30.12.2012 Views

On the Pan-African transition of the Arabian–Nubian Shield from ...

On the Pan-African transition of the Arabian–Nubian Shield from ...

On the Pan-African transition of the Arabian–Nubian Shield from ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>On</strong> <strong>the</strong> <strong>Pan</strong>-<strong>African</strong> <strong>transition</strong> <strong>of</strong> <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong> <strong>from</strong> compression to<br />

extension: The post-collision Dokhan volcanic suite <strong>of</strong> Kid-Malhak region,<br />

Sinai, Egypt<br />

Mohammed Z. El-Bialy<br />

Geology Department, Faculty <strong>of</strong> Science, Suez Canal University, Port Said, Egypt<br />

article info<br />

Article history:<br />

Received 19 March 2009<br />

Received in revised form 18 May 2009<br />

Accepted 12 June 2009<br />

Available online 21 June 2009<br />

Keywords:<br />

<strong>Pan</strong>-<strong>African</strong><br />

<strong>Arabian–Nubian</strong> <strong>Shield</strong><br />

Volcanic suite<br />

Petrography<br />

Geochemistry<br />

Tectonic setting<br />

1. Introduction<br />

abstract<br />

The Egyptian basement complex <strong>of</strong> <strong>the</strong> Eastern Desert and Sinai<br />

consist <strong>of</strong> Neoproterozoic juvenile crust developed in <strong>the</strong> nor<strong>the</strong>astern<br />

part <strong>of</strong> <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong> (Stern, 1994, 2002). The most<br />

prominent feature <strong>of</strong> this crust is <strong>the</strong> presence <strong>of</strong> dismembered<br />

ophiolites, metamorphosed volcano-sedimentary successions and<br />

calc-alkaline I-type intrusive complexes. Current idea on <strong>the</strong> tectonic<br />

evolution <strong>of</strong> <strong>the</strong>se orogenic terrains points to an essential role <strong>of</strong><br />

convergent processes, through <strong>the</strong> formation <strong>of</strong> intra-oceanic island arc<br />

system, subsequent ocean closure, amalgamation <strong>of</strong> <strong>the</strong> arc complexes<br />

and accretion to continental crust, followed by crustal thickening<br />

(Bentor, 1985; Abdel-Rahman and Martin, 1987; Kröner et al., 1988;<br />

Stern, 1994, 2002).These tectonic events took place during <strong>the</strong> <strong>Pan</strong>-<br />

<strong>African</strong> orogenic event between 900 and 614 Ma (Stern and Hedge,<br />

E-mail address: mzbialy@yahoo.com.<br />

Gondwana Research 17 (2010) 26–43<br />

Contents lists available at ScienceDirect<br />

Gondwana Research<br />

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

The <strong>Pan</strong>-<strong>African</strong> Kid-Malhak Dokhan volcanic suite (609 ±12 Ma) is exposed in <strong>the</strong> nor<strong>the</strong>rnmost part <strong>of</strong> <strong>the</strong><br />

<strong>Arabian–Nubian</strong> <strong>Shield</strong>. The suite consists <strong>of</strong> non-metamorphosed varicolored alternating succession <strong>of</strong><br />

porphyritic lava flows <strong>of</strong> commonly felsic composition (rhyolite–dacite) interlayered with compositionally<br />

equivalent pyroclastic beds (dominantly ignimbrites). These Dokhan volcanics are quite evolved (SiO2≈ 65–<br />

77 wt.%), with strong high-K calc-alkaline affinity and are characterized by relative enrichment in total<br />

alkalis, Ba, Y, Zr and total REEs, depletion in Sr, and a LREE-enriched REE patterns with significant negative Eu<br />

anomalies. The Kid-Malhak Dokhan lavas display geochemical characteristics <strong>of</strong> both orogenic arc-type and<br />

anorogenic within-plate environments, suggesting eruption in a <strong>transition</strong>al “post-collisional tectonic<br />

setting. The ages <strong>of</strong> emplacement <strong>of</strong> <strong>the</strong> Dokhan volcanics in Egypt including that <strong>of</strong> Kid-Malhak region (580–<br />

620 Ma) coincide with end <strong>of</strong> <strong>the</strong> documented collision between <strong>the</strong> juvenile <strong>Arabian–Nubian</strong> crust and<br />

Saharan Metacraton and <strong>the</strong> subsequent extensional collapse event. This post-collision <strong>transition</strong> <strong>from</strong><br />

compression to extension is explained by <strong>the</strong> extensional collapse following continental collision, which was<br />

controlled mainly by lithospheric delamination and slab break<strong>of</strong>f (passive rifting). Various trace element<br />

characteristics discussed herein have indicated that <strong>the</strong> studied Dokhan magma was highly likely generated<br />

<strong>from</strong> crustal sources and that assimilation–fractional crystallization (AFC) and crustal contamination have<br />

played a major role and are most probably superimposed on fractional crystallization during <strong>the</strong> magmatic<br />

evolution <strong>of</strong> Kid-Malhak Dokhan volcanic suite. The eruption <strong>of</strong> <strong>the</strong> high-K calc-alkaline post-collisional<br />

Dokhan volcanics in Egypt defines a tectono-magmatic <strong>transition</strong> between <strong>the</strong> older calc-alkaline arc-related<br />

and <strong>the</strong> subsequent alkaline magmatism in <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong>.<br />

© 2009 Published by Elsevier B.V. on behalf <strong>of</strong> International Association for Gondwana Research.<br />

1342-937X/$ – see front matter © 2009 Published by Elsevier B.V. on behalf <strong>of</strong> International Association for Gondwana Research.<br />

doi:10.1016/j.gr.2009.06.004<br />

1985; Kröner et al., 1992; Beyth et al., 1994; Stern, 1994). The final stage<br />

(between 614 and 550 Ma) in <strong>the</strong> <strong>Pan</strong>-<strong>African</strong> crustal evolution in Egypt<br />

was characterized by <strong>the</strong> eruption <strong>of</strong> K-rich volcanic rocks (Dokhan<br />

volcanics) and emplacement <strong>of</strong> shallow level felsic intrusions (Egyptian<br />

Younger Granites). The term “Dokhan volcanics” refers to varicolored<br />

thick sequence <strong>of</strong> lava flows and pyroclastics <strong>of</strong> predominantly andesitic<br />

to rhyolitic composition in association with ignimbritic rhyolites (Basta<br />

et al., 1980; Heikal et al., 1980; Stern and Gottfried, 1986). The Dokhan<br />

volcanic sequences are widely distributed and extensively studied in <strong>the</strong><br />

Eastern Desert <strong>of</strong> Egypt, whereas <strong>the</strong>y are much less known in Sinai<br />

(Fig. 1). However, some workers consider some comparable nonmetamorphosed<br />

volcanic sequences in Sinai (e.g. G. Khashabi, G. Fierani,<br />

W. Kid, W. Meknas and o<strong>the</strong>rs) as representatives and equivalent to <strong>the</strong><br />

Dokhan volcanics <strong>of</strong> <strong>the</strong> Eastern Desert (e.g. Bentor, 1985; Hassan and<br />

Hashad, 1990; El Metwally et al., 1999; Azzaz et al., 2000; Hassan et al.,<br />

2001; El-Bialy and El Omlla, 2007).The various studies carried out on <strong>the</strong><br />

geochemistry <strong>of</strong> <strong>the</strong> Dokhan volcanics in <strong>the</strong> Egyptian Eastern Desert<br />

revealed that <strong>the</strong>y have medium- to high-K calc-alkaline affinities. There<br />

is general agreement that fractional crystallization <strong>of</strong> basaltic magma


Fig. 1. Distribution <strong>of</strong> <strong>the</strong> Dokhan volcanics in Eastern Desert and Sinai (modified after<br />

Abdel-Rahman (1996), and location <strong>of</strong> <strong>the</strong> study area.<br />

coupled with minor crustal contamination controlled <strong>the</strong>ir magmatic<br />

evolution (Stern and Gottfried,1986; El Gaby et al.,1989; Abdel-Rahman,<br />

1996; Mohamed et al., 2000; Saleh, 2003; Moghazi, 2003; El Sayed et al.,<br />

2004; Eliwa et al., 2006). However, <strong>the</strong>re is considerable discussion and<br />

no consensus has been achieved concerning <strong>the</strong> tectonic setting for <strong>the</strong><br />

genesis <strong>of</strong> <strong>the</strong>se magmas. The controversy is centered around whe<strong>the</strong>r<br />

Dokhan volcanics have been formed (1) in a subduction environment<br />

(Hassan and Hashad, 1990; El Gaby et al., 1990; Abdel-Rahman, 1996;<br />

Hassan et al., 2001; Saleh, 2003), (2) in association with extension after<br />

crustal thickening (Stern et al., 1984, 1988; Stern and Gottfried, 1986;<br />

Mohamed et al., 2000), or (3) during <strong>transition</strong> between subduction and<br />

extension (Ressetar and Monard, 1983; Moghazi, 2003; El Sayed et al.,<br />

2004; Eliwa et al., 2006). Much <strong>of</strong> <strong>the</strong> available reliable ages on Dokhan<br />

volcanics (580–620 Ma), were obtained <strong>from</strong> <strong>the</strong> Eastern Desert ra<strong>the</strong>r<br />

than <strong>from</strong> Sinai (El Shazly et al., 1973; Stern, 1979; Ries et al., 1983;<br />

Ressetar and Monard,1983; Gillespie and Dixon, 1983; Stern and Hedge,<br />

1985; Abdel-Rahman and Doig, 1987; Wilde and Youssef, 2000;<br />

Breitkreuz et al., 2008). In Sinai, Bentor (1985) concluded that a<br />

comparable period to <strong>the</strong> Dokhan volcanism <strong>of</strong> intense and apparently<br />

continuous volcanism occurred between 620 and 580 Ma ago. Also,<br />

similar Rb–Sr ages were obtained on orogenic calc-alkaline andesite–<br />

rhyolite rocks <strong>from</strong> Sinai, such as 587±9 Ma for Wadi Rutig andesite–<br />

rhyolite and 609±12 Ma for rhyolites–dacites <strong>from</strong> Wadi Madsus and<br />

Wadi Qabila (<strong>the</strong> investigated volcanics) by Bielski (1982).<br />

The volcanics, under investigation, are usually included within<br />

<strong>the</strong> o<strong>the</strong>r surrounding metavolcano–sedimentary complex, known<br />

as Kid Group <strong>of</strong> Shimron (1980), in many research works (i.e.<br />

Shimron, 1980, 1983; Reymer, 1983; Reymer and Yogev, 1983; Furnes<br />

et al., 1985; Hassanen, 1992). This paper presents new geologic and<br />

geochemical data on <strong>the</strong> Dokhan volcanic rocks <strong>of</strong> Kid-Malhak<br />

region, South Sinai, Egypt. The aim is to integrate field, petrographic<br />

and geochemical data to correlate <strong>the</strong>m with <strong>the</strong> Dokhan volcanics <strong>of</strong><br />

<strong>the</strong> Eastern Desert and to contribute to <strong>the</strong> understanding <strong>of</strong> <strong>the</strong>ir<br />

origin and tectonic setting.<br />

M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

2. Geological setting<br />

2.1. Regional geology<br />

The basement rocks exposed in <strong>the</strong> Kid-Malhak region (Fig. 2) can<br />

be categorized into <strong>the</strong> following major units (<strong>from</strong> <strong>the</strong> oldest to<br />

youngest): gneisses, metasediments, metavolcanics (basalts and<br />

andesites), metagabbro–diorite complex, old granites (tonalities–<br />

quartz diorites), Dokhan volcanics (rhyolites and dacites) and younger<br />

granites (monzo-syenogranites). The investigated Dokhan volcanics<br />

are believed to be extruded between <strong>the</strong> emplacement <strong>of</strong> <strong>the</strong><br />

preceding old and post-dating younger granites.<br />

Gneissic rocks consist <strong>of</strong> foliated and lineated diorites and tonalites<br />

with occasional mafic-rich xenoliths. Development <strong>of</strong> augen structure<br />

is obvious in some outcrops. These gneisses were referred to as Qenaia<br />

Formation by Bentor and Eyal (1987) and are considered <strong>of</strong> plutonic<br />

origin (Brooijmans et al., 2003; Be'eri-Shlevin et al., 2009b).<br />

Metasediments are <strong>the</strong> major exposed rock unit, covering about half<br />

<strong>of</strong> <strong>the</strong> basement in <strong>the</strong> mapped area. El Metwally et al. (1999)<br />

distinguished <strong>the</strong>m as meta-calcpelites and meta-psammopelites.<br />

Meta-calcpelites comprise banded para-amphibolites, and hornblende<br />

and chlorite schists, while meta-psammopelites include garnetiferous<br />

biotite schist and phyllite. The aforementioned metasediments<br />

make up <strong>the</strong> Umm Zariq Formation (Furnes et al., 1985), Heib<br />

Formation and most <strong>of</strong> Malhak Formations (Shimron, 1980). The<br />

metasediments crop out around <strong>the</strong> downstream <strong>of</strong> Wadi Kid (sou<strong>the</strong>rn<br />

part <strong>of</strong> <strong>the</strong> mapped area; Fig. 2) are quite different <strong>from</strong> <strong>the</strong> above<br />

mentioned one. They are dominantly a metasedimentary sequence<br />

consisting <strong>of</strong> low-grade schists, metaconglomerates, metamudstones<br />

and metasandstones. The conglomerates contain granitic, andesitic,<br />

and pelitic pebbles, whereas psammitic and pelitic sequences display<br />

occasionally sedimentary structures, such as graded bedding and low<br />

angle structures (Blasband et al., 1997, 2000; Brooijmans et al., 2003).<br />

This sedimentary sequence was referred to as Tarr Formation<br />

(Shimron, 1980, 1983). The metavolcanics are greenschist to amphibolite<br />

facies metamorphosed lava flows, ranging in composition <strong>from</strong><br />

basalts to andesites (Furnes et al., 1985). In hand specimen, <strong>the</strong>y are<br />

ei<strong>the</strong>r aphyric or fairly porphyritic with scattered plagioclase phenocrysts<br />

(andesites). The lavas may be massive, structureless flows N5 m<br />

in thickness, or in some cases with pillow structures. The lower lava<br />

flow sheets in this unit are interbedded with and overlying <strong>the</strong><br />

abovementioned thick metasedimentary succession. The grade <strong>of</strong><br />

metamorphism decrease upward in <strong>the</strong>se metavolcanics, and <strong>the</strong><br />

rocks <strong>of</strong> <strong>the</strong> upper horizons are relatively fresh or weakly<br />

metamorphosed to <strong>the</strong> lower greenschist facies (Hassanen, 1992).<br />

The metagabbro-diorite complex occurs as a single large intrusive<br />

body in <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> Kid-Malhak region. It was called as<br />

“Sharira gabbro and diorite complex” by Furnes et al. (1985) who<br />

described it as a layered gabbroic and dioritic layered intrusive<br />

massif <strong>of</strong> at least 2000 m thick and was dated at 570 ± 4 Ma by<br />

Moghazi et al. (1998). Old granites are found as several isolated<br />

bodies in <strong>the</strong> most southwestern corner <strong>of</strong> <strong>the</strong> mapped area along<br />

<strong>the</strong> nor<strong>the</strong>rn bank <strong>of</strong> W. Yahmid. Old granites are quite heterogeneous,<br />

and several varieties <strong>of</strong> different composition and/or<br />

texture are recognized. These rocks are grey colored, ranging in<br />

composition <strong>from</strong> quartz diorite to tonalite, and in texture <strong>from</strong><br />

granular to porphyritic. They <strong>of</strong>ten host rounded to sub-rounded<br />

microgranular mafic enclaves <strong>of</strong> mela-quartz diorite composition.<br />

Younger granites represent <strong>the</strong> last magmatic activity in <strong>the</strong> region,<br />

bordering it <strong>from</strong> <strong>the</strong> north, west and south and expanding laterally<br />

outside <strong>the</strong> mapped area. They are cross-cut by numerous dykes <strong>of</strong><br />

variable composition and directions. Aplitic dykes and quartz veins<br />

traversing <strong>the</strong>se rocks are ra<strong>the</strong>r common. They have a dominant<br />

granite to granodiorite composition and biotite as <strong>the</strong> main mafic<br />

mineral. Moreover, <strong>the</strong>y vary in texture <strong>from</strong> fine to coarse-grained<br />

granular to porphyritic with pink alkali feldspar megacrysts.<br />

27


28 M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

2.2. Field observations<br />

Fig. 2. Simplified geological map <strong>of</strong> Wadi Kid-Wadi Malhak region, Sinai (modified after Shimron (1980) and El Metwally et al. (1999)).<br />

The investigated Dokhan volcanics in Kid-Malhak region form<br />

brown to dark purple landscape <strong>of</strong> moderate relief, covering an area<br />

<strong>of</strong> approximately 110 km 2 <strong>of</strong> <strong>the</strong> entire mapped area (495 km 2 ). The<br />

studied Dokhan volcanics are exposed along Wadi Kid, Wadi Malhak,<br />

and Wadi Qabila (Fig. 2). They extrude and overlie o<strong>the</strong>r older<br />

metamorphosed basement units including <strong>the</strong> gneisses, metasediments<br />

and metavolcanics, and are intruded by <strong>the</strong> younger granites<br />

at Wadi Qabila with sharp contacts. A younger to contemporaneous<br />

rocks to <strong>the</strong>se volcanics are thick interbedded epiclastic conglomerate<br />

beds (100–150 m), separating <strong>the</strong>se volcanics into upper and<br />

lower stratigraphic sequences, and enclose frequent pebbles <strong>of</strong> <strong>the</strong>m<br />

along with those <strong>of</strong> o<strong>the</strong>r older basement. These interbedded<br />

epiclastic conglomerates are most probably stratigraphically equivalent<br />

to <strong>the</strong> Hammamat Group (Akaad and Noweir, 1980) clastic<br />

sediments <strong>of</strong> <strong>the</strong> Egyptian Eastern Desert. The Dokhan volcanics in<br />

<strong>the</strong> study area consist <strong>of</strong> non-metamorphosed varicolored alternating<br />

succession <strong>of</strong> dominantly porphyritic flows <strong>of</strong> commonly felsic<br />

composition interlayered with compositionally equivalent pyroclastic<br />

beds. The later is dominated by welded ash tuffs (ignimbrites)<br />

and less commonly crystal-lithic ash tuffs. Individual flows in this<br />

area average less than 10 m thick. The very lowermost part <strong>of</strong> <strong>the</strong><br />

succession is characterized by minor andesitic flows, while <strong>the</strong> rest<br />

<strong>of</strong> <strong>the</strong> succession upward is dominated by felsic lava flows and<br />

pyroclastics <strong>of</strong> <strong>the</strong> rhyolite–rhyodacite–dacite compositional range.<br />

Mafic cognate volcanic xenoliths are occasionally encountered in <strong>the</strong><br />

rhyolite–dacite lava flows.<br />

3. Petrography<br />

Based on mineralogical composition, textural and field characteristics<br />

and geochemical data (see below), <strong>the</strong> investigated volcanics<br />

can be divided into two units, namely: lava flows and ignimbrites.<br />

Twenty eight volcanic samples were studied in <strong>the</strong> present work,<br />

twenty three <strong>of</strong> <strong>the</strong>m are lava flows, and <strong>the</strong> remainders are welded<br />

tuffs (ignimbrites).<br />

3.1. Lava flows<br />

Lava flows predominate in this volcanic suite, comprising chiefly<br />

rhyolites with subordinate dacites and trachydacites (Fig. 3). Lava<br />

flows are strongly porphyritic, with up to 41% phenocrysts set in an<br />

originally partly glassy, now devitrified groundmass. Groundmass <strong>of</strong><br />

almost all lavas is fine-grained anhedral granular displaying flow<br />

banding or trachytic textures. Rhyolites contain quartz, sodic plagioclase<br />

and K-feldspar phenocrysts with occasional biotite and/or<br />

hornblende microphenocrysts (b0.5 mm). Quartz phenocrysts occur<br />

as highly corroded and embayed crystals 1–3 mm across. Fur<strong>the</strong>rmore,<br />

coarse secondary quartz forming veins and nests is widespread. Such<br />

nests are up to 7 mm in size and are composed <strong>of</strong> anhedral interlocking<br />

grains, 0.1–2 mm across. Plagioclase phenocrysts (An10–19) are lath- or


Fig. 3. Geochemical classification <strong>of</strong> <strong>the</strong> studied volcanics using total alkalis vs. silica<br />

