0% found this document useful (0 votes)
12 views322 pages

Tectonics - Recent Advances

This chapter provides background information on oceanic core complexes (OCCs) found along slow-spreading mid-ocean ridges. OCCs are characterized by tectonized lithosphere exposing altered gabbros and serpentinized mantle. The largest OCCs are found in the Mid-Atlantic Ridge and Philippine Sea. The chapter focuses on the Sierra Leone area of the MAR between 5-7°N, which contains evidence of OCC formation including serpentinized peridotites and gabbros. However, unlike typical OCCs, the altered rocks in Sierra Leone are found mainly in graben depressions rather than along the ridge axis. The chapter aims to analyze geological data from this

Uploaded by

Jr Assis
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
12 views322 pages

Tectonics - Recent Advances

This chapter provides background information on oceanic core complexes (OCCs) found along slow-spreading mid-ocean ridges. OCCs are characterized by tectonized lithosphere exposing altered gabbros and serpentinized mantle. The largest OCCs are found in the Mid-Atlantic Ridge and Philippine Sea. The chapter focuses on the Sierra Leone area of the MAR between 5-7°N, which contains evidence of OCC formation including serpentinized peridotites and gabbros. However, unlike typical OCCs, the altered rocks in Sierra Leone are found mainly in graben depressions rather than along the ridge axis. The chapter aims to analyze geological data from this

Uploaded by

Jr Assis
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 322

Chapter 1

Cyclic Development of Axial Parts of


Slow-Spreading Ridges: Evidence from
Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N

E.V. Sharkov

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/39372

1. Introduction
Slow- and ultraslow-spreading mid-ocean ridges became come to attention of researchers
in recent years again because identification of so-called oceanic core complexes (OCCs).
These complexes are characterized by tectonized and heterogeneous lithosphere, large
yields of altered gabbros and serpentinized mantle at the oceanic bottom and presence of
large deep-sea hydrothermal fields and mineralization (Conference…, 2010). For example,
OCCs are quite common in the slow-spreading Mid-Atlantic Ridge (MAR) where they
make up ~30% of its length (Escartín et al., 2008; Smith et al., 2008; MacLeod et al., 2009
and references therein). OCC form about 50% of ultra-slow South-West Indian Ridge
length (Cannat, 2010); the most studied site here is Atlantis Bank (Thy, 2003; Schwartz et
al., 2009). OCCs are known in back-arc seas too, for example in the Philippine Sea (Ohara
et al., 2001).

The largest of the OCC is the Godzilla Mullion in the Philippine Sea. The second in the
world and the largest in the MAR is the St. Peter and St. Paul complex about 90 km long and
up to 4000 m in height, located near the axial zone of MAR in the equatorial region, south of
the Sierra Leone area. A feature of this OCC is a dissected topography, with its most
elevated blocks even reach the ocean surface to form the St. Peter and St. Paul Rocks. They
are composed mainly serpentinized often sheared mantle hornblende (metasomatized)
peridotites, containing hornblendite schlierens and veins (Roden et al., 1984; Hékinian et al.,
2000). Such peridotites are commonly found as xenoliths in intraplate (plume-related)
basalts of oceans and continents, representing fragments of the cooled upper parts of mantle
plume heads above its melting zone (Magmatic ..., 1988); so, that is tectonic block of the
upper edge of a mantle plume moved out to the surface here.

© 2012 Sharkov, licensee InTech. This is an open access chapter distributed under the terms of the Creative
Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits unrestricted use,
distribution, and reproduction in any medium, provided the original work is properly cited.
4 Tectonics – Recent Advances

The name for OCCs was given by analogy to metamorphic core complexes (metamorphic
cores), located in core (inner) parts of orogens on continents. In essence, such complexes are
represented by exposed metamorphosed deep-crustal rocks, which underwent by viscous-
plastic and brittle deformation. The same situation is typical for OCCs, which are outcrops
of tectonized and altered deep-seated crustal and mantle rocks in the axial parts of mid-
oceanic ridges. Because of characteristic striated surface (mullion structure), such complexes
often referred as megamullions (Tucholke, 1998).

According to the commonly accepted (Penrose) model of plate tectonics, the occurrence of
mid-ocean ridge (MOR) is associated with uprising of hot deep mantle material, which,
reaching shallow depths, begins to melt due to adiabatic decompression. It is assumed that
formation of new oceanic crust occurs here, which symmetrical spreads to both sides of the
ridge due to convection currents in underlying mantle. Resulted excess of the crust is
absorbed in subduction zones beneath island arcs and active continental margins. The axial
part of the MORs, where crust is generated (constructive plate boundaries), considered as
centers or zones of oceanic spreading.

From such positions, outcrops of plutonic rocks in the spreading zones do not fit in the
traditional model of plate tectonics. According to numerous studies of OCCs, axial parts
of ridges are uplifted relative to their average height, and often have asymmetrical
structure, where outcrops of plutonic rocks are disposed outside of axial valleys, where
neovolcanic hills are located (Ildefonse et al., 2007; Smith et al., 2008; MacLeod et al., 2009,
etc.). Modern volcanism is practically absent, however, numerous hydrothermal vents
occurred.

In this regard, it was suggested that oceanic core complex results from activity of an oceanic
detachment fault (Conference…, 2010). This fault is a large-offset normal fault formed at or in
the vicinity of a mid-ocean ridge axes that accommodates a significant fraction of the plate
separation (Fig. 1); offsets range from kilometers to tens kilometers or more. According to
this model, oceanic detachment faults may initiate as steep normal faults at depth, and turn
into shallow low angle extensional faults through rotation of the footwall. It is suggested
that this type of spreading should be recognized as a fundamentally distinct mode of
seafloor spreading that does not result in a classical Penrose model of oceanic crustal
structure. However, many elements and details of this hypothesis of “one-side spreading”
are poorly justified (i.e., unknown fault geometry at depth, structure of magmatic systems,
route of hydrothermal currents, etc.), as well as motives of appearance of such detachment
faults, which absent in fast-spreading ridges.

Identification of the OCCs set to geologists a number of problems which solution is possible
only using the complex of geological, petrological and geochemical studies. Such work was
done on example of Sierra Leone area, located in axis of the MAR (5-7°N). It was based on
materials dredged during the cruises of R/V "Akademik Ioffe" (10th cruise, 2001-2002) and
"Professor Logachev" (22th cruise, 2003) (Sharkov et al., 2005, 2007, 2008; Savelieva et al.,
2006; Simonov et al., 2009; Aranovich et al., 2010). Judging on presence here of serpentinites
upon mantle peridotites and altered tectonized lower-crustal gabbros, as well as widespread
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 5

development of extensional structures, including normal faults, the Sierra Leone area can be
determined as OCC. However, unlike of typical OCCs, the altered deep-seated rocks were
found mainly in the bottom and slopes of deep graben-like depressions, whereas surface of
the ridge is covered by the flows of fresh pillow lavas with chilled glassy crusts; i.e. a kind of
structural discordance occurred in the area. Marked asymmetry in structure of the ridge is
not found here as well as clear evidence of oceanic detachment fault existence. In this
context, studied area is of great interest as a possible example of the transition from the
typical OCCs to regions of the ridge between them where only basalts developed and
spreading is symmetrical.

Figure 1. Scheme of oceanic core complex with oceanic detachment fault. After: Conference ..., 2010.

The aim of this paper, based on our data from the Sierra Leone area and published
information, to discuss diverse processes, occurred in axes of the modern slow-spreading
MORs, and give a new way to interpret of geological, petrological and geodynamical data
both in their spreading zones and in underlying mantle.

2. Brief description of geological background


The studied segment of the MAR with strongly dissected relief is located in the vicinity of
non-transformed Sierra Leone Fault, between Bogdanov Fracture Zone (710' N) and 500' N
(Fig. 2). South to area, from 500' N and to the Strakhov FZ, the MAR represents leveled
basaltic plateau, crossed by narrow meridionally-oriented axial rift valley. Geological
structure of the area is showed in (Pushcharovsky et al, 2004).
6 Tectonics – Recent Advances

A - location of the Sierra Leone area in the Atlantic; B – Bathymetric map of the Sierra Leone area: 1- sites of dredgins; 2
– sites of Rosetta module; 3-6 – profiles; C – Markov Deep

Figure 2. Sierra Leone area, the Central Atlantic

Feature of morphostructural image of the area is lack of transform faults and the spreading
zone is represented here by en echelon system of graben-like depressions (valleys) of 4-5 km
depth from ocean surface. As it mentioned above, altered deep-seated rocks found mainly in
the sides of rift valleys and on their floor, at that outcrops of plutonic rocks are traced for
about 60 km along the MAR axis. Flows of fresh pillow lavas cover top of the ridge and
partly fill bottom of some rift valleys. Thickness of these flows is small because within the
area of their distribution are found outcrops of altered plutonic rocks. Despite the uneven
sampling, we can say with confidence that both sides of the rift valleys formed by the same
complex of rocks that characterize the entire section of oceanic crust.

