Sie sind auf Seite 1von 16

Precambrian Research 107 (2001) 179 194

www.elsevier.com/locate/precamres

Post-collisional shortening in the late Pan-African Hamisana


high strain zone, SE Egypt: field and magnetic fabric
evidence
Helga de Wall a,*, Reinhard O. Greiling a, Mohamed Fouad Sadek b
a

Geologisch-Palaontologisches Institut, Ruprecht-Karls-Uni6ersitat, Im Neuenheimer Feld 234, 69120 Heidelberg, Germany


b
Egyptian Geological Sur6ey and Mining Authority, 3 Salah Salem Street, Abbassiyah, Cairo, Egypt
Received 7 April 2000; accepted 8 September 2000

Abstract
The Hamisana zone (HZ) is one of the major high strain zones of the Pan-African (Neoproterozoic) Arabian Nubian shield (ANS). It trends broadly NS from northern Sudan into southeastern Egypt and meets the present Red
Sea coast at :23N. The HZ has been the subject of controversy with regard to its importance for the Pan-African
structural evolution. Interpretations range from a suture zone, a regional shear zone, or a large-scale transpressional
wrench fault system. In this study, we characterize the nature of the high strain deformation by applying the
anisotropy of magnetic susceptibility method along with field and microstructural investigations. These investigations
demonstrate that deformation in the HZ is dominated by pure shear under upper greenschist/amphibolite grade
metamorphic conditions, producing EW shortening, but with a strong N S-extensional component. This deformation also led to folding of regional-scale thrusts (including the base of ophiolite nappes such as Gabal Gerf and Onib).
Consequently, the high strain deformation is younger than ophiolite emplacement and suturing of terranes. A weak
subsequent overprint was mostly non-coaxial. It took place under considerably lower temperature and led to a minor
NESW-trending, dextral wrench fault. Although it is of only local importance this fault may be itself a conjugate
relative to the prominent NWSE-trending sinistral Najd faults in the northern ANS. Therefore, the HZ is dominated
by late orogenic compressional deformation and cannot be related to either large-scale transpressional orogeny or
major escape tectonics. 2001 Elsevier Science B.V. All rights reserved.
Keywords: Hamisana zone; ArabianNubian shield; Late-orogenic compression; Magnetic susceptibility

1. Introduction
The Pan-African orogenic cycle has long been
recognized as a period of major crustal accretion,
* Corresponding author. Tel.: + 49-6221-544843; fax: +496221-545503.
E-mail address: hr8@ix.urz.uni-heidelberg.de (H. de Wall).

where continental, island-arc, and oceanic terranes were brought together to form the crystalline basement of the African continent as part
of a late Neoproterozoic supercontinent (Unrug,
1997, for review). Whereas some parts of the
Pan-African orogen are characterized by continental collisional tectonics (Burke and Sengor,

0301-9268/01/$ - see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 3 0 1 - 9 2 6 8 ( 0 0 ) 0 0 1 4 1 - 8

180

H. de Wall et al. / Precambrian Research 107 (2001) 179194

1986; Stern, 1994) others are typical for accretionary orogens (Windley, 1992, for definition).
During the last two decades, the Arabian Nubian shield (ANS), the northeastern part of
Africa Arabia, has been established as a typical
example of an accretionary orogen with lateral
accretion and suturing (Gass, 1981; Kroner, 1985;
Kroner et al., 1987; Stern, 1994). In the course of
orogenic activity, numerous sutures and other
major deformational zones originated all across
this shield. Subsequent to the assembly of terranes, late deformation zones developed as the
expression of lateral escape tectonics or other
transtensional or transpressional episodes during
the final stages of continental crust formation
(Burke and Sengor, 1986; Stern, 1985, 1994;
Johnson and Kattan, 1999). Contrasting models
of transpressional orogeny (Sanderson and Marchini, 1984) have been proposed for the ANS
(OConnor et al., 1994; Abdelsalam et al., 1998),
some of which relate terrane assembly to a single,
major event of transpression.
In order to distinguish between these models, it
is essential to re-examine the evolution of the
orogenic structures in the ANS. Towards that
aim, the present authors studied a major PanAfrican deformational zone, the Hamisana zone
(HZ), which is exposed in the Red Sea Hills of
southeastern Egypt and northeastern Sudan (Fig.
1). The NS-trending HZ is spectacularly visible
on satellite images, where it can be traced for
more than 300 km (Miller and Dixon, 1992). It is
parallel with a zone of similar size on the SudanEthiopia border, the Baraka zone (Dixon et al.,
1987) and the less prominent Oko zone, which is
located 100 km to the east/southeast of the HZ
(Abdelsalam, 1994, Fig. 1).
Well-defined ophiolites and ophiolite belts have
been mapped on both sides of the HZ and interpreted as a suture zone between major PanAfrican terranes. East of the HZ this belt was
called the Onib-Sol Hamed suture (Fitches et al.,
1983; Dixon et al., 1987; Kroner et al., 1987;
Stern et al., 1989, 1990, Fig. 1). The HZ, cuts the
Onib-Sol Hamed suture and is, consequently,
younger. Although this structural relationship has

been well documented (Stern et al., 1990; Miller


and Dixon, 1992), a contrasting model of the HZ
as a major transpressional zone during terrane
accretion has been proposed recently, comprising

Fig. 1. Major deformation zones in the Pan-African basement


of SE Egypt and NE Sudan and location of the study area in
the northern HZ. Compiled from Kroner et al. (1987); Stern et
al. (1990); Greiling et al. (1994); Greiling et al. (1988);
EGSMA (1996); EGSMA (1999); OConnor (1996); Abdelsalam et al. (1998); Rashwan and Greiling (1999) and A.
Makhlouf (unpublished data).

H. de Wall et al. / Precambrian Research 107 (2001) 179194

181

2. Evolution of the Arabian Nubian shield and


the Hamisana zone

2.1. ArabianNubian shield

Fig. 2. Geological sketch map of the study area. For location


see Fig. 1. The right-lateral strike-slip fault in the SE corner of
the map is taken from Stern et al. (1990).

a relatively simple evolution from oblique convergence and suturing towards transpressional shear
zones as a final stage of orogeny (OConnor et al.,
1994; Nasr et al., 1998; Smith et al., 1999).
As it is clear from this controversy, the time
sequence and the general tectonic significance of
these structures at a regional scale are not yet
fully known. Based on recent mapping by the
Egyptian Geological Survey, a part of the HZ
(Figs. 1 and 2) was studied in the field and
sampled for detailed structural studies. In addition to the usual structural geological investigations, the anisotropy of magnetic susceptibility
(AMS) of the rocks in the HZ was studied. This
latter method has been shown to be most sensitive
to document the type and style of deformation
even in places where no or hardly any deformation is visible macroscopically (Tarling and
Hrouda, 1993). In that way, it is possible not only
to characterize the deformation within the HZ but
also to determine a subsequent overprint, which
may escape detection by other methods. This
application of the AMS method is the first major
study of this kind in the ANS, apart from minor
pilot studies (Greiling, 1997; de Wall et al., 1998,
1999; El Eraki and Greiling, 1998).