(TAS) diagram (Le Bas et al., 1986), with <strong>the</strong> discriminating boundary between alkaline<br />

and subalkaline fields after Irvine and Baragar (1971). For comparison, a field is shown<br />

for Dokhan volcanics (DV) <strong>from</strong> o<strong>the</strong>r areas (Basta et al., 1980; Ressetar and Monard,<br />

1983; Stern and Gottfried, 1986; Bentor and Eyal, 1987; Abdel-Rahman, 1996; Azzaz<br />

et al., 2000; Hassan et al., 2001; Moghazi, 2003; Eliwa, 2006).<br />

tabular-shaped, euhedral to subhedral, and range up to 1 cm in size,<br />

tending sometimes to cluster in glomerocrysts. Interiors <strong>of</strong> some <strong>of</strong> <strong>the</strong><br />

phenocrysts are highly sericitized with relatively coarse secondary<br />

muscovite scales. Moreover, some plagioclase phenocrysts frequently<br />

contain irregular patches <strong>of</strong> K-feldspar within <strong>the</strong>m. The textural<br />

relation <strong>of</strong> <strong>the</strong>se patches with <strong>the</strong> plagioclase host indicates intergrowth<br />

ra<strong>the</strong>r than exsolution. Some rhyolite samples lack plagioclase<br />

as a phenocryst phase. Alkali feldspar phenocrysts are common in<br />

most rhyolites, represented by subhedral to resorbed perthitic<br />

sanidine and occasional fine-cross-hatch twinned anorthoclase. Biotite<br />

and hornblende microphenocrysts (b1 mm) are present in some<br />

rhyolite samples. The groundmass consists <strong>of</strong> quartz, K-feldspar and<br />

M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

plagioclase with intergranular biotite, opaques, chlorite, sericite and<br />

accessory apatite and zircon. Devitrification textures are quite<br />

common in <strong>the</strong> groundmass, including granophric intergrowths,<br />

micro-poikilitic quartz and bow-tie spherulites. Dacites consist <strong>of</strong><br />

abundant plagioclase and quartz and mafic phenocrysts that are set in<br />

a felty textured microcrystalline groundmass <strong>of</strong> quartz, feldspars,<br />

biotite and opaques. Plagioclase phenocrysts have albite–oligoclase<br />

composition (An8–22) and are found ei<strong>the</strong>r independent or as<br />

glomerocrysts. They are frequently altered to sericite, particularly in<br />

<strong>the</strong>ir cores. In contrast to rhyolites, quartz is rare and alkali feldspars<br />

are absent as phenocryst phases. Alkali feldspars and plagioclase are<br />

found in sub-equal amounts in <strong>the</strong> groundmass. Trachydacites are less<br />

porphyritic compared with rhyolites and dacites. They consist <strong>of</strong> sparse<br />

sodic plagioclase and alkali feldspar phenocrysts and pyroxene<br />

microphenocrysts enclosed in a pilotaxitic fine-grained groundmass<br />

<strong>of</strong> sanidine, quartz and plagioclase microlites impregnated with tiny<br />

chlorite flakes and fine opaque dust.<br />

3.2. Ignimbrites<br />

Ignimbrites are welded vitric tuffs, with few and smaller crystals<br />

and lithic fragments (b10%). They are characterized by <strong>the</strong> predominance<br />

<strong>of</strong> glass shards and pumaceous fragments and displaying<br />

features indicative <strong>of</strong> pyroclastic flow origin. The degree <strong>of</strong> welding<br />

and nature <strong>of</strong> <strong>the</strong> principle juvenile materials (Pumice and glass<br />

shards) can readily distinguish <strong>the</strong>m into eutaxitic textured and thinlaminated<br />

vitric tuffs. Eutaxitic-textured vitric tuffs are highly welded,<br />

consisting <strong>of</strong> flattened and stretched pumaceous fragments <strong>of</strong><br />

lenticular shape, sometimes with crenulated ends (fiamme) that lie<br />

parallel and beside one ano<strong>the</strong>r, showing a well defined lineation.<br />

Fiammes can be clearly distinguished <strong>from</strong> <strong>the</strong> enclosing matrix by<br />

<strong>the</strong>ir lenticular form and higher degree <strong>of</strong> devitrification. Crystals and<br />

lithic fragments may occasionally cause divergence and bending <strong>of</strong> <strong>the</strong><br />

fiamme around <strong>the</strong>m. The linear fabric resulted <strong>from</strong> stretching and<br />

welding <strong>of</strong> <strong>the</strong> fiamme indicates secondary mass flowage <strong>of</strong> <strong>the</strong> tuff<br />

during welding (Cas and Wright, 1987). Relics <strong>of</strong> glassy shards are<br />

Table 1<br />

Major elements composition (wt.%), calculated CIPW normative minerals and some elemental ratios <strong>of</strong> Kid-Malhak area Dokhan volcanics.<br />

Sample no. M2 M7 Q1 M6 Q10 Q7 Q11 Q6 Q8 K6 Q5 K5 K1 K2 K11 K12 Average<br />

Normalized major elements (wt.%)<br />

SiO2 76.61 75.39 72.02 64.87 75.84 72.21 70.63 71.15 74.03 64.66 68.9 69.34 74.43 70.12 75.46 66.95 71.41<br />

TiO2 0.19 0.17 0.33 1.11 0.32 0.37 0.45 0.4 0.32 1.1 1.02 1.01 0.3 0.65 0.25 1.05 0.57<br />

Al2O3 11.7 13.83 15.61 16.01 13.28 15.23 15.31 15.33 13.77 16.23 14.44 14.13 13.54 14.59 13.15 15.22 14.46<br />

FeO* 1.8 0.72 1.38 5.14 1.53 1.66 2.27 2.16 1.93 5.06 5.5 5.13 1.83 3.34 1.5 5.21 2.89<br />

MnO 0.06 0.06 0.05 0.09 0.08 0.03 0.05 0.05 0.09 0.09 0.11 0.13 0.06 0.08 0.07 0.11 0.07<br />

MgO 0.69 0.15 0.26 1.48 0.27 0.17 0.45 0.54 0.44 1.47 1.02 1.01 0.5 1.1 0.39 1.21 0.7<br />

CaO 1 0.63 0.94 2.28 1.03 0.64 2.44 1.63 0.62 2.37 2.1 2.26 1.22 1.85 0.82 2.26 1.51<br />

Na2O 2.09 4.56 5.94 4.83 4.78 6.07 4.68 4.37 2.26 4.86 3.67 3.71 3.75 4.21 3.42 4.29 4.22<br />

K2O 5.84 4.45 3.43 3.8 2.83 3.56 3.58 4.25 6.52 3.78 2.93 3.01 4.31 3.88 4.91 3.38 4.03<br />

P2O5 0.02 0.03 0.04 0.4 0.04 0.06 0.14 0.12 0.02 0.4 0.29 0.27 0.06 0.18 0.03 0.32 0.15<br />

Total 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100<br />

Na2O+K2O 7.92 9.01 9.37 8.63 7.61 9.63 8.25 8.62 8.78 8.63 6.6 6.72 8.06 8.09 8.33 7.67 8.25<br />

Na2O/K2O 0.36 1.03 1.73 1.27 1.69 1.71 1.31 1.03 0.35 1.28 1.25 1.23 0.87 1.09 0.7 1.27 1.13<br />

FeO /MgO 2.6 4.74 5.26 3.48 5.7 9.8 5.07 3.96 4.42 3.44 5.4 5.07 3.66 3.04 3.85 4.31 4.61<br />

A/CNK 1.01 1.03 1.03 0.99 1.04 1.02 0.96 1.04 1.16 0.99 1.11 1.05 1.04 1.01 1.06 1.03 1.04<br />

CIPW norm<br />

Quartz 37.63 29.84 21.16 13.02 33.65 20.74 22.94 24.08 32.64 12.58 27.52 27.3 31.65 23.49 33.45 19.94 25.73<br />

Anorthite 4.83 2.93 4.4 8.7 4.85 2.78 10.19 7.3 2.95 9.14 8.52 9.45 5.66 8 3.87 9.12 6.42<br />

Diopside 0 0 0 0 0 0 0.85 0 0 0 0 0 0 0 0 0 0.05<br />

Hypers<strong>the</strong>ne 4.57 1.44 2.54 10.74 2.86 2.7 3.89 4.44 4 10.61 10.41 9.79 3.97 7.48 3.25 10.34 5.82<br />

Albite 17.69 38.59 50.26 40.87 40.45 51.36 39.6 36.98 19.12 41.12 31.05 31.39 31.73 35.62 28.94 36.3 35.69<br />

Orthoclase 34.51 26.3 20.27 22.46 16.72 21.04 21.16 25.12 38.53 22.34 17.32 17.79 25.47 22.93 29.02 19.98 23.81<br />

Apatite 0.05 0.07 0.09 0.93 0.09 0.14 0.32 0.28 0.05 0.93 0.67 0.63 0.14 0.42 0.07 0.74 0.35<br />

Ilmenite 0.36 0.32 0.63 2.11 0.61 0.7 0.85 0.76 0.61 2.09 1.94 1.92 0.57 1.23 0.47 1.99 1.07<br />

Corundum 0.17 0.44 0.51 0.76 0.58 0.37 0 0.86 1.92 0.79 2.11 1.31 0.63 0.53 0.79 1.16 0.81<br />

Magnetite 0.15 0.06 0.12 0.42 0.13 0.13 0.19 0.17 0.16 0.41 0.45 0.42 0.15 0.28 0.12 0.42 0.23<br />

Zircon 0.09 0.03 0.06 0.06 0.09 0.07 0.03 0.03 0.06 0.06 0.06 0.06 0.06 0.06 0.06 0.06 0.06<br />

Diff. index 89.83 94.73 91.69 76.34 90.82 93.14 83.69 86.17 90.29 76.04 75.88 76.48 88.85 82.04 91.4 76.22 85.23<br />

29


30 M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

devitrified into axiolites and fan spherulites, while some o<strong>the</strong>r glassy<br />

shards are devitrified into an assemblage <strong>of</strong> spherulites, granophyres<br />

and very fine aggregates <strong>of</strong> recrystallized quartz. Crystals and crystal<br />

fragments consist mostly <strong>of</strong> juvenile interatelluric quartz, sanidine<br />

and plagioclase, while lithic fragments are predominated by cognate<br />

volcanic rocks. Thin-laminated vitric tuffs are composed <strong>of</strong> alternations<br />

<strong>of</strong> thin devitrified microcrystalline laminae, which are mostly<br />

large stretched fiamme, and thicker opaque-rich cryptocrystalline<br />

laminae contain scattered crystal and lithic fragments. Both are<br />

usually parallel and even, but <strong>the</strong> former laminae may diverge and<br />

warp-round different crystals and lithic fragments included in <strong>the</strong>m.<br />

The striking difference between <strong>the</strong> studied Dokhan volcanics and<br />

o<strong>the</strong>rs in Sinai (El Metwally et al., 1999; Azzaz et al., 2000, Hassan<br />

et al., 2001) and those <strong>of</strong> <strong>the</strong> Eastern Desert <strong>of</strong> Egypt, is <strong>the</strong>ir felsic<br />

compositional range in comparison with <strong>the</strong> “bimodal” or continuous<br />

spectral composition <strong>of</strong> <strong>the</strong> Dokhan volcanics in <strong>the</strong> Eastern Desert<br />

(e.g. Basta et al., 1980; Ressetar and Monard, 1983; Stern and Gottfried,<br />

1986; Abdel-Rahman, 1996; Mohamed et al., 2000; Saleh, 2003;<br />

Moghazi, 2003; El Sayed et al., 2004; Eliwa, 2006).<br />

4. Geochemistry<br />

4.1. Analytical techniques<br />

Concentrations <strong>of</strong> major and most trace elements (XRF) for 16<br />

representative lava flows samples and <strong>of</strong> REEs plus Cs, Hf and Ta (ICP-<br />

MS) for 5 selected samples <strong>of</strong> <strong>the</strong>m were determined at <strong>the</strong><br />

GeoAnalytical Laboratory, Washington State University, Pullman,<br />

WA, USA. Major elements and <strong>the</strong> trace elements Cr, Sc, V, Ba, Sr, Rb,<br />

Zr, Y, Nb, Ga, Cu, Zn, Pb, Th and U were determined on fused beads<br />

(3.5 g sample + 7 g dilithium tetraborate flux) by (ThermoARL<br />

Advant'XP+) X-ray fluorescence spectrometer using a rhodium (Rh)<br />

target operated at 50 kV/50 mA with full vacuum and a 25 mm mask<br />

for all <strong>the</strong> previous elements. The trace element concentrations (ppm)<br />

Table 2<br />

Trace and REE elements contents (ppm), calculated crystallization temperatures and some elemental ratios <strong>of</strong> Kid-Malhak area Dokhan volcanics.<br />

Sample no. M2 M7 Q1 M6 Q10 Q7 Q11 Q6 Q8 K6 Q5 K5 K1 K2 K11 K12 Average<br />

Ni 1 0 0 4 0 0 1 0 0 0 16 14 1 13 0 9 3.7<br />

Cr 2 2 1 2 4 1 4 3 4 1 18 10 4 21 3 8 5.55<br />

Sc 5 3 6 11 8 6 4 3 8 9 11 8 6 8 6 11 7.08<br />

V 3 2 11 51 10 10 32 28 13 47 91 82 15 44 7 68 32.09<br />

Cs 0.77 0.74 2.21 0.98 2.26 1.39<br />

Ba 1390 716 1076 744 913 1301 1039 914 992 693 706 641 1028 926 1003 712 924.66<br />

Rb 89 103 90 72 64 81 86 112 137 68 80 77 88 87 99 75 88.01<br />

Sr 129 100 151 321 149 160 528 380 133 297 311 298 221 282 127 308 243.47<br />

Hf 11.11 9.65 4.92 8.57 7.71 8.39<br />

Zr 415 121 301 293 413 337 181 164 282 272 283 304 331 280 309 291 286.09<br />

Y 49 18 33 36 39 29 16 14 32 33 35 29 35 32 35 34 31.04<br />

Ta 1.95 0.84 0.71 1.23 1.09 1.16<br />

Nb 21.9 11.3 13.2 13 12.3 11.1 7.9 7.7 11.9 12.3 22.4 19 13.98 14.1 14.4 16.7 13.95<br />

Ga 17 15 16 20 17 17 18 19 21 17 20 17 18 19 18 18 17.91<br />

Cu 1 1 0 5 0 2 2 0 0 6 41 43 0 11 1 25 8.65<br />

Zn 19 19 18 85 6 21 35 35 38 79 92 88 21 49 21 86 44.38<br />

Pb 8 20 8 9 6 11 13 12 10 7 12 11 9 10 11 10 10.48<br />

Th 12 15 17 11 15 17 12 12 14 7 8 8 11 12 14 9 12.04<br />

U 4 4 5 3 5 5 3 3 4 3 2 2 4 4 4 2 3.46<br />

Rb/Sr 0.69 1.03 0.6 0.22 0.43 0.51 0.16 0.3 1.03 0.23 0.26 0.26 0.4 0.31 0.78 0.24 0.46<br />

Ba/Sr 10.75 7.15 7.1 2.32 6.15 8.15 1.97 2.4 7.44 2.33 2.27 2.15 4.65 3.28 7.9 2.31 4.9<br />

K/Rb 547 357 316 439 368 366 344 314 395 465 303 325 407 370 412 374 381<br />

Ba/Rb 15.69 6.93 11.92 10.35 14.29 16.13 12.06 8.13 7.23 10.26 8.8 8.32 11.68 10.64 10.13 9.49 10.75<br />

Y/Nb 2.21 1.58 2.46 2.76 3.15 2.57 1.99 1.82 2.66 2.71 1.56 1.53 2.5 2.27 2.44 2.03 2.27<br />

Zr/Nb 18.95 10.71 22.77 22.56 33.59 30.35 22.92 21.31 23.71 22.15 12.63 16 23.68 19.86 21.52 17.41 21.26<br />

Zr/Hf 37.34 42.81 33.35 38.62 36.32<br />

Nb/Ta 11.23 14.64 10.85 11.37 12.94<br />

T(C) 882 766 841 821 881 850 785 786 854 814 841 843 857 830 854 831 833.5<br />

La 47.29 31 60 35 48.85 54 29 30.3 33 34 27 30 42.44 38.25 40 32 38.5<br />

Ce 100.65 59 114 77 100.65 107 55 59.37 76 73 65 69 86.92 79.45 84 71 79.79<br />

Pr 12.16 12.25 6.59 10.39 10.28 0 10.34<br />

Nd 46.2 24 44 39 46.95 44 22 22.85 36 38 35 36 38.55 39.24 38 38 36.7<br />

Sm 9.63 9.48 3.96 7.6 7.92 7.72<br />

Eu 1.14 2.02 0.84 1.39 1.54 1.38<br />

Gd 8.2 7.86 2.91 6.34 6.53 6.37<br />

Tb 1.41 1.26 0.45 1.02 1.04 1.04<br />

Dy 8.85 7.37 2.56 6.25 6.19 6.24<br />

Ho 1.86 1.51 0.51 1.31 1.27 1.29<br />

Er 5.2 3.81 1.42 3.48 2.89 3.36<br />

Tm 0.79 0.58 0.23 0.57 0.52 0.54<br />

Yb 5.05 3.56 1.49 3.39 3.16 3.33<br />

Lu 0.79 0.56 0.25 0.55 0.52 0.54<br />

REE 249.23 246.72 133.76 210.2 198.8 207.74<br />

Eu/Eu* 0.39 0.71 0.75 0.61 0.65 0.62<br />

Ce/Yb 19.95 28.27 39.76 25.64 25.14 27.75<br />

La/Yb 9.37 13.72 20.29 12.52 12.1 13.6<br />

Sm/Nd 0.21 0.2 0.17 0.2 0.2 0.2<br />

Eu/Sm 0.12 0.21 0.21 0.18 0.19 0.18<br />

(La/Yb)N 6.37 9.32 13.78 8.5 8.22 9.24<br />

(La/Sm)N 3.07 3.22 4.78 3.49 3.02 3.51<br />

(Gd/Yb)N 1.31 1.79 1.58 1.51 1.67 1.57<br />

(La/Lu)N 6.21 8.98 12.35 8.01 7.64 8.64<br />

Cs, Hf, Ta and REE concentrations <strong>of</strong> <strong>the</strong> underlined samples are determined by ICP-MS, o<strong>the</strong>rwise by XRF.


are un-normalized, where original major element concentrations<br />

determined are normalized to 100% on a volatile-free basis, with total<br />

iron expressed as FeO. REE, Cs, Hf and Ta elements were analyzed by a<br />

(Sciex Elan-250) ICP-Mass Spectrometer. Samples were first subjected<br />

to complex combined fusion-dissolution procedure to ensure <strong>the</strong><br />

complete dissolution <strong>of</strong> zircons and o<strong>the</strong>r refractory phases. Dissolved/diluted<br />

samples (60 ml final volume; 1:240 final dilution) are<br />

introduced into <strong>the</strong> argon plasma at 1.0 ml/min using a peristaltic<br />

pump and an automatic sampler. Plasma power is 1500 W. The<br />

instrument is run in “multi-element” mode averaging 10 repeats <strong>of</strong><br />

0.5 s/element for a total integrated count time <strong>of</strong> 5 s/element. Full<br />

details on sample preparation, analytical procedures, precision and<br />

accuracy are given by Johnson et al. (1999) and Knaack et al. (1994),<br />

also available on WSU Department <strong>of</strong> Geology website at (http://<br />

www.wsu.edu/~geolab/note.html).<br />

4.2. Results<br />

Major, trace and REE data <strong>of</strong> <strong>the</strong> investigated Dokhan volcanics are<br />

listed in Tables 1 and 2 in addition to calculated normative minerals,<br />

different geochemical ratios and parameters as well as crystallization<br />

temperature estimates. Table 3 shows a comparison between data <strong>of</strong><br />