The structure of bedrocks on the eastern slope of the deepest (~5 km) Markov Deep can be
seen on Fig. 3, which were finding by marine acoustic complex (sidescan sonar) GBO MAK-
1M during the 22th cruise of R/V "Professor Logachev". The crust has a well-defined
subhorizontal layered structure, partially masked by sediments, and looks like structure of
the Kane OCC (23°30 'N) (Dick et al., 2008). Numerous steep-dipping normal faults are
clearly visible here; one of them (at the left), apparently filled with dolerite dike which,
probably, represents a lava flow’s feeder.
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 7

Clearly visible layering of the lithosphere, partially overlapped by sediments (gray). Numerous steeply dipping
normal faults are visible; one of them (on the left side of the figure) seems to be filled with dolerite dikes and,
probably, was a feeder of lava flow.

Figure 3. Structure of the eastern slope of the Markov Deep by data of remote sensing obtained
using marine acoustic complex (sidescan sonar) GBO MAK-1M from board of R/V "Professor Loga-
chev" (22th cruise, 2003).
8 Tectonics – Recent Advances

3. Dredged rocks from the Sierra Leone area


The spectrum of dredged rocks at the area is typical for slow-spreading ridges (Sharkov et
al., 2005, 2008;. Savelieva et al., 2006):

1. strongly serpentinized ultrabasites (depleted lherzolite and harzburgite, rare dunite);


most of them are mantle restites, but in some samples relics of cumulate structures pre-
served, attesting their intrusive origin;
2. two types of altered tectonized gabbros of lower oceanic crust: (i) primitive magnesian
gabbros (troctolite, olivine gabbro, and gabbro), which related to MORB and (ii) fer-
rogabbros (Fe-Ti-oxides-bearing gabbronorite, hornblende-bearing gabbro and gabbro-
diorite), related to specific siliceous Fe-Ti-oxide series (see Section 2.3);
3. small veins and nests of plagiogranites (trondhjemite);
4. dolerite dike and/or sill complexes, including ilmenite- and hornblende-bearing varie-
ties also;
5. basalts – fragments of pillow and massive lavas; very fresh varieties with chilled glassy
crusts predominate among them.

Most of the rocks were undergone to secondary alterations. Magmatic minerals (olivine,
pyroxene, and plagioclase) often display deformation textures resulted from early high-
temperature cataclasis, associated with plastic flow of solidified, but still hot rocks. Judging by
the Ti-zircon thermometry, it occurred within a temperature range from 815oC to 710oC (Zinger
et al, 2010). During pervasive low-temperature alterations, peridotites underwent strong
serpentinization, while gabbros and some basalts – amphibolization with appearance of fibrous
actinolite upon pyroxenes and thin veins of prehnite, carbonate and chlorite along the fractures.
In some cases rocks were schistozed and brecciated, and underwent by metasomatic processes;
the thickest metasomatic zone bear veinlet-disseminated sulfide mineralization (see Section 3).

3.1. Features of the fresh basalts


Most of studied fresh basalts with chilled glassy crusts often have porphyritic structure with
phenocrysts of three major types: Ol±Chr, Ol+Pl±Chr and Pl+Cpx, which is typical for
MORB (Langmuir et al., 1992). Equilibrium cumulates in transitional magma chambers have
to correspond with dunite, troctolite, olivine gabbro and gabbro, typical for many layered
mafic-ultramafic intrusions on continents (Sharkov, 2006).

Sometimes partly-melted xenocrysts were found in basalts and volcanic glass: olivine Fo88-89,
similar in composition to the olivine of mantle restite, and plagioclase An83-86 (Fig. 4), similar
in composition to the plagioclases of lower-crustal primitive gabbro. This is evidence that
the basaltic melts crossed rocks of the shallow lithosphere on their way to surface.

All studied fresh lavas are commonly oceanic plateau basalts (T-MORB) and more rare close
to E-MORB in composition. They are characterized by the same level of REE with typical for
MORB flat character of distribution; the Ce/Yb ratio ranging from 0.95 to 1.69. Judging on
#mg (56-63) and mineral compositions, they are not primary mantle-derived melts and
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 9

underwent by crystallizing differentiation in transitional (intrusive) magma chambers. It is


in a good agreement with small negative Eu-anomaly which reflects the fractionation of
plagioclase in intermediate magma chamber (Sharkov et al., 2005).

Diamonds denote basalts and their glasses (glasses are shown in the inset by open symbols), boxes denote gabbros,
and circles are trondhjemites. Dashed lines indicate the position of the data point of metasomatite replacing gabbro.
Hypothetical sources: (1) HIMU, (2) EM2. Continuous line outlines the field of the MAR basalts between 3° and 46°S
(Fontingie and Schilling, 1996), dashed line denotes MAR basalts between 30° and 50°N (Yu et al., 1997), dotted line
outlines basalts of Sao Migel Island, Azores (Widom et al., 1997).

Figure 4. Dissolution of plagioclase xenocrysts and textural–compositional heterogeneity of the chilled


glass Sample I1052/38. Image in back-scattered electrons.
10 Tectonics – Recent Advances

Fresh basalts of the Sierra Leone area in terms of Sr-Nd isotopic characteristics fall into
central part of the field of modern MORB for the southern hemisphere and occupy an
intermediate position between the most depleted basalts of the MAR (87Sr/86Sr <0.7025, εNd
>12) and enriched basalts of high latitudes both the northern and southern hemisphere
(87Sr/86Sr > 0.7030, εNd <8) (Fig. 5). Variations of the isotopic characteristics within a
relatively small (less than 300 km in the meridional direction) studied area is comparable in
scale with variations along the 15-20 times more extended segments of the MAR (Sharkov et
al., 2008). The points form an elongated box on the diagram which suggests the presence
here of two finite member (depleted and enriched) mixing in different proportions.

Figure 5. Sr–Nd isotope diagram for the studied basalts and their glasses dredged at the Sierra Leone
area, Mid-Atlantic ridge, 5°–7°N.

Significant nonsystematic differences in 87Sr/86Sr ratio and less significant differences in εNd
value between basalts and their chilled glassy crusts were firstly found in some samples
(Sharkov et al., 2008). Higher Sr isotopic ratios can be observed both in the glasses and the
basalts at the same lava fragments (Fig. 5, inset), at that isotope and geochemical
characteristics of the samples show no essential correlation. So, seawater did not affect to the
Sr and Nd isotope system in the chilled crusts of the studied pillow lavas. It is suggested
that such isotopic differences are related to a small-scale heterogeneity of the melts which
had no time to homogenized during their rapid ascent to the surface. The heterogeneity was
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 11

presumably related to the partial contamination of basaltic melts by older plutonic rocks
material (especially, lower-crustal gabbros) (Fig. 3).

3.2. Primitive gabbro


The magnesian primitive gabbros are represented by troctolite, olivine gabbro and gabbro,
which dominated among lower-crustal rocks of the area. The same gabbros are widespread
in rocks of the lower oceanic crust and traditionally viewed as intrusive equivalents of
MORB (Pearce, 2002; Ildefonce et al., 2007; Dick et al., 2008, etc.). Absence of reactionary
subsolidus pyroxene-spinel rims between Mg-olivine and Ca-plagioclase, typical for deep-
seated gabbros of continents evidence, that crystallization of parental melts occurred under
pressure, essential low 5 kbar (Sharkov, 2006).

According to our geochemical data, these gabbros have lower contents of REE, Th, Nb, Ta,
Zr and Hf as compared to the fresh basalts of the area (Sharkov et al., 2005). From this
follows that the fresh basalts and older primitive gabbros were formed from some different
mantle sources.

3.3. Ferrogabbro
Different ferrogabbros play essential role among lower-crustal rocks (about 1/3 of the
gabbros’ samples). They are represented by melanocratic troctolite, norite, gabbro-norite,
and gabbronorite-diorite, enriched in Fe-Ti oxides (ilmenite and magnetite) and often by
brown primary-magmatic hornblende (kaersutite) (Sharkov et al., 2005). Subvolcanic
analogues of ferrogabbro are represented by hornblende Fe-Ti-oxide dolerites, and very rare
basaltic flows with essential amount of Fe-Ti oxides, mainly ilmenite.

