incorporation of foliated wallrock slivers into the quartz veins<br />

and replacement of foliation seams adjacent to veins by hydrothermal<br />

minerals suggest that the mineralized quartz veins have developed coeval<br />

with formation of the shear zones (e.g., Robert and Brown, 1986).<br />

Structural control on mineralization is further evident by anisotropy<br />

of magnetic susceptibility measurements (Nano et al., 2002).<br />

Association of gold and pyrite is observed in all mineralized quartz<br />

samples. However, occurrence of gold specks in the hematite veins in<br />

the altered host rocks points toward the possible contribution of sulfidation<br />

of iron-rich rocks as a gold deposition mechanism. Presence<br />

of invisible gold in pyrrhotite and pyrite is evident from the electron<br />

microprobe data and LA-ICP-MS. The LA-ICP-MS data demonstrated<br />

that the elevated contents of As in Fe-sulfides are not coupled with<br />

an increase in their Au and Ag contents. A correlation exists between<br />

Au and Cu values in the measured pyrites. Similar to hydrothermal<br />

pyrites in orogenic and Carlin-style gold deposits (e.g., Large et al.,<br />

2009), pyrite from the in the investigated samples shows simple<br />

trace element geochemistry, including only low values of Au, As, Ni,<br />

Cu, Zn, Co and Pb. The presence of (10s–100s ppm-levels) refractory<br />

gold in sulfides disseminated in gold-quartz lodes was described in<br />

other gold deposits in the Eastern Desert of Egypt (Hilmy and<br />

Osman, 1989; <strong>Zoheir</strong>, 2008), elsewhere in important orogenic gold<br />

deposits in the greenstone belts (e.g., Hutti deposit in India, Saha<br />

and Venkatesh, 2002). Solid solution may have been responsible for<br />

invisible gold, whereas, free gold deposition is a function of remobilization,<br />

reconstitution and concentration of the earlier phase. Thus, scarcity<br />

of visible gold in the quartz lodes in Wadi El <strong>Beida</strong>–Wadi Khashab area<br />

may be explained by the relatively low contents of invisible gold in<br />

sulfides.<br />

Infiltration of gold-bearing hydrothermal fluids into the shear zones<br />

led to pervasive alteration progressed from a quartz–sericite–pyrite<br />

assemblage in the vicinity of Au-quartz lodes, outwards to a quartz–<br />

chlorite–calcite assemblage. In places, both alteration assemblages have<br />

later been partly obliterated by argillic alteration (kaolinite + clay<br />

minerals). The systematic increase in SiO 2,K 2O and 3 K/Al and depletion<br />

in Na2O is suggestive of pervasive silicification and sericitization during<br />

evolution of the alteration. While, the erratic distribution of S in the<br />

wallrocks and granophyric dikes implies that sulfidation was selectively<br />

intense in the Fe-rich rocks. Phillips and Groves (1983) argued that the


B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

Fig. 7. Histogram presentations of microthermometric data from fluid inclusions in mineralized quartz veins from Wadi El <strong>Beida</strong>–Wadi Khashab area. (a) Melting temperatures of<br />

CO 2 in CO 2-bearing inclusions type-II and type-III, (b) partial homogenization temperatures of CO 2 (T hCO2 )inCO 2-bearing inclusions, (c) clathrate melting temperatures (T m clath)in<br />

CO 2-bearing inclusions, (d) total homogenization temperatures (T h total) of all inclusion types measured, (e) calculated salinities of the three types of inclusions, (f) T h total versus<br />

salinity (eq. wt.% NaCl) plot of the fluid inclusions in the mineralized quartz veins used to test for fluid mixing. Notice the arbitrary trends of all inclusion types meet at the high<br />

temperature, low salinity area.<br />

95


96 B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

depositional mechanism for gold in lode deposits hosted by iron-rich<br />

rocks was sulfidation of the host rock. Removal of S from the ore fluid<br />

destabilizes gold-thio complexes and leads to gold deposition. The frequent<br />

association of hematite and variably oxidized Fe–Cu-sulfides<br />

(i.e. pyrite, chalcopyrite) with gold specks suggests that sulfidation<br />

was important for gold deposition at the Wadi El <strong>Beida</strong>–Wadi Khashab<br />