<strong>the</strong> studied Dokhan volcanics average with o<strong>the</strong>r averages <strong>of</strong> Dokhan<br />

volcanics <strong>from</strong> different occurrences in <strong>the</strong> Egyptian Eastern Desert<br />

and Sinai.<br />

4.2.1. Whole-rock geochemical characteristics<br />

Geochemical classification using total alkalis versus silica (TAS)<br />

diagram (Le Bas et al., 1986) shows that <strong>the</strong> studied volcanics<br />

(SiO2≈65–77 wt.%) are classified predominantly as rhyolites and<br />

much less as dacites and trachydacites (Fig. 3). Fur<strong>the</strong>r, <strong>the</strong>y plot<br />

within <strong>the</strong> field constructed on <strong>the</strong> TAS diagram for <strong>the</strong> Dokhan<br />

volcanics using some published data. All <strong>the</strong> lava samples plot in <strong>the</strong><br />

M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

subalkaline field <strong>of</strong> Irvine and Baragar (1971) on <strong>the</strong> TAS diagram<br />

(Fig. 3). Moreover, all <strong>the</strong> samples have clear calc-alkaline affinity, as<br />

shown in <strong>the</strong> AFM diagram (Fig. 4A). K2O contents <strong>of</strong> <strong>the</strong> studied<br />

volcanics are plotted vs. SiO 2 in <strong>the</strong> diagram <strong>of</strong> Le Maitre (1989)<br />

shown in Fig. 4B. In this diagram, most <strong>of</strong> <strong>the</strong> samples exhibit apparent<br />

strong high-K affinity, with few medium-K samples. This is consistent<br />

with <strong>the</strong> prevalent high potassic nature <strong>of</strong> Dokhan volcanics <strong>from</strong><br />

o<strong>the</strong>r areas, particularly in <strong>the</strong> more evolved dacite to rhyolite lavas<br />

(e.g. Mohamed et al., 2000; Moghazi, 2003; El Sayed et al., 2004;<br />

Eliwa, et al., 2006). With <strong>the</strong> exception <strong>of</strong> one diopside–normative<br />

sample (Q7), all samples are corundum–normative, and most <strong>of</strong> <strong>the</strong>m<br />

have molar A/CNK ratios greater than one, indicating a dominant<br />

peraluminous character <strong>of</strong> <strong>the</strong>se volcanics (Table 1). Likewise, felsic<br />

members <strong>of</strong> <strong>the</strong> Dokhan volcanic suites in <strong>the</strong> Eastern Desert have<br />

been demonstrated to be peraluminous (Ressetar and Monard, 1983).<br />

The Dokhan volcanics <strong>of</strong> <strong>the</strong> Eastern Desert occurrences form a<br />

compositionally continuous suite <strong>from</strong> basalt to rhyolite (Basta et al.,<br />

1980; Abdel-Rahman, 1996; Mohamed et al., 2000; Saleh, 2003;<br />

Moghazi, 2003; El Sayed et al., 2004; Eliwa, et al., 2006) and<br />

occasionally a bimodal suite with silica gap between mafic and felsic<br />

rocks at <strong>the</strong> sou<strong>the</strong>rnmost Dokhan localities (Ressetar and Monard,<br />

1983; Stern and Gottfried, 1986; Stern et al., 1988). In contrast, <strong>the</strong><br />

investigated Dokhan volcanics and o<strong>the</strong>rs in Sinai (see Table 3) have<br />

narrow evolved compositional range (rhyolite–dacite). Fur<strong>the</strong>r, <strong>the</strong><br />

Dokhan volcanics <strong>from</strong> Eastern Desert have average dacitic composition<br />

with average silica (63–68%) and total alkalis (6.7–7.5 wt.%) while<br />

those <strong>from</strong> Sinai, including <strong>the</strong> studied volcanics, have an average<br />

rhyolitic composition (see Table 3).<br />

The studied Dokhan volcanics are relatively evolved with <strong>the</strong>ir SiO 2<br />

contents ranging <strong>from</strong> 64.66 to 76.61 wt.% and <strong>the</strong> differentiation<br />

index (DI, sum <strong>of</strong> normative Q+Or+Ab: Thornton and Tuttle, 1960)<br />

ranging <strong>from</strong> 94.73 to 76.03 (Table 1). Also, <strong>the</strong>y are marked by high<br />

contents <strong>of</strong> total alkalis (6.6–9.63 wt.%; average–8.25 wt.%) and a<br />

Table 3<br />

Comparison <strong>of</strong> average major and trace element concentrations <strong>of</strong> Kid-Malhak Dokhan volcanics with o<strong>the</strong>r Dokhan volcanics <strong>from</strong> Sinai and Eastern Desert <strong>of</strong> Egypt: 1 and 2 — Gabal<br />

Khashabi (Bentor and Eyal, 1987; Azzaz et al., 2000 respectively); 3 — Wadi Meknas (Hassan et al., 2001); 4 — Mount El-Kharaza (Abdel-Rahman, 1996); 5 — Wadi Um Sidra-Um<br />

Asmer (Eliwa et al, 2006); 6 — South Safaga area (Moghazi, 2003); 7 — Wadi Fatira and Qena-Safaga road (Ressetar and Monard, 1983); 8 — Fatira area (Mohamed et al., 2000).<br />

Location Sinai Eastern Desert Continental crust<br />

Reference This study 1 2 3 4 5 6 7 8 9<br />

No. <strong>of</strong> samples 16 5 7 18 24 35 26 59 34 –<br />

SiO2 71.41 71.64 73.46 68.43 66.21 65.31 64.68 63.46 67.98 59.1<br />

TiO2 0.57 0.54 0.42 0.52 0.65 0.87 0.86 1.14 0.66 0.7<br />

Al2O3 14.46 15.47 14.33 15.11 16.01 15.55 15.61 15.6 14.99 15.8<br />

FeO* 2.89 2.11 2 3.66 4.82 5.31 4.92 6.16 4.29 6.6<br />

MnO 0.07 0.6 0.48 0.11 1.45 0.08 0.11 2.82 0.08 0.11<br />

MgO 0.7 0.06 0.07 1.65 0.1 1.83 2.38 0.1 2.01 4.4<br />

CaO 1.51 1.02 0.88 1.78 3.07 3.41 3.64 4.02 3.04 6.4<br />

Na2O 4.22 4.65 3.77 4.43 4.36 4.3 4.42 4.14 3.64 3.2<br />

K2O 4.03 3.78 4.48 4.1 3.16 3.12 3.21 2.57 3.16 1.9<br />

P2O5 0.15 0.12 0.11 0.21 0.17 0.22 0.18 0.16 0.2<br />

Total 100 100 100 100 100 100 100 100 100 98.41<br />

Na2O+K2O 8.25 8.43 8.25 8.53 7.52 7.42 7.63 6.71 6.8 5.1<br />

Ni 3.7 21.72 31 44.5 55 24.68 51<br />

Cr 5.55 103 38.33 69 90.13 67 75.74 91<br />

V 32.09 30 33.83 112 112.88 203 60.68 131<br />

Ba 924.66 883 801 532 888.57 630 705.83 660 582 390<br />

Rb 88.01 108 112 121.22 72.78 84 85.67 57.8 59.43 58<br />

Sr 243.47 359 266 191 330.58 524 433.33 723 404 325<br />

Hf 8.39 4.76 5.21 3.7<br />

Zr 286.09 112 279 236 202.13 239 232.71 258 155 123<br />

Y 31.04 21 31.83 20.98 24 30.29 27.6 16.68 20<br />

Nb 13.95 15 16.43 11.43 15.63 10.58 14.7 10.21 12<br />

Zn 44.38 39 64.78 37.21 73<br />

Pb 10.48 15.72 12.6 17.21 12.6<br />

Th 12.04 6.89 8.2 7.5 16.61 5.6<br />

U 3.46 8.52 2.4 4.14 2.43 1.42<br />

∑REE 207 155 160 104 103.05<br />

Average composition <strong>of</strong> <strong>the</strong> continental crust (reference 9) is after Rudnick and Fountain (1995).<br />

31


32 M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

Fig. 4. (A) AFM diagram for <strong>the</strong> studied volcanics showing distinction between tholeiitic<br />

and calc-alkaline suites (Irvine and Baragar, 1971). (B) SiO 2 vs. K 2O diagram (Le Maitre,<br />

1989). Field for Dokhan volcanics (DV) as in Fig. 3.<br />

wide range <strong>of</strong> Na2O/K2O ratios (0.35–1.73). Overall, <strong>the</strong>y are<br />

characterized by ra<strong>the</strong>r wide range <strong>of</strong> <strong>the</strong> major oxides; TiO 2, (0.19–<br />

1.11%), Al2O3 (11.7–16.23%), FeO* (0.72–5.5%), MgO (0.15–1.48%), MnO<br />

(0.03–0.13%) and CaO (0.62–2.44%) with <strong>the</strong> higher contents in<br />

dacites and <strong>the</strong> lower in <strong>the</strong> more evolved rhyolites. Major elements<br />

Harker diagrams (Fig. 5), show a tight compositional gap between<br />

64.87 and 66.95 wt.% SiO2 values. The studied volcanics, exhibit a<br />

general negative correlation between silica and Al 2O 3, FeO*, MgO,<br />

P2O5, CaO and TiO2. These negative trends vary between fairly linear<br />

(MgO, CaO and P 2O 5), and inflected or segmented trends (Al 2O 3, TiO 2<br />

and FeO*). The depicted similar inflections in TiO2 and FeO* occur<br />

around silica value <strong>of</strong> 70%, CaO and MgO display much broader scatter<br />

towards <strong>the</strong> SiO2-rich end <strong>of</strong> <strong>the</strong> diagrams, which is attributed most<br />

likely to mixing processes. Al 2O 3 shows an exclusive zigzag pattern<br />

with intermediary period <strong>of</strong> alumina augmentation with silica rise,<br />

which can be attributed to a transitory dominant role <strong>of</strong> feldspar<br />

fractionation.<br />

4.2.2. Trace elements<br />

The trace element compositions <strong>of</strong> 16 representative samples<br />

<strong>from</strong> <strong>the</strong> Kid-Malhak Dokhan volcanics are given in Table 2. The<br />

investigated volcanics are relatively enriched in Ba (706–1390,<br />

average; 925 ppm), Y (14–49, average; 31 ppm), Zr (121–415,<br />

average; 286 ppm), and total REEs (134–249, average; 207 ppm)<br />

compared to o<strong>the</strong>r Dokhan volcanics, and also to average continental<br />

crustal rocks (Table 3). In contrast, <strong>the</strong>y are fairly depleted in Sr<br />

(100–528, average; 243 ppm). The most striking feature <strong>of</strong> <strong>the</strong><br />

studied volcanics is <strong>the</strong>ir serious depletion in <strong>the</strong> <strong>transition</strong> metals<br />

Cr (1–21 ppm) and Ni (0–16 ppm), even regarded as incompatible<br />

elements in felsic magmas. Silica variation diagrams <strong>of</strong> trace elements<br />

are illustrated in Fig. 6. Rb, Zr and Y show scattered linear to<br />

curvilinear trends, typified by slight enrichment with increasing <strong>of</strong><br />

silica. The HFS pair; Zr and Y show exceedingly coherent convex<br />

trends, in which both remain slightly varying through differentiation<br />

(SiO 2 rise), <strong>the</strong>n begin to increase gradually at about 75% silica value.<br />

This coherence may be explained in terms <strong>of</strong> <strong>the</strong> typical concentration<br />

<strong>of</strong> Y in accessory minerals, commonly zircon, as it replaces Zr.<br />

Nb, as alumina does, exhibit a zigzag-shaped plot, with a midway<br />

segment <strong>of</strong> rapid Nb declination with increasing silica. Ba show a<br />

wavy pattern, in which it reaches a climax <strong>of</strong> concentration around<br />

72 wt.% silica, <strong>the</strong>n start to fall again with increasing silica. This<br />

behavior <strong>of</strong> Ba, also observed in o<strong>the</strong>r fractionating calc-alkaline<br />

Dokhan volcanics in <strong>the</strong> Eastern Desert, has been documented by<br />

several authors (Mohamed et al., 2000; Moghazi, 2003; Eliwa et al.,<br />

2006). Sr concentrations are almost steady in dacites (bSiO 2 70 wt.%)<br />

but in contrary decrease over <strong>the</strong> entire rhyolite compositional range.<br />

The combination <strong>of</strong> linear, curvilinear and scattered trends on <strong>the</strong><br />

variation diagrams indicates that several processes influenced <strong>the</strong><br />

compositions <strong>of</strong> <strong>the</strong> rocks.<br />

Table 2 reveals that <strong>the</strong> studied rocks display a quite wide range <strong>of</strong><br />

Rb/Sr ratios (0.22–1.03) with an average <strong>of</strong> 0.46. Many <strong>of</strong> <strong>the</strong><br />

calculated Rb/Sr values are quite higher than lower and middle<br />

continental crust which have Rb/Sr ratios <strong>of</strong> 0.12 and 0.22 respectively<br />

(Rudnick and Fountain, 1995; Wedepohl,1995), indicating a role <strong>of</strong> <strong>the</strong><br />

upper continental crust in <strong>the</strong>ir genesis, as <strong>the</strong> crust can be regarded<br />

as being vertically zoned with Rb/Sr increasing upward (Blevin and<br />

Chappell, 1995). Y/Nb ratios <strong>of</strong> <strong>the</strong> studied volcanics vary between a<br />

minimum <strong>of</strong> 1.53 and a maximum <strong>of</strong> 2.76 (Table 2). Zr/Hf and Nb/Ta<br />

ratios for five samples are given in Table 2. With <strong>the</strong> exception <strong>of</strong><br />

sample Q10, <strong>the</strong> rest <strong>of</strong> <strong>the</strong> samples have Zr/Hf ratios within <strong>the</strong><br />

narrow range <strong>of</strong> 33 to 40 typifying most igneous rocks including <strong>the</strong><br />

values <strong>of</strong> chondrites (36) (Jochum et al., 1986; Dostal and Chatterjee,<br />

2000). Such deviation <strong>from</strong> chondritic value is rare and in this sample<br />

may be attributed to crystal fractionation involving accessory phases<br />

(Wolff, 1984).<br />

Incompatible trace-element spider diagram for <strong>the</strong> Kid-Malhak<br />

Dokhan volcanics, normalized to primitive sources (MORB <strong>of</strong> Taylor and<br />

McLennan (1985) is shown in Fig. 7. In this diagram, <strong>the</strong> pattern have<br />

ra<strong>the</strong>r steep slope, with obvious tilting <strong>of</strong> <strong>the</strong> trace element patterns up<br />

to left due to selective enrichment <strong>of</strong> <strong>the</strong> incompatible elements at <strong>the</strong><br />

left-hand side (LIL, Th, U and LREE) over <strong>the</strong> more compatible one to <strong>the</strong><br />

right (Y and HREE). HRRE are <strong>the</strong> most depleted, reaching levels<br />

comparable or even lower than those <strong>of</strong> MORB. Also, <strong>the</strong> pattern is<br />

characterized by serious negative Nb, Sr and Ti anomalies. The negative<br />

Nb anomaly is most characteristic <strong>of</strong> subduction zone volcanic rocks or<br />

typical continental crust. The negative anomalies for Sr and Ti may be<br />

attributed to fractionation <strong>of</strong> plagioclase and ilmenite and/or rutile<br />

respectively during petrogenesis. Marked positive spikes occur at Ba, U,<br />

La and Zr. Such spiked pattern may be a characteristic <strong>of</strong> subductionrelated<br />

magmas (Wilson, 1989). The more mobile Ba, U and La elements<br />

relative enrichment attests to <strong>the</strong> involvement <strong>of</strong> subduction zone fluids<br />

rich in <strong>the</strong>m (Wilson, 1989; Rollinson, 1993). The less mobile Zr<br />

enrichment is largely controlled by <strong>the</strong> chemistry <strong>of</strong> <strong>the</strong> source,<br />

suggesting derivation <strong>from</strong> a crustal source.


4.2.3. Rare earth elements<br />

The concentrations <strong>of</strong> rare earth elements (REE) <strong>of</strong> five Dokhan<br />

volcanic samples in ppm are given in Table 2. The analyzed samples<br />

have comparatively high abundances <strong>of</strong> total REEs (average;<br />

207 ppm) (Tables 2 and 3). They are characterized by moderate<br />

degree <strong>of</strong> REE fractionation, {(La/Yb) N=13.78–6.37}. La concentrations<br />

(31–60 ppm) vary between about 130 and 257 times chondrite<br />

value, while Lu (0.25–0.79 ppm) ranges between 10 and 32 times<br />

chondrite value. The degree <strong>of</strong> LREE fractionation is ra<strong>the</strong>r high with<br />

reasonably steep slope pattern {(La/Sm) N=3.07–4.78}, while <strong>the</strong><br />

heavy REE are weakly fractionated with approximately flat pattern<br />

{(Gd/Yb) N=1.31–1.79}. The REE patterns have reasonable concordance,<br />

with some difference in attitude due to variation in absolute<br />

abundances (Fig. 8). All <strong>the</strong> REE patterns have negative Eu anomalies<br />

and display concave upward shapes <strong>of</strong> quite negative slopes (Ce/<br />

Yb=19.95–39.76 and La/Yb=9.37–20.29) due to light REE enrich-<br />

M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

Fig. 5. Harker variation diagrams <strong>of</strong> some major oxides against SiO 2 (wt.%) for <strong>the</strong> Kid-Malhak Dokhan volcanics.<br />

ment relative to middle and heavy REE. The heavy REE are not highly<br />

depleted, but in turn are comparable or even sometimes slightly<br />

enriched relative to middle REE (i.e. sample Q6), suggesting absence<br />

<strong>of</strong> garnet in <strong>the</strong> source, since heavy REE are highly compatible in<br />

garnet (Wilson, 1989). The REE patterns <strong>of</strong> <strong>the</strong> studied samples<br />

display negative Eu anomalies <strong>of</strong> notably variable sizes (Eu/<br />

Eu*=0.39–0.75). These negative Eu anomalies are ascribed to <strong>the</strong><br />

removal <strong>of</strong> feldspar <strong>from</strong> <strong>the</strong> melt by crystal fractionation or <strong>the</strong><br />

partial melting <strong>of</strong> a rock in which feldspar is retained in <strong>the</strong> source<br />

(Wilson, 1989; Rollinson, 1993). Explanation <strong>of</strong> Eu anomalies solely on<br />

<strong>the</strong> basis <strong>of</strong> feldspars is an oversimplification, because o<strong>the</strong>r minerals<br />

have a role, although minor relative to feldspars, in producing Eu<br />

anomalies (Rollinson, 1993). Hornblende fractionation may contributes<br />

in part for reduction <strong>of</strong> <strong>the</strong> negative Eu anomaly sizes, because<br />

although REE are compatible in hornblende in felsic and intermediate<br />

liquids, Eu has a considerably lower partition coefficient relative to<br />

33


34 M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

most middle and heavy REE. The REE diagram in Fig. 8 shows that <strong>the</strong><br />

five REE patterns are nearly parallel LREE-enriched with negative Eu<br />

anomalies <strong>of</strong> variable sizes. The heaviest MREE and <strong>the</strong> HREE show<br />

steady decrease with increasing atomic number, giving rise to fairly<br />

flat segments <strong>of</strong> <strong>the</strong> patterns. Compared to o<strong>the</strong>r Dokhan volcanics<br />

<strong>from</strong> various localities in <strong>the</strong> Eastern Desert and Sinai, <strong>the</strong> REE<br />

patterns <strong>of</strong> <strong>the</strong> investigated Dokhan volcanics are very similar to those<br />

<strong>of</strong> <strong>the</strong> o<strong>the</strong>r Dokhan volcanics in <strong>the</strong> Eastern Desert and Sinai (Fig. 8).<br />

5. Discussion<br />

5.1. Zircon saturation <strong>the</strong>rmometry<br />

Fig. 6. Harker variation diagrams <strong>of</strong> some trace elements against SiO 2 (wt.%) for <strong>the</strong> Kid-Malhak Dokhan volcanics.<br />

Experimental studies have demonstrated that zircon, when<br />

crystalline, has low solubility in crustal melts and fluids (e.g., Watson,<br />

1979; Waston and Harrison, 1983; Ayers and Watson, 1991).<br />

Zirconium (Zr) commonly forms a separate phase, <strong>the</strong> mineral zircon<br />