The ANS evolved during the Neoproterozoic,


Pan-African episode from a number of island-arc
and minor continental terranes. They were accreted onto an East Sahara craton, which was
previously affected in the Neoproterozoic and is
now mostly covered by Phanerozoic sediments
and is presumed to extend westwards from the
Keraf suture (Abdelsalam et al., 1998). Lithologically, this tectonic evolution from oceanic lithosphere to island-arcs and continental lithosphere
is well documented by the rock sequences building
up the ANS: ophiolite suites, ocean floor (cherts)
and passive margin (quartzo-feldspathic and carbonate) sedimentary sequences, island arc, calc-alkaline igneous rocks, both intrusive and
supracrustal, together with syn-orogenic sediments derived from these igneous rocks. Finally,
syn-collisional and late- to post-tectonic magmatic activity produced some gabbroic but mostly
granitoid intrusives and related volcanics. Lateorogenic, molasse-type sediments are only found
in some intramontane basins, mostly in the NW
of the ANS (Akaad and El Ramly, 1958; Osman
et al., 1993). Structurally, at least parts of the
tectonic evolution are also documented, for example subduction-related tectonic melange or fossil
accretionary prism (Wadi Ghadir; El Sharkawy
and El Bayoumi, 1979; Shackleton et al., 1980) or
early, syn-island-arc deformation (Wadi Hafafit;
El Ramly et al., 1984; Rashwan, 1991). Subsequent structures comprise both extensional features such as fault-bounded (molasse) basins (Rice
et al., 1993) and metamorphic core complexes
(Sturchio, et al., 1983; Fritz et al., 1996), largescale thrusts (Ries et al., 1983; El Ramly et al.,
1984), and major wrench fault systems (Stern,
1985; Sultan et al., 1988; Shimron, 1990; OConnor et al., 1994; Abdelsalam et al., 1998; Greiling
et al., 1998). Apparently, both extensional and
compressional deformation, as they are documented in the ANS, represent post-collisional
structures and are succeeded by and associated
with wrench faulting (Abdeen et al., 1992; Greiling et al., 1994; Johnson and Kattan, 1999).

182

H. de Wall et al. / Precambrian Research 107 (2001) 179194

The most important of such faults in the northwestern ANS are the NW SE-oriented Najd system, including Wadi Kharit-Wadi Hodein fault
zones, the broadly N S-oriented HZ and the Oko
Shear zone (Abdelsalam, 1994; Kroner et al.,
1987; Stern, 1985, 1994; Greiling et al., 1994; Fig.
1).

2.2. The Hamisana zone


The northern part of the HZ cuts across sequences of gneisses, which are structurally overlain by ophiolitic nappes and intruded by a
number of granitoids (EGSMA, 1999, Fig. 2).
The probably oldest rocks exposed are a sequence
of variegated gneisses, containing both metasedimentary and igneous components. The latter comprise both mafic, amphibolitic layers as well as
felsic, fine-grained and coarse-grained para- and
ortho-gneisses and, locally, close to younger intrusives, also migmatites. The gneissic sequences
form a basal tier (tier 1, Bennet and Mosley, 1987;
Greiling et al., 1994) and are structurally overlain
by a number of major nappe units, which are
dominated by ophiolite fragments (Gerf nappe,
Kroner et al., 1987) and form a second tier (tier
2). Both these major units are folded and intruded
by a number of syn- to late-tectonic granitoid
plutons as well as some dyke swarms (Stern et al.,
1990; Miller and Dixon, 1992).
The ophiolites are c. 730 Ma old (Zimmer et al.,
1995) and the tier 1 gneisses originated about c.
660 Ma ago, during and after the formation and
collision of island-arcs (Stern et al., 1989). Activity along the HZ may have begun as early as 660
Ma ago, culminating in intense thermal activity
550 580 Ma ago (Stern et al., 1989). Undeformed, post-tectonic granites dated at 510 Ma
ago define an upper age limit of the deformation
(Stern et al., 1989). Whereas tiers 1 and 2 are
generally flat-lying at a regional scale, they are
overprinted by the N S-trending HZ, up to a few
tens of km wide, where foliations are generally
steep (Fig. 3) and strain appears to be higher and
deformation more complex than in the surrounding domains (Stern et al., 1990; Miller and Dixon,
1992).

An apparent offset of the Onib-Sol Hamed


suture zone to the east of the HZ against the
Allaqi suture zone to the west led earlier authors
to presume a horizontal, strike-slip movement
along the HZ (Kroner et al., 1987; Stern et al.,
1989). However, recent results from the Allaqi
suture zone (Sadek, 1995; EGSMA, 1996; our
observations) show that this suture is curving
from the Wadi Allaqi towards the southeast and
that related ophiolite fragments occur to the south
of the Allaqi-Heiani Belt of ophiolite nappes,
which was earlier thought to represent the suture
zone. Therefore, an alternative location for the
Allaqi suture is tentatively shown on Fig. 1.

3. Field observations
The gneissic rocks of the HZ and the surrounding area are characterized by a distinct metamorphic banding (s1), which is particularly clear
between felsic or intermediate layers intercalated
with amphibolite (Fig. 9(A) (C)). On both sides
of the HZ this banding is flat-lying at a regional
scale with gentle folds at km-scale wavelengths
(EGSMA, 1999).
Towards the HZ, the NS-trending folds become tighter, so that the metamorphic banding

Fig. 3. Schematic section across the HZ (a) and structural field


data from the study area, (b) orientation of foliation showing
a girdle distribution. The best-fit great circle and its pole
(p-pole) are shown, (c) orientation of mineral and stretching
lineations. Diagrams (b) and (c) are equal area lower hemisphere stereonets.

H. de Wall et al. / Precambrian Research 107 (2001) 179194

dips more steeply (Fig. 9(A) (B)). At the western


margin of the HZ, the orientation of the metamorphic banding rotates successively from moderate to steep westerly dips to almost vertical at the
HZ itself. There, a zone c. 5 10 km wide is
characterized by steep to vertical dips, striking
broadly NS (Figs. 2 and 3(b)), with subhorizontal mineral and stretching lineations in N S directions (Fig. 3(c)). At lithologic contacts in the HZ,
a penetrative foliation (s2) can be observed, which
cuts the metamorphic banding at a low angle.
Towards east, the dip shallows again and shows a
transition to flat-lying ophiolite nappes, including
the Wadi Onib ophiolite fragment (Fitches et al.,
1983), overlying granitoid and other gneissic
sequences.
Consequently, the HZ appears as a major N Strending, upright antiformal structure, with a high
strain zone broadly along and parallel with its
axial surface. In some parts, minor folds with
amplitudes and wavelengths in a cm-scale around
N S-trending axes can be recognized (El Kady,
unpublished Diploma thesis, and own observations). Apparently, the foliation (s2) is the axial
surface cleavage of these folds.
Locally, a subhorizontal to moderately plunging NS lineation is well developed, marked by
stretched plagioclase and a hornblende mineral
lineation in the amphibolites (Fig. 9(D)) and
metavolcanic rocks and by a quartz and feldspar
stretching lineation in the orthogneisses. The lineation orientation forms a well-defined cluster on
the stereonet with a mean of 354/05 (Fig. 3(c)). Its
significance is discussed in the context of the
microstructural observations.