area. Removal of S from the ore fluid through fluid–wallrock interaction<br />

could have destabilized gold-thio complexes and contributed significantly<br />

in gold deposition (e.g., Cox et al., 1995).<br />

The distribution and compositional characteristics of the different<br />

aqueous and aqueous–carbonic fluid inclusions (types-I, II and III) are<br />

best explained by mixing of H2O- with CO2-rich fluids. The fact that<br />

most fluid inclusions homogenize to the liquid phase and have overlapping<br />

salinity values excludes the possibility that unmixing of H2O–<br />

CO 2–NaCl occurred during trapping or that significant reworking of<br />

the fluid inclusions subsequent to trapping has occurred (e.g.,<br />

Ramboz et al., 1982). Heterogeneous entrapment and fluid mixing<br />

in the investigated Au-quartz lodes occurred at depths of 3–9km<br />

(i.e. T h total vs. eq. wt.% NaCl plot, Fig. 7f). Mixing with aqueous fluid<br />

enriched the ore fluids in oxygen and decreased sulfur activity (e.g.,<br />

Loucks and Mavrogenes, 1999). Fluid mixing as a gold-deposition<br />

mechanism in lode-gold deposits is described in the Eastern Desert<br />

Au deposits (e.g., Harraz, 2000) and elsewhere in important gold deposits<br />

worldwide (e.g., Boiron et al., 2003; Olivo et al, 2006, and references<br />

therein).<br />

Unlike the genetic model proposed by Arehart (1996) for Carlintype<br />

deposits, emplacement of orogenic lode gold mineralization generally<br />

occurs at crustal depths in excess of several kilometers. This<br />

makes direct interaction of a hot and reduced fluid with a more oxidized<br />

fluid of purely meteoric origin is less likely and requires an alternative<br />

process to initiate fluid mixing, such as the fault valve<br />

mechanism (e.g., Sibson and Scott, 1998). In this model, repeated<br />

lock-up and failure of second-order faults result in fluid migration<br />

into the surrounding wallrocks during cyclic pressure built-up, followed<br />

by fault failure and reversal of the pressure head, and hydrothermal<br />

sealing of the dilational zone. During penetration of, and<br />

interaction with the wallrock, the hydrothermal auriferous fluid<br />

evolves toward more reduced conditions before reacting with more<br />

oxidized fluids within the active fault zone, causing destabilization<br />

of gold complexes and gold precipitation (e.g., Cox et al., 1995).<br />

The new geochemical and mineralogical data presented in this<br />

study combined with the extensive gossan along the shear zones, together<br />

with the available geophysical data (Sultan et al., 2009) most<br />

likely indicate potential gold ore bodies. Accordingly, this area and<br />

areas of discernible hydrothermal alteration along major convergent<br />

wrench structures in the Eastern Desert of Egypt warrant extensive<br />

assay and pitting programs for commercial evaluation.<br />

Acknowledgments<br />

This work is financed by Egyptian Science and Technology Development<br />

Fund (STDF), grant no. 150. The author highly appreciates<br />

the intimate cooperation between the STDF team and Benha<br />

University. Professors B. Lehmann (TU-Clausthal, Germany), P.<br />

Weihed (Lulea University of Technology, Sweden) and Richard<br />

Goldfarb (US Geological Survey, Denver) are thanked for allowing<br />

the different laboratory faculties during the course of undertaking<br />

this study. The manuscript was much improved by reviews of Dr.<br />

Agustin Martin-Izard (<strong>JGE</strong> Associate Editor) and two anonymous<br />

reviewers.<br />

References<br />

Abdeen, M.M., Greiling, R.O., 2005. A quantitative structural study of late Pan-African<br />

compressional deformation in the central Eastern Desert (Egypt) during Gondwana<br />

assembly. Gondwana Research 8, 457–471.<br />

Abdeen, M.M., Sadek, M.F., Greiling, R.O., 2008. Thrusting and multiple folding in the<br />

Neoproterozoic Pan-African basement of Wadi Hodein area, south Eastern Desert,<br />