(ZrSiO4). In felsic rocks, Zr and Hf are usually incompatible and are<br />

typically concentrated in residual silicate liquids until zircon saturation<br />

occurs (Watson, 1979; Waston and Harrison, 1983). Melt<br />

temperature estimates <strong>of</strong> <strong>the</strong> studied Dokhan volcanics are calculated<br />

and listed in Table 2 according to <strong>the</strong> experimental works <strong>of</strong> Waston<br />

and Harrison (1983) on Zr saturation in hydrous, low-temperature,<br />

intermediate and felsic magmas. They investigated <strong>the</strong> relation<br />

between zircon crystallization and melt composition which is given<br />

by <strong>the</strong> solubility model:<br />

ln Dzr Zircon=melt = f−3:80−½0:85ðM−1ÞŠg +12; 900 = T<br />

where: DzrZircon/melt is <strong>the</strong> concentration ratio <strong>of</strong> Zr in <strong>the</strong> stoichiometric<br />

zircon to Zr in <strong>the</strong> melt, T is <strong>the</strong> absolute temperature and M is<br />

<strong>the</strong> cation ratio (Na+K+2Ca)/(Al*Si). This geo<strong>the</strong>rmometer is based


on <strong>the</strong> fact that, in felsic rocks, Zr is an essential structural component<br />

in <strong>the</strong> accessory mineral zircon and its concentration depends on <strong>the</strong><br />

solubility <strong>of</strong> zircon <strong>the</strong> magma, which is a function <strong>of</strong> temperature,<br />

given a certain melt composition. Thus, assuming magma saturation<br />

in Zr, magma temperature can be calculated <strong>from</strong> a chemical analysis<br />

(Barrie, 1995; Rosa et al., 2006). Temperature values inferred range<br />

between a minimum <strong>of</strong> 766 °C and a maximum <strong>of</strong> 882 °C, with<br />

average temperature <strong>of</strong> 833 °C for <strong>the</strong> whole samples. However, recent<br />

sensitive high-resolution ion microprobe (SHRIMP) U–Pb dating <strong>of</strong><br />

zircons <strong>from</strong> <strong>the</strong> Dokhan volcanic suite at <strong>the</strong> type locality <strong>of</strong> Gabal<br />

Dokhan and at Gabal Urf, Wadi Qaa and Gabal Kharaza, Eastern Desert<br />

<strong>of</strong> Egypt (Wilde and Youssef, 2000; Breitkreuz et al., 2008 respectively)<br />

revealed a significant inherited zircon component. Therefore,<br />

and since some <strong>of</strong> <strong>the</strong> Zr in <strong>the</strong> studied lavas may have been in<br />

inherited zircon, <strong>the</strong> calculated temperatures are to be considered<br />

maximum temperatures <strong>of</strong> crustal fusion.<br />

5.2. Petrogenesis<br />

5.2.1. Fractional crystallization vs. crustal contamination<br />

It has been previously argued that <strong>the</strong> evolved Dokhan volcanic<br />

rocks (andesites, dacites and rhyolites), at many areas in <strong>the</strong> Eastern<br />

Desert <strong>of</strong> Egypt, were generated <strong>from</strong> more primitive basaltic<br />

precursors by crystal fractionation (Stern and Gottfried, 1986;<br />

Abdel-Rahman, 1996; Mohamed et al., 2000; Saleh, 2003; Moghazi,<br />

2003; El Sayed et al, 2004). Moreover, it is highly likely that crystal<br />

fractionation was coupled with assimilation <strong>of</strong> crustal materials (Stern<br />

and Gottfried, 1986, Moghazi, 2003).<br />

The systematic variation <strong>of</strong> <strong>the</strong> major and trace element contents<br />

<strong>of</strong> <strong>the</strong> investigated Dokhan lavas can reasonably be interpreted<br />

qualitatively in terms <strong>of</strong> fractional crystallization. This is indicated<br />

by <strong>the</strong> curvilinear trends <strong>of</strong> some elements in Harker diagrams,<br />

especially Zr and Y, and by <strong>the</strong> common turnover or change <strong>of</strong> slopes<br />

<strong>of</strong> <strong>the</strong> trends <strong>of</strong> many elements including; Al2O3 TiO2, FeO*, Ba, Sr, and<br />

Nb versus SiO 2 (Figs. 5 and 6). Also, major and trace element<br />

abundances vary along general trends <strong>of</strong> decreasing Al2O3, CaO, MgO,<br />

FeO*, P 2O 5, TiO 2 and Sr and increasing Rb, Y, and Zr (Figs. 5 and 6) with<br />

increasing SiO2. However, much <strong>of</strong> <strong>the</strong> broad scatter readily<br />

recognized in most <strong>of</strong> <strong>the</strong>se diagrams, is presumably attributed to<br />

crustal contamination processes. Additionally, <strong>the</strong> comparable and<br />

parallel configuration <strong>of</strong> <strong>the</strong> normalized REE patterns (Fig. 8) coupled<br />

with increasing total REE contents with increasing SiO2 (Tables 1 and<br />

2), imply a major role <strong>of</strong> crystal fractionation during <strong>the</strong> evolution <strong>of</strong><br />

<strong>the</strong>se rocks. Petrographically, <strong>the</strong> general absence <strong>of</strong> textural and<br />

compositional evidence for disequilibrium, mixing, or contamination<br />

supports derivation through fractional crystallization. The abovementioned<br />

characteristics support previous suggestions that <strong>the</strong><br />

Fig. 7. MORB-normalized multi-element patterns for Kid-Malhak Dokhan volcanics.<br />

Normalization values are <strong>from</strong> Taylor and McLennan (1985).<br />

M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

Fig. 8. Chondrite normalized REE diagram for Kid-Malhak Dokhan volcanics. C1chondrite<br />

normalization values are <strong>from</strong> McDonough and Sun (1995). For comparison, a<br />

compiled pattern is shown for Dokhan volcanics (DV) <strong>from</strong> o<strong>the</strong>r areas (Abdel-Rahman,<br />

1996; Mohamed et al, 2000; Moghazi, 2003; Eliwa, 2006).<br />

Dokhan volcanics evolved predominantly through fractional crystallization<br />

<strong>of</strong> <strong>the</strong> petrographically observed phenocryst assemblage,<br />

which is plagioclase+K-feldspar+biotite+hornblende+magnetite.<br />

<strong>On</strong> <strong>the</strong> o<strong>the</strong>r hand, it is significant to evaluate whe<strong>the</strong>r or not <strong>the</strong><br />

studied Dokhan volcanics have undergone crustal contamination. This<br />

is because mafic rocks are absent and all <strong>of</strong> <strong>the</strong>se felsic lavas are<br />

porphyritic indicating that <strong>the</strong>y might have resided in crustal magma<br />

chambers prior to eruption and would thus have had adequate<br />

opportunity to interact with continental crust. For this intention,<br />

evidences <strong>from</strong> trace element ratios for <strong>the</strong> examined volcanics are<br />

considered. During closed system fractionation, <strong>the</strong> ratio <strong>of</strong> strongly to<br />

moderately incompatible elements will remain constant or slightly<br />

increase, whereas during partial melting process, this ratio dramatically<br />

increases with decreasing degree <strong>of</strong> melting. Davidson et al.<br />

(1988) argued that ratios, such as K/Rb, Ba/Nb and Rb/Zr, do not<br />

significantly change by simple fractional crystallization, whereas<br />

variations in <strong>the</strong>se ratios are preferably related to crustal contamination<br />

by assimilation–fractional crystallization (ACF; De Paolo, 1981)<br />

processes. Examination <strong>of</strong> <strong>the</strong> investigated rocks (Fig. 9) shows that<br />

<strong>the</strong>re is a substantial variation between <strong>the</strong> samples in <strong>the</strong>se ratios,<br />

which may reach five orders <strong>of</strong> magnitude (i.e. Rb/Zr and Ba/Nb). This<br />

implies that crustal contamination did play a major role in <strong>the</strong><br />

evolution <strong>of</strong> <strong>the</strong> studied Dokhan volcanics. Additionally, <strong>the</strong>se rocks<br />

feature distinct negative Nb anomalies which suggest involvement <strong>of</strong><br />

crustal materials and crustal contamination. Moreover, a useful index<br />

to differentiate between crystal fractionation and crustal contamination<br />

is given in Fig. 10 using trace element ratios such as Nb/Y vs. Rb/Y<br />

(Chazot and Bertrand, 1995). In such diagram, crustal compositions<br />

and crustal melts are distinguished by high Rb/Nb ratios (N1) (Chazot<br />

and Bertrand, 1995). In <strong>the</strong> Nb/Y vs. Rb/Y diagram, all <strong>the</strong> Kid-Malhak<br />

lavas have exceedingly high Rb/Nb ratios (N3) and plot relatively close<br />

to lower and upper crustal values (Rudnick and Fountain, 1995).<br />

Therefore, crustal contamination is most probably superimposed on<br />

crystal fractionation in <strong>the</strong> evolution <strong>of</strong> Kid-Malhak Dokhan volcanics.<br />

Fur<strong>the</strong>rmore, this possible contributions <strong>from</strong> <strong>the</strong> crust to magma<br />

compositions (e.g. assimilation–fractional crystallization (AFC) process<br />

(De Paolo,1981), as well as crustal contamination phenomena are<br />

fur<strong>the</strong>r investigated using <strong>the</strong> Th/Nb elemental ratios plotted vs. Zr<br />

(Fig. 11). In this diagram a LILE/HFSE ratio (Th/Nb), which should not<br />

be significantly affected by fractional crystallization, are plotted<br />

against an incompatible trace element (Zr) whose abundance<br />

increases as fractionation increases. Such a diagram allows <strong>the</strong><br />

distinction between magmatic evolutions controlled by simple<br />

fractional crystallization in closed systems, that will produce an<br />

almost horizontal trend with increasing fractionation, and those<br />

influenced by additional processes. Kid-Malhak Dokhan volcanics<br />

35


36 M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

Fig. 9. SiO 2 vs. Ba/Nb, Rb/Zr and K/Rb for <strong>the</strong> Kid-Malhak Dokhan volcanics showing <strong>the</strong><br />

variation <strong>of</strong> Ba/Nb, Rb/Zr and K/Rb values with increasing SiO 2.<br />

define trends characterized by fairly elevated Th/Nb elemental ratios<br />

(Fig. 11), implying that assimilation–fractional crystallization (AFC)<br />

and crustal contamination have played a major role during magmatic<br />

evolution <strong>of</strong> Kid-Malhak Dokhan volcanics. After all, <strong>the</strong> recognition <strong>of</strong><br />

680 Ma xenocrysts in andesite samples <strong>of</strong> <strong>the</strong> Dokhan Volcanics at <strong>the</strong><br />

type locality area at Gabal Dokhan (Wilde and Youssef, 2000)<br />

additionally supports <strong>the</strong> idea <strong>of</strong> crustal contamination during<br />

evolution <strong>of</strong> <strong>the</strong>se rocks.<br />

5.2.2. Source material: crustal- vs. mantle-derived components<br />

Source materials <strong>from</strong> which <strong>the</strong> studied Dokhan magma were<br />

generated are highly likely crustal sources, with little contribution<br />

<strong>from</strong> mantle, which might acted mainly as a source <strong>of</strong> heat required<br />

for fusion <strong>of</strong> <strong>the</strong> crustal precursor. This conclusion is supported by<br />

<strong>the</strong>ir prevalent peraluminous nature, enrichment in alkalis, Ba and Th,<br />

Fig. 10. Nb/Y vs. Rb/Y diagram for <strong>the</strong> Kid-Malhak Dokhan volcanic rocks (modified after<br />

Chazot and Bertrand, 1995). The upper and lower crustal compositions are <strong>from</strong><br />

Rudnick and Fountain (1995).<br />

and depletion in Nb and HREE. Fur<strong>the</strong>r, MORB normalized trace<br />

element patterns (Fig. 7) show significant negative Sr, Nb and Ti. The<br />

negative Nb and Ti anomalies might suggest a derivation <strong>from</strong> a<br />

crustal source or crustal contamination coupled with advanced<br />

degrees <strong>of</strong> differentiation. Table 2 reveals that <strong>the</strong> studied rocks<br />

display a quite wide range <strong>of</strong> Rb/Sr ratios (0.22–1.03) with an average<br />

<strong>of</strong> 0.46. Taking into account that mantle materials have very low Rb/Sr<br />

ratios, 0.1–0.01, (Taylor and McLennan, 1985; H<strong>of</strong>mann, 1988) while<br />

lower and middle continental crust have Rb/Sr ratios <strong>of</strong> 0.12 and 0.22<br />

respectively (Rudnick and Fountain, 1995; Wedepohl, 1995), <strong>the</strong> Rb/Sr<br />

ratios <strong>of</strong> <strong>the</strong> studied Dokhan volcanics negate mantle material<br />

contribution and imply genuine crustal components on <strong>the</strong>ir genesis.<br />

Y/Nb ratios <strong>of</strong> <strong>the</strong> studied volcanics vary between a minimum <strong>of</strong> 1.53<br />

and a maximum <strong>of</strong> 2.76 (Table 2). This provides unambiguous<br />

evidence <strong>of</strong> <strong>the</strong>ir exclusive derivation <strong>from</strong> crustal material, in view<br />

<strong>of</strong> <strong>the</strong> fact that mantle sources have Y/Nb ratios less than 1.2, while Y/<br />

Nb ratios greater than 1.2 characterizes materials <strong>of</strong> crustal origin<br />

(Eby, 1990, 1992). A typical primitive mantle value for Nb/Ta ratio is<br />

accepted to be 17.5±2.0 (H<strong>of</strong>mann, 1988; Green, 1995). However,<br />

<strong>the</strong>re is a controversy whe<strong>the</strong>r a typical continental crust value is<br />

within <strong>the</strong> range <strong>of</strong> <strong>the</strong> mantle-derived melts (H<strong>of</strong>mann et al., 1986;<br />

Jochum et al., 1986) or is significantly lower (~11–12; Taylor and<br />

Fig. 11. Zr vs. Th/Nb plot for Kid-Malhak Dokhan volcanic rocks. Schematic trends<br />

reflecting increasing fractional crystallization (FC), assimilation–fractional crystallization<br />

(AFC), and bulk assimilation (BA) are <strong>from</strong> Nicolae and Saccani (2003).


McLennan, 1985; Green, 1995; Barth et al., 2000). The author favors<br />

<strong>the</strong> second opinion, typical continental crust Na/Ta ratio approximates<br />

11, which implies that <strong>the</strong>se two elements in <strong>the</strong> studied<br />

samples (Nb/Ta; 10.85–14.64) were fractionated <strong>from</strong> each o<strong>the</strong>r in<br />

<strong>the</strong> continental crust. I suggest that, earlier accreted arc sequences <strong>of</strong><br />

<strong>the</strong> juvenile <strong>Arabian–Nubian</strong> <strong>Shield</strong> crust probably melted to give rise<br />

to this calc-alkaline suite, a view supported by <strong>the</strong> presence <strong>of</strong><br />

inherited zircons and older xenocrysts (680 Ma) in <strong>the</strong> Dokhan<br />

volcanics <strong>from</strong> <strong>the</strong> Eastern Desert (Wilde and Youssef, 2000;<br />

Breitkreuz et al., 2008).<br />

5.3. Tectonic setting<br />

The Kid-Malhak Dokhan volcanics display many <strong>of</strong> <strong>the</strong> chemical<br />

characteristics <strong>of</strong> <strong>the</strong> magmas formed at subduction-related settings.<br />

This includes <strong>the</strong>ir; calc-alkaline nature, strong enrichment <strong>of</strong> LIL<br />

elements (Cs, Rb Ba, Th and U) and depletion in HFS elements (Zr, Hf,<br />

Nb, Y, Ti) relative to typical MORB values, negative Nb anomalies, light<br />

REE-enriched patterns and weak fractionation <strong>of</strong> MREE and HREE.<br />

Also, <strong>the</strong> high-K calc-alkaline affinity strongly points to <strong>the</strong>ir eruption<br />

at continental (active continental margin), ra<strong>the</strong>r than oceanic setting<br />

(Pearce et al., 1984). However, <strong>the</strong>y have slightly higher contents <strong>of</strong><br />

alkalis, Nb, Zr and Y than typical island arc or continental margins<br />

volcanic suites. The studied Dokhan volcanic samples are plotted on a<br />

series <strong>of</strong> tectonic discrimination diagrams (Fig. 12), which are<br />

appropriate for magmas <strong>of</strong> felsic composition. Nb is one <strong>of</strong> <strong>the</strong> most<br />

effective elements to identify whe<strong>the</strong>r an acid igneous rock has a<br />

within-plate (high Nb) or volcanic arc (low Nb) character. The studied<br />

Dokhan lavas are sub-equally plotted in both volcanic arc field and <strong>the</strong><br />

overlap zone between <strong>the</strong> volcanic arc and within plate settings on <strong>the</strong><br />

binary SiO 2–Nb diagram <strong>of</strong> Pearce and Gale (1977) (Fig. 12A),<br />

suggesting a <strong>transition</strong>al tectonic setting. Pearce (1982) suggested<br />

<strong>the</strong> use <strong>of</strong> a simple bivariant plot <strong>of</strong> Ti against Zr in intermediate and<br />

acid volcanic rocks for recognition volcanic arc and within plate<br />

settings. He also stated that <strong>the</strong> volcanic arc lavas never reach <strong>the</strong><br />

higher Ti contents <strong>of</strong> <strong>the</strong> within plate lavas for any given Zr<br />

concentration, due to <strong>the</strong> early crystallization <strong>of</strong> magnetite. Fig. 12B<br />

shows <strong>the</strong> relationship between <strong>the</strong>se two immobile elements, where<br />

<strong>the</strong> studied volcanics are dually plotted in <strong>the</strong> within-plate and<br />

volcanic arc lava fields. The Dokhan samples plot in <strong>the</strong> “volcanic arc<br />

granite” field on <strong>the</strong> binary Y+Nb–Rb diagram <strong>of</strong> Pearce et al. (1984)<br />

(Fig. 12C), except for two samples, which fall in <strong>the</strong> “within-plate”<br />

field. Accordingly, <strong>the</strong> Kid-Malhak Dokhan lavas have geochemical<br />

characteristics <strong>of</strong> both subduction-related and within-plate settings, a<br />

phenomenon previously noted by many authors and has led to a<br />

dispute on <strong>the</strong> convergent margin versus within-plate geotectonic<br />

setting for <strong>the</strong> Dokhan volcanics. Since determination <strong>of</strong> <strong>the</strong> tectonic<br />

environment for <strong>the</strong> Dokhan volcanic rocks is one <strong>of</strong> <strong>the</strong> main<br />

objectives <strong>of</strong> this work, previous tectonic models will be debated in<br />

order to explain <strong>the</strong>ir implication and relationship to <strong>the</strong> emplacement<br />

<strong>of</strong> <strong>the</strong> Dokhan volcanics in Egypt.<br />

5.3.1. Arc-related model<br />

In accordance with <strong>the</strong>ir geochemical characteristics, several<br />

workers (Basta et al., 1980; Hassan and Hashad, 1990; El Gaby et al.,<br />

1990; Abdel-Rahman, 1996; Blasy, 2000; Hassan et al., 2001; Saleh,<br />

2003) show that <strong>the</strong> Dokhan volcanics are calc-alkaline in nature and<br />

<strong>the</strong>ir trace element and REE contents resemble those <strong>of</strong> arc-related<br />

orogenic suites and thus <strong>the</strong>y were erupted during <strong>the</strong> final stages <strong>of</strong><br />

active subduction beneath <strong>the</strong> Eastern Desert <strong>of</strong> Egypt. Never<strong>the</strong>less,<br />

<strong>the</strong>re is some regional evidence that supports <strong>the</strong> termination <strong>of</strong><br />

subduction processes before <strong>the</strong>ir eruption:<br />

1- The Younger and <strong>the</strong> alkaline granites intrusions in Egypt are<br />

massive and un-deformed post-orogenic to anorogenic granites<br />

with many <strong>of</strong> <strong>the</strong> former and exclusively all <strong>of</strong> <strong>the</strong> later having<br />