Ferrogabbros, like primitive gabbros, are characterized by low concentrations of light REE;
however, they are enriched in ore components – Zn, Sn and Mo, have elevated contents of
Cu and Pb, and low – Ni and Cr. In contrast to the primitive gabbro, ferrogabbros have
positive anomalies of Nb and Ta. Study of melt inclusions in chromites from rocks of this
series showed that their composition vary from Fe-Ti basalt to andesite (icelandite) and
dacite (Simonov et al., 2009). The ion-microprobe study of the melt inclusions yielded direct
evidence for elevated water content (up to 1.24–1.77 wt %) in the melts that produced
ferrogabbros; small globules of Fe-Ni sulfides were found in them also. So, these rocks from
one hand are saturated and supersaturated by SiO2 and have increased H2O content, which
typical for subduction-related magmas, and on the other hand have high contents of Ti, Fe,
Nb, Та and Р, typical for magmas of plume origin.

Ferrogabbros are obligatory component of the lower-crustal sections of rocks in OCCs,


where they play essential role. Many people thought that the ferrogabbros were produced
by fractional crystallization of the MORB-type melts (Dick et al., 1992; Thy, 2002 and
references therein). However, they often intruded primitive gabbros (Thy, 2002) and their
quantity usually exceed possibility of the MORB crystallizing differentiation.
12 Tectonics – Recent Advances

We believe that these rocks belong to magmatic siliceous Fe-Ti-oxide series, specific for
the oceanic environment, which origin related to melting of hydrated oceanic lithosphere
by action of a new mantle plume (Sharkov et al., 2005). Newly formed mantle-derived
melts passed through the upper cooled part of the plume head, accumulated at the
mantle-crust boundary and produced a magma chamber, which started to ascend
according to the zone refinement mechanism, i.e. by melting the roof and crystallizing at
the bottom (Fig. 6). The melt was continuously enriched in components not only melted
rocks of the chamber’s roof but also from the partly melted rocks at the heated peripheries
of the melting zone, where processes of anatexis occurred (see Section 2.4), as well as
fluid material from the heated rocks on the distant periphery. Obviously, unusual
characteristics of these melts, like their enrichment in SiO2, H2O, and some ore
components, typical for hydrothermal activity (Pb, Cu, Zn, etc.), can be explained by such
features of melting process.

Figure 6. Hypothetical scheme of the melts of the siliceous Fe–Ti-oxide series genesis
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 13

D. Pearce (2002) drawn attention to the fact that, unlike the fast-spreading ridges, in the
slow-spreading ridges volcanic equivalents ferrogabbros very rare. Truly, there are only two
small fragments of hornblende basalt with ilmenite, as well as a sample of variolite
(Krassivskaya et al., 2010) in our collection. Apparently, the main cause of the limited
vertical mobility of such melts is their water saturation, which decreases sharply at a
pressure about 1 kbar (Fluids..., 1991). This leads to separation and removal of water,
following increase of the solidus temperature and, as a result, to rapid solidification of melts
at depth; therefore, the volcanic eruption of such magmas are rare, how it is clearly seen in
the most OCCs and, particularly, in studied area.

3.4. Oceanic plagiogranites


As in all OCCs, small quantities of plagiogranites (tonalites and trondhjemite) are found on
the area. Their origin is usually attributed to later stages of magmatic crystallization.
However, according to our data, formation of such melts can be explained by anatexis of
hydrated lithospheric rocks near intrusive contacts (Aranovich et al, 2010). Special role in
this process belongs to "metamorphosed" sea-water from subfloor hydrosphere, enriched in
NaCl due to absorption of pure part of sea-water during formation of the secondary
hydrous minerals (serpentine, chlorite, actinolite, etc.) in the bedrocks.

4. Metasomatism and ore mineralization


Hydrothermal metasomatic zone with rich sulphide mineralization was found in the
Markov Deep (Sharkov et al., 2007 and references herein). According to results of
dredging, at least two zones of intense tectonic deformation and metasomatism (at depth
4400-4600 and 3700 m) occur here, extending in NW direction with gentle angles (30o- 40o)
dip to east (Fig. 7). These zones are formed by brecciated and schistosed ferrogabbros to
thin-foliated cataclasites upon them of chlorite-amphibole-epidote-clinozoisite
composition. The presence of chaotic plication, striation, grooves, slickensides and slip-
scratches on the surface of the clasts as well as fragments of small folds with distinct axis,
which are oriented along lineation, point to the fact that tectonic movements have evolved
under shear conditions.

Sulfide mineralization represented by quartz-sulfide and prehnite-sulfide veins, sulfide


dissemination and massive ore deposits. Mineral composition of ores is represented by
pyrite, chalcopyrite, sphalerite, pyrrhotite, bornite, and atacomite as well as native Cu, Pb,
Zn, Au and Sn and intermetallides (isoferroplatinum, tetraferroplatinum, and brass).

According to our data, hydrothermal-metasomatic processes occurred under low pressure


(0.5-1 kbar) and started at temperature ~750оС; however, the major ore-forming metasomatic
processes occurred in range 400-160oC. Sm-Nd isotopic data and δ34S value evidence that
ore-bearing fluid initially had magmatic origin and then were progressively diluted with
sea-water of oceanic subfloor hydrosphere.
14 Tectonics – Recent Advances

1 - zones of tectonic deformation and metasomatism, 2 - faults, 3 - geological boundaries; 4-7 – fields of preferential
distribution: 4 - basalts, 5 - gabbro, 6 - ultrabasic, 7 - sedimentary cover; 8 - zone of hydrothermally altered rocks; 9 -
sulfide mineralization in the bedrock.

Figure 7. Scheme of the geological structure of the eastern edge of the Markov depression. According to
Beltenev et al. (2004).
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 15

The main source of ore-bearing fluids could be an intrusion (Fig. 8), formed by water-
saturated melt of siliceous Fe-Ti-oxide series, which contains sulfur and ore components (see
Section 2.3). Addition of the ore material can be leached from the host gabbros on the way
up, and part of the sulfur been introduced by sea water. Separation of fluids from such
magmas commonly occurs at a pressure of about 1 kbar, when the solubility of water in
them decreases sharply (see Section 2.3), i.e. at the depth of 3-4 km below the seabed, where,
apparently, was located solidifying intrusion.

1 – vents of hydrothermal systems at the ocean bottom o ("black smokers"), 2 – solidifying intrusion of siliceous Fe-Ti-
oxide series with lenses of the residual melt; 3 – place of formation of plagiogranite melts; 4 – chilled zone of the
intrusion, 5 - fractured rocks of hydrated oceanic lithosphere.

Figure 8. Scheme of the ore-bearing hydrothermal system structure

The appearance of relatively high-temperature metasomatites on the oceanic floor indicates


that these rocks were moved to the surface after their formation at the depth. Judging by
presence of atakamite and weak oxidation of sulfides, it happened very recently, apparently
in the process of ongoing formation of the present Markov Deep. We suggest that this ore-
bearing zone is fragment of feeding system of an extinct black smoker.

5. Results of U-Pb dating of zircon from gabbros by SHRIMP-II


The U-Pb SHRIMP-II dating of zircons from the gabbros of the area showed that these rocks,
which are in the present-day oceanic spreading axis of the MAR, were formed earlier, in the
Holocene-Pleistocene, 0.7-2.3 Ma; above, presence of zircon grains with age up to Mesoarchean
(older than ~ 87 Ma up to 3117 Ma) were established (Bortnikov et al., 2008) (Fig. 9). The
magmatic nature of young zircon with thin oscillatory zoning and sectorial structure suggests
that its age defines the crystallization age of the host magmatic rocks; “old” zircon are defined
as xenocrysts. Later the presence of uneven-aged zircon grains were found not only in the
gabbros of the Sierra-Leone area, but in gabbros as well as ultramafics, plagiogranites, diorites,
and even basalts of other parts of the MAR, i.e., presence of ancient xenogenic zircon grains in
oceanic bedrocks is widespread (Skolotnev et al, 2010 and references herein)
16 Tectonics – Recent Advances
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 17

A. Microimages of zircons from gabbronorites of the Markov Deep (dredge sites I1028, L1153, and L1097). Hereinafter: (a)
natural appearance, (b) cathodoluminescence image (CL); index of dredge site: (I) R/V Akademik Ioffe and (L) R/V Professor
Logatchev. Spot numbers of U–Pb age determination are as in Tables 1 and 2. Age is shown for xenogenic zircons. (1)
Euhedral zircon with weakly corroded prism surface and diopside inclusion (dark); sectorial and oscillatory zonings are
seen; (2) bipyramidal zircon with corroded prism face; oriented light fragments of bands (reflection of high-temperature
cataclasis) are seen in the CL images; (3) zircon with corroded prism and pyramid faces; the CL images demonstrate
sectorial zoning, resorption of prism and pyramid faces, and formation of colloform shell; (4) prismatic zircon with weakly
corroded surface; the Cl image shows coarse zoning, weak resorption of prism and pyramid faces, and thin shell; (5)
euhedral grain with well-expressed oscillatory zoning; (6) zircon with corroded surface; the CL image shows core with
fragments of light bands (reflection of high-temperature cataclasis) and colloform shell; (7) subhedral zircon with corroded
prism surface; the CL image demonstrates fragments of sectorial zoning and small inclusions of plagioclase (dark) of
irregular shape (poikilitic structure); (8) growth of secondary small pyramidal zircons due to redeposition of matter on the
other side of the crystal; oriented light bands produced by high temperature cataclasis are seen clearly in the CL; (9)
fragment of long-prismatic zircon with corroded pyramidal termination; deformation-related light bands are observed in
the CL image; (10) rounded zircon with coarse zoning in the core and thin shell; (11) analogue of zircon 10, but with a
wider shell; (12) fragment of prismatic zircon lacking internal structure in the CL.
B. Microimages of zircons from troctolite, Site I1069-19. (1) Prismatic grain with corroded surface; fragments of coarse
zoning and shell are seen in the CL; (2) analogue of 1, with wider shell; (3) subhedral zircon with coarse concentric
zoning, elements of sectorial zoning, and fragments of thin shell; (4) subhedral grain with weak corrosion of one
pyramid; no internal structure was identified in the CL.