4. Magnetic fabric analyses

4.1. Anisotropy of magnetic susceptibility (AMS)


methods
Anisotropic magnetic behaviour of low field
susceptibility has been established as a method for
the detection of fabric anisotropies (Tarling and
Hrouda, 1993; Borradaile and Henry, 1997). The
advantage of this method, compared to conventional macroscopic and microscopic and X-ray

183

diffraction techniques, is its high sensitivity to


weak differences in the crystallographic orientation of paramagnetic minerals (such as phyllosilicates and amphiboles) and in the grain shape and
distribution of ferrimagnetic minerals (magnetite).
AMS studies have been found to be particularly
useful in constraining very low strains both related to regional deformation events (Borradaile,
1988) and pluton emplacements (Bouchez et al.,
1990, 1997). In deformation fabrics there is a
correlation between the rock foliation and the
magnetic foliation and rock lineation and magnetic lineation, respectively (Tarling and Hrouda,
1993; Siegesmund et al., 1995).
For detailed magnetic fabric analyses by the
AMS method (Tarling and Hrouda, 1993) various
rock types from 35 sites distributed over an area
of approximately 500 km2 were sampled (Fig. 2).
Handspecimens were oriented in the field and
from every sample up to 6 AMS standard cylinders (1 in. in diameter, 0.8 in. in height) were
drilled in the laboratory. A total of 109 specimens
were investigated. The measurements were performed in low fields (300 A/m) with a Kappabridge (KLY-2) manufactured by AGICO
(Jelinek, 1980). Thermomagnetic investigations
for magneto-mineralogy used a CS-2 furnace apparatus of AGICO (Hrouda, 1994).
The magnetic susceptibility (k) is a material
property which describes the ability to acquire a
magnetization (M) in an applied field (H). As
both M and H expressed in SI units (Syste`me
Internationale), are measured in A/m, the constant k is dimensionless. The AMS is a second
rank tensor which is expressed by the AMS-ellipsoid with principal axes smax; sint; smin. The volume susceptibility (s= (smax + sint + smin)/3),
shape and anisotropy factors are calculated from
directional measurements taken in different sample positions using the standard software for the
KLY-2 equipment (Hrouda et al., 1990). The
shape of the AMS-ellipsoid is described by the
parameter T: (2(ln sint ln smin)/(ln smax ln
sint) 1) which is \ 0 to + 1 for oblate, 0 for
neutral and B 0 to 1 for prolate geometry
(Jelinek, 1981). The anisotropy of the ellipsoid is
expressed by the P% value, the so-called corrected
anisotropy factor: exp (2(ln smax ln s)2 + 2(ln

H. de Wall et al. / Precambrian Research 107 (2001) 179194

184

Table 1
Magnetic data for the individual samples
Sample

Rock type

sa

Linb

a95

Folc

a95

Td

P%e

P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P

orthogneiss
orthogneiss
orthogneiss
orthogneiss
orthogneiss
amphibolite
orthogneiss
orthogneiss
biotite-schist
amphibolite
orthogneiss
orthogneiss
leucogranite
amphibolite
amphibolite
amphibolite
orthogneiss
amphibolite
biotite-schist
leucogranite
orthogneiss
orthogneiss
orthogneiss
orthogneiss
amphibolite
amphibolite
amphibolite
orthogneiss
orthogneiss
amphibolite
orthogneiss
amphibolite
amphibolite
amphibolite
paragneiss

1
2
2
2
4
3
4
2
4
4
3
3
4
4
4
3
4
5
3
4
4
4
4
3
3
2
3
3
3
2
4
3
1
1
4

850
447
395
395
6849
377
9900
2120
4743
733
324
4297
20
405
740
717
229
737
290
7208
303
6398
61
1112
3159
14370
805
9027
55
3605
2406
663
977
427
82

181/11
223/70
011/01
011/01
213/16
041/16
351/04
248/52
169/24
015/30
028/45
353/06
026/47
020/39
341/53
347/11
008/29
120/38
168/13
179/05
170/10
346/13
359/20
060/04
198/03
172/04
018/15
190/06
198/21
049/37
354/21
175/34
006/37
357/42
171/10

4.3
2.2
2.2
3.8
1.2
3.2
6.8
1.5
4.7
1.7
1.9
18.0
4.5
3.1
4.6
4.1
3.0
2.0
7.5
2.6
5.5
2.9
3.0
2.6
3.3
6.7
4.6
2.6
11.2
1.9
5.1

1.9

094/76
281/79
101/84
101/84
133/59
315/75
079/62
264/51
252/75
295/22
360/45
065/19
049/43
293/87
261/84
262/65
076/57
054/63
078/89
263/40
089/47
067/55
270/88
138/15
289/76
106/10
092/44
279/77
246/29
109/56
314/27
264/86
281/84
298/61
221/17

4.4
5.2
5.2
3.7
1.3
6.1
4.8
1.3
4.5
1.2
2.5
6.5
2.3
4.3
4.7
10.8
2.4
1.9
2.7
3
4.1
2.6
2.8
10.1
2.7
5.7
5.1
16.6
1.6
1.1
3.5

2.2

0.22
0.88
0.84
0.84
0.14
0.70
0.56
0.34
0.33
0.17
0.41
0.12
0.38
0.64
0.51
0.83
0.15
0.88
0.46
0.73
0.25
0.27
0.70
0.34
0.11
0.18
0.68
0.26
0.54
0.20
0.10
0.46
0.17
0.83
0.69

1.31
1.13
1.11
1.11
1.27
1.12
1.50
1.24
1.60
1.07
1.11
1.23
1.06
1.04
1.03
1.11
1.02
1.07
1.11
1.40
1.17
1.51
1.23
1.59
1.19
1.34
1.10
1.28
1.12
1.20
1.23
1.11
1.10
1.15
1.10

1
2a
2b
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
25
26
27
29
30
34
36
37
38
39
40

s= volume susceptibility in 106 SI.


Lin trend and plunge of the magnetic lineation and a95 confidence angle.
c
Fol dip direction and angle of dip of the magnetic foliation and a95 confidence angle.
d
T = mean shape factor of the AMS-ellipsoid.
e
P%= mean corrected anisotropy factor for the AMS-ellipsoid.
a

sint ln s)2 + 2(ln smin ln s)2)1/2 (Jelinek,


1981). From the determined geographic orientation for the principal axes of the susceptibility
ellipsoids, the mean values for the magnetic lineation and the magnetic foliation are calculated
based on Fisher distribution analyses. The mag-

netic lineation refers to the direction of maximum


susceptibility (smax), the direction of minimum
susceptibility (smin) is the pole to the magnetic
foliation plane. Directional data are presented in
stereonets (Figs. 5 and 6(a) and (b)). All results of
the AMS analyses are compiled in Table 1.

H. de Wall et al. / Precambrian Research 107 (2001) 179194

185

4.2. Volume susceptibility


In polycrystalline rocks as they are characteristic for the HZ the volume susceptibility is the sum
of susceptibility and volume content of the various mineralogical components. Since different
rock types occur in the HZ a strong variation in
volume susceptibility due to their variations in
mineral composition is observed. Lowest values
are in the order of 10 5 SI for leucogranites, the
highest susceptibilities exceed 5 10 2 SI in magnetite-bearing amphibolites (Table 1). In the samples with low susceptibility, paramagnetic
minerals such as biotite in the paragneisses and
biotite schists and hornblende in the amphibolites
are dominating the magnetic fabric. In the samples with higher susceptibilities (s \10 3) the
magnetic behaviour is determined by the ferrimagnetic mineral fabric. Thermomagnetic curves
reveal a strong decrease in susceptibility at temperatures of 590C indicating that the ferrimagnetic fabric is represented by pure magnetite (Fig.
4).

4.3. Orientation and interpretation of the AMS


directional data
A compilation of the orientation of the AMS
data from all samples measured is given in the
stereographic projection in Fig. 5. The magnetic

Fig. 4. Examples for thermomagnetic curves, s(T), for a


paramagnetic sample (HAMS 8) and a ferrimagnetic sample
(HAMS 10). The curves show the evolution of the magnetic
susceptibility during heating. See the text for further discussion.