M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

typical A-type characteristics (Stern and Gottfried, 1986; Abdel-<br />

Rahman and Martin, 1990; Hassan and Hashad, 1990; Moghazi,<br />

1999; El-Bialy and El Omlla, 2007; El-Bialy and Streck, 2009;<br />

Farahat et al, 2007; Katzir et al., 2007; Moussa et al., 2008). The<br />

reported ages for <strong>the</strong>se post-orogenic to anorogenic granitoid<br />

assemblages range between 610 and 475 Ma. (Hashad, 1980;<br />

Fig. 12. Discrimination diagrams illustrating tectonic setting <strong>of</strong> <strong>the</strong> studied volcanics:<br />

(A) SiO 2 vs. Nb diagram (Pearce and Gale,1977); (B) Zr vs. TiO 2 (Pearce,1980); (C) Yb+Nb<br />

vs. Rb (Pearce et al., 1984). The A-type granites field in (C) is after Whalen et al. (1987).<br />

37


38 M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

Bielski, 1982; Bentor, 1985; Stern and Hedge, 1985; Abdel-Rahman<br />

and Martin, 1990; Beyth et al., 1994; Moghazi, 1999; Katzir et al.<br />

2007; Moussa et al., 2008; Ali et al., 2009; Be'eri-Shlevin et al.,<br />

2009b). The range <strong>of</strong> emplacement ages <strong>of</strong> <strong>the</strong> spatially associated<br />

Dokhan volcanics (620–580 Ma) matches and falls within that <strong>of</strong><br />

those granitoids. Assumption <strong>of</strong> <strong>the</strong> eruption <strong>of</strong> <strong>the</strong> Dokhan<br />

volcanics in orogenic arc-related setting implies two different<br />

tectonic environments for a single province during <strong>the</strong> same time<br />

period, which is impossible.<br />

2- Geological and geochronological studies carried out at different<br />

parts in <strong>the</strong> Egyptian basement complex have documented two<br />

tectono-magmatic episodes in <strong>the</strong> Neoproterozoic evolution <strong>of</strong><br />

Egypt. The first tectono-magmatic episode (between 700 and<br />

720 Ma) represents <strong>the</strong> age <strong>of</strong> ophiolite obduction and island arc<br />

terrain accretion (Kröner et al., 1992; Stern, 1994; Kröner et al.,<br />

1994; Harms et al., 1994; Stern, 2002; Kusky et al., 2003), while <strong>the</strong><br />

accreted terrane collided with <strong>the</strong> East Saharan Metacraton during<br />

<strong>the</strong> second episode (between 650 and 620 Ma) (Kröner et al.,1994;<br />

Sultan et al., 1994; Stern, 1994; Stern, 2002; Abdelsalam and Stern,<br />

1996). Greiling et al. (1994) suggested that collision ceased at 615–<br />

600 Ma and <strong>the</strong> extensional collapse commenced sometime<br />

between <strong>the</strong> 600 and 575 Ma time period, followed by a stage <strong>of</strong><br />

transpressional tectonism along major shear zones until 530 Ma.<br />

The un-deformed character <strong>of</strong> <strong>the</strong> Dokhan volcanics along with<br />

<strong>the</strong>ir emplacement ages (580–620 Ma; see Section 1), which<br />

synchronize to post-date <strong>the</strong> end <strong>of</strong> <strong>the</strong> collisional stage, suggest<br />

that <strong>the</strong>ir eruption had taken place at <strong>the</strong> <strong>transition</strong> between<br />

collision and extension (post-collisional).<br />

3- Plotting <strong>of</strong> <strong>the</strong> Dokhan lava samples on <strong>the</strong> Y+Nb–Rb diagram <strong>of</strong><br />

Pearce et al. (1984), shows that <strong>the</strong> majority <strong>of</strong> <strong>the</strong> samples fall in<br />

<strong>the</strong> field <strong>of</strong> A-type granite <strong>of</strong> Whalen et al. (1987) on <strong>the</strong> same<br />

diagram (Fig. 12C). Moreover, plots <strong>of</strong> <strong>the</strong> samples on <strong>the</strong> on Zr+<br />

Nb+Ce+Y versus (K2O+Na 2O)/CaO and FeO*/MgO diagrams<br />

(Whalen et al., 1987) reveal that <strong>the</strong> majority <strong>of</strong> <strong>the</strong> studied<br />

samples fall in <strong>the</strong> A-type granite field (not shown here). The same<br />

phenomenon is also reported for Dokhan volcanics <strong>from</strong> different<br />

parts <strong>of</strong> <strong>the</strong> Eastern Desert (e.g. Mohamed et al., 2000; Eliwa et al.,<br />

2006). The prevalent A-type character <strong>of</strong> <strong>the</strong> Dokhan volcanics<br />

indicate that <strong>the</strong>y were emplaced in “non-orogenic” environment,<br />

ei<strong>the</strong>r post-collision or true anorogenic within-plate setting (Eby,<br />

1992), and positively not in compressional arc-related one.<br />

4- The emplacement <strong>of</strong> <strong>the</strong> Dokhan volcanics in <strong>the</strong> Eastern Desert<br />

and Sinai (580–620 Ma) were almost contemporaneous with <strong>the</strong><br />

predominance <strong>of</strong> dyke swarms (Stern et al., 1983; Stern and Hedge,<br />

1985; Stern and Gottfried, 1986; Stern and Manton, 1987; Stern<br />

et al., 1988; Friz-Töpfer, 1991; El-Sayed, 2006), <strong>the</strong> major postaccerationary<br />

(640–540) Najd strike-slip fault system and shear<br />

zones in <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong> (Stacey and Agar, 1985; Stern,<br />

1985; Sultan et al. 1993; Abdelsalam and Stern, 1996, Fritz et al.,<br />

1996; Kusky and Matsah, 2003), and exhumation <strong>of</strong> metamorphic<br />

core complexes at about 600–580 Ma (Fritz et al., 1996; Fritz et al.,<br />

2002; Blasband et al., 1997; Blasband et al., 2000; Brooijmans et al,<br />

2003) preclude <strong>the</strong> prevalence <strong>of</strong> compression tectonic regime and<br />

indicate that extensional tectonics were active during <strong>the</strong> Dokhan<br />

volcanics formation.<br />

5.3.2. Continental rifting model<br />

Several workers considered <strong>the</strong> A-type character <strong>of</strong> <strong>the</strong> felsic<br />

members <strong>of</strong> Dokhan volcanics, linked with <strong>the</strong> synchronous emplacement<br />

<strong>of</strong> <strong>the</strong> Dokhan volcanics with <strong>the</strong> high level A-type granite<br />

intrusions, toge<strong>the</strong>r with <strong>the</strong> bimodal nature <strong>of</strong> some Dokhan volcanic<br />

suites in <strong>the</strong> Eastern Desert as evidences <strong>of</strong> strong crustal extension in<br />

an anorogenic within-plate rifting environment, analogous to that <strong>of</strong><br />

<strong>the</strong> Neoproterozoic Oslo Rift <strong>of</strong> Norway (Stern et al., 1984; Stern et al.,<br />

1988; Stern and Gottfried, 1986; Willis et al, 1988; Mohamed et al.,<br />

2000). Never<strong>the</strong>less, <strong>the</strong>re are various oppositions to <strong>the</strong> emplace-<br />

ment <strong>of</strong> <strong>the</strong> Dokhan volcanics in au<strong>the</strong>ntic anorogenic rift-related<br />

settings.<br />

1- The bimodal nature <strong>of</strong> Dokhan lava suites, which is used as a criteria<br />

to substantiate <strong>the</strong> anorogenic rifting model for <strong>the</strong> genesis <strong>of</strong> this<br />

rock unit is not <strong>the</strong> norm, but only restricted to <strong>the</strong> sou<strong>the</strong>rnmost<br />

Dokhan localities(i.e. Safaga-Qena and S. Safaga areas) (Eliwa et al.,<br />

2006). In most <strong>of</strong> <strong>the</strong> localities in <strong>the</strong> Eastern Desert, Dokhan<br />

volcanics form a compositionally continuous suite (Basta et al.,<br />

1980; Abdel-Rahman, 1996; Blasy, 2000; Eliwa et al., 2006; Saleh,<br />

2003; Moghazi, 2003; El Sayed et al., 2004). From o<strong>the</strong>r side, this<br />

bimodality, if accepted as a typical nature <strong>of</strong> Dokhan volcanics,<br />

cannot be applied as an explicit evidence <strong>of</strong> anorogenic rifting, since<br />

it is now evident that many arc magmatism is bimodal in<br />

composition, and lacks andesitic magmas (Tamura and Tatsumi,<br />

2002; Leat et al., 2003; Smith et al., 2003).<br />

2- The A-type character <strong>of</strong> <strong>the</strong> felsic members <strong>of</strong> <strong>the</strong> Dokhan volcanic<br />

suites does not crucially imply anorogenic rift-related setting. It is<br />

now widely accepted that <strong>the</strong> A-type magmas can be generated in<br />

post-orogenic and anorogenic settings (Whalen et al., 1987; Eby,<br />

1990, 1992; Whalen et al., 1996). According to Eby (1990, 1992), Atype<br />

magmas are divided into A1 and A2 groups. The A1 group (true<br />

anorogenic) refers to differentiates <strong>of</strong> magmas derived <strong>from</strong> sources<br />

like those <strong>of</strong> oceanic–island basalts (OIB) but emplaced in continental<br />

rifts or during intra-plate magmatism. The A2 group (post-orogenic<br />

or post-collisional), on <strong>the</strong> o<strong>the</strong>r hand, represents magmas derived<br />

<strong>from</strong> continental crust or underplated crust that has been through a<br />

cycle <strong>of</strong> continent–continent collision or island–arc magmatism.<br />

3- Correspondence between <strong>the</strong> tectonic setting <strong>of</strong> <strong>the</strong> Dokhan<br />

volcanics and Oslo Rift is an oversimplification. Basaltic lavas <strong>of</strong><br />

Oslo rift, and o<strong>the</strong>r continental rift systems, have major and trace<br />

element concentrations comparable to ocean island basalts (OIB),<br />

indicating <strong>the</strong> participation <strong>of</strong> as<strong>the</strong>nospheric mantle in <strong>the</strong>ir<br />

genesis (Neumann, 1994; Neumann et al., 2002). Conversely and<br />

for example, <strong>the</strong> studied Dokhan lavas have major and trace<br />

element abundances and characteristics, which are very similar to<br />

subduction-related magmas and imply greater contribution <strong>from</strong><br />

crustal sources ra<strong>the</strong>r than as<strong>the</strong>nospheric mantle.<br />

5.3.3. Post-collision extensional collapse (<strong>the</strong> suggested model)<br />

From <strong>the</strong> previous discussion, it is obvious, that <strong>the</strong> Kid-Malhak<br />

Dokhan lavas have geochemical characteristics <strong>of</strong> both orogenic arctype<br />

and anorogenic within-plate environments, suggesting eruption<br />

in a <strong>transition</strong>al tectonic setting. Fig. 12C reveals that all samples were<br />

emplaced in a “post-collisional” regime ra<strong>the</strong>r than volcanic arc or<br />

within-plate settings. The same conclusion can also be achieved<br />

through <strong>the</strong>ir prevalent A-type character combined with <strong>the</strong>ir Nb/Y<br />

ratios (N1.2), which characterize <strong>the</strong> post-orogenic (post-collision) A2<br />

subtype magmas (Eby, 1990, 1992). Post-collision magmatism is<br />

linked to rapid post-closure uplift, which may take place some time<br />

after <strong>the</strong> collision itself, and which is also linked to <strong>the</strong> subsequent<br />

collapse <strong>of</strong> <strong>the</strong> orogenic belt (Pearce, 1996). For clarification <strong>of</strong> <strong>the</strong><br />

post-collision stage <strong>of</strong> <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong> (ANS) evolution,<br />

during which <strong>the</strong> Dokhan volcanics had been emplaced, a summary <strong>of</strong><br />

<strong>the</strong> <strong>Shield</strong>'s tectonic evolution is given below:<br />

1- Rifting <strong>of</strong> Rodinia super-continent led to <strong>the</strong> formation <strong>of</strong> <strong>the</strong><br />

Mozambique Ocean. Intra-oceanic subduction gave rise to <strong>the</strong><br />

development <strong>of</strong> island-arcs, back-arc basins and ophiolitic sequences<br />

(Fig. 13A) (El Gaby et al., 1984; Bentor, 1985; Pallister et al., 1988;<br />

Stern, 1994) at approximately 900–750 Ma in <strong>the</strong> Mozambique<br />

ocean (Stern, 1994; Abdelsalam and Stern, 1996; Abdelsalam et al.,<br />

2008; Rino et al., 2008; Vaughan and <strong>Pan</strong>khurst, 2008).<br />

2- Subduction at continental margins led to accretion <strong>of</strong> island-arcs<br />

and ophiolite remnants in <strong>the</strong> Mozambique Ocean and <strong>the</strong><br />

amalgamation <strong>of</strong> accreted terrains onto east Gondwana continental<br />

block. Terrane accretion in <strong>the</strong> ANS took place along north to east


trending arc–arc sutures developed between 700 and 800 Ma. The<br />

arc accretion marked <strong>the</strong> closure <strong>of</strong> Mozambique Ocean, and was<br />

responsible for fast continental crustal growth and lithosphere<br />

thickening <strong>of</strong> <strong>the</strong> juvenile ANS crust. (Pallister et al., 1988; Kröner<br />

M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

Fig. 13. Cartoons displaying <strong>the</strong> major stages <strong>of</strong> <strong>the</strong> evolution <strong>of</strong> <strong>the</strong> Arabian Nubian <strong>Shield</strong> (A and B), with special emphasis on <strong>the</strong> generation <strong>of</strong> <strong>the</strong> Dokhan volcanics during <strong>the</strong><br />

post-collisional stage (C and D) (modified after Blasband et al., 2000 and Davies and von Blanckenburg, 1995). The oceanic phase is shown in [A], with island-arcs developing in <strong>the</strong><br />

Mozambique ocean. Oceanic crust was formed at <strong>the</strong> MOR, in fore-arcs and back-arcs. At stage [B], “<strong>of</strong>f-shore amalgamation” or arc-accretion and subduction at <strong>the</strong> continental<br />

margins took place and accretion <strong>of</strong> <strong>the</strong> island-arcs and superterranes started. In [C]; <strong>the</strong> start <strong>of</strong> extensional collapse and slab break<strong>of</strong>f <strong>of</strong> <strong>the</strong> oceanic lithosphere as a consequence <strong>of</strong><br />

<strong>the</strong> collision between <strong>the</strong> accreted juvenile <strong>Arabian–Nubian</strong> <strong>Shield</strong> crust with <strong>the</strong> Pre-Neoproterozoic Saharan Metacraton to <strong>the</strong> west after <strong>the</strong> closure <strong>of</strong> Mozambique Ocean.<br />

Following subduction, rifting <strong>of</strong> <strong>the</strong> oceanic lithosphere allows as<strong>the</strong>nosphere to rise and fill <strong>the</strong> rift; slab weakening follows with underplating <strong>of</strong> continental crust. In [D]; slab<br />

detaches and sinks away. High temperatures maintained at <strong>the</strong> base <strong>of</strong> lithosphere by deep return flow causing partial melting <strong>of</strong> <strong>the</strong> underplate to form Dokhan magmas, which<br />

ascend to <strong>the</strong> surface through <strong>the</strong> crust.<br />

et al., 1992, 1994; Stern, 1994, 2002; Abdelsalam and Stern, 1996;<br />

Rino et al., 2008; Condie et al., 2009).<br />

3- The juvenile accreted ANS crust collided with pre-Neoproterozoic<br />

continental blocks <strong>of</strong> west Gondwana (Saharan Metacraton;<br />

39


40 M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

Abdelsalam et al., 2002) at 750–650 Ma along arc–continental<br />

sutures (Fig. 13b) (Kröner et al., 1994; Sultan et al., 1994; Stern,<br />

1994, 2002; Abdelsalam and Stern, 1996; Stoeser and Frost, 2006;<br />

Stern, 2008).<br />

4- Collision ceased at ca. 615–600 Ma and <strong>the</strong> extensional collapse <strong>of</strong><br />

<strong>the</strong> thickened lithosphere took place at ca. 600–560 Ma. Extensional<br />

or gravitational collapse led to extension and thinning <strong>of</strong> <strong>the</strong><br />

<strong>Arabian–Nubian</strong> <strong>Shield</strong> crust, which continued until approximately<br />

530 Ma. (Greiling et al.,1994; Blasband et al.,1997, 2000; Brooijmans<br />

et al., 2003).<br />

The ages <strong>of</strong> emplacement <strong>of</strong> <strong>the</strong> Dokhan volcanics in Egypt<br />

including that <strong>of</strong> Kid-Malhak region (609±12 Ma; Bielski, 1982)<br />

coincide with end <strong>of</strong> collision and <strong>the</strong> extensional collapse event. The<br />

post-collision <strong>transition</strong> <strong>from</strong> a compressional arc-accretion setting to<br />

an extensional regime is best explained by <strong>the</strong> extensional collapse<br />

model (Dewey, 1988). Extensional collapse follows continental<br />

collision and is controlled mainly by lithospheric delamination and<br />

slab break<strong>of</strong>f (passive rifting) and not by rifting controlled by <strong>the</strong><br />

ascent <strong>of</strong> <strong>the</strong> as<strong>the</strong>nospheric mantle material (active rifting) as in <strong>the</strong><br />

Oslo rift in Norway (see Fig. 13C, D). The extensional collapse model<br />

has been argued and documented in several parts <strong>of</strong> <strong>the</strong> world (e.g.<br />

Dewey, 1988; Davies and von Blanckenburg, 1995; Cosca et al, 1999;<br />

A<strong>the</strong>rton and Ghani, 2002; Clift et al., 2004; Abdelsalam et al., 2008;<br />

Casini and Oggiano, 2008; Sun et al, 2008; Jöns et al., 2009) and also in<br />

<strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong> (e.g. Blasband et al.,<br />

1997, 2000; Brooijmans et al., 2003; Moghazi, 2003).<br />

“Slab break<strong>of</strong>f” is <strong>the</strong> buoyancy-driven detachment <strong>of</strong> subducted<br />

dense oceanic lithosphere <strong>from</strong> <strong>the</strong> light continental lithosphere that<br />

follows it during continental collision (Davies and von Blanckenburg,<br />

1995). This situation <strong>of</strong> opposing buoyancy forces leads to extensional<br />

deformation (narrow rifting) in <strong>the</strong> subducted slab. As a result<br />

<strong>of</strong> <strong>the</strong> rifting during break<strong>of</strong>f, <strong>the</strong> as<strong>the</strong>nosphere upwells into <strong>the</strong><br />

narrow rift, and following break<strong>of</strong>f it impinges on <strong>the</strong> mechanical<br />

lithosphere <strong>of</strong> <strong>the</strong> overriding plate (Davies and von Blanckenburg,<br />

1995; Sperner et al., 2001; Coulon et al., 2002; Heidbach et al 2008;<br />

Regard et al, 2008). The generation <strong>of</strong> <strong>the</strong> Dokhan Volcanic magma<br />

can thus be regarded as a consequence <strong>of</strong> continental collision<br />

between <strong>the</strong> juvenile arc-accreted crust <strong>of</strong> <strong>the</strong> <strong>Arabian–Nubian</strong><br />

<strong>Shield</strong> and <strong>the</strong> pre-Neoproterozoic Saharan Metacraton to <strong>the</strong> west<br />

(Fig. 13C). As <strong>the</strong> oceanic slab falls away (Fig. 13D), more <strong>of</strong> <strong>the</strong><br />

enriched mantle would be exposed, in a manner similar to a plume<br />

impinging on lithosphere, and may lead to limited <strong>the</strong>rmal erosion <strong>of</strong><br />

<strong>the</strong> lithosphere (Monnereau et al., 1993). The overlying continental<br />

lithosphere is heated by conduction as <strong>the</strong> as<strong>the</strong>nosphere impinges<br />

on its base leading to melting <strong>of</strong> <strong>the</strong> metasomatised, veined and<br />

hydrated layers (A<strong>the</strong>rton and Ghani, 2002). This part <strong>of</strong> lithosphere<br />

may have obtained unique trace element characteristics owing to its<br />

interaction with fluids and melts driven <strong>of</strong>f <strong>from</strong> <strong>the</strong> subducting slab<br />

during previous long subduction events (i.e. between 900 and<br />

620 Ma; Stern, 1994) in <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong> (Moghazi, 2003).<br />

Expected melts, according to Davies and von Blanckenburg (1995),<br />

could be alkaline to ultrapotassic <strong>from</strong> small degree melts or calcalkaline<br />