Egypt. Journal of African Earth Sciences 52 (1–2), 21–29.<br />

Abd El-Wahed, M.A., 2008. Thrusting and transpressional shearing in the Pan-African<br />

nappe southwest El-Sibai core complex, Central Eastern Desert, Egypt. Journal of<br />

African Earth Sciences 50, 16–36.<br />

Abd El-Wahed, M., Kamh, S.Z., 2010. Pan-African dextral transpressive duplex and<br />

flower structure in the Central Eastern Desert of Egypt. Gondwana Research 18,<br />

315–336.<br />

Anderson, M.R., Rankin, A.H., Spiro, B., 1992. Fluid mixing in the generation of<br />

mesothermal gold mineralization in the Transvaal Sequence, Transvaal, South Africa.<br />

European Journal of Mineralogy 4, 933–948.<br />

Ashmawy, M.H., 1987. The ophiolitic mélange of the south Eastern Desert of Egypt: remote<br />

sensing fieldwork and petrographic investigations. Ph.D. Thesis, Berliner<br />

Geowissenschaft Abhefte, Berlin 84(A), 135 p.<br />

Arehart, G.B., 1996. Characteristics and origin of sediment-hosted disseminated gold<br />

deposits: a review. Ore Geology Reviews 11, 383–403.<br />

Bakker, R.J., 1997. Clathrates: computer programs to calculate fluid inclusion V–X properties<br />

using clathrate melting temperatures. Computer Geosciences 23, 1–18.<br />

Bakker, R.J., Brown, P.E., 2003. Computer modeling in fluid inclusion research. In: Samson,<br />

I., Anderson, A., Marshall, D. (Eds.), Fluid Inclusions: Analysis and Interpretation:<br />

Mineralogical Association of Canada. Short Course, 32, pp. 175–212.<br />

Bodnar, R.J., 1993. Revised equation and table for determining the freezing point depression<br />

of H 2O–NaCl solution. Geochimica Acta 57, 683–684.<br />

Bonnemaison, M., Marcoux, E., 1990. Auriferous mineralization in some shear zones. A<br />

three stage model of metallogenesis. Mineralium Deposita 25, 96–104.<br />

Boiron, M.C., Cathelineaua, M., Banks, D.A., Fourcade, S., Vallance, J., 2003. Mixing of<br />

metamorphic and surficial fluids during the uplift of the Hercynian upper crust:<br />

consequences for gold deposition. Chemical Geology 194, 119–141.<br />

Brown, P.E., Lamb, W.M., 1989. P–V–T properties of fluids in the system H 2O–CO 2–<br />

NaCl: new graphical presentations and implications for fluid inclusion studies.<br />

Geochimica et Cosmochimica Acta 53, 1209–1221.<br />

Cassidy, K.F., Bennett, J.M., 1993. Gold mineralisation at the Lady Bountiful Mine, Western<br />

Australia: an example of a granitoid-hosted Archaean lode gold deposit. Mineralium<br />

Deposita 28, 388–408.<br />

Clendenin, C.W., 1993. Small-scale positive flower structures: an ore control identifiedintheViburnumTrend,southeastMissouri,USA.MineraliumDeposita28,<br />

22–27.<br />

Cox, S.F., Sun, S.S., Etheridge, M.A., Wall, V.J., Potter, T.F., 1995. Structural and geochemical<br />

controls on the development of turbidite-hosted gold quartz vein deposits,<br />

Wattle Gully mine, central Victoria. Economic Geology 90, 1722–1746.<br />

Cox, S.F., Wall, V.J., Etheridge, M.A., Potter, T.F., 1991. Deformational and metamorphic<br />

processes in the formation of mesothermal vein-hosted gold deposits—examples<br />

from the Lachlan Fold Belt in central Victoria, Australia. Ore Geology Reviews 6,<br />

391–423.<br />

Crawford, M.L., Hollister, L.S., 1986. Metamorphic fluids: the evidence from fluid inclusions.<br />

In: Walther, J.V., Wood, B.J. (Eds.), Fluid–Rock Interactions During Metamorphism.<br />

Springer-Verlag, New York, pp. 1–35.<br />

Dabo, M., Gueye, M., Ngom, P.M., Diagne, M., 2008. Orogen-parallel tectonic transport:<br />

transpression and strain partitionning in the Mauritanides of NE Senegal. In: Ennih,<br />