Figure 9. Zircons from gabbros of the Markov Deep. After N.S. Bortnikov et al. (2008).

We believe that such xenogenic zircon could initially belong to fragments of material from
the "slab graveyards" in the deep mantle, captured by mantle plume, which moved from the
core-mantle boundary (Bortnikov et al, 2008). Such "graveyards" may contain rocks of
different ages and backgrounds, including the Precambrian gneisses and sedimentary rocks
involved in subduction zones. A detailed study of exhumed slabs presented by ultrahigh-
pressure complexes of Kazakhstan, China, Norway and others, which were formed at P >
2.8-4 GPa (and possibly up to 8.5 GPa) and T = 600-900°C, showed that zircon could persist
even under these conditions (Ernst, 2001 and references herein). Apparently, these PT-
18 Tectonics – Recent Advances

conditions preserved in rocks "graveyard slabs" also, because, according to seismic


tomography (Karason, van der Hilst, 2000), they form a great bodies of hundreds kilometers
thick; billions years are required to warm up them by conductive heat.

Possibility of existence of buried subducted material beneath the Central Atlantic also
evidence from results of study of lithium isotopy in basaltic glasses at 12-16oN (Casey et al.,
2010). Like in Sierra Leone area, these basalts in composition are intermediate between E-
and T-MORB and are characterized by positive Ta- and Nb-, as well as Ti-, Sr- and Eu-
anomalies. They set the lowest of recorded value of δ7Li, indicating presence in magma
components of refractory rutile-bearing eclogite.

According to geophysical data, asthenosphere beneath the MAR is represented by lens-like


body about 200-300 km thick (Fig, 10) (Anderson et al., 1992; Ritsema, Allen, 2003). So, our
finding of ancient zircons in gabbros and Li-isotopic data support the idea about existence of
сolder mantle beneath the axial part of the MAR, which has penetrated by mantle plumes.

Figure 10. Tomographic profile along the axis of the Mid-Atlantic Ridge, showing that the highest
speed anomalies of transverse waves are localized under the "hot spots" (triangles): 2 - Tristan, 6 - As-
cension, 14 - Azores and 17 – the Iceland;; the latter can be traced to the lower mantle and possible to
the mantle-liquid core boundary (after Ritsema & Allen, 2003).
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 19

6. Processes of formation of shallow lithosphere (oceanic crust and


lithospheric mantle) in the Sierra Leone area
According to the classical model, the occurrence of mid-ocean ridges associated with localized
upwelling of deep hot mantle material, which melts due to adiabatic decompression
producing specific MORB volcanism. However, it was found that asthenosphere beneath the
MAR has lens-like shape up to 200-300 km thick, which is located between colder and dense
material of shallow lithosphere and underlied mantle (see section 5).).

As it was shown above, oceanic crust of the Sierra Leone area was formed at least three
independent episodes of magmatic activity: the modern, attributed with the eruptions of
fresh pillow-lavas, and two previous ones, which led to formation of lower-crustal gabbros
(altered primitive gabbros and ferrogabbros consequently). Accordingly, the fresh basalts
are not genetically related to the altered lower crust and question arises about it’s origin.

6.1. Origin of lower oceanic crust: Evidence from Sierra Leone area
There is a consensus that it exists between the mantle and the upper oceanic crust and
composed of various gabbros, often alternated with peridotites. It implies existence between
mantle and upper oceanic crust transitional magmatic chambers (intrusions), which
solidification provide formation of the lower crust. At the same time mechanism of this crust
formation is open to question because its geological study in present-day oceans is concerned
with serious technical difficulties. About its composition and structure we can judge only by
random samples, gave by dredges, inhabited submarine devices, or "pinpricks" of deep-water
drilling wells. According to Pearce (2002), two main points of view on the origin of the oceanic
lower crust dominate now: (1) its formation during crystallization in a single melt lens
followed by the flow of crystal mush down and away from the ridge (model “gabbro glacier",
Quick, Denlinger, 1993) and (2) through grows from series of sill-like bodies throughout the
crust (the “Christmas tree” model). However, the situation is stayed uncertain.

From this point of view essential help for understanding processes of formation and
development of lower crust of the modern oceans can provide gabbro complexes of
ophiolites – fragments of ancient oceanic or back-arc seas lithosphere, find in orogens
(Knipper et al, 2001; Dilek, Furnes, 2011 and references therein). In contrast to modern
oceanic floor, about which structure we can judge only by fragmental data, they are
available for direct geological studies.

Of particular interest in this regard is well-preserved Voikar (Voikar-Syninsky) ophiolite


assemblage in the Polar Urals (Russia). Its gabbro complex consists on two major
megarhythms (Fig. 11), generally similar in structure to large layered mafic-ultramafic
intrusions, formed in calm tectonic settings, were found there above the mantle restite
complex (Sharkov et al, 2001). At that for the lower megarhythm are typical primitive
gabbro and olivine gabbro, and for the upper – mainly gabbro-norite, sometimes
hornblende-bearing, often with increased concentrations of ilmenite and titanomagnetite,
which resemble the rocks of the siliceous Fe-Ti-oxide series. All rocks of the assemblage are
20 Tectonics – Recent Advances

cut by diabase dikes. Thus, as is the case of the Sierra Leone area, there are two independent
sets of intrusive rocks recorded here, successively formed by different magmas.
In contrast to majority of continental large layered mafic-ultramafic intrusions, almost all
plutonic rocks of the Voikar’s ophiolites were undergone shearing. Like the site of Sierra
Leone, it started with ductile flow of rocks at high temperatures and changed by the brittle-
plastic and brittle deformation under conditions of the greenschist facies during cooling.
This led to a strong serpentinization of ultramafic rocks and amphibolization of gabbros
with extensive development of fibrous amphibole upon pyroxenes. Cumulative structures
remain rare, although the overall shape of rocks indicates their intrusive origin.
The absence of cryptic layering in cross-section of the megarhythms suggests that their
formation occurred by crystallization of the transitional magma chambers accompanied by
replenishment of fresh melts under conditions of open magmatic system (Sharkov et al.,
2001). At such circumstances, solidifying from the bottom up magma chamber could be a
lens, gradually moving up and leaving a "tail" of hardened hot material. In other words,
although these intrusions were not necessarily initially large, but could gradually grew with
the arrival of new portions of melt.
According to Sm-Nd and Re-Os isotopy data, significant differences between material of the
lower and upper megarhythms as well as mantle section occur in the Voikar ophiolite
assemblage. Thus, presence of ancient material determined in the rocks of the upper
megarhythm, where the 187Os/188Os ratio is 6.5-7.1, which is much higher than in the mantle
rocks of the assemblage and two times higher then in diabases of the sheeted complex dikes
(Sharma et al., 1995, 1998). These data indicate that: (1) formation of the gabbro complex was
happened at two stages, i.e., a two-stage build-up of the lower crust occurred here; (2)
judging from the relatively well-preserved sections of the complex, formation of each of
them happened during the relative calm of tectonic processes; (3) still hot rocks, soon after
their solidification, were involved in processes of plastic flow, gradually changed by plastic-
brittle and brittle deformations; and (4) there are marked differences in isotopic
characteristics between major constituents of the assemblage: mantle rocks, dikes sheeted
complex, as well as two megarhythms of the gabbro complex.
The data available on the lower crust of the Sierra Leone area and others OCCs (see above) are
in good agreement with data on the Voikar gabbro complex. The presence in the area’s lower-
crustal gabbros relic cumulative structures and elements of the primary magmatic layering
(Fig. 3) suggests that this crust is formed by large layered mafic-ultramafic intrusive bodies of
different age and origin. Very likely, that its formation happened mainly through
underplating, i.e. building up from below, through accumulation of newly formed basaltic
melts at crust-mantle boundary how it established on the continents (Rudnick, 2000).
Presence in the lower crust of the area both primitive gabbro, derived from MORB-type
melts, and ferrogabbros of siliceous Fe-Ti-oxide series, shows that, like in Voikar, at least
two different types of layered intrusions occurred here.
Appearance not numerous relatively fresh gabbros, olivine gabbros and troctolites among
dominated altered gabbro can be considered them to the recent cycle of activity. In any case,
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 21

judging on the phenocrysts in the fresh basalts (see Section 2.2), those rocks were probably
formed in transitional chambers of young magmatic systems.
Strongly serpentinized mantle peridotites, like in the most ophiolites including Voikar’s,
represented here by typical mantle restites – harzburgites and subordinate depleted
lherzolite and dunites (Savelieva et al., 2006). Some of these peridotites, judging on rare
good preserved samples, have cataclastic structure (Fig. 11) evidenced about their involving
in deformation processes.