<strong>from</strong> slightly higher degree melting <strong>of</strong> more fertile or<br />

hydrated peridotite layers. These magmas will rise into <strong>the</strong> crust,<br />

pass through and induce crustal melting to produce granitic magma.<br />

Crustal melting will be enhanced by <strong>the</strong>rmotectonic processes, e.g.<br />

heat conduction after shallow break<strong>of</strong>f, or heating and decompression<br />

on rapid unro<strong>of</strong>ing (Davies and von Blanckenburg, 1995). In<br />

agreement with Moghazi (2003), <strong>the</strong> calc-alkaline nature <strong>of</strong> <strong>the</strong><br />

Dokhan volcanics can be explained by high degree partial melting <strong>of</strong><br />

<strong>the</strong> juvenile <strong>Arabian–Nubian</strong> <strong>Shield</strong> crust. Even though this model<br />

needs to be evaluated geodynamically, it seems to be <strong>the</strong> most<br />

conceivable for <strong>the</strong> stratigraphic position and geochemical characteristics<br />

<strong>of</strong> <strong>the</strong> Dokhan volcanics in Egypt and o<strong>the</strong>r comparable<br />

magmatic rocks formed at <strong>the</strong> post-collisional <strong>transition</strong>al stage<br />

between arc-accretion and crustal extension in <strong>the</strong> <strong>Arabian–Nubian</strong><br />

<strong>Shield</strong>.<br />

5.4. Dokhan volcanics: a <strong>transition</strong> between calc-alkaline and alkaline<br />

magmatism?<br />

The gradual <strong>transition</strong> <strong>from</strong> an “orogenic” calc-alkaline magmatism<br />

to a continental intra-plate alkaline suite has been described in several<br />

regions worldwide (e.g. North Africa; El Bakkali et al., 1998; Coulon<br />

et al., 2002; Turkey and Anatolia; Wilson et al., 1997; Aldanmaz et al.,<br />

2000; Agostini et al., 2007; Sou<strong>the</strong>rn India; Rajesh, 2004; Sardinia:<br />

Lustrino et al., 2007; Carpathian <strong>Pan</strong>nonian Region; Harangi, 2001;<br />

Konecny et al., 2002; Seghedi et al., 2005; Central Africa Belt;<br />

Tetsopgang et al., 2008). This <strong>transition</strong> may follow ei<strong>the</strong>r a continental<br />

collision (e.g. Turkey, Iran), or <strong>the</strong> cessation <strong>of</strong> subduction at a<br />

previously convergent margin, for instance, as a result <strong>of</strong> a ridge–<br />

trench collision (e.g. California; Cole and Basu, 1992). In <strong>the</strong> nor<strong>the</strong>rnmost<br />

part <strong>of</strong> <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong>, in Sinai, Jordan and<br />

sou<strong>the</strong>rn Israel, such <strong>transition</strong> is also recorded (Beyth et al., 1994;<br />

Jarar et al., 2003; Stern, 2008; Ali et al., 2009; Be'eri-Shlevin et al.,<br />

2009a,b). The late evolutionary stages <strong>of</strong> <strong>the</strong> nor<strong>the</strong>rn <strong>Arabian–Nubian</strong><br />

<strong>Shield</strong> (650–545 Ma) document <strong>the</strong> growth and maturation <strong>of</strong> <strong>the</strong><br />

continental crust <strong>from</strong> an orogen to a craton. During this period<br />

magmatism changed <strong>from</strong> calc-alkaline to alkaline and <strong>the</strong> tectonic<br />

setting in which <strong>the</strong>se magmas emplaced, have changed <strong>from</strong> collision<br />

to post-collision, to within-plate extension and finally to a stable<br />

platform setting. The nature <strong>of</strong> <strong>transition</strong> <strong>from</strong> calc-alkaline to alkaline<br />

magmatism, whe<strong>the</strong>r abrupt or gradual, and its correlation to field<br />

evidence for initiation <strong>of</strong> extension in <strong>the</strong> nor<strong>the</strong>rn <strong>Arabian–Nubian</strong><br />

<strong>Shield</strong> are poorly resolved. Beyth et al. (1994) and Ali et al. (2009)<br />

referred to <strong>the</strong> <strong>transition</strong> between calc-alkaline and within-plate<br />

alkaline magmatism in sou<strong>the</strong>rn Israel and Sinai respectively, and<br />

suggested that it occurred at c. 610 Ma, which is concordant with <strong>the</strong><br />

emplacement ages <strong>of</strong> <strong>the</strong> Dokhan volcanics in Egypt.<br />

The nor<strong>the</strong>rnmost <strong>Arabian–Nubian</strong> shield comprising <strong>the</strong> Eastern<br />

Desert <strong>of</strong> Egypt, Sinai, sou<strong>the</strong>rn Jordan and Israel is characterized by<br />

vast and prolonged (~820–570 Ma) intrusive and extrusive magmatism.<br />

Magmatic rocks were divided to (1) syn- to late-orogenic I-type<br />

and (2) anorogenic, A-type, <strong>the</strong> <strong>transition</strong> between <strong>the</strong> two types<br />

occurring at ~610–600 Ma (Bentor, 1985; Beyth et al., 1994; Stern,<br />

1994; Jarar et al., 2003; Moussa et al., 2008; Ali et al., 2009; Be'eri-<br />

Shlevin et al., 2009a,b). The recognition <strong>of</strong> a widespread high-K calcalkaline<br />

suite, formed at about <strong>the</strong> time <strong>of</strong> this <strong>transition</strong>, and partially<br />

overlapping <strong>the</strong> beginning <strong>of</strong> alkaline A-type magmatism (Be'eri-<br />

Shlevin et al., 2009a,b), redefines different stages <strong>of</strong> calc-alkaline<br />

magma production in this region. The magmatic calc-alkaline pulses<br />

thus include (a) early medium-K calc-alkaline plutons (now gneisses)<br />

associated with metavolcanics and metasediments, and older synorogenic<br />

I-type granitoids, all <strong>of</strong> arc-type affinity that formed at ~820–<br />

740 Ma, (b) a late syn-collisional medium-K to high-K calc alkaline<br />

suite (many <strong>of</strong> <strong>the</strong> younger granites <strong>of</strong> Egypt) formed during late<br />

stages <strong>of</strong> accretion (~670 Ma to 635–625 Ma) and (c) post-collisional<br />

mainly high-K calc-alkaline suite <strong>of</strong> 620–580 Ma (<strong>the</strong> Dokhan<br />

volcanics and latest phases <strong>of</strong> <strong>the</strong> Egyptian younger granites) that<br />

overlaps <strong>the</strong> alkaline suite <strong>of</strong> ~600–470 Ma (post-collisional and<br />

anorogenic A-type granites) for ~20 m.y. The temporal <strong>transition</strong> <strong>from</strong><br />

medium-K through high-K calc-alkaline to alkaline magmatism is<br />

correlated with <strong>the</strong> change in tectonic regime (Bentor, 1985; Stern and<br />

Hedge, 1985; Beyth et al., 1994; Jarar et al., 2003; Moussa et al., 2008,<br />

Ali et al., 2009; Be'eri-Shlevin et al., 2009a,b) and implies ei<strong>the</strong>r a<br />

change in magma sources or a change in <strong>the</strong> processes that formed<br />

magmas during 820–580 Ma. Therefore, <strong>the</strong> eruption <strong>of</strong> <strong>the</strong> high-K<br />

calc-alkaline post-collisional Dokhan volcanics defines a tectonomagmatic<br />

<strong>transition</strong> between <strong>the</strong> older calc-alkaline arc-related<br />

magmatism and <strong>the</strong> subsequent alkaline magmatism in <strong>the</strong> nor<strong>the</strong>rn<br />

part <strong>of</strong> <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong>.


6. Conclusions<br />

The concluding remarks, based on <strong>the</strong> field, petrographic and<br />

geochemical investigations presented here, are as follows:<br />

1- The <strong>Pan</strong>-<strong>African</strong> Dokhan volcanic suite in Kid-Malhak region (609±<br />

12 Ma; Bielski, 1982), extrudes and overlies o<strong>the</strong>r older metamorphosed<br />

basement units including <strong>the</strong> gneisses, metasediments and<br />

metavolcanics, and conversely intruded by <strong>the</strong> younger granites. A<br />

younger to contemporaneous thick interbedded epiclastic conglomerate<br />

beds (100–150 m), separate <strong>the</strong>ir succession into upper and<br />

lower stratigraphic sequences, and are most probably stratigraphically<br />

equivalent to <strong>the</strong> Hammamat Group (Akaad and Noweir, 1980)<br />

clastic sediments <strong>of</strong> <strong>the</strong> Egyptian Eastern Desert.<br />

2- The Dokhan volcanics in <strong>the</strong> study area consist <strong>of</strong> non-metamorphosed<br />

varicolored alternating succession <strong>of</strong> porphyritic lava flows<br />

<strong>of</strong> commonly felsic composition interlayered with compositionally<br />

equivalent pyroclastic beds (dominantly ignimbrites). Lava flows<br />

predominate in this volcanic suite, comprising chiefly rhyolites<br />

with subordinate dacites and trachydacites.<br />

3- The studied Dokhan volcanics are quite evolved (SiO 2≈65–77 wt.<br />

%), with strong high-K calc-alkaline affinity and are characterized<br />

by relative enrichment in total alkalis, Ba, Y, Zr and total REEs, and<br />

fair depletion in Sr. Also, <strong>the</strong>y are characterized by a LREE-enriched<br />

REE patterns with significant negative Eu anomalies.<br />

4- The striking difference between <strong>the</strong> studied Dokhan volcanics and<br />

o<strong>the</strong>rs in Sinai (El Metwally et al., 1999; Azzaz et al., 2000; Hassan<br />

et al., 2001) and those <strong>of</strong> <strong>the</strong> Eastern Desert <strong>of</strong> Egypt, is <strong>the</strong>ir<br />

evolved felsic compositional range (rhyolite-dacite) in comparison<br />

with <strong>the</strong> “bimodal” or continuous spectral composition <strong>of</strong> <strong>the</strong><br />

Dokhan volcanics in <strong>the</strong> Eastern Desert.<br />

5- The Kid-Malhak Dokhan lavas have geochemical characteristics <strong>of</strong><br />

both orogenic arc-type and anorogenic within-plate environments,<br />

suggesting eruption in a <strong>transition</strong>al “post-collisional tectonic setting.<br />

The ages <strong>of</strong> emplacement <strong>of</strong> <strong>the</strong> Dokhan volcanics in Egypt including<br />

that <strong>of</strong> Kid-Malhak region (580–620 Ma) coincide with end <strong>of</strong> <strong>the</strong><br />

documented collision between <strong>the</strong> juvenile <strong>Arabian–Nubian</strong> crust and<br />

Saharan Metacraton and <strong>the</strong> subsequent extensional collapse event.<br />

The post-collision <strong>transition</strong> <strong>from</strong> a compressional arc-accretion<br />

setting to an extensional regime is explained by <strong>the</strong> extensional<br />

collapse following continental collision and is controlled mainly by<br />

lithospheric delamination and slab break<strong>of</strong>f (passive rifting).<br />

6- Trace element characteristics (Nb depletion and <strong>the</strong> variation <strong>of</strong><br />

<strong>the</strong> K/Rb, Ba/Nb, Rb/Nb, Th/Nb and Rb/Zr ratios), strongly indicate<br />

that assimilation–fractional crystallization (AFC) and crustal<br />

contamination have played a major role and are most probably<br />

superimposed on fractional crystallization during <strong>the</strong> magmatic<br />

evolution <strong>of</strong> Kid-Malhak Dokhan volcanics.<br />

7- Source materials <strong>from</strong> which <strong>the</strong> studied Dokhan magmas were<br />

generated are highly likely crustal sources. This conclusion is<br />

supported by <strong>the</strong>ir prevalent peraluminous nature, enrichment in<br />

alkalis Ba and Th, depletion in Nb, Ti and HREE, high Rb/Sr and Y/<br />

Nb and low Nb/Ta ratios comparable to crustal rock, and finally <strong>the</strong><br />

presence <strong>of</strong> inherited zircons and older xenocrysts in o<strong>the</strong>r Dokhan<br />

volcanics in <strong>the</strong> Eastern Desert.<br />

8- Application <strong>of</strong> zircon saturation <strong>the</strong>rmometry revealed temperatures<br />

ranging between a minimum <strong>of</strong> 766 °C and a maximum <strong>of</strong><br />

882 °C. These calculated temperatures cannot simply regarded as<br />

crystallization temperatures, but are to be considered maximum<br />

temperatures <strong>of</strong> crustal fusion, since inherited zircon components<br />

are reported in o<strong>the</strong>r Dokhan volcanic suites <strong>from</strong> Eastern Desert.<br />

9- The eruption <strong>of</strong> <strong>the</strong> high-K calc-alkaline post-collisional Dokhan<br />

volcanics defines a tectono-magmatic <strong>transition</strong> between <strong>the</strong> older<br />

calc-alkaline arc-related magmatism and <strong>the</strong> subsequent alkaline<br />

anorogenic magmatism in <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> <strong>Arabian–Nubian</strong><br />

<strong>Shield</strong>.<br />

M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

Acknowledgments<br />

Sincere thanks are extended to Dr. I. H. Khalifa (Geology Department,<br />

Suez Canal University, Ismailia, Egypt) for assistance during field work. I<br />

greatly appreciate <strong>the</strong> comments and suggestions <strong>of</strong> Pr<strong>of</strong>essor M. Santosh<br />

(Editor-in-Chief; Gondwana research). Also, I gratefully acknowledge <strong>the</strong><br />

thorough and constructive comments <strong>of</strong> Dr. H.M. Rajesh and <strong>of</strong> an<br />

anonymous reviewer that seriously improved <strong>the</strong> manuscript.<br />

References<br />

Abdel-Rahman, A.M., 1996. <strong>Pan</strong>-<strong>African</strong> volcanism: petrology and geochemistry <strong>of</strong> <strong>the</strong><br />

Dokhan Volcanic suite in <strong>the</strong> nor<strong>the</strong>rn Nubian <strong>Shield</strong>. Geological Magazine 133,17–31.<br />

Abdel-Rahman, A.M., Doig, R., 1987. The Rb–Sr geochronological evolution <strong>of</strong> <strong>the</strong> Ras<br />

Gharib segment <strong>of</strong> <strong>the</strong> nor<strong>the</strong>rn Nubian shield. Journal <strong>of</strong> <strong>the</strong> Geological Society <strong>of</strong><br />

London 144, 577–586.<br />

Abdel-Rahman, A.M., Martin, R.F., 1987. Late <strong>Pan</strong>-<strong>African</strong> magmatism and crustal<br />

development in nor<strong>the</strong>astern Egypt. Geological Journal 22, 281–301.<br />

Abdel-Rahman, A.M., Martin, R.F., 1990. The Mount Gharib A-type granite, Nubian<br />

<strong>Shield</strong>: petrogenesis and role <strong>of</strong> metasomatism at <strong>the</strong> source. Contribution to<br />

Mineralogy and Petrology 104, 173–183.<br />

Abdelsalam, M.G., Stern, R.J., 1996. Sutures and shear zones in <strong>the</strong> Arabian Nubian<br />

<strong>Shield</strong>. Journal <strong>of</strong> <strong>African</strong> Earth Sciences 23, 289–310.<br />

Abdelsalam, M.G., Li'egeois, J., Stern, R.J., 2002. The Saharan metacraton. Journal <strong>of</strong><br />

<strong>African</strong> Earth Sciences 34, 119–136.<br />

Abdelsalam, M.G., Tsige, L., Yihunie, T., Hussien, B., 2008. Terrane rotation during <strong>the</strong><br />

East <strong>African</strong> Orogeny: evidence <strong>from</strong> <strong>the</strong> Bulbul Shear Zone, south Ethiopia.<br />

Gondwana Research 14, 497–508.<br />

Agostini, S., Doglioni, C., Innocenti, F., Manetti, P., Tonarini, S., Yilmaz Savas_cin, M.,<br />

2007. The <strong>transition</strong> <strong>from</strong> subduction-related to intraplate Neogene magmatism in<br />

<strong>the</strong> Western Anatolia and Aegean area. In: Beccaluva, L., Bianchini, G., Wilson, M.<br />

(Eds.), Cenozoic Volcanism in <strong>the</strong> Mediterranean Area: Geological Society <strong>of</strong><br />

America Special Paper, 418, pp. 1–15.<br />

Akaad, M.K., Noweir, A.M., 1980. Geology and lithostratigraphy <strong>of</strong> <strong>the</strong> Arabian Desert<br />

orogenic belt <strong>of</strong> Egypt between latitudes 25° 35′ N and 26° 30′ N. Institute Applied<br />

Geology Jeddah Bulletin 3, 127–135.<br />

Aldanmaz, E., Pearce, J.A., Thirlwall, M.F., Mitchell, J.G., 2000. Petrogenetic evolution <strong>of</strong><br />

late Cenozoic, post-collision volcanism in western Anatolia, Turkey. Journal <strong>of</strong><br />

Volcanology and Geo<strong>the</strong>rmal Research 102, 67–95.<br />

Ali, B.H., Wilde, S.A., Gabr, M.M.A., 2009. Granitoid evolution in Sinai, Egypt, based on<br />

precise SHRIMP U–Pb zircon geochronology. Gondwana Research 15, 38–48.<br />

A<strong>the</strong>rton, M.P., Ghani, A.A., 2002. Slab break<strong>of</strong>f: a model for Caledonian, Late Granite<br />

syn-collisional magmatism in <strong>the</strong> orthotectonic (metamorphic) zone <strong>of</strong> Scotland<br />

and Donegal. Ireland. Lithos 62, 65–85.<br />

Ayers, J.C., Watson, E.B., 1991. Solubility <strong>of</strong> apatite, monazite, zircon and rutile in<br />

supercritical aqueous fluids with implications for subduction zone geochemistry.<br />

Philosophical Transactions. Royal Society <strong>of</strong> London 335, 365–375.<br />

Azzaz, S.A., El Baroudy, A.F., Abd Allah, S.E., 2000. Volcano-sedimentary association and<br />

Dokhan volcanics <strong>of</strong> <strong>the</strong> northwestern Sharm El-Sheikh, Sinai, Egypt. Egyptian<br />

Mineralogist 12, 217–245.<br />

Barrie, C.T., 1995. Zircon <strong>the</strong>rmometry <strong>of</strong> high-temperature rhyolites near volcanicassociated<br />

massive sulfide deposits Abitibi subprovince, Canada. Geology 23,169–172.<br />

Barth, M.G., McDonough, W.F., Rudnick, R.L., 2000. Tracking <strong>the</strong> budget <strong>of</strong> Nb and Ta in<br />

<strong>the</strong> continental crust. Chemical Geology 165, 197–213.<br />

Basta, E.Z., Kotb, H., Awadalla, M.F., 1980. Petrochemical and geochemical characteristics<br />

<strong>of</strong> <strong>the</strong> Dokhan Formation at <strong>the</strong> type locality, Jabal Dokhan, Eastern Desert, Egypt.<br />

Institute Applied Geology Jeddah Bulletin 3, 121–140.<br />

Be'eri-Shlevin, Y., Katzir, Y., Whitehouse, M.J., 2009a. Post-collisional tectono-magmatic<br />

evolution in <strong>the</strong> nor<strong>the</strong>rn Arabian Nubian <strong>Shield</strong> (ANS): time constraints <strong>from</strong> ionprobe<br />

U–Pb dating <strong>of</strong> zircon. Journal <strong>of</strong> <strong>the</strong> Geological Society <strong>of</strong> London 166, 1–15.<br />

Be'eri-Shlevin, Y., Katzir, Y., Valley, J.W., 2009b. Crustal evolution and recycling in a<br />

juvenile continent: oxygen isotope ratio <strong>of</strong> zircon in <strong>the</strong> nor<strong>the</strong>rn Arabian Nubian<br />