N., Liégeois, J.P. (Eds.), The Boundaries of West African Craton: Geological Society,<br />

London, Spec. Pub., 297, pp. 483–497.<br />

de Wall, H., Greiling, R.O., Sadek, M.F., 2001. Post-collisional shortening in the late Pan-<br />

African Hamisana high strain zone, SE Egypt: field and magnetic fabric evidence.<br />

Precambrian Research 107 (3–4), 179–194.<br />

Dubé, B., Gosselin, P., 2007. Greenstone-hosted quartz-carbonate vein deposits. In:<br />

Goodfellow, W.D. (Ed.), Mineral Deposits of Canada: A Synthesis of Major<br />

Deposit-Types, District Metallogeny, the Evolution of Geological Provinces, and Exploration<br />

Methods: Geological Association of Canada, Mineral Deposits Division,<br />

Special Publication, 5, pp. 49–73.<br />

EGSMA, 1992. Baranis Quadrangle Map, Scale, 1:250,000. Egyptian Geological Survey<br />

and Mining Authority, Abbassia-Cairo, Egypt.<br />

El Amawy, M.A., Wetait, M.A., El Alfy, Z.S., Shweel, A.S., 2000. Geology, geochemistry<br />

and structural evolution of Wadi <strong>Beida</strong> area, south Eastern Desert, Egypt. Journal<br />

of the Geological Society of Egypt 44 (1), 65–84.<br />

Eisenlohr, B.N., Groves, D., Partington, G.A., 1989. Crustal-scale shear zones and their<br />

significance to Archaean gold mineralization in Western Australia. Mineralium<br />

Deposita 24, 1–8.<br />

Fritz, H., Dalmeyer, D.R., Wallbrecher, E., Loizenbauer, J., Hoinkes, G., Neumayr, P., Khudeir,<br />

A.A., 2002. Neoproterozoic tectonothermal evolution of the Central Eastern<br />

Desert, Egypt: a slow velocity tectonic process of core complex exhumation. Journal<br />

of African Earth Sciences 34, 543–576.<br />

Fritz, H., Wallbrecher, E., Khudeir, A.A., Abu El Ela, F., Dallmeyer, D.R., 1996. Formation<br />

of Neoproterozoic metamorphic core complexes during oblique convergence<br />

(Eastern Desert, Egypt). Journal of African Earth Sciences 23, 311–323.<br />

Goldfarb, R.J., Groves, D.I., Gardoll, S., 2001. Orogenic gold and geologic time; a global<br />

synthesis. Ore Geology Reviews 18, 1–75.<br />

Goldstein, R.H., Reynolds, T.J., 1994. Systematics of fluid inclusions in diagenetic minerals.<br />

SEPM Short Course. . 31 pp.<br />

Grant, J.A., 1986. The isocon diagram — a simple solution to Gresens' equation for metasomatic<br />

alteration. Economic Geology 81, 1976–1982.<br />

Greiling, R.O., 1987. Directions of Pan-African thrusting in the Eastern Desert of Egypt<br />

derived from lineation and strain data. In: Matheis, G., Schandelmeier, H. (Eds.),<br />

Current Research in African Earth Sciences. Balkema, Rotterdam, pp. 83–86.<br />

Greiling, R.O., Abdeen, M.M., Dardir, A.A., El Akhal, H., El Ramly, M.F., Kamal El Din,<br />

G.M., Osman, A.F., Rashwan, A.A., Rice, A.H.N., Sadek, M.F., 1994. A structural


synthesis of the Proterozoic Arabian–Nubian shield in Egypt. Geologische<br />

Rundschau 83, 484–501.<br />

Harding, T.P., Vierbuchen, R.C., Christie-Blick, N., 1985. Structural styles, plate-tectonic<br />

settings, and hydrocarbon traps of divergent (transtensional) wrench faults: the Society<br />

of Economic Paleontologists and Mineralogists Special. Publication 37, 51–77.<br />

Harraz, H.Z., 2000. A genetic model for a mesothermal Au deposit: evidence from fluid<br />

inclusions and stable isotopic studies at El Sid gold mine, Eastern Desert, Egypt.<br />