Figure 11. Geological section of gabbro (layered) complex Voykar ophiolite assemblage (Polar Urals),
by (after Sharkov et al, 2001).
22 Tectonics – Recent Advances

Thus, according to data available, the most OCCs (Escartín et al., 2008; Smith et al., 2008;
MacLeod et al., 2009: Silantyev et al., 2011 and references therein), as well as ophiolite
assemblages (Knipper et al., 2002; Dilek and Furnes, 2011), have the similar structure and
composition of lower crust and lithospheric mantle. So, the structure of the studied area
represents common type of the shallow oceanic lithosphere and we can discuss some
general problems of origin and functioning of slow-spreading ridges on its example.

7. Discussion
7.1. Origin of oceanic core complexes: Evidence from Sierra Leone area
As it follows from study of typical OCCs (see Introduction), they are parts of slow-
spreading ridges with asymmetric structure with one-way sliding of crustal material; it is
suggested that their origin is attributed to activity of hypothetical oceanic detachment faults
(Fig. 1). It assumes that these faults develop as a result of strain focusing around
rheologically strong gabbro plutons hosted in weaker serpentinized lithospheric mantle;
hence it deduced that OCCs were formed during periods of relatively enhanced magma
supply. However, as mentioned above, even sticklers of this hypothesis of "one-sided
spreading" recognize that many of its elements and details are still poorly substantiated
(Conference ..., 2010). In fact, it is determined only asymmetry of the structure of these parts
of mid-ocean ridges with exposed altered deep-seated rocks and presence there gently
sloping and normal faults.

There are two main hypotheses of the OCCs origin exist now. Predominant model of their
appearance is considered with activity of oceanic detachment faults during periods of
reduced magmatic activity or its absence ("dry spreading») (Tucholke, Lin, 1994; Tucholke et
al., 2008; Dick et al., 2008; Escartin et al, 2008, etc.). According to another view, based on
widespread gabbro in such structures, the OCCs were formed at period of relatively
depressed (but not reduced) magmatism, realized as large plutons from overlapping access
of magma to the surface by oceanic detachment faults (Ildefonse et al., 2007). Sort of these
conceptions is model of “life cycle of OCC” (MacLeod et al., 2009). According to this model,
spreading becomes markedly asymmetric when the core complex is active, and volcanism is
suppressed or absent; when the asymmetry is such that the detachments accommodate
more than half the total plate separation, the active faults migrate across the axial valley. As
a consequence magma is emplaced into and captured by the footwall of the detachment
fault rather than being injected into the hanging wall, explaining the frequent presence of
gabbro bodies and other melt relicts at oceanic OCC. Core complexes are ultimately
terminated when sufficient magma is emplaced to overwhelm the detachment fault.

However, a numbers of problems remain unsolved in context of these models: motives of


ascending of older altered lithospheric rocks at high hypsometric levels, lack of genetic
interrelations between fresh basalts and older lithospheric rocks, presence essential quantity of
rocks of the siliceous Fe-Ti-oxide series (ferrogabbros) among them, etc. Above all, OCCs, in
essence, represent outcrops of shallow oceanic lithosphere, which formation has not
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 23

considered with hypothetical oceanic detachment faults: the latter can only expose them, but
not create, especially because this lithosphere was formed much earlier and at greater depths.
So, the main problems are origin of this lithosphere, reason for its local ascending in axial parts
of slow- and ultraslow ridges, and how these ridges functionate under condition of lens-like
asthenosphere beneath them, i.e. what the reasons for the oceanic spreading there?

7.1.1. What has occurred in slow-spreading ridges in geomechanics terms?


Since H. Hess (1962) times, the most researches believe that process of oceanic spreading
associated with ascending of hot mantle material to the axial part of mid-oceanic ridges, its
adiabatic melting accompanied by formation of oceanic crust, growth of the plate, and their
motion to both sides of ridges under influence of deep convection. The most complete
geomechanical aspects of the model was considered by D. Turcotte and J. Schubert (2002
and references herein).

It is known that shallow lithosphere in the MAR axis is in a position of mechanical


instability, how it evidence from presence of constant shallow earthquakes, caused by
processes of stretching, discontinuity and delamination of bedrocks which indicate uplift
and spreading of its axial part. According to (Turcotte, Schubert, 2002), upwelling of mantle
rocks is accompanied by heat loss, occurred through molecular heat conduction; as a result,
they attached to the base of separating plates, becoming part of them. Because material of
heated plastic asthenospheric material can flow like a liquid in geological time scale under
influence of external forces, increase of load promote flow of this material to the ridge axis,
ensuring its stable triangular shape over time.

In accordance with Turcotte-Schubert model, the triangular shape of the ridge should lead
to gravitational instability of the system, causing sliding (slumping) of material along its
slopes. Mathematical simulation of such process, performed on example of the MAR,
revealed that the force of the ridge push are sufficient to implement such a mechanism
(Scheidegger, 1987). As a result, the crustal material should slide from uprising dome-
shaped part of the ridge axis (tectonic erosion), exhuming of deep-seated rocks on the
oceanic floor (Fig. 12).

Such structure of mid-oceanic ridges in terms of geomechanics is typical for piercment


structures which formation determined by introduction of plastic less dense and less viscous
layer into overlying more dense layer under gravity influence (Scheidegger, 1987 and references
herein). According to the theory, penetrating masses, being less dense than overlying rocks will
tend moving upwards, regardless of else tectonic forces. Though classical theory of piercing
structure formation developed on example of salt domes, we have a close situation in the MAR:
heated plastic asthenospheric material and overlied it cold dense shallow lithosphere. In this
case, due to tectonic erosion, pressure above growing asthenospheric crest falls and as a result
of adiabatic decompression (decrease in the solidus temperature with decreasing pressure) it
led to melting of the material (Fig. 13). According to calculations (Girnis, 2003), smelting of
MORB begins at pressure ~15 kbar, however, the mass-melting occurred at pressures 8-10 kbar,
i.e. at the depths 28-35 km, where major melting zone has to situated.
24 Tectonics – Recent Advances

Figure 12. Micrograph cataclased harzburgite restite; large grains (porfiroclast) deformed orthopyrox-
ene surrounded by small neoblasts of olivine and pyroxene. Sample 1063-39, polarized light (collection
of E.V. Sharkov).

.
1 – sediments; 2 - mantle plumes penetrating into the asthenospheric mantle, and partly or completely mixed with it; 3
- asthenospheric lens under the MAR, bordered by cooling zone in contact with lithospheric mantle; 4 - melting zone in
the upper part of the asthenospheric lens; 5 - transitional magma chamber; 6 - depleted lithospheric mantle (restite
from a previous episode of melting), transformed into the lithospheric mantle; 7 - oceanic crust formed by gabbros and
basalts; 8 - oceanic lithospheric mantle; 9 - direction of movement of material.

Figure 13. The proposed scheme of the deep structure of the MAR
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 25

Appearance of the melting zone brings further contribution in ascending of the ridge’s axial
part because melting of silicate rocks leads to decrease of density of material in the magma
generation zone by 11-13% (Handbook ..., 1969). This led to development of fractures in the
overlying lithospheric peridotites, promoted to their serpentinization under influence of
subfloor sea-water which reduced their density by 35-38%. All of this stimulate further
growth of the dome and strengthen the tectonic erosion of its axis.

In essence, geological sense of oceanic spreading as well as formation of new lithosphere


plates lies in this complex of processes. From such standpoint, constructive plate
boundaries, at least in the slow-spreading ridges, should be formed by a collage of tectonic
slices of shallow oceanic lithosphere and basaltic covers, i.e. these lithospheric plates in
geomechanical sense are not monolithic as suggested by Hess (1962). It is in a good
agreement with results of study of ophiolite assemblages which are formed by packages of
tectonic slices of similar crustal rocks and mantle restites.