<strong>Shield</strong>. Lithos 107, 169–184.<br />

Bentor, Y.K., 1985. The crustal evolution <strong>of</strong> <strong>the</strong> Arabo-Nubian Massif with special<br />

reference to <strong>the</strong> Sinai Peninsula. Precambrian Research 28, 1–74.<br />

Bentor, Y.K., Eyal, M., 1987. The geology <strong>of</strong> Sou<strong>the</strong>rn Sinai, its implication for <strong>the</strong><br />

evolution <strong>of</strong> <strong>the</strong> <strong>Arabian–Nubian</strong> Massif Volume 1: Jebel Sabbagh Sheet. Israel<br />

Academy <strong>of</strong> Sciences and Humanities, Jerusalem. 484p.<br />

Beyth, M., Stern, R.J., Al<strong>the</strong>rr, R., Kröner, A., 1994. The late Precambrian Timna igneous<br />

complex, sou<strong>the</strong>rn Israel: evidence for comagmatic-type sanukitoid monzodiorite<br />

and alkali granite magma. Lithos 31, 103–124.<br />

Bielski, M., 1982. Stages in <strong>the</strong> <strong>Arabian–Nubian</strong> Massif in Sinai. Ph.D. Thesis, Hebrew<br />

University, Jerusalem. 155 pp (unpublished).<br />

Blasband, B., Brooijmans, P., Dirks, P., Visser, W., White, S., 1997. A <strong>Pan</strong>-<strong>African</strong> core<br />

complex in <strong>the</strong> Sinai, Egypt. Geologie en Mijnbouw 73, 247–266.<br />

Blasband, B., White, S., Brooijmans, P., De Boorder, H., Visser, W., 2000. Neoproterozoic<br />

extensional collapse in <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong>. Journal <strong>of</strong> <strong>the</strong> Geological Society<br />

<strong>of</strong> London 157, 615–628.<br />

Blasy, M., 2000. Petrogenesis <strong>of</strong> Safaga Dokhan volcanics, Eastern Desert, Egypt.<br />

Egyptian Journal <strong>of</strong> Geology 44, 101–113.<br />

Blevin, P.L., Chappell, B.W., 1995. Chemistry, origin and evolution <strong>of</strong> mineralized<br />

granites in <strong>the</strong> Lachlan Fold Belt Australia: <strong>the</strong> metallogeny <strong>of</strong> I- and S-type granites.<br />

Economic Geology 90, 1604–1619.<br />

41


42 M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

Breitkreuz, C., Eliwa, H., Khalaf, I., El Gameel, K., Sergeev, S., Larionov, A., H<strong>of</strong>fmann, U.,<br />

2008. Neoproterozoic SHRIMP U–Pb zircon ages <strong>of</strong> Dokhan volcanics in <strong>the</strong><br />

nor<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> Eastern Desert, Egypt. 33rd International Geological Congress<br />

Oslo 2008 August 6–14th. Abstract.<br />

Brooijmans, P., Blasband, B., White, S.H., Visser, W.J., Dirks, P., 2003. Geo<strong>the</strong>rmobarometric<br />

evidence for a metamorphic core complex in Sinai, Egypt. Precambrian<br />

Research 123, 249–268.<br />

Cas, R.A.F., Wright, J.V., 1987. Volcanic Successions, Modern and Ancient. Allen and<br />

Unwin, London. 528 pp.<br />

Casini, L., Oggiano, G., 2008. Late orogenic collapse and <strong>the</strong>rmal doming in <strong>the</strong> nor<strong>the</strong>rn<br />

Gondwana margin incorporated in <strong>the</strong> Variscan Chain: a case study <strong>from</strong> <strong>the</strong> Ozieri<br />

Metamorphic Complex, nor<strong>the</strong>rn Sardinia, Italy. Gondwana Research 13, 396–406.<br />

Chazot, G., Bertrand, H., 1995. Genesis <strong>of</strong> silicic magmas during Tertiary continental<br />

rifting in Yemen. Lithos 36, 69–83.<br />

Clift, P.D., Dewey, J.F., Drautt, A.E., Chew, D.M., Mange, M., Ryan, P.D., 2004. Rapid<br />

tectonic exhumation, detachment faulting and orogenic collapse in <strong>the</strong> Caledonides<br />

<strong>of</strong> western Ireland. Tectonophysics 384, 91–113.<br />

Cole, R.B., Basu, A., 1992. Middle Tertiary volcanism during ridge–trench interactions in<br />

western California. Science 258, 793–796.<br />

Condie, K.C., Belousova, E., Griffin, W.L., Sircombe, K.N., 2009. Granitoid events in space and<br />

time: constraints <strong>from</strong> igneous and detrital zircon age spectra. Gondwana Research 15,<br />

228–242.<br />

Cosca,M.A.,Shimron,A.,Caby,R.,1999.LatePrecambrianmetamorphismandcooling<br />

in <strong>the</strong> <strong>Arabian–Nubian</strong> <strong>Shield</strong>: petrology and 40 Ar/ 39 Ar geochronology <strong>of</strong><br />

metamorphic rocks <strong>of</strong> <strong>the</strong> Elat area (sou<strong>the</strong>rn Israel). Precambrian Research 98,<br />

107–127.<br />

Coulon, C., Megartsi, M.H., Fourcade, S., Maury, R.C., Bellon, H., Louni-Hacini, A., Cotten, J.,<br />

Coutelle, A., Hermitte, D., 2002. Post-collisional <strong>transition</strong> <strong>from</strong> calc-alkaline to<br />

alkaline volcanism during <strong>the</strong> Neogene in Oranie (Algeria): magmatic expression <strong>of</strong><br />

a slab break<strong>of</strong>f. Lithos 62, 87–110.<br />

Davidson, J.P., Dungan, M.A., Ferguson, K.M., Colucci, M.T., 1988. Crust–magma interactions<br />

and <strong>the</strong> evolution <strong>of</strong> arc magmas: <strong>the</strong> San Pedro–Pellado Volcanic complex, sou<strong>the</strong>rn<br />

Chilean Andes. Geology 15, 443–446.<br />

Davies, J.H., Von Blanckenburg, F., 1995. Slab break<strong>of</strong>f: a model <strong>of</strong> lithosphere<br />

detachment and its test in <strong>the</strong> magmatism and deformation <strong>of</strong> collisional orogens.<br />

Earth and Planetary Science Letters 129, 85–102.<br />

De Paolo, D.J., 1981. Trace element and isotopic effects <strong>of</strong> combined wall-rock<br />

assimilation and fractional crystallization. Earth and Planetary Science Letters 53,<br />

189–202.<br />

Dewey, J.F., 1988. Extensional collapse <strong>of</strong> orogens. Tectonics 7, 1123–1139.<br />

Dostal, J., Chatterjee, A.K., 2000. Contrasting behaviour <strong>of</strong> Nb/Ta and Zr/Hf ratios in a<br />

peraluminous granitic pluton (Nova Scotia, Canada). Chemical Geology 163, 207–218.<br />

Eby, G.N., 1990. The A-type granitoids: a review <strong>of</strong> <strong>the</strong>ir occurrence and chemical<br />

characteristics and speculations on <strong>the</strong>ir petrogenesis. Lithos 26, 115–134.<br />

Eby, G.N., 1992. Chemical subdivision <strong>of</strong> <strong>the</strong> A-type granitoids: petrogenetic and<br />

tectonic implications. Geology 20, 641–644.<br />

El Bakkali, S., Gourgaud, A., Bourdier, J.L., Gundogdu, N., 1998. Post-collision Neogene<br />

volcanism <strong>of</strong> <strong>the</strong> Eastern Rif (Morocco): magmatic evolution through time. Lithos<br />

45, 523–543.<br />

El Gaby, S., El Nady, O., Khudeir, A., 1984. Tectonic evolution <strong>of</strong> <strong>the</strong> basement complex in<br />

<strong>the</strong> Central Eastern Desert <strong>of</strong> Egypt. Geologische Rundschau 73, 1019–1036.<br />

El Gaby, S., Khudeir, A.A., El Taky, M., 1989. The Dokhan Volcanics <strong>of</strong> Wadi Queih area,<br />

central Eastern Desert, Egypt. Proceedings <strong>of</strong> <strong>the</strong> 1st Conference on Geochemistry.<br />

Alexandria University, Egypt, pp. 42–62.<br />

El Gaby, S., List, F.K., Tehrani, R., 1990. The basement complex <strong>of</strong> <strong>the</strong> Eastern Desert and<br />

Sinai. In: Said, R. (Ed.), The Geology <strong>of</strong> Egypt. Balkema, Rotterdam, pp. 175–184.<br />

El Metwally, A.A., El Aasy, I.E., Ibrahim, M.E., Essawy, M.A., El Mowafy, A.A., 1999.<br />

Petrological, structural and geochemical studies on <strong>the</strong> basement rocks <strong>of</strong> Gabal Umm<br />

Zariq-Wadi Kid area, Sou<strong>the</strong>astern Sinai. Egyptian Journal <strong>of</strong> Geology 43, 147–180.<br />

El Sayed, M.M., Obeid, M.A., Furnes, H., Moghazi, A.M., 2004. Late Neoproterozoic<br />

volcanism in <strong>the</strong> sou<strong>the</strong>rn Eastern Desert, Egypt: petrological, structural and<br />

geochemical constraints on <strong>the</strong> tectonic-magmatic evolution <strong>of</strong> <strong>the</strong> Allaqi Dokhan<br />

volcanic suite. Neues Jahrbuch für Mineralogie Abhandlungen 180, 261–286.<br />

El Shazly, E.M., Hashad, A.H., Sayyah, T.A., Bassyuni, F.A., 1973. Geochronology <strong>of</strong> Abu<br />

Swayel area, South Eastern Desert. Egyptian Journal <strong>of</strong> Geology 17, 1–18.<br />

El-Bialy, M.Z., El Omlla, M.M., 2007. Source and tectonic setting <strong>of</strong> Sharm El-Sheikh alkaline<br />

A-type granitoids, Sinai, Egypt. Annals. Geological Survey <strong>of</strong> Egypt 29, 135–157.<br />

El-Bialy, M.Z., Streck, M.J., 2009. Late Neoproterozoic alkaline magmatism in <strong>the</strong><br />

<strong>Arabian–Nubian</strong> <strong>Shield</strong>: <strong>the</strong> postcollisional A-type granite <strong>of</strong> Sahara-Umm Adawi<br />

pluton, Sinai, Egypt. Arabian Journal <strong>of</strong> Geosciences 2, 151–174.<br />

Eliwa, H.A., Kimura, J.I., Itaya, T., 2006. Late Neoproterozoic Dokhan Volcanics, North Eastern<br />

Desert, Egypt: geochemistry and petrogenesis. Precambrian Research 151, 31–52.<br />

El-Sayed, M.M., 2006. Geochemistry and petrogenesis <strong>of</strong> <strong>the</strong> post-orogenic bimodal<br />

dyke swarms in NW Sinai, Egypt: constraints on <strong>the</strong> magmatic–tectonic processes<br />

during <strong>the</strong> late Precambrian. Chemie der Erde 66, 129–141.<br />

Farahat, E.S., Mohamed, H.A., Ahmed, A.F., El Mahallawi, M.M., 2007. Origin <strong>of</strong> I- and Atype<br />

granitoids <strong>from</strong> <strong>the</strong> Eastern Desert <strong>of</strong> Egypt: Implications for crustal growth in<br />

<strong>the</strong> nor<strong>the</strong>rn <strong>Arabian–Nubian</strong> <strong>Shield</strong>. Journal <strong>of</strong> <strong>African</strong> Earth Sciences 49, 43–58.<br />

Fritz, H., Wallbrecher, E., Khudeir, A.A., Abu El Ela, F., Dallmeyer, D.R., 1996. Formation <strong>of</strong><br />

Neoproterozoic metamorphic core complexes during oblique convergence (Eastern<br />

Desert, Egypt). Journal <strong>of</strong> <strong>African</strong> Earth Sciences 23, 311–329.<br />

Ressetar, R., Monard, J.R., 1983. Chemical composition and tectonic setting <strong>of</strong> <strong>the</strong><br />

Dokhan Volcanic formation, Eastern Desert, Egypt. Journal <strong>of</strong> <strong>African</strong> Earth Sciences<br />

1, 103–112.<br />

Fritz, H., Dallmeyer, D.R, Wallbrecher, E., Loizenbauer, J., Hoinkes, G., Neumayer, P.,<br />

Khudeir, A.A., 2002. Neoproterozoic tectono<strong>the</strong>rmal evolution <strong>of</strong> <strong>the</strong> Central<br />

Eastern Desert, Egypt: a slow velocity tectonic process <strong>of</strong> core complex exhumation.<br />

Journal <strong>of</strong> <strong>African</strong> Earth Sciences 34, 137–155.<br />

Friz-Töpfer, A., 1991. Geochemical characterization <strong>of</strong> <strong>Pan</strong>-<strong>African</strong> dyke swarms in<br />

sou<strong>the</strong>rn Sinai: <strong>from</strong> continental margin to intraplate magmatism. Precambrian<br />

Research 49, 281–300.<br />

Furnes, H., Shimron, A.E., Roberts, D., 1985. Geochemistry <strong>of</strong> <strong>Pan</strong>-<strong>African</strong> volcanic arc<br />

sequences in sou<strong>the</strong>astern Sinai Peninsula and plate tectonic implications.<br />

Precambrian Research 29, 359–382.<br />

Gillespie, G.J., Dixon, T.H., 1983. Lead isotope systematics <strong>of</strong> some igneous rocks <strong>from</strong><br />

<strong>the</strong> Egyptian <strong>Shield</strong>. Precambrian Research 20, 63–77.<br />

Green, T.H., 1995. Significance <strong>of</strong> Nb/Ta as an indicator <strong>of</strong> geochemical processes in <strong>the</strong><br />

crust–mantle system. Chemical Geology 120, 347–359.<br />

Greiling, R.O., Abdeen, M.M., Dardir, A.A., El Akhal, H., El Ramly, M.F., Kamal El Din, G.M.,<br />

Osman, A.F., Rashwan, A.A., Rice, A.H.N., Sadek, M.F.,1994. A structural syn<strong>the</strong>sis <strong>of</strong> <strong>the</strong><br />

Proterozoic <strong>Arabian–Nubian</strong> <strong>Shield</strong> in Egypt. Geologische Rundschau 83, 484–501.<br />

Harangi, S., 2001. Neogene to Quaternary Volcanism <strong>of</strong> <strong>the</strong> Carpathian <strong>Pan</strong>nonian<br />

Regionça review. Acta Geologica Hungarica 44, 223–258.<br />

Harms, U., Darbyshire, D.B.F., Denkler, T., Hengst, M., Schandelmeier, H., 1994. Evolution <strong>of</strong><br />

<strong>the</strong> Neoproterozoic Delgo suture zone and crustal growth in Nor<strong>the</strong>rn Sudan:<br />

geochemical and radiogenic isotope constraints. Geologische Rundschau 83, 591–603.<br />

Hashad, A.H., 1980. Present status <strong>of</strong> geochronological data on <strong>the</strong> Egyptian basement<br />

complex. Institute <strong>of</strong> Applied Geology Bulletin, Jeddah 3, 31–46.<br />

Hassan, M.A., Hashad, A.H., 1990. Precambrian <strong>of</strong> Egypt. In: Said, R. (Ed.), The Geology <strong>of</strong><br />

Egypt. Alkema, Rotterdam, pp. 201–245.<br />

Shimron, A.E., 1980. Proterozoic island arc volcanism and sedimentation in Sinai.<br />

Precambrian Research 12, 437–458.<br />

Hassan, I.S., Khalifa, I.H., Ibrahim, S.K., 2001. Petrogenesis <strong>of</strong> late Precambrian Dokhan<br />

Volcanics suite at Wadi Meknas, sou<strong>the</strong>astern Sinai, Egyptian. Journal <strong>of</strong> Geology 45,<br />

309–323.<br />

Hassanen, M.A., 1992. Geochemistry and petrogenesis <strong>of</strong> <strong>the</strong> late Proterozoic Kid<br />

Volcanics; evidence relevant to arc-intra-arc rifting volcanism in sou<strong>the</strong>rn Sinai,<br />

Egypt. Journal <strong>of</strong> <strong>African</strong> Earth Sciences 14, 131–145.<br />

Heidbach, O., Müller, B., Peters, G., Buchmann, T., Oth, A., Nuckelt, A., 2008. Farewell<br />

signals <strong>of</strong> slab break-<strong>of</strong>f in Vrancea, Romania <strong>from</strong> observation and numerical<br />

modeling. Geophysical Research Abstracts 10 EGU2008-A-09083.<br />

Heikal, M.A., Higazy, M.H., El Rahmany, M.M., 1980. Ignimbritic rhyolite in <strong>the</strong> Wassif<br />

area, Eastern Desert, Egypt. Institute <strong>of</strong> Applied Geology Bulletin, Jeddah 3, 107–114.<br />

H<strong>of</strong>mann, A.W.,1988. Chemical differentiation <strong>of</strong> <strong>the</strong> Earth: <strong>the</strong> relationship between mantle,<br />

continental crust and oceanic crust. Earth and Planetary Science Letters 90, 297–314.<br />

H<strong>of</strong>mann, A.W., Jochum, K.P., Seufert, M., White, W.M., 1986. Nb and Pb in oceanic basalts:<br />

new constraints on mantle evolution. Earth and Planetary Science Letters 79, 33–45.<br />

Irvine, T.N., Baragar, W.R.A., 1971. A guide to <strong>the</strong> chemical classification <strong>of</strong> <strong>the</strong> common<br />

volcanic rocks. Canadian Journal <strong>of</strong> Earth Sciences 8, 523–548.<br />

Jarar, G., Stern, R.J., Saffarini, G., Al-Zubi, H., 2003. Late- and post orogenic Neoproterozoic<br />

intrusions <strong>of</strong> Jordan: implications for crustal growth in <strong>the</strong> nor<strong>the</strong>rnmost segment <strong>of</strong><br />

<strong>the</strong> East <strong>African</strong> Orogen. Precambrian Research 123, 295–319.<br />

Jochum, K.P., Seufert, H.M., Spettel, B., Plame, H., 1986. The solar-system abundances <strong>of</strong><br />

Nb, Ta and Y and <strong>the</strong> relative abundances <strong>of</strong> refractory lithophile elements in<br />

differentiated planetary bodies. Geochimica Cosmochimica Acta 50, 1173–1183.<br />

Johnson, D.M., Hooper, P.R., Conrey, R.M.,1999. XRF analysis <strong>of</strong> rocks and minerals for major<br />

and trace elements on a single low dilution Li-tetraborate fused bead, pp. 843–867.<br />

Jöns, N., Emmel, B., Schenk, V., Razakamanana, T., 2009. From orogenesis to passive<br />

margin — <strong>the</strong> cooling history <strong>of</strong> <strong>the</strong> Bemarivo Belt, a multi-<strong>the</strong>rmochronometer<br />

approach. Gondwana Research 16, 72–81.<br />

Katzir, Y., Eyal, M., Litvinovsky, B.A., Jahn, B.M., Zanvilevich, A.N., Valley, J.W., Beeri, Y.,<br />

Pelly, I., Shimshilashvili, E., 2007. Petrogenesis <strong>of</strong> A-type granites and origin <strong>of</strong><br />

vertical zoning in <strong>the</strong> Ka<strong>the</strong>rina pluton, Gebel Mussa (Mt. Moses) area, Sinai, Egypt.<br />

Lithos 95, 208–228.<br />

Knaack, C., Cornelius, S., Hooper, P., 1994. Trace element analyses <strong>of</strong> rocks and minerals<br />

by ICP-MS. Washington State University, Geology Department Open File Report,<br />

Pullman, WA. 18 pp.<br />

Kröner, A., Reischmann, T., Wust, H.J., Rashwan, A., 1988. Is <strong>the</strong>re any pre-<strong>Pan</strong>-<strong>African</strong><br />

(N950 Ma) basement in <strong>the</strong> Eastern Desert <strong>of</strong> Egypt? In: El Gaby, S., Greiling, R.<br />

(Eds.), The <strong>Pan</strong>-<strong>African</strong> Belt <strong>of</strong> Nor<strong>the</strong>ast Africa and Adjacent Areas. Vieweg and<br />

Sohn, Braunschweig/Wiesbaden, Braun, pp. 93–119.<br />

Sultan, M., Tucker, R.D., El Alfy, Z., Attia, R., Ragab, A.G., 1994. U–Pb (zircon) ages for <strong>the</strong><br />

gneissic terrane west <strong>of</strong> <strong>the</strong> Nile, sou<strong>the</strong>rn Egypt. Geologische Rundschau 83, 514–522.<br />

Kröner, A., Todt, W., Hussein, I.M., Mansour, M., Rashwan, A., 1992. Dating <strong>of</strong> late<br />

Proterozoic ophiolites in Egypt and <strong>the</strong> Sudan using single zircon evaporation<br />

technique. Precambrian Research 59, 15–32.<br />

Sultan, M., Tuckler, R.D., Gharbawi, R.I., Ragab, A.I., Alfy, Z., 1993. <strong>On</strong> <strong>the</strong> location <strong>of</strong> <strong>the</strong><br />

boundary between <strong>the</strong> Nubian <strong>Shield</strong> and <strong>the</strong> old <strong>African</strong> continent: inferences<br />

<strong>from</strong> U–Pb (zircon) and common Pb data. In: Thorweihe, U., Schandelmeier, H.<br />

(Eds.), Geoscientific Research in NE Africa. Balkema, Amsterdam, pp. 75–77.<br />

Kröner, A., Kruger, J., Rashwan, A.A, 1994. Age and tectonic setting <strong>of</strong> granitoid gneisses<br />

in <strong>the</strong> Eastern Desert <strong>of</strong> Egypt and south-west Sinai. Geologische Rundschau 83,<br />

502–513.<br />

Kusky, T.M., Matsah, M., 2003. Neoproterozoic dextral faulting on <strong>the</strong> Najd fault system,<br />

Saudi Arabia, preceded sinistral faulting and escape tectonics related to closure <strong>of</strong> <strong>the</strong><br />

Mozambique Ocean. Geological Society <strong>of</strong> London Special Publication 206, 327–361.<br />

Tamura, Y., Tatsumi, Y., 2002. Remelting <strong>of</strong> andesitic crust as a possible origin <strong>of</strong> rhyolitic<br />

magma in oceanic arcs: an example <strong>from</strong> <strong>the</strong> Izu-Bonin arc, Japan. Journal <strong>of</strong> Petrology<br />

43, 1029–1047.<br />

Kusky, T.M., Abdelsalam, M.G., Tucker, R.D., Stern, R.J., 2003. Evolution <strong>of</strong> <strong>the</strong> East<br />

<strong>African</strong> and related orogens, and <strong>the</strong> assembly <strong>of</strong> Gondwana. Precambrian Research<br />

123, 81–337.