Fig. 5. Orientation of magnetic fabric elements in geographic


orientation. Compilation of all field measurements in
stereonets (lower hemisphere). smin represents the poles of the
magnetic foliation, which show a girdle distribution. A best-fit
great circle and its pole (p-pole) are indicated. smax represents
the orientation of the magnetic lineation. The mean direction
is shown.

foliation poles (smin) are distributed along a great


circle, documenting a rotation of the foliation
surfaces around a horizontal NS trending axis
(p-pole). This geometry is in agreement with the
geometry of a regional-scale concentric fold structure with a horizontal, NS trending i-axis (Fig.
3). The p-pole (002/08) constructed from the
magnetic foliation poles is subparallel with the
magnetic lineation (Fig. 5(b)) with a mean orientation of 179/01.
The general pattern of magnetic fabric correlates well with the field-measured foliation and
lineation data, as presented in Fig. 3(b) and (c).
However, whilst the lineations measured in the
field appear to have a clear unimodal distribution
with a single maximum around a horizontal
mean, the magnetic lineations tend to scatter towards moderate to steep inclinations. The
stronger scatter in the magnetic lineation data
relative to the field data reveals that the fabric in
the HZ is more complex than expected from the
analysis of the field data alone. Therefore, we
analyzed the AMS data in more detail: the directional data are presented in foliation and lineation
maps, respectively, with added stereographic projections of the structural elements (Fig. 6(A) and
(B)). In both maps, a subdivision into two structural domains is evident. In domain I, the magnetic foliation trends generally NS. However,
dip angles vary between 10 and 89. The stereo

186

H. de Wall et al. / Precambrian Research 107 (2001) 179194

graphic projection exhibits a distribution of the


magnetic foliation poles along a great circle with a
nearly horizontal p-pole, which is trending N S,
and subparallel with the strike of the HZ. This
pattern reflects the general geometry in the HZ
and corresponds to the field observations (Fig.
3(b)). The magnetic lineations trend N S with
gentle to moderate plunges towards the S and N,
respectively, but the lineations are generally
steeper in the western part of the HZ. This orientation accords with the field data presented in Fig.
3(c), where the magnetic lineation is parallel with
the best-fit p-pole for the magnetic foliation. So,
generally, the magnetic fabric in structural domain I is symmetrical with a girdle distribution of
the magnetic foliation around a N S-trending
axis, which is parallel with the observed lineation.
A second structural domain (domain II) forms
a SWNE trending sigmoidal trace across the HZ
where the magnetic foliations and lineations are
reoriented towards a NE SW direction. The angle of plunge for the magnetic lineations is variable and ranges from low (03) to high (70)
values.

4.4. Shape of AMS-ellipsoids


The AMS measurements provide detailed information on fabric anisotropies and give strain and
mineralogical information in addition to conventional structural data (Borradaile and Henry,
1997). However, for a reliable interpretation of
the magnetic fabric a detailed knowledge of the
components of the rock fabric and their respective
contributions to the final AMS values is necessary
(Borradaile, 1988; Rochette, 1987). In the present
study various rock types were investigated. As a
consequence, a strong variation in shape (T) ranging between + 0.88 and 0.56, and anisotropy
(P%) reaching from 1.02 up to 1.60 (Table 1) is
observed. In the Jelinek-diagram (P% vs. T) (Jelinek, 1981) this variation is obvious for both
structural domains (Fig. 7(a)). This may indicate
a general fabric inhomogeneity for the samples
from the HZ with both prolate (stretched: smax \
\ sint = smin) and oblate (flattened: smax = sint \
\ smin) subfabrics. However, it could also be a
consequence of single crystal magnetic anisotropy
of the paramagnetic mineral components. AMS in

Fig. 6. (A) Magnetic foliation map and (B) magnetic lineation map, covering the same area as Fig. 2. Inset stereonets show the
orientation means as calculated for each sampling site (data in Table 1). Note two domains (I and II, respectively), which are distinct
by their fabric orientation.

H. de Wall et al. / Precambrian Research 107 (2001) 179194

187

5. Microstructures
Rocks in the HZ are characterized by a metamorphic fabric developed under amphibolite facies conditions. Amphibolites and hornblendebearing gneisses, derived from volcanic rocks and
gabbros contain coarse-grained green hornblende,
biotite, feldspar and quartz. In the amphibolites,
pronounced linear as well as distinct planar fabrics occur, which are defined by elongated
feldspar aggregates and the preferred orientation
of hornblende (Fig. 9(D)). Aspect ratios for prolate feldspar porphyroclasts in amphibolite, with a
well-developed linear fabric, reach up to 4.
Coarse-grained biotite and muscovite produce
the metamorphic foliation in the orthogneisses
and metagranitoids. Occasionally, a stretching lineation is marked by elongated quartz grains and
ribbon quartz. Inclusions of a continuous mica
foliation within large anhedral quartz grains and

Fig. 7. Anisotropy data for structural domains I (squares) and


II (triangles), respectively. Plotted are the mean values for each
sampling site (data in Table 1). (A) Shape factor (T) and
anisotropy (P%) of the AMS-ellipsoids, presented in a Jelinekdiagram (Jelinek, 1981). (B) Shape factor (T) versus bulk
susceptibility (smean).

paramagnetic minerals is a function of single crystal anisotropy and degree of preferred orientation
(Siegesmund et al., 1995; Hrouda et al., 1997).
However, from Fig. 7 it is obvious that scattering
of the shape factor values is indicated both in
samples with low (mainly paramagnetic minerals)
and with high (ferrimagnetic minerals) susceptibilities. Therefore, we conclude that beside mineralogical influences the observed variations in
ellipsoid shape are related to differences in fabric
geometry across the HZ. However, a significant
relationship of the shape factors to a particular
structural domain or a systematic variation in the
shape along and across the HZ was not detected.
In general, the AMS results agree with the field
observations, where also variations in the rock
fabrics have been detected.

Fig. 8. Map showing magnetic lineation trajectories in the


investigated part of the HZ. The area is the same as for Fig. 2
and Fig. 6. Domain II is representing a late-stage dextral shear
zone (NE SW) that is overprinting the N S oriented structural grain of the HZ. This late dextral shear zone may be
related to another dextral fault mapped by Stern et al. (1990),
part of which is shown in the SE corner of the present map.

188

H. de Wall et al. / Precambrian Research 107 (2001) 179194

irregular phase boundaries between quartz and


feldspar indicate secondary grain growth (Fig.
9(E)). The quartz grains are overprinted by a
subsequent deformation, which produced undulatory extinction and deformation bands.
In samples displaying a gneissic fabric, S C
geometries are observed, which are defined by
elongated grains of quartz, mica or hornblende.
In structural domain I, this fabric is not very
pronounced and only occasionally a shear sense
determination is possible. The angle between the
oblique fabric (S) and the shear foliation (C)
ranges between 30 and 45 and indicates a component of left-lateral shear for the ductile deformation (Fig. 9(E)).
In structural domain II the strain is concentrated within discrete micro-shear zones that can
be observed locally between quartz and feldspar
phase boundaries. In these zones, quartz grains
show dynamic recrystallisation and anisometric
subgrains with long axes oriented oblique (40)
to the shear zone boundaries (Fig. 9(F)). In the
feldspar, sets of intracrystalline microfractures
were observed and deformation twinning and
discontinuous undulatory extinction is abundant.
Occasionally, subgrain development and recrystallisation is also observed within narrow zones
in the feldspar grains. Biotite recrystallized during this deformation episode. Apart from the
distinct shear zones, the fabric was only weakly
affected by deformation, resulting in undulatory
extinction in quartz and twinning in feldspar.
These observations indicate that in the domain II
deformation was distinctly localized. The high
temperature deformation fabric observed in
structural domain I was overprinted by lower
temperature deformation in structural domain II.
The microstructures in samples P 23 (Fig. 9(F))
and P 6 from this domain indicate a dextral
movement for this low temperature shear deformation which is opposite to the left-lateral movement under amphibolite facies metamorphic
conditions at the HZ as a whole (Fig. 8). A
strong indication of this deformation episode is
also observed in the samples P 9, P 11 and P 12
at the western part of the study area, where the
magnetic lineations are rotated from N S (structural domain I) towards NW SE (structural do-

main II). Here, quartz is strongly affected by


undulosity and subgrain development and in
feldspar densely spaced deformation twinning
and fracturing is observed.
Almost all samples experienced a minor overprint under lower greenschist facies metamorphic
conditions. This caused fracturing of feldspar
and hornblende and, occasionally, was accompanied by retrograde mineral reactions such as formation of chlorite along fractures in hornblende,
as well as the sericitization of feldspar.