Thus, in contrast to the generally accepted views, we do not attribute oceanic spreading
with hypothetical convection currents in asthenosphere, but with processes of gravitational
instability at the lithosphere-asthenosphere boundary in axes of the slow-spreading ridges,
resulting in sliping of rocks on their slopes. Gabbros and restite ultrabasites as well as
basalts, genetically related to this tectonomagmatic episode, are involved in this process.
Sliding of the rocks is accompanied by their delamination, formation of different faults,
tectonic slices, etc., that creates a characteristic "seismic noise".

7.1.2. Processes of the OCCs formation: Evidence from the Sierra Leone area
How it evidence from study of the Sierra Leone area, located in spreading zone of the MAR,
formation of its structure occurred at least in two stages. The first stage attributed to
formation of shallow oceanic lithosphere (lower crust and restite mantle peridotites) and the
second, modern – unconformably overlying them flows of fresh basalts.

Between these stages there was occurred rise of the lithosphere dome, accompanied by the
sliding of material (tectonic erosion), exhumed deep-seated rocks to the oceanic floor and
the appearance of numerous extensional structures, ensures the existence of pallets
(subfloor) hydrosphere and the ways for hydrothermal fluids ascent; remains of a former
hydrothermal systems were found in the Markov Deep (see section 3). This stage of the area
development can be defined as a formation of oceanic core complex (OCC).

The second (current) stage of the ridge development on the studied area is also
characterized by extensive development of extensional structures up to appearance of the
rift graben-like structures and a fairly powerful basaltic volcanism. These melts come from
intermediate magma chambers, where they were subjected to fractional crystallization, and,
before reaching the oceanic floor, passed through the ancient lithospheric rocks, partially
assimilating its material.

Inasmuch as situation at the lithosphere-asthenosphere boundary in the slow-spreading


MAR before the OCCs formation was in a state of unstable equilibrium (see Section 6.1.1),
26 Tectonics – Recent Advances

such development of events demanded of a trigger to start ascent of the asthenospheric crest
in studied area. Most likely this role played a mantle plume, which reached the boundary
and lifted it, thereby disturbed the unstable equilibrium and initiated rise of the dome (Fig.
14). Its recent existence here follows from isotopic data (Schilling et al., 1994), as well as the
general uplifting of the territory and composition of fresh pillow lavas (mainly T-MORB
(oceanic plateau basalts) to E-MORB); such characteristic of basalts typical for sites on the
ridges next to manifestations of intraplate (plume) magmatism (Basaltic ..., 1981).

1 – mantle plume; 2 – its cooling borders; 3 – asthenosphere; 4 – shallow lithosphere; 5 – basaltic melt

Figure 14. Scheme of cyclic evolution of tectonomagmatic processes in axial part of the MAR

Most likely, the typical for OСCs hydrothermal fields are also associated with magmatic
systems, generated by mantle plumes. This is particularly true for water-saturated melts of
siliceous Fe-Ti-oxide series, which have limited mobility in the vertical, and usually do not
reach the surface in the thickness of the crust hardens in the form of intrusions (see Section
2.3, Fig. 8). It is in a good agreement with our data on the Sierra Leone area, where we found
recently extincted ore-bearing hydrothermal-metasomatic system, attributed to such magma
(see Section 3). Very likely, that typical for OOCs phenomenon of wide development of
hydrothermal fields under conditions of “dry spreading” (i.e. by practically absence of
volcanic eruptions) can be successfully explained by this circumstance.

From this view, appearance of OCC reflects the first stage of the crest uplifting, followed
"squeezing-out" by plume head of cold rigid lithosphere as a dome at relatively high
hypsometric levels and starting process of tectonic erosion on its axial part, which leads to
the appearance (exhumation) of deep-seated rocks on the ocean floor. The head of the plume
was, in general, asymmetrical and often provided outside of the ridge axial zone that led to
the emergence of on-side spreading. Fragments of the plume heads, as it shown on example
of the St. Peter and St. Paul complex (see Introduction), can be found sometimes. Perhaps,
they are encountered more often, but it is difficult diagnostics because of strong
serpentinization.

Widely represented in various OCCs surfaces with corrugations and striations (mullion
structures) are usually interpreted as evidence for the existence of oceanic detachment
faults, namely its footwall (see Introduction). However, mullion structures are a common
pattern under joint flowage of very different on viscosity tectonic plates (Allaby & Allaby,
1999), in this case – the serpentinite and gabbro, and does not carry specific information
about existence of oceanic detachment fault here.
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 27

From all this it follows that formation of the OCCs is likely represent the first stages of dome
growth due to appearance a new mantle plume that disturbs the unstable equilibrium at the
lithosphere-asthenosphere boundary. This led to uplifting of the area, to beginning of
tectonic erosion and provided specific magmatism of siliceous Fe-Ti-oxide series and related
hydrothermal activity.

7.1.3. Cyclic character of tectonomagmatic processes in the axial zone of the slow-spreading
ridges
One of important points for understanding situation in the ridges axes are processes of
adiabatic melting. As a result of the OCC appearance and beginning of sliding of
material from the ridge axis, asthenosheric material began to inflow there and the ridge,
because of mechanical reasons, gradually got a symmetrical structure. Due to
concomitant reduction of pressure in the frontal part of the crest, process of melting had
to strengthened. Since position of the solidus isotherm in this case is determined by
lithostatic pressure, a melting zone should has form of a flattened lens at the top of the
asthenospheric crest. It led to mixing of asthenosphere’s and plume’s materials and
appearance melts of T- and E-MORB composition. Judging on the Sierra Leone area, this
change of the melting regime accompanied by temporal interruption of tectonomagmatic
processes, after that flows of newly formed basalts began to overlap altered rocks of the
former OCC. The next stage of the dome ascending should be already vast eruptions
under conditions of a "normal" bilateral spreading, how it occurred to the south and
north of the studied area.

As a result of melting, density of asthenospheric lherzolite would gradually decrease,


mainly due to removing of Ca and Fe with basaltic melt. Simple calculations show that
decrease in the density of material in this case could reach 8-10%, because strongly
depleted mantle material (mainly harzburgite) consists mostly of relatively light
magnesian olivine and orthopyroxene. Because of this, restite material will accumulate in
upper part of the melting zone, forming a separate layer, which cannot be involve in
processes of the asthenospheric convection. Formation of such layer of light refractory
material should lead to the cessation of melting of the mantle beneath the ridge axis, and,
as a result, it become part of shallow lithosphere and situation returned to state of
unstable equilibrium.

Thus, there are three main stages of the cyclic development of spreading zones: (i) initial
- OCC (often one-sided spreading) → (ii) intermediate, such as Sierra Leone (the
transition to a bilateral spreading) → (iii) normal (bilateral spreading). Each of these
three types of spreading are observed in different segments of the present-day MAR,
suggesting that these sites are various stages of development. In general, once started,
the processes in the axial zone of the ridge are mutually self-sustaining conditioned,
resulting in almost continuous growth of the oceanic lithosphere in the slow-spreading
ridges axes.
28 Tectonics – Recent Advances

8. Processes of slow-spreading ridges formation and development:


Evidence from the MAR
8.1. Interaction of asthenosphere and mantle plumes
In contrast to MORB, derived from moderate-depleted mantle material, magmatism, related to
mantle plumes, is presented by geochemically-enriched Fe-Ti picrites and basalts, evidence
about rather different melting source. According to Anderson et al. (1992), the most centers of
intraplate (plume) magmatism in the Atlantic are localized within the MAR. From this it
follows that slow-spreading mid-ocean ridges are an areas of joint manifestations of
asthenospheric and plume activity, and relationships between them is a key for understanding
the functioning and development of slow-spreading mid-oceanic ridges.
Continuous smelting of basalts from the asthenospheric material had to inevitably affect to
its composition in terms of increasing degree of depletion. However, this has not happened,
and composition of the melting substrate as a whole remains practically the same during at
least the latest 140 Myr in case of the Central Atlantic. Because asthenosphere beneath the
MAR is a lens-shape body of 200-300 km thick (see Section 4), it requires a constant feed of it
by geochemically-enriched material. Evidently, it can be the material of mantle plumes,
constantly rejuvenating the composition of the asthenosphere after removing from it the
low-melting components (basalts).
How in particular an interaction of the asthenosphere and mantle plumes could be occurred?
It is known that under conditions of rigid continental lithosphere plume-related magmatic
systems form isolated localities. However, in the case under consideration the situation is quite
different: both asthenospheric and plume material have close visco-plastic consistency.
Accordingly, only the largest and most stable in time plumes like Iceland, Azores, Tristan, etc.,
can cross such thick lens. They lose much of their material, which mixes with the
asthenosphere matter, leading to the appearance of T-MORB (oceanic plateau basalts) and E-
MORB in the adjacent parts of the spreading zone (Basaltic ..., 1981). Only such plumes can
pass through the asthenosphere lens and products of their melting reach the surface, forming
oceanic islands and seamounts. The existence of less powerful plumes may indicate the
appearance of mantle-crustal magmas of the siliceous Fe-Ti-oxide series (see Section 2.3); still
weaker plumes "damped" in the thick asthenospheric lens. In this connection, attention is
drawn to that this lens itself is not a single uniform body, and is subdivided into several
segments (Fig. 10). It is also confirmed by the results of geochemical studies of basalts
throughout the MAR length (Fig. 15) (Dmitriev et al, 1999; Silantyev, Sokolov, 2010).
Mantle plumes, penetrating the asthenospheric lens, should contribute to forced convective
mixing of its material and lead to practical levelling of its composition. Obviously, for this
reason, geochemical and isotopic-geochemical characteristics of MORB, both in the Sierra
Leone area and all over the MAR, are close to each other. From this it follows that material
of the asthenosphere is a mixture of moderately depleted lherzolites and geochemically-
enriched material of mantle plumes as finite compositions. Asthenospheric plastic material,
in contrast to the melt, is mixed substantially worse, which are evidence from variations of
isotopic and geochemical characteristics of the fresh basalts (see Section 2.1).
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 29