Le Bas, M.J., Le Maitre, R.W., Streckeisen, A., Zanetin, B., 1986. A chemical classification <strong>of</strong><br />

volcanic rocks based on <strong>the</strong> total alkali–silica diagram. Journal <strong>of</strong> Petrology 27, 745–750.<br />

Le Maitre, R.W., 1989. A Classification <strong>of</strong> Igneous Rocks and Glossary <strong>of</strong> Terms. Blackwell,<br />

Oxford. 193 pp.<br />

Leat, P.T., Smellie, J.L., Millar, I.L., Larter, R.D., 2003. Magmatism in <strong>the</strong> south sandwich<br />

arc, intra-oceanic subduction systems: tectonic and magmatic processes. Geological<br />

Society <strong>of</strong> London Special Publications 219, 285–313.<br />

Lustrino, M., Morra, V., Fedele, L., Serracino, M., 2007. The <strong>transition</strong> between ‘orogenic’<br />

and ‘anorogenic’ magmatism in <strong>the</strong> western Mediterranean area: <strong>the</strong> Middle<br />

Miocene volcanic rocks <strong>of</strong> Isla del Toro (SW Sardinia, Italy). Terra Nova 19, 148–159.<br />

McDonough, W.F., Sun, S., 1995. The composition <strong>of</strong> <strong>the</strong> Earth. Chemical Geology 120,<br />

223–253.<br />

Moghazi, A.M., 1999. Magma source and evolution <strong>of</strong> Late Neoproterozoic granitoids in<br />

<strong>the</strong> Gabal El-Urf area, Eastern Desert, Egypt: geochemical and Sr–Nd isotopic<br />

constraints. Geological Magazine 136, 285–300.<br />

Moghazi, A.M., 2003. Geochemistry and petrogenesis <strong>of</strong> a high-K calc-alkaline Dokhan<br />

Volcanic suite, South Safaga area, Egypt: <strong>the</strong> role <strong>of</strong> late Neoproterozoic crustal<br />

extension. Precambrian Research 125, 161–178.<br />

Moghazi, A.M., Andersen, T., Oweiss, G.A., El Bouseily, A.M., 1998. Geochemical and<br />

Sr–Nd–Pb isotopic data bearing on <strong>the</strong> origin <strong>of</strong> <strong>Pan</strong>-<strong>African</strong> granitoids in <strong>the</strong> Kid<br />

area, sou<strong>the</strong>ast Sinai, Egypt. Journal <strong>of</strong> <strong>the</strong> Geological Society <strong>of</strong> London 155,<br />

697–710.<br />

Mohamed, F.H., Moghazi, A.M., Hassanen, M.A., 2000. Geochemistry, petrogenesis and<br />

tectonic setting <strong>of</strong> late Neoproterozoic Dokhan-type volcanic rocks in <strong>the</strong> Fatira<br />

area, eastern Egypt. International Journal <strong>of</strong> Earth Sciences 88, 764–777.<br />

Monnereau, M., Rabinowicz, M., Arquis, E., 1993. Mechanical erosion and reheating <strong>of</strong> <strong>the</strong><br />

lithosphere: a numerical model for hotspot swells. Journal <strong>of</strong> Geophysical Research 98,<br />

809–823.<br />

Moussa, E.M.M., Stern, R.J., Manton, W.I., Ali, K.A., 2008. SHRIMP zircon dating and Sm/<br />

Nd isotopic investigations <strong>of</strong> Neoproterozoic granitoids, Eastern Desert, Egypt.<br />

Precambrian Research 160, 341–356.<br />

Neumann, E.-R., 1994. The Oslo rift: P–T relations and lithospheric structure.<br />

Tectonophysics 240, 159–172.<br />

Neumann, E.-R., Dunworth, E.A., Sundvoll, B.A., Tollefsrud, J.I., 2002. B1 basaltic lavas in<br />

Vestfold–Jeløya area, central Oslo rift: derivation <strong>from</strong> initial melts formed by<br />

progressive partial melting <strong>of</strong> an enriched mantle source. Lithos 61, 21–53.<br />

Nicolae, I., Saccani, E., 2003. Petrology and geochemistry <strong>of</strong> <strong>the</strong> Late Jurassic calcalkaline<br />

series associated to Middle Jurassic ophiolites in <strong>the</strong> South Apuseni<br />

Mountains (Romania). Schweizerische Mineralogische und Petrographische Mitteilungen<br />

83, 81–96.<br />

Pallister, J.S., Stacey, J.S., Fischer, L.B., Premo, W.R., 1988. Precambrian ophiolites <strong>of</strong><br />

Arabia: geologic settings, U–Pb geochronology, Pb-isotope characteristics, and<br />

implications for continental accretion. Precambrian Research 38, 1–54.<br />

Pearce, J.A., 1982. Trace element characteristics <strong>of</strong> lavas <strong>from</strong> destructive plate boundaries.<br />

In: Thorpe, R.S. (Ed.), Andesites, Orogenic Andesites and Related Rocks. John Wiley,<br />

New York, pp. 525–548.<br />

Pearce, J.A., 1996. Sources and settings <strong>of</strong> granitic rocks. Episodes 19, 120–125.<br />

Pearce, J.A., Gale, G.E., 1977. Identification <strong>of</strong> ore deposition environment <strong>from</strong> trace<br />

element geochemistry. Geological Society <strong>of</strong> London Special Publication 7, 4–24.<br />

Pearce, J.A., Harris, N.B.W., Tindle, A.G., 1984. Trace element discrimination diagrams for<br />

<strong>the</strong> tectonic interpretation <strong>of</strong> granitic rocks. Journal <strong>of</strong> Petrology 25, 959–983.<br />

Rajesh, H.M., 2004. The igneous charnockite – high-K alkali-calcic I-type granite –<br />

incipient charnockite association in Trivandrum Block, sou<strong>the</strong>rn India. Contributions<br />

to Mineralogy and Petrology 147, 346–362.<br />

Regard, V., Faccenna, C., Bellier, O., Martinod, J., 2008. Laboratory experiments <strong>of</strong> slab<br />

break-<strong>of</strong>f and slab dip reversal: insight into <strong>the</strong> Alpine Oligocene reorganization.<br />

Terra Nova 20, 267–273.<br />

Reymer, A.P.S., 1983. Metamorphism and tectonics <strong>of</strong> a <strong>Pan</strong>-<strong>African</strong> terrain in<br />

sou<strong>the</strong>astern Sinai. Precambrian Research 19, 225–238.<br />

Reymer, A.P.S., Yogev, A., 1983. Stratigraphy and tectonic history <strong>of</strong> <strong>the</strong> sou<strong>the</strong>rn Wadi<br />

Kid metamorphic complex, Sou<strong>the</strong>astern Sinai. Israel Journal <strong>of</strong> Earth Sciences 32,<br />

105–116.<br />

Ries, A.C., Shackelton, R.M., Graham, R.H., Fitches, W.R., 1983. <strong>Pan</strong>-<strong>African</strong> structures,<br />

ophiolites and melanges in <strong>the</strong> Eastern Desert <strong>of</strong> Egypt: a traverse at 26°N. Journal<br />

<strong>of</strong> <strong>the</strong> Geological Society <strong>of</strong> London 140, 75–95.<br />

Rino, S., Kon, Y., Sato, W., Maruyama, S., Santosh, M., Zhao, D., 2008. The Grenvillian and<br />

<strong>Pan</strong>-<strong>African</strong> orogens: world's largest orogenies through geologic time, and <strong>the</strong>ir<br />

implications on <strong>the</strong> origin <strong>of</strong> superplume. Gondwana Research 14, 51–72.<br />

Rollinson, H.R., 1993. Using Geochemical Data; Evaluation, Presentation, Interpretation.<br />

Longman Group Ltd., London. 343 pp.<br />

Rosa, D.R.N., Inverno, C.M.C., Oliveira, V.M.J., Rosa, C.J.P., 2006. Geochemistry and<br />

geo<strong>the</strong>rmometry <strong>of</strong> volcanic rocks <strong>from</strong> Serra Branca, Iberian Pyrite Belt, Portugal.<br />

Gondwana Research 10, 328–339.<br />

Rudnick, R.L., Fountain, D.M., 1995. Nature and composition <strong>of</strong> <strong>the</strong> continental crust: a<br />

lower crustal perspective. Reviews <strong>of</strong> Geophysics 33, 267–309.<br />

Saleh, G.M., 2003. Neoproterozoic volcanism at Um Shilman-Um Dubr area, Sou<strong>the</strong>ast<br />

Aswan, Egypt: geology, geochemistry and tectonic environment <strong>of</strong> <strong>the</strong> Dokhan<br />

Volcanic Formation. Neues Jahrbuch für Mineralogie. Abhandlungen 177, 321–347.<br />

M.Z. El-Bialy / Gondwana Research 17 (2010) 26–43<br />

Seghedi, I., Downes, H., Harangi, S., Mason, P.R.D., Pecskay, Z., 2005. Geochemical<br />

response <strong>of</strong> magmas to Neogene Quaternary continental collision in <strong>the</strong> Carpathian<br />

<strong>Pan</strong>nonian region: a review. Tectonophysics 410, 485–499.<br />

Shimron, A.E., 1983. The Tarr Complex — revisited: folding, thrusts and melanges in <strong>the</strong><br />

sou<strong>the</strong>rn Wadi Kid region, Sinai Peninsula. Israel Journal <strong>of</strong> Earth Sciences 32,123–148.<br />

Smith, I.E.M, Worthington, T.J., Stewart, R.B., Price, R.C., Gamble, J.A., 2003. Felsic<br />

volcanism in <strong>the</strong> Kermadic arc, SW Pacific: crustal recycling in an oceanic setting.<br />

Intra-oceanic-Subduction Systems: Tectonic and Magmatic Processes. Geological<br />

Society <strong>of</strong> London Special Publications 219, 99–118.<br />

Sperner, B., Lorenz, F., Bonjer, K., Hettel, S., Mueller, B., Wenzel, F., 2001. Slab break-<strong>of</strong>f±<br />

abrupt cut or gradual detachment? New insights <strong>from</strong> <strong>the</strong> Vrancea Region (SE<br />

Carpathians, Romania). Terra Nova 13, 172–179.<br />

Stacey, J.S., Agar, R.A., 1985. U–Pb isotopic evidence for <strong>the</strong> accretion <strong>of</strong> a continental<br />

microplate in <strong>the</strong> Zalm region <strong>of</strong> <strong>the</strong> Saudi Arabian <strong>Shield</strong>. Journal <strong>of</strong> <strong>the</strong> Geological<br />

Society <strong>of</strong> London 142, 1189–1204.<br />

Stern, R.J., 1979. Late Precambrian ensimatic volcanism in <strong>the</strong> central Eastern Desert <strong>of</strong><br />

Egypt. Ph.D. Thesis. California University, San Diego, 210 pp.<br />

Stern, R.J., 1994. Arc assembly and continental collision in <strong>the</strong> Neoproterozoic East<br />

<strong>African</strong> Orogen: implications for <strong>the</strong> consolidation <strong>of</strong> Gondwanaland. Annual<br />

Reviews <strong>of</strong> Earth and Planetary Sciences 22, 319–351.<br />

Stern, R.J., 2002. Crustal evolution in <strong>the</strong> East <strong>African</strong> Orogen: a neodymium isotopic<br />

perspective. Journal <strong>of</strong> <strong>African</strong> Earth Sciences 34, 109–117.<br />

Stern, R.J., 2008. Neoproterozoic crustal growth: <strong>the</strong> solid Earth system during a critical<br />

episode <strong>of</strong> <strong>the</strong> Earth history. Gondwana Research 14, 33–50.<br />

Stern, R.J., Gottfried, D., 1986. Petrogenesis <strong>of</strong> late Precambrian (575–600 Ma) bimodal<br />

suite in nor<strong>the</strong>ast Africa. Contribution to Mineralogy and Petrology 92, 492–501.<br />

Stern, R.J., Hedge, C.E., 1985. Geochronologic and isotopic constraints on Late<br />

Precambrian crustal evolution in <strong>the</strong> Eastern Desert <strong>of</strong> Egypt. American Journal <strong>of</strong><br />

Science 285, 97–127.<br />

Stern, R.J., Manton, W.I., 1987. Age <strong>of</strong> Feiran basement rocks, Sinai: implications for late<br />

Precambrian crustal evolution in <strong>the</strong> nor<strong>the</strong>rn <strong>Arabian–Nubian</strong> <strong>Shield</strong>. Journal <strong>of</strong><br />

<strong>the</strong> Geological Society <strong>of</strong> London 144, 569–575.<br />

Stern, R.J., Gottfried, D., Hedge, C.E., 1984. Late Precambrian rifting and crustal evolution<br />

in <strong>the</strong> nor<strong>the</strong>ast Desert <strong>of</strong> Egypt. Geology 12, 168–172.<br />

Stern, R.J., Sellers, G., Gottfried, D., 1988. Bimodal dyke swarms in <strong>the</strong> North Eastern<br />

Desert <strong>of</strong> Egypt: significance for <strong>the</strong> origin <strong>of</strong> late Precambrian “A-type” granites in<br />

nor<strong>the</strong>rn Afro-Arabia. In: El Gaby, S., Greiling, R.O. (Eds.), The <strong>Pan</strong>-<strong>African</strong> Belt <strong>of</strong><br />

Nor<strong>the</strong>ast Africa and Adjacent Areas. Vieweg and Sohn, Braunschweig/Wiesbaden,<br />

Braun, pp. 147–177.<br />

Stoeser, D.B., Frost, C.D., 2006. Nd, Pb, Sr, and O isotopic characterization <strong>of</strong> Saudi<br />

Arabian <strong>Shield</strong> terranes. Chemical Geology 226, 163–188.<br />

Sun, L., Wang, Y., Fan, W., Zi, J., 2008. Post-collisional potassic magmatism in <strong>the</strong><br />

Sou<strong>the</strong>rn Awulale Mountain, western Tianshan Orogen: petrogenetic and tectonic<br />

implications. Gondwana Research 14, 383–394.<br />

Taylor, S.R., McLennan, S.M., 1985. The continental crust: its composition and evolution.<br />

Geoscience Texts. Blackwell Scientific Publishers, Oxford. 312 pp.<br />

Tetsopgang, S., Suzuki, K., Njonfang, E., 2008. Petrology and CHIME geochronology <strong>of</strong><br />

<strong>Pan</strong>-<strong>African</strong> high K and Sr/Y granitoids in <strong>the</strong> Nkambe area, Cameroon. Gondwana<br />

Research 14, 686–699.<br />

Thornton, C.P., Tuttle, O.F., 1960. Chemistry <strong>of</strong> igneous rocks, part 1: differentiation<br />

index. American Journal <strong>of</strong> Science 280, 664–684.<br />

Vaughan, A.P.M., <strong>Pan</strong>khurst, R.J., 2008. Tectonic overview <strong>of</strong> <strong>the</strong> West Gondwana margin.<br />

Gondwana Research 13, 150–162.<br />

Waston, E.B., Harrison, T.M., 1983. Zircon saturation revisited: temperature and<br />

composition effects in a variety <strong>of</strong> crustal magma types. Earth and Planetary Science<br />

Letters 64, 295–304.<br />

Watson, E.B.,1979. Zircon saturation in felsic liquids: experimental results and applications<br />

to trace element geochemistry. Contribution to Mineralogy and Petrology 70, 407–419.<br />

Wedepohl, K.H., 1995. The composition <strong>of</strong> <strong>the</strong> continental crust. Geochimica et<br />

Cosmochimica Acta 59, 1217–1232.<br />

Whalen, J.B., Currie, K.I., Chappell, B.W., 1987. A-type granites: geochemical characteristic,<br />

discrimination and petrogenesis. Contribution to Mineralogy and Petrology<br />

95, 407–419.<br />

Whalen, J.B., Jenner, G.A., Longstaffe, F.J., Robert, F., Garie'py, C.,1996. Geochemical and isotopic<br />

(O, Nd, Pb and Sr) constraints on A-type granite petrogenesis based on <strong>the</strong> Topsails<br />

Igneous Suite Newfoundland Appalachians. Journal <strong>of</strong> Petrology 37, 1463–1489.<br />

Wilde, S.A., Youssef, K., 2000. Significance <strong>of</strong> SHRIMP U–Pb dating <strong>of</strong> <strong>the</strong> Imperial<br />

Porphyry and associated Dokhan Volcanics, Gabal Dokhan, nor<strong>the</strong>rn Eastern Desert,<br />

Egypt. Journal <strong>of</strong> <strong>African</strong> Earth Sciences 31, 403–413.<br />

Willis, K.M., Stern, R.J., Clauer, N.,1988. Age and geochemistry <strong>of</strong> late Precambrian sediments<br />

<strong>of</strong> <strong>the</strong> Hammamat Series <strong>from</strong> <strong>the</strong> nor<strong>the</strong>astern Desert <strong>of</strong> Egypt. Precambrian Research<br />

42, 173–187.<br />

Wilson, M., 1989. Igneous petrogenesis. Unwin Hyman Ltd., London. 466 pp.<br />

Wilson, M., Tankut, A., Gulec, N., 1997. Tertiary volcanism <strong>of</strong> <strong>the</strong> Galatia province, northwest<br />

Central Anatolia, Turkey. Lithos 42, 105–121.<br />

Wolff, J.A., 1984. Variation in Nb/Ta during differentiation <strong>of</strong> phonolitic magmas,<br />

Tenerife, Canary Islands. Geochimica et Cosmochimica Acta 48, 1345–1348.<br />

43

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!