6. Interpretation of AMS data and


microstructures
The axes of the AMS-ellipsoids are generally
subparallel with the fabric axes with smax parallel
to the mineral or stretching lineation and smin
normal to the metamorphic foliation (Figs. 3 and
5). In samples with high susceptibility this anisotropy is due to anisometric magnetite grains
elongated in the direction of lineation (Gregoire
et al., 1995), and to an anisotropy in distribution
of magnetite (Hargraves et al., 1991). In samples
dominated by a paramagnetic fabric the coincidence is caused by the preferred orientation of
biotite and hornblende, which are minerals with
high single crystal susceptibility anisotropies
(Friedrich, 1995; Siegesmund et al., 1995;
Hrouda et al., 1997). The magnetic fabric, therefore, can provide information about the orientation of the finite strain in the HZ. However, it
has to be taken into account, that the AMS
method integrates across the whole fabric and,
therefore, may reflect resultants of interfering
subfabrics (Rochette and Vialon, 1984; Hrouda
and Potfaj, 1993). For example in a SC geometry (sensu Berthe et al., 1979), deflections between the magnetic foliation (S) and the shear
foliation (C-plane) may occur (Aranguren et al.,
1995). However, in the samples investigated here,
such a shear fabric is only weakly developed and
the resulting deflections are considered to be negligible. Another problem interpreting AMS data
can be caused by mineralogical components that
produce an inverse magnetic fabric as it is occasionally observed for hornblende with smax per-

H. de Wall et al. / Precambrian Research 107 (2001) 179194

189

Fig. 9. Structural features of the HZ: A, B Field photographs from the western margin of the HZ, looking S (A) and SW (B),
respectively, from a point at the southern end of Gabal Um Rasein (Fig. 2 for location). The photographs are arranged so as to
show the steep dip of the foliation within the HZ (cliffs in A) and the shallowing dips away from the HZ (towards W, B). The central
part of the slope in B is build up of the lbib gneiss, the top of the Gabal Gerf ophiolite nappe (compare map, Fig. 2). C Example
of banded gneiss with amphibolite layers (black), showing (broadly symmetric) boudinage, which indicates N S extension and E W
shortening. East side of Um Rasein, Fig. 2 for location; hammer for scale (above central boudin). D Handspecimen of amphibolite
from the HZ (P 26) cut parallel with a mineral lineation marked by elongated feldspar porphyroclasts (structural X direction; XY
section) and normal to the mineral lineation (YZ section), respectively. The amphibolite shows a pronounced stretching fabric. E,
F Photomicrographs of orthogneisses from the HZ, showing examples from domain I (E) and II (F), respectively: (E) coarse-grained,
elongated quartz with lobate boundaries but little internal deformation defines an early foliation (S). Younger foliation surfaces (C),
defined by mica minerals, cut the early foliation and are interpreted here as shear surfaces. This S C fabric indicates a sinistral sense
of shear (sample P 21, X polarizers). (F) Dynamically recrystallized quartz. Quartz grains and subgrains are oriented oblique to the
earlier foliation and indicate a dextral shear deformation (sample P 23, X polarizers).

pendicular to the long mineral axis (Rochette et


al., 1992; Friedrich, 1995). For the investigated
amphibolites and hornblende-bearing gneisses, a
parallel orientation of the magnetic lineation
with the macroscopically visible mineral lineation was deduced so that a contribution of
inverse fabric elements can be excluded.
Microstructural analyses and AMS-directional

data and shape parameters indicate a variation


in deformation fabric along the HZ ranging
from strongly oblate to strongly prolate geometries. Simple shear deformation is concentrated in discrete zones. Between these zones the
fabric shows mostly a pronounced stretching lineation but locally also flattening fabrics with no
macroscopically visible mineral lineations.

190

H. de Wall et al. / Precambrian Research 107 (2001) 179194

As a consequence of strong simple shear deformation a divergence between the high strain and
low strain geometries should be expected, since
the angle between the shear zone boundary and
the X-axis of the strain ellipsoid decreases from
45 towards a subparallel orientation with progressive simple shear deformation. Such oblique
magnetic fabrics have been observed in regions
were granite emplacement is related to strike slip
fault zones (Djouadi et al., 1997) or to transform
geometry (Bouillin et al., 1993).
However, in the present study, the lineations in
high strain and low strain areas are parallel, for
example 352/06 for the nearly undeformed metagranitoid (P 10) and 359/20 for a strongly deformed orthogneiss (P 21). This supports the
results of microstructural observations that noncoaxial deformation is a minor component of the
bulk deformation in the HZ.
The pronounced sigmoidal deflection of the
magnetic lineation in structural domain II (Fig. 6)
is clear on the map in Fig. 9, where the trend of
the traces of the magnetic lineation trajectories is
presented. This deflection and progressive rotation of the magnetic lineation in the SW-part of
the study area are accompanied by low temperature deformation as described in the chapter
above (fractures in feldspar, strong undulose extinction and localized recrystallization in quartz).
For some samples a sense of shear could be
determined and is indicated on the map (Fig. 9).

7. Discussion
From the regional tectonic context and from
geologic maps, a strike-slip movement along the
HZ has been inferred (Vail, 1985, 1988; Kroner et
al., 1987; Greiling et al., 1994; Nasr et al., 1998).
However, at least some field-based structural
studies (Stern et al., 1990; Miller and Dixon,
1992) failed to detect traces of such a large-scale
deformation and in particular those of a strikeslip episode along the north-central part of the
HZ. These latter results are partly in accordance
with our observations in the field that the deformation appears to be relatively weak in places.
But detailed studies revealed a more complex

story: the amount and geometry of strain show


strong variations across the HZ.
Macroscopic fabric elements and the magnetic
fabric in the study area show a good correlation.
For example, magnetic lineations are subparallel
with the stretching or mineral lineations and magnetic foliations are subparallel with the metamorphic banding. AMS and field data can thus be
combined for an interpretation of the deformation
regime in the zone of concentrated deformation.
In structural domain I (Figs. 6 and 9), the major
part of the HZ, the foliation poles form a girdle
distribution around the orientation of a welldefined subhorizontal NS trending lineation that
is developed locally as a stretching lineation. This
geometry is attributable to a constrictional
regime, in contrast to a flattening regime, where a
well-clustered foliation and a more scattered lineation should be expected. These results are in
conflict with the interpretation of Miller and
Dixon (1992) who described a down dip direction
for the lineation and inferred a predominance of
crustal shortening by folding and thrusting. However, we agree with these authors that the ductile
deformation in the HZ is mainly coaxial and
strike slip deformation is limited.
The curvature and bending of pre-existing
structural units towards the HZ and into parallelism with the HZ-trend (Fig. 1) may be suggestive of a drag, as it is produced by (strike-slip)
fault movement. However, this drag is broadly
symmetrical on both sides of the HZ and thus not
indicative of any significant relative movement
across the HZ. Instead, our data imply that this
outcrop pattern is that of an antiform, which
becomes progressively tighter at the HZ. Although this antiform is almost horizontal at a
local scale (Fig. 3), it plunges gently towards the
south at a regional scale, thus producing the
observed outcrop pattern of fold limbs bending
away from the tight fold closure at the HSZ
towards the NW and NE (Fig. 1).
No deviation of magnetic lineation orientations
between the different lithologies is observed. This
holds also for the deformed granitoids. Therefore,
the intrusion of the granitoids is considered to
have taken place before or during the deformation
along the HZ. Consequently, such high strain