1 – TOR-1; 2 – TOR-1 + TOR-Fe; 3 – TOR-2; 4 – TOR-1 + TOR-2; 5 - TOR-K. I, II and III - segments of the 1st order: I -
South Atlantic, II - Southern Region of the North Atlantic, III - Northern Region of the North Atlantic (TOR – varieties
of tholeiites of ocean ridges), 1 - 4 – segments of the 2nd order: 1 – Equatorial ; 2 – Central; 3 – Azores; 4 – Icelandic.
Fractures zones: JM – Jan Mayen, CG – Charlie Gibbs, P – Pico, D – Oceanographer, K – Kane, CV - Cape Verde, CH –
Cheyne, 26o - 26o S, BV – Bouvet. TOR –tholeiite of oceanic ridges..

Figure 15. Scheme of tectonomagmatic segmentation of the Mid-Atlantic Ridge. After (Dmitriev et al, 1999).

As it was mentioned above, according to geophysical data, a colder mantle occurs beneath
asthenospheric lens (Fig. 10). It is supported by finds of ancient xenogenic zircon grains,
which, very likely, were trapped by rising plumes from the " slab graveyards" in the mantle
beneath the ridge (see Section 4). Apparently, the impurities, trapped by plumes from deep
mantle, as well as material of the shallow lithosphere, trapped by basalts on their way to the
surface (see Section 2.1), play essential role in scatter of points on the Sr-Nd diagrams and
the appearance of various "mantle reservoirs," in particular, HIMU.
30 Tectonics – Recent Advances

Thus, mantle plumes bring to the geochemically-depleted oceanic asthenosphere


considerable amount of fresh hot geochemically-enriched material, providing forced
convection of the material. This leads to more or less effective mixing of the two types of
finite materials, as well as to general leveling of the asthenosphere composition and
temperature, thereby support their sustainable dynamic equilibrium in considerable time, at
least ~140 Myr in case of the Central Atlantic.

8.2. How formation and development of the oceanic asthenospheric lens occurred?
Obviously, constant addition of a new (plume) material had to cause to increasing of the
asthenospheric lens size, leading to its extention in both directions: from the ridge axis
(spreading itself) and along it (propagation of ridge). As a consequence of the lens
extending, it becomes a "trap" for the plumes, rising in the neighborhood, which became
parts of the asthenosphere’s supply system and contribute to further widening of the ocean
floor in width and length.

However, it remains unclear how the asthenospheric lens, which promoted oceanic
spreading, was initially formed. Perhaps, its occurrence was attributed to an elongated area
of concentrated manifestation of mantle plumes activity. Apparently, in this case the
extended heads of neighboring large plumes came in contact with each other and merged
(coalescence) into a single body. This body will grow due to involvement of plumes in the
neighborhood, gradually increasing in size and can gradually developed into a zone of
oceanic spreading. Possible examples of initial stages of the process are ultraslow-spreading
ridges (Knipovich Gakkel, Monze, Lena Trough), which develope in the North Atlantic and
the Arctic Ocean, where the MAR propagates (Snow, Edmonds, 2007).

Modern example of an elongated area of mantle plumes activity may be the Trans-Eurasian
Belt of tectonomagmatic activation, which stretches out along the whole of Eurasia from the
Atlantic to the Pacific and appeared after closure of the Mesozoic Tethys Ocean (Sharkov,
2011). If the plumes are distributed uniformly, a large igneous province, like the Permian-
Triassic Siberian Traps, formed instead of an ocean.

Thus, data available on the Sierra Leone area and other OCCs allow to complement the existing
models of the structure and development of slow-spreading ridges and liberalize present-day
views on processes, occurred in their axes. Besides, they provide opportunity to discuss
problem of structure of the mantle beneath the slow-spreading ridges. As shown above, the
shallow oceanic lithosphere is composed mainly by plutonic rocks and high depleted mantle
and qualitatively different from the underlying asthenospheric lens, formed by moderately
depleted material. Located beneath the lens deep mantle also differs significantly from the
asthenospheric material. From this it follows that all three components of the mantle under the
slow-spreading MAR have an independent origin, and whole-mantle convection is absent here.

The latest data on the geology, petrology, geochemistry, isotopy and geophysics of the
oceanic bedrocks takes into account in the proposed model. These data were not known a
half century ago, when the basic conceptions of plate tectonics were elaborated. Gradually it
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 31

has revealed details that allow to take a new look at the nature and mechanisms of oceanic
spreading. In general, our data suggest that real tectonomagmatic processes in the axis of
slow-spreading MAR are essential different from existing views on processes and
mechanisms of oceanic spreading. Obviously, there is still much uncertainty, but it is also
clear that new approaches to the study of geology and petrology of ocean are necessary.

9. Conclusions
1. Sierra Leone area, located in the axial part of the MAR (5o-7oN), is characterized by
outcrops of extensive deformed and metamorphosed plutonic rocks of the shallow oce-
anic lithosphere. The area are characterized by wide development of extensional struc-
tures, including rift valleys and variably oriented faults. These features of the structure
and composition of rocks can define the area as a kind of oceanic core complex (OCC).
However, unlike typical OCCs, outcrops of altered gabbros and serpentinites occur on-
ly on valleys’ slopes and floors, while surface of the ridge is overlapped by flows of
fresh pillow lavas with chilled glassy crusts, i.e., a kind of "structural unconformity" oc-
curs here.
2. Fresh basalts are close in composition to E- and T-MORB (oceanic plateau basalts); judg-
ing by #mg and composition of phenocrysts, they are not primary mantle-derived
melts, and underwent by crystallization differentiation in the intermediate (intrusive)
chambers. On the way to the surface, they crossed the mantle peridotite and lower-
crustal gabbros, and were partly contaminated by their material.
3. The lower crust in the area is composed by gabbros of two types: (1) primitive, magne-
sian, derived from MORB, and (2) often hornblende ferrogabbros derived from melts of
siliceous Fe-Ti-oxide series. These melts, on the one hand, were saturated and supersat-
urated in silica and characterized by elevated water content, which is typical for su-
prasubduction magmas, and on the other hand – have a high content of Ti, Fe, Nb, Ta
and P, which are typical for magmas of intraplate (plume) origin. It suggests that for-
mation of such specific melts was attributed to melting of hydrated oceanic shallow
lithosphere under influence of new mantle plumes. Minor trondhjemite occurrences are
observed in form of veins and small bodies; their origin is considered to anatexis of hy-
drated rocks of the lithosphere by influence of mafic intrusions.
4. Sulfide mineralization, found in the Markov Deep, is confined to a zone of hydrother-
mal-metasomatic processing in cataclasites upon ferrogabbros and, apparently, was at-
tributed to fluids of magmatic origin, gradually diluted by sea-water from the subfloor
hydrosphere. The source of these fluids could be shallow intrusions of siliceous Fe-Ti-
oxide series. This mineralized zone is probably a piece of a Extinct39 feeder system of
former "black smoker".
5. SHRIMP-II U-Pb dating of zircon grains, extracted from gabbros of the area, revealed
the two groups: (i) "young", primary magma, with the age of 0.7-2.3 Ma, and (ii) "an-
cient", xenogenic, with age from 87 to 3117 Ma; at that zircons of different ages may be
found in the same samples. It is assumed that the grains of the "ancient" group belong
32 Tectonics – Recent Advances