Journal of African Earth Sciences 30, 267–282.<br />

Hilmy, H.M., Osman, A., 1989. Remobilization of gold from a chalcopyrite–pyrite mineralization<br />

Hamash gold mine, Southeastern Desert, Egypt. Mineralium Deposita<br />

24, 244–249.<br />

Hodgson, C.J., 1989. The structure of shear-related, vein-type gold deposits: a review.<br />

Ore Geology Reviews 4, 635–678.<br />

Huizenga, J.M., Touret, J.L.R., 1999. Fluid inclusions in shear zones: the case for the<br />

Umwindsi shear zone in the Harare–Shamva–Bindura greenstone belt, NE Zimbabwe.<br />

European Journal of Mineralogy 11, 1079–1090.<br />

Huston, D.L., Vandenberg, L.C., Wygralak, A., Mernagh, T., Bagas, L., Crispe, A., Lambeck,<br />

L., Cross, A., Fraser, G., Williams, N., Worden, K., Meixner, A., Goleby, B., Jones, L.,<br />

Lyons, P., Maidment, D., 2007. Lode gold mineralization in the Tanami region,<br />

northern Australia. Mineralium Deposita 42, 175–204.<br />

Jacobs, G.K., Kerrick, D.M., 1981. Methane: an equation of state with application to the ternary<br />

system H 2O–CO 2–CH 4. Geochimica et Cosmochimica Acta 45, 607–614 D.M.<br />

Kontny, A., Sadek, M.F., Abdallah, M., Marioth, R., Greiling, R.O., 1999. First investigation<br />

on shear zone related gold mineralization at El <strong>Beida</strong> area, South Eastern Desert,<br />

Egypt. In: de Wall, H., Greiling, R.O. (Eds.), Aspects of Pan-African Tectonics: International<br />

Cooperation, Bilateral Seminars International Bureau, Forschungszentrum<br />

Julich, 32, pp. 91–97.<br />

Kröner, A., Greiling, R., Reischmann, T., Hussein, I.R.M., Stern, R.J., Dürr, S., Krüger, J.,<br />

Zimmer, M., 1987. Pan-African crustal evolution in the Nubian segment of northeast<br />

Africa. In: Kroner, A. (Ed.), Proterozoic Lithosphere Evolution: American Geophysical<br />

Union Geodynamics Series, 17, pp. 235–257.<br />

Kusky, T.M., Ramadan, T.M., 2002. Structural controls on Neoproterozoic mineralization<br />

in the South Eastern Desert, Egypt: an integrated field, Landsat TM, and SIR-<br />

C/X SAR approach. Journal of African Earth Sciences 35, 107–121.<br />

Large, R., Danyushevsky, L., Hollit, C., Maslennikov, V., Meffre, S., Gilbert, S., Bull, S.,<br />

Scott, R., Emsbo, P., Thomas, H., Singh, B., Foster, J., 2009. Gold and trace element<br />

zonation in pyrite using a laser imaging technique: implications for the timing of<br />

gold in orogenic and Carlin-style sediment hosted deposits. Economic Geology<br />

104, 635–668.<br />

Leitch, C.H.B., Lentz, D.R., 1994. The Gresens approach to mass balance constraints of alteration<br />

systems: methods, pitfalls, examples. In: Lentz, D.R. (Ed.), Alteration and<br />

Alteration Processes Associated With Ore-Forming Systems. Geological Association<br />

of Canada, Short Course Notes, pp. 161–192.<br />

Loizenbauer, J., Wallbrecher, E., Fntz, H., Neumayr, P., Khudeir, A.A., Kloetzil, U., 2001.<br />

Structural geology, single zircon ages and fluid inclusion studies of the Meatiq<br />

metamorphic core complex. Implications for Neoproterozoic tectonics in the Eastern<br />

Desert of Egypt. Precambrian Research 110, 357–383.<br />

Loucks, R.R., Mavrogenes, J.A., 1999. Gold solubility in supercritical hydrothermal<br />

brines measured in synthetic fluid inclusions. Science 284, 2159–2163.<br />

Makroum, F.M., 2001. Pan-African tectonic evolution of the Wadi El Mayit area and its<br />

environs, central eastern Desert, Egypt. The 2nd International Conference on the<br />