H. de Wall et al. / Precambrian Research 107 (2001) 179194

zones as the HZ are not only important late- to


post-tectonic elements but may have played a
major role also during and after the accretionary
orogenic evolution, probably also during the ascent of granitoid plutons (Hutton and Reavy,
1992; DLemos et al., 1992; Castro et al., 1999).
The observed lower temperature overprint in
the NE SW-trending structural domain II by
dextral shear deformation is geometrically related
to the NE SW-trending dextral shear zone, described by Stern et al. (1990) (Figs. 1 and 9) as
late structures truncating the N S-trending fabric
of the HZ. These authors interpreted such structures as shear components of a transpressive deformation with shortening normal to the HZ,
which developed contemporaneously or at a late
stage of HZ deformation. However, our study
shows that the structures in the HZ are related to
two different episodes of deformation. An early,
mainly coaxial phase, with E W shortening and
pronounced NS extension under amphibolite facies metamorphic conditions (structural domain I)
and a subsequent, mostly non-coaxial overprint
under considerably lower temperature, greenschist
facies metamorphic conditions (structural domain
II).
The late overprint could be related to stress
fields active during the break-up of the Red Sea,
where continental high strain zones may have
been reactivated (Dixon et al., 1987; Talbot and
Ghebreab, 1997). However, earlier work on postPan-African basement deformation showed a
more brittle faulting during Red Sea tectonics
(Greiling et al., 1988). Therefore, it appears to be
more likely that also the domain II deformation is
related to a latest Pan-African deformational
episode. Consequently, the NE SW-trending,
dextral strike-slip zone at domain II is interpreted
as a conjugate zone relative to the NW SE-trending, sinistral fault zones of Wadi Hodein or Najd
(Stern, 1985, 1994; Greiling et al., 1994). In addition to Stern et al. (1990) our results show that
such dextral, conjugate strike-slip zones may be
more widespread than it was known previously,
although their regional importance is limited.
In conclusion, the present results show the HZ
as a zone of high strain but with only minor
components of wrench faulting. Deformation is

191

dominated by pure shear producing EW-shortening, with a strong NS-extensional component,


resulting in constrictional strain. It is only a subsequent overprint, which produced minor dextral
wrench movement along a discrete, NE SWtrending zone (domain II on Fig. 9).
Since these structures overprint the ophiolite
nappes of the Allaqi-Sol Hamed suture, it can be
inferred that the EW-shortening which produced
the HZ took place after suturing of the Gabgaba
terrane with the terrane adjacent towards north.
In a wider regional context, the present results
provide an important complement to the tectonic
history of the ANS: the Allaqi-Sol Hamed suture,
trending broadly EW, is overprinted by EW
shortening during the formation of the HZ. Such
an EW shortening has been proposed during the
suturing of the Gabgaba terrane with older continental crust towards W (Keraf suture, Abdelsalam et al., 1998, Fig. 1). As a consequence, the
accretion of the ANS in northern Sudan southern Egypt, can be confirmed as polyphase or at
least a two-phase event with early suturing by
NS convergence (formation of Allaqi-Sol
Hamed suture) and a later stage with c. EW
convergence leading to collision of the earlier
assembled terranes with a craton towards west
and the formation of the Keraf suture (Abdelsalam et al., 1998). These results preclude earlier
views of the Pan-African evolution as a broadly
continuous episode of transpressional orogeny
(OConnor et al., 1994). Neither are there traces
of large-scale escape movements in the present
area to the south of the Najd Faults (Stern, 1985,
1994) in a wider sense.

Acknowledgements
This work is a part of a cooperation between
the Geologisch-Palaontologisches Institut, Heidelberg, and the Egyptian Geological Survey and
Mining Authority (EGSMA), Cairo, supported by
the BMBW through the Forschungszentrum
Julich, International Bureau. The financial assistance is gratefully acknowledged. We thank the
Chairmen of EGSMA, G.M. Naim and M. El
Hinnawy for their generous support. Many

192

H. de Wall et al. / Precambrian Research 107 (2001) 179194

thanks to L. Nano for his assistance in laboratory


work and L.N. Warr, A. Kontny and A. Maklouf
for helpful discussions. M. El Eraki provided
some of the samples. We thank reviewers J. Meert
and R.J. Stern for detailed comments and suggestions that considerably improved the quality of
the paper.

References
Abdeen, M.M., Dardir, A.A., Greiling, R.O., 1992. Geological
and structural evolution of the Wadi Queih Area (N
Quseir), Pan African Basement of the Eastern Desert of
Egypt. Zbl. Geol. Palaont. Teil. 1, 26532659.
Abdelsalam, M.G., 1994. The Oko Shear Zone, Sudan: postaccretionary deformation in the ArabianNubian shield. J.
Geol. Soc. Lond. 151, 767776.
Abdelsalam, M.G., Stern, R.J., Copeland, P., Elfaki, E.M.,
Elhur, B., Ibrahim, F.M., 1998. The Neoproterozoic Keraf
Suture in NE Sudan: sinistral transpression along the
eastern margin of West Gondwana. J. Geol. 106, 133147.
Akaad, M.K., El Ramly, M.F., 1958. Seven new occurences of
the Igla Formation in the Eastern Desert of Egypt. Geol.
Surv. Egypt, Paper 3, 37 pp.
Aranguren, A., Cuevas, J., Tubia, J.M., 1995. Composite
magnetic fabric in SC mylonites from a granodioritc
pluton. Terra Abstracts 7, 142.
Bennet, J.D., Mosley, P.N., 1987. Tiered-tectonics and evolution, E Desert and Sinai, Egypt, Current Res. African
Earth Sciences, in: Matheis, G., Schandelmeier, H. (Eds.),
Geoscientific Research in Northeast Africa, Balkema, 79
82.
Berthe, D., Choukroune, P., Jegouzo, P., 1979. Orthogneiss,
mylonite and non-coaxial deformation of granites: the
example of the South Armoricain shear zone. J. Struct.
Geol. 1, 31 42.
Borradaile, G.J., 1988. Magnetic susceptibility, petrofabrics
and strain. Tectonophysics 156, 120.
Borradaile, G.J., Henry, B., 1997. Tectonic applications of
magnetic susceptibility and its anisotropy. Earth Sci. Rev.
42, 49 93.
Bouchez, J.L., Gleizes, G., Djouadi, M.T., Rochette, P., 1990.
Microstructure and magnetic susceptibility applied to emplacement kinematics of granites: the example of the Foix
pluton (French Pyrenees). Tectonophysics 184, 157171.
Bouchez, J.L., Hutton, D.H.W., Stephens, W.E. (Eds.), 1997.
Granite: From Segregation of Melt to Emplacement Fabrics. Kluwer Academic Publishers, Dordrecht, 358 pp.
Bouillin, J.-P., Bouchez, J.L., Lespinasse, P., Pecher, A., 1993.
Granite emplacement in an extensional setting: an AMS
study of the magmatic structures of Monte Capanne (Elba,
Italy). Earth Planet Sci. Lett. 118, 263279.
Burke, K., Sengor, C., 1986. Tectonic escape in the evolution
of the continental crust. AGU Geodynam. Ser. 14, 4153.