to material of " slab graveyards", which fragments were captured by ascended mantle
plumes in the deep mantle.
6. It is shown that structure of the MAR in geomechanical terms is an example of a pierc-
ing structure, appeared due to the introduction of plastic less dense and less viscous
layer (asthenosphere) in a more dense layer (shallow lithosphere) under gravity. The
system is in unstable equilibrium state until appearance of mantle plume, which lifted
the dense lithosphere at higher hypsometric level in shape of a dome. This is causing
sliding of material on slopes of the dome (tectonic erosion), often one-sided, resulting in
exhumation of the deep-seated rocks on ocean floor and forming OCC. The interaction
of the plume head with the hydrated lithosphere, led to appearance of melts of siliceous
Fe-Ti-oxide series which providing hydrothermal activity.
7. As a result of tectonic erosion, pressure at the ridge axis decreased, which led to starting
of asthenospheric crest ascent; because of decompression, a zone of adiabatic melting
appeared at the top of the crest and role of basaltic volcanism gradually increasing. The
ridge axis gradually got a stable triangular shape and slumping of material becomes bi-
lateral. As a result, intermediate structures, such as Sierra Leone area, with magmatism
of E- and T-MORB are formed, and then the stage of vast eruptions of MORB comes.
The process goes to end when the melting zone was overfill by light restite material,
which is not involved in convection; the restites became a part of the lithosphere, and
the system returns to a state of unstable equilibrium.
8. Geological sense of the oceanic spreading, evidently, is a combination of thermal and
geomechanical processes at the lithosphere-asthenosphere boundary, starting with the
formation of domes and slumping newly formed material (new lithospheric plates) on
their slopes. From this standpoint constructive plate boundaries, at least in slow-
spreading ridges, should be represented by a collage of tectonic slices of shallow ocean-
ic lithosphere and basaltic sheets. So, process of spreading in the MAR has a cyclic char-
acter. It begins from appearance of OCCs, often characterized by “one-side spreading”
and numerous hydrothermal fields, and via structure type Sierra Leone area pass to
normal bilateral spreading with vast basaltic eruptions.
9. Based on these data and taking into account the published materials of seismic tomogra-
phy, it is developed a new model of oceanic spreading in the MAR. It is shown that the
long-term existence of the MAR’s oceanic spreading (at least 140 Myr) and stability of
composition of basalts can be explained by dynamic equilibrium between permanent re-
moval from asthenosphere of newly formed basalt and replenishment of new geochemi-
cally-enriched material of mantle plumes. The constant injections of a hot plume material
in asthenospheric lens provide forced convection its material and prevents it from freez-
ing; moreover, it also promote expansion of the asthenosphere both across a ridge axis
(oceanic spreading) and along its axis (propagation of the ridge).
10. How it is evidence from the MAR, three independent components in structure of its
mantle occur: (i) shallow lithosphere (including basaltic upper crust), (ii) asthenosheric
lens beneath the ridge, and (iii) deep mantle with “graveyards of slabs”. Each of them,
how it was shown above, has own origin and composition. From this evidently follows
that the total convection in the oceanic mantle is absent.
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 33

11. Thus, the processes in the slow-spreading ridge axes are mutually conditioned self-
sustaining character, resulting in almost continuous growth of the oceanic lithosphere
in its different parts, and supported the process of spreading, as evidenced by the pres-
ence of symmetrical magnetic anomalies.
12. Slow- and ultraslow-spreading ridges are a special class of oceanic spreading, characterized
by widespread development of oceanic core complexes and absence of subduction on pe-
riphery of the oceans, where developed passive margins. Appearance of such ridges is as-
sociated with the elongated areas of concentrated manifestation of sustainable mantle
plume activity. Apparently, extended heads of large plume came into contact with each
other, merging (coalescenced) in almost single body asthenospheric lenses.

Author details
E.V. Sharkov
Institute of Geology of Ore Deposits, Petrography,
Mineralogy and Geochemistry (IGEM), Russian Academy of Sciences, Moscow, Russia

Acknowledgement
I am very thanks to N.S. Bortnikov, S.A. Silantyev and A.V. Girnis for stimulated
discussions and comments.

10. References
Allaby, A. & Allaby, M., 2011. A Dictionary of Earth Sciences. 1999. Encyclopedia.com. 9 Sep., pdf.
Anderson, D.L., Tanimoto, T. & Shang, Y., 1992. Plate tectonics and hot spots: the third
dimension. Science, 256: 1645-1650.
Aranovich, L.Ya., Bortnikov, N.S., Serebryakov, N.S., & Sharkov, E.V., 2010. Conditions of
the Formation of Plagiogranite from the Markov Trough, Mid-Atlantic Ridge, 5°52′–
6°02′N. Doklady Earth Sciences, 434 (1): 1257–1262.
Basaltic Volcanism Study Project 1981. Basaltic volcanism on the terrestrial planets, 1981.
New York: Pergamon Press, 1286 p.
Beltenev, V., Ivanov, V., Rozhdestvenskaya, I. et al., 2009. New data about hydrothermal
fields on the Mid-Atlantic Ridge between 11o-14oN: 32th Cruise of R/V «Professor
Logachev». InterRidge News, 18: 14-18.
Bortnikov, N. S., Sharkov, E. V. , Bogatikov, O. A., Zinger, T.F., Lepekhina, E.N., Antonov,
A.V. & Sergeev, S.A., 2008. Founding of young an anicient zircons in gabbros of Markov
Deep, Mid-Atlantic Ridge, 5o30.4’ – 5o32.4’ N (results of SHRIMP-II U-Pb dating):
meaning for understanding of deep-seated geodynamics of the modern oceans. Dokl.
Earth Sci. 421(5): 859–868.
Cannat, M., Escartín, J., Lavier, L. et al., 2010. Lateral and Temporal Variations in the Degree
of mechanical weakening in the footwall of oceanic detachment faults. AGU Chapman
Conference “Detachments in Oceanic Lithosphere: Deformation, Magmatism, Fluid Flow, and
Ecosystems” Agros, Cyprus 8-15 May 2010. Conference Report: 38-39.
34 Tectonics – Recent Advances

Casey, J.F., Gao, Y., Benavidez, R. & Dragoi, C., 2010. The lowest δ7Li yet recorded in MORB
glasses: the connection with oceanic core complex formation, refractory rutile-bearing
eclogitic mantle sources and melt supply. 2010 AGU Fall Meeting. 13-17 December 2010.
San Francisco, California. Abstract V11A-2245, pdf.
Condie, K.C., 2005. High field strength element ration in Archean basalts: a window to
evolving sources of mantle plumes? Lithos, 79: 491-504.
Conference Outline. AGU Chapman Conference “Detachments in Oceanic Lithosphere:
Deformation, Magmatism, Fluid Flow, and Ecosystems” Agros, Cyprus 8-15 May 2010.
Conference Report: 20-21.
Dick, H.J.B., Meyer, P.S., Bloomes S.H. et al., 1991. Lithostratigraphic Evolution of an in Situ
Section of Oceanic Layer 3 . Eds. R.P. Von Herzen, P.T. Robinson et al. Proc. ODP, Sci.
Results 118: 439–538.
Dick, H.J.B., Robinson, P.T. & Meyers, P.S., 1992. The plutonic foundation of a low-
spreading ridge. Amer. Geophys. Union monograph 70: 1-39.
Dick, H.J.B., Tivey, M.A. & Tucholke, B.E., 2008. Plutonic foundation of a slow-spread ridge
segment: Oceanic core complex at Kane Megamullion, 23°30’N, 45°20’W. Geochem.,
Geophys., Geosyst., 9, Q05014, doi:05010.01029/02007GC001645.
Dilek Y. & Furnes H., 2011.Ophiolite genesis and global tectonics: Geochemical and tectonic
fingerprinting of ancient oceanic lithosphere. GSA Bull., 123(¾): 387-411.
Dmitriev, L.V., Sokolov, S.Yu., Melson, V.G. & O’Hirn, T., 1999. Plume and spreading
association of basalts and their reflection in petrological and geochemical parameters in
northern part of the Mid-Atlantic ridge. Russian J. Earth Sci., 1(6): 457-476.
Escartín, J., Smith, D.K., Cann, J. et al., 2008.Central role of detachment faults in accretion of
slow-speading oceanic lithosphere. Nature, 4559: 790-795.
Ernst, W.G., 2001. Subduction, ultrahigh-pressure metamorphism and regurgitation of
byont crustal slices – implications for arcs and continental growth. Phys. Earth and
Planet. Inter. 127: 253-275.
Fluids and redox reactions in magmatic systems. A.A. Kadik (ed.). 1991. Moscow, Nauka
Publ., 256 p.
Fontingie, D. & Schilling ,J.-G., 1996. Mantle heterogeneities beneath the South Atlantic: a
Nd-Sr-Pb isotope study along Mid-Atlantic Ridge (3°S-46°S). Earth Planet. Sci. Lett.,
142(1/2): 209-221.
Gillis, K.M. & Coogan, L.A., 2002. Anatectic migmatites from the roof of an ocean ridge
magma chamber. J. Petrol., 43: 2075–2095.
Girnis, A. V., 2003. Olivine-Orthopyroxene-Melt Equilibrium