Geology of Africa, Assiut University, Egypt. Bulletin 2, 219–233.<br />

Nano, L., Kontny, A., Sadek, M.F., Greiling, R.O., 2002. Structural evolution of metavolcanics<br />

in the surrounding of the gold mineralization at El <strong>Beida</strong>, South Eastern Desert,<br />

Egypt. Annals Geological Survey of Egypt 25, 11–22.<br />

Obeid, M.A., Hussein, A.M., Abdallah, M.A., 2001. Shear zone-hosted gold mineralization in<br />

the Late Proterozoic rocks of El <strong>Beida</strong> area, south Eastern Desert, Egypt. Transaction of<br />

Institution of Mining and Metallurgy (Sect. B: Applied Earth Sciences) 110, 192–204.<br />

Olivo, G.R., Chang, F., Kyser, T.K., 2006. Formation of the Auriferous and Barren North<br />

Dipper Veins in the Sigma Mine, Val d'Or, Canada: constraints from structural,<br />

mineralogical, fluid inclusion, and isotopic data. Economic Geology 101,<br />

607–631.<br />

Phillips, G.N., Groves, D.I., 1983. The nature of Archaean gold-bearing fluids as deduced<br />

from gold deposits of Western Australia. Journal of the Geological Society Australia<br />

30, 25–39.<br />

Pichavant, M., Ramboz, C., Weisbrod, A., 1982. Fluid immiscibility in natural processes:<br />

use and misuse of fluid inclusion data. Part I. Phase equilibria analysis. A theoretical<br />

and geometrical approach. Chemical Geology 37, 1–27.<br />

B.A. <strong>Zoheir</strong> / Journal of Geochemical Exploration 114 (2012) 82–97<br />

Ramadan, T. M., 1994. Geological and geochemical studies on some basement rocks at<br />

Wadi Hodein area, South Eastern Desert, Egypt. Ph. D. Thesis, Faculty of Science, Al<br />

Azhar University, Cairo, 188 pp.<br />

Ramadan, T.M., Kontny, A., 2004. Mineralogical and structural characterization of alteration<br />

zones detected by orbital remote sensing at Shalatein District, SE Desert,<br />

Egypt. Journal of African Earth Sciences 40, 89–99.<br />

Ramboz, C., Pichavant, M., Weisbrod, A., 1982. Fluid immiscibility in natural processes:<br />

use and misuse of fluid inclusion data. Part. II: Interpretation of fluid inclusion data<br />

in terms of immiscibility. Chemical Geology 37, 29–48.<br />

Robert, F., Brown, A.C., 1986. Archaean gold-bearing quartz veins at the Sigma mine, Abitibi<br />

greenstone belt. Quebec. Part II. Vein paragenesis and hydrothermal alteration.<br />

Economic Geology 81, 593–616.<br />

Roberts, R.G., 1987. Archean lode gold deposits. Geoscience Canada 14, 1–19.<br />

Roedder, E., 1984. Fluid inclusions, Mineralogical Society of America. Reviews in Mineralogy,<br />

12. Mineralogical Society of America, Washington, pp. 644–648.<br />

Roedder, E., Bodnar, R.J., 1980. Geologic pressure determinations from fluid inclusion<br />

studies. Annals of Review Earth and Planetary Sciences 8, 263–301.<br />

Sadek, M.F., Tolba, M.I., Youssef, M.M., Abdel Gawad, G.M., Salem, S.M., Atia, S.A., 1996.<br />

Geology of Wadi Kreiga–Gabal Korbiai Area, South Eastern Desert. Egypt, Internal<br />

report, Geological Survey of Egypt, Cairo. 67 pp.<br />

Saha, I., Venkatesh, A.S., 2002. Invisible gold within sulfides from the Archean Huttti–Maski<br />

schist belt, southern India. Journal of Asian Earth Sciences 20, 449–457.<br />

Selverstone, J., Morteani, G., Staude, J.M., 1991. Fluid channeling during ductile shearing:<br />

transformation of granodiorite into aluminous schist in the Tauern Window, Eastern<br />

Alps. Journal of Metamorphic Geology 9, 419–431.<br />

Shalaby, A., 2010. The northern dome of Wadi Hafafit culmination, Eastern Desert,<br />