Castro, A., Fernandez, S., Vigneresse, J.L. (Eds.) 1999. Understanding granites: Integrating new and classical techniques.
Geol. Soc. Lond. Spec. Publ., 168, pp. 278.
de Wall, H., Warr, L.N., Kamal El-Din, G.M., Greiling, R.O.,
1998. Spatorogene Entwicklung an steilen Deformationszonen pure shear oder simple shear? Panafrikanische
Atalla Zone, Eastern Desert, A8 gypten. Freiberger
Forschungsheft C471, 55 57.
de Wall, H., Greiling, R.O., El Kady, M., Sadek, M.F., 1999.
Magnetic fabric study in the Um Rasein area, Northern
Hamisana zone, Eastern Desert, Egypt. In: de Wall, H.,
Greiling, R.O. (Eds.), Aspects of Pan-African Tectonics,
International Cooperation, Bilateral Seminars Int Bureau,
Forschungszentrum Julich. 32, 85 89.
Dixon, T.H., Stern, R.J., Hussein, I.M., 1987. Control of Red
Sea rift geometry by Precambrian structures. Tectonics 6,
551 571.
Djouadi, M.T., Gleizes, G., Ferre, E., Bouchez, J.L., Caby, R.,
Lesquer, A., 1997. Oblique magmatic structures of two
epizonal granite plutons, Hoggar Algeria: late orogenic
emplacement in a transcurrent orogen. Tectonophysics
279, 351 374.
DLemos, R.S., Brown, M., Strachan, R.A., 1992. Granite
magma generation, ascent and emplacement within a transpressional orogen. J. Geol. Soc. Lond. 149, 487 490.
EGSMA, 1996. Geologic map of Wadi Jabjabah Quadrangle,
Egypt, NF36K. Geological Survey of Egypt, Cairo.
EGSMA, 1999. Geologic map of Jabal Ilban Quadrangle,
Egypt, NF37: Geological Survey of Egypt, Cairo.
El Eraki, M.A., Greiling, R.O., 1998. Application of AMS
approach to the study of a major continental shear zone
(Hamisana Shear Zone, Arabian Nubian shield). Freib.
Forschungsheft C471, 66 67.
El Ramly, M.F., Greiling, R., Kroner, A., Rashwan, A.A.,
1984. On the tectonic evolution of the Wadi Hafafit area
and evirons, Eastern Desert of Egypt, Fac. Earth Sci.,
King Abdulaziz Univ. Jeddah Bull. 6, 113 126.
El Sharkawy, M.A., El Bayoumi, R.M., 1979. The ophiolites
of Wadi Ghadir area, Eastern Desert. Egypt. Annals Geol.
Surv. Egypt 9, 125 135.
Fitches, W.R., Graham, R.H., Hussein, I.M., Ries, A.C.,
Shackleton, R.M., Price, R.C., 1983. The late Proterozoic
ophiolite of Sol Hamed. NE Sudan, Precambrian Res. 19,
385 411.
Friedrich, D., 1995. Gefugeuntersuchungen an Amphiboliten
der Bohmischen Masse unter besonderer Berucksichtigung
der Anisotropie der magnetischen Suszeptibilitat. Geotekt.
Forsch. 82, 1 118.
Fritz, H., Wallbrecher, E., Khudeir, A.A., Abu el Ala, F.,
Dallmeyer, D.R., 1996. Formation of Neoproterozoic
metamorphic core complexes during oblique convergence
(Eastern Desert, Egypt.). J. African Earth Sci. 23, 311
329.
Gass, I.G., 1981. Pan-African (Upper Proterozoic) plate tectonics of the Arabian Nubian shield. In: Kroner, A. (Ed.),
Precambrian Plate Tectonics, Developments in Precambrian Geology. Elsevier, Amsterdam, 387 405.

H. de Wall et al. / Precambrian Research 107 (2001) 179194


Greiling, R.O., 1997. Thrust tectonics in crystalline domains:
The origin of a gneiss dome. Proc. Ind. Acad. Sci. (Earth
Planet Sci.) 106, 209220.
Greiling, R.O., El Ramly, M.F., El Akhal, H., Stern, R.J.,
1988. Tectonic evolution of the northwestern Red Sea
margin as related to basement structures. Tectonophysics
153, 179 191.
Greiling, R.O., Abdeen, M.M., Dardir, A.A., El Akhal, H., El
Ramly, M.F., Kamal El Din, G.M., Osman, A.F., Rashwan, A.A., Rice, A.H.N., Sadek, M.F., 1994. A structural
synthesis of the Proterozoic ArabianNubian shield in
Egypt. Geol. Rdsch. 83, 484501.
Greiling, R.O., de Wall, H., Warr, L.N., Naim, G.M., Hussein, A.A., Sadek, M.F., Abdeen, M.M., ElKady, M.F.,
Makhlouf, A., 1998. Basement structure in Eastern Egypt:
quantitative perspectives for the second century. Geol.
Surv. Egypt Spec. Pub. 75, 289302.
Gregoire, V., Sant Blanquat, M., Nedelec, A., Bouchez, J.L.,
1995. Shape anisotropy versus magnetic interactions of
magnetite grains: experiments and application to AMS in
granitic rocks. Geophys. Res. Lett. 22, 27652768.
Hargraves, R.B., Johnson, D., Chan, C.Y., 1991. Distribution
anisotropy: the cause of AMS in igneous rocks? Geophys.
Res. Lett. 18, 2193 2196.
Hutton, D.H.W., Reavy, R.J., 1992. Strike-slip tectonics and
granite petrogenesis. Tectonics 11, 960967.
Hrouda, F., 1994. A technique for the measurement of thermal
changes of magnetic susceptibility of weakly magnetic
rocks by the CS-2 apparatus and KLY-2 Kappabridge.
Geophys. J. Int. 118, 604612.
Hrouda, F., Jelinek, V., Hruskova, L., 1990. A package of
programs for statistical evaluation of magnetic data using
IBM-PC computers. EOS Trans. Am. Geophys. Union 71,
43, 1289.
Hrouda, F., Potfaj, M., 1993. Deformation of sediments in the
post-orogenic Intra-Carpathian Paleogene Basin as indicated by magnetic anisotropy. Tectonophysics 224, 425
434.
Hrouda, F., Schulmann, K., Suppes, M., Ullemayer, K., de
Wall, H., Weber, K., 1997. Quantitive Relationship between Low-Field AMS and Phyllosilicate Fabric: A Review. Phys. Chem. Earth 22, 1531560.
Jelinek, V., 1980. Kappabridge KLY-2. A precision laboratory
bridge for measuring magnetic susceptibility of rocks and
its application (including anisotropy). Geofyzika Brno, 4 p.
Jelinek, V., 1981. Characterization of the magnetic fabrics of
rocks. Tectonophysics 79, 6367.
Johnson, P.R., Kattan, F., 1999. The Ruwah Ar Rika, and
Halaban-Zarghat
fault
zones:
northwest-trending
Neoproterozoic brittle-ductile shear zones in west-central
Saudi Arabia, in: de Wall, H., Greiling, R.O. (Eds.),
Aspects of Pan-African Tectonics, International Cooperation, Bilateral Seminars Int. Bureau, Forschungszentrum
Julich 32, 75 79.
Kroner, A., 1985. Ophiolites and the evolution of tectonic
boundaries in the late Proterozoic ArabianNubian shield
of northeast Africa and Arabia. Precambrian Res. 27,
277 300.

193

Kroner, A., Greiling, R., Reischmann, T., Hussein, I.M.,


Stern, R.J., Durr, S., Kruger, J., Zimmer, M., 1987. PanAfrican crustal evolution in northeast Africa. In: Kroner,
A., (Ed.), Proterozoic Lithospheric Evolution. Geodyn.
Ser., 17, 235 257.
Miller, M.M., Dixon, T.H., 1992. Late Proterozoic evolution
of the northern part of the Hamisana zone, northeast
Sudan: constraints on Pan-African accretionary tectonics.
J. Geol. Soc. Lond. 149, 743 750.
Nasr, B.B., Beniamin, N.Y., Sadek, M.F., Abd Elmigid, E.,
Smith, M., 1998. The geological and tectonic study of the
Gerf ultrabasic complex, South Eastern Desert, Egypt.
Geol. Surv. Egypt Pub. 75, 635 643.
OConnor, E.A., 1996. One hundred years of geological partnership: British-Egyptian geoscientific collaboration,
1896 1996. British Geological Survey, Keyworth.
OConnor, E.A., Bennet, J.D., Rashwan, A.A., Nasr, B.B.,
Mansour, M.M., Romani, R.F., Sadek, M.F., 1994.
Crustal growth in the Nubian shield of south eastern
Egypt. Geol. Surv. Egypt Spec. Pub. 69, 189 195.
Osman, A.F., Greiling, R.O., Warr, L.N., Hilmy, M.E.,
Ragab, A.I., El Ramly, M.F., 1993. Distinction of different
syn- and late orogenic Pan-African sedimentary sequences
by metamorphic grade and illite crystallinity, W. Zeidun,
Central Eastern Egypt. In: Thorweihe, U., Schandelmeier,
H. (Eds.) Geoscientific Research in Northeast Africa.
Balkema 21 25.
Rashwan, A.A., 1991. Petrography, Geochemistry and Petrogenesis of the Migif-Hafafit gneisses at Hafafit mine Area,
Egypt. Forschungszentrum Julich GmbH. Scientific Ser.
Int. Bureau 5, 359.
Rashwan, A.A., Greiling, R.O., 1999. Pan-African basal
gneisses of the SE desert of Egypt: lithology, geochemistry
and tectonic setting. In: de Wall, H., Greiling, R.O. (Eds.),
Aspects of Pan-African Tectonics, International Cooperation, Bilateral Seminars Int. Bureau, Forschungszentrum
Julich 32, 63 71.
Rice, A.H.N., Osman, A.F., Sadek, M.F., Ragab, A.I., Abdeen, M.M., 1993. Preliminary comparison of six late- to
post- Pan-African molasse basins, E. Desert, Egypt. In:
Thorweihe, U., Schandelmeier, H. (Eds.). Geoscientific Research in Northeast Africa. Balkema, 41 45.
Ries, A.C., Shackleton, R.M., Graham, R.H., Fitches, W.R.,
1983. Pan-African structure, ophiolites and melange in the
EDE: A traverse at 26N. J. Geol. Soc. Lond. 140, 75 95.
Rochette, P., 1987. Magnetic susceptibility of the rock matrix
related to magnetic fabric studies. J. Struct. Geol. 9, 1015
1020.
Rochette, P., Vialon, P., 1984. Development of planar and
linear fabrics in Dauphinois shales and slates (French
Alps) studied by magnetic anisotropy and its mineralogical
control. J. Struct. Geol. 6, 33 38.
Rochette, P., Jackson, M., Aubourg, C., 1992. Rock magnetism and the interpretation of anisotropy of magnetic
susceptibility. Rev. Geophys. 30, 209 226.
Sadek, M.F., 1995. Geology, Geochemistry and Structure of
Gabal Muqsim Area and Environs, South Eastern Desert,

194

H. de Wall et al. / Precambrian Research 107 (2001) 179194


Stern, R.J., 1994. Arc assembly and continental collison in the
Neoproterozoic East African orogen: Implications for the
consolidation of Gondwanaland. Ann. Rev. Earth Planetary Sci. 22, 319 351.
Sturchio, N.C., Sultan, M., Batiza, R., 1983. Geology and
origin of the Meatiq dome, Egypt: A Precambrian metamorphic core complex? Geology 11, 72 76.
Sultan, M., Arvidson, R.E., Duncan, I.J., Stern, R.J., 1988.
Extension of the Najd Shear System from Saudi Arabia to
the central Eastern Desert of Egypt based on integrated
field and Landsat observations. Tectonics 7, 1291 1306.
Talbot, C.J., Ghebreab, W., 1997. Red Sea detachment and
basement core complexes in Eritrea. Geology 25, 655 659.
Tarling, D.H., Hrouda, F., 1993. The Magnetic Anisotropy of
Rocks, Chapman & Hall, 217 p.
Unrug, R., 1997. Rodinia to Gondwana: the geodynamic map
of Gondwana supercontinent assembly. GSA Today 7,
1 6.
Vail, J.R., 1985. Pan-African (late Precambrian) tectonic terrains and the reconstructions of the Arabian Nubian
shield. Geology 13, 839 842.
Vail, J.R., 1988. Tectonics and evolution of the Proterozoic
Basement of NE Africa. In: El Gaby, S., Greiling, R.O.
(Eds.) The Pan-African Belt of NE Africa and Adjacent
Areas. Earth Evol. Sci., Vieweg, Wiesbaden, 195 226.
Windley, B., 1992. Proterozoic collisional and accretionary
orogens. In: Condie, K.C. (Ed.), Proterozoic Crustal Evolution. Elsevier, Amsterdam, pp. 419 446.
Zimmer, M., Kroner, A., Jochum, K.P., Reischmann, T.,
Todt, W., 1995. The Gabal Gerf complex: A Precambrian
N-Morb ophiolite in the Nubian Shield, NE Africa. Chem.
Geol. 123, 29 51.

Egypt. Forschungszentrum Julich GmbH. Scientific Ser.


Int. Bureau 32, 280.
Sanderson, D.J., Marchini, W.R.D., 1984. Transpression. J.
Struct. Geol. 6, 449 458.
Shackleton, R.M., Ries, A.C., Graham, R.H., Fitches, W.R.,
1980. Late Precambrian ophiolitic melange in the Eastern
Desert of Egypt. Nature 285, 472474.
Shimron, A.E., 1990. The Red Sea line a Late Proterozoic
transcurrent fault. J. African Earth Sci. 11, 95112.
Siegesmund, S., Ullemeyer, K., Dahms, M., 1995. Control of
magnetic rock fabrics by mica preferred orientation: a
quantitative approach. J. Struct. Geol. 17, 16011613.
Smith, M., OConnor, E., Nasr, B.B., 1999. Transpressional
flower structures and escape tectonics: A new look at the
Pan-African collision in the Eastern Desert, Egypt. In: de
Wall, H., Greiling, R.O. (Eds.) Aspects of Pan-African
Tectonics, International Cooperation, Bilateral Seminars
Int. Bureau, Germany: Forschungszentrum Julich 32, 81
82.
Stern, R.J., 1985. The Najd Fault System, Saudi Arabia and
Egypt: a late Precambrian rift-related transform system?
Tectonics 4, 497 511.
Stern, R.J., Kroner, A., Manton, W.I., Reischmann, T., Mansour, M., Hussein, I.M., 1989. Geochronology of the late
Precambrian Hamisana shear zone, Red Sea Hills, Sudan
and Egypt. J. Geol. Soc. Lond. 146, 10171029.
Stern, R.J., Nielsen, K.C., Best, E., Sultan, M., Arvidson,
R.E., Kroner, A., 1990. Orientation of late Precambrian
sutures in the ArabianNubian shield. Geology 18, 1103
1106.

Das könnte Ihnen auch gefallen