Egypt: structural setting in tectonic framework of a scissor-like wrench corridor.<br />

Journal of African Earth Sciences 57 (3), 227–241.<br />

Shalaby,A.,Stüwe,K.,Makroum,F.,Fritz,H.,Kebede,T.,Klotzli,U.,2005.TheWadi<br />

Mubarak belt, Eastern Desert of Egypt: a Neoproterozoic conjugate shear system<br />

in the Arabian–Nubian Shield. Precambrian Research 136, 27–50.<br />

Shepherd, T.J., Rankin, A.H., Alderton, D.H.M., 1985. A Practical Guide to Fluid Inclusion<br />

Studies. Blackie & Son Ltd, Glasgow & London. 240 pp.<br />

Sibson, R.H., Scott, J., 1998. Stress/fault controls on the containment and release of<br />

overpressured fluids: examples from gold–quartz vein systems in Juneau, Alaska;<br />

Victoria, Australia and Otago, New Zealand. Ore Geology Reviews 13, 293–306.<br />

Simpson, C., Schmidt, S.W., 1983. An evaluation of criteria to deduce the sense of movement<br />

in sheared rocks. Bulletin of the Geological Society of America 91, 1281–1288.<br />

Stern, R.J., 1985. The Najd fault system, Saudi Arabia and Egypt: a Late Precambrian<br />

rift-related transform system. Tectonics 4, 497–511.<br />

Stern, R.J., 1994. Arc assembly and continental collision in the Neoproterozoic East African<br />

Orogen: implications for the consolidation of Gondwanaland. Annual Reviews<br />

of Earth and Planetary Science 22, 319–351.<br />

Stern, R.J., Kröner, A., Manton, W.I., Reischmann, T., Mansour, M., Hussein, I.M., 1989.<br />

Geochronology of the late Precambrian Hamisana shear zone, Red Sea Hills,<br />

Sudan and Egypt. Journal of the Geological Society of London 146, 1017–1029.<br />

Stipp, M., Stünitz, H., Heilbronner, R., Schmid, S.M., 2002. The eastern Tonale fault zone:<br />

a ‘natural laboratory’ for crystal plastic deformation of quartz over a temperature<br />

range from 250 to 700 °C. Journal of Structural Geology 24, 1861–1884.<br />

Sultan, S.A., Mansour, S.A., Santos, F.M., Helaly, A.S., 2009. Geophysical exploration for<br />

gold and associated minerals, case study: Wadi El <strong>Beida</strong> area, South Eastern Desert,<br />

Egypt. Journal of Geophysics and Engineering 6, 345–356.<br />

Touret, J.L.R., 2001. Fluids in metamorphic rocks. Lithos 55, 1–25.<br />

Waldron, J.W.F., Roselli, C., Johnston, S.K., 2007. Transpressional structures on a Late<br />

Palaeozoic intracontinental transform fault, Canadian Appalachians. Geological<br />

Society, London, Special Publications 290, 367–385.<br />

Wallbrecher, E., Fritz, H., Khudeir, A.A., Farahad, F., 1993. Kinematics of Panafrican thrusting<br />

and extension in Egypt. In: Thorweihe, U., Schandelmeier, H. (Eds.), Geoscientific<br />

Research in Northeast Africa. Rotterdam, Proceedings of the International Conference<br />

on Geoscientific Research in Northeast Africa Balkema, pp. 27–30.<br />

Wilcox, R.E., Harding, T.P., Seely, D.R., 1973. Basic wrench tectonics. American Association<br />

of Petroleum Geologists Bulletin 57, 74–96.<br />

<strong>Zoheir</strong>, B.A., 2008. Characteristics and genesis of shear zone-related gold mineralization in<br />

Egypt: a case study from the Um El Tuyor mine, south Eastern Desert. Ore Geology<br />

Reviews 34, 445–470.<br />

<strong>Zoheir</strong>, B.A., 2011. Transpressional zones in ophiolitic mélange terranes: potential<br />

exploration targets for gold in the South Eastern Desert, Egypt. Journal of Geochemical<br />

Exploration 111, 23–38.<br />

97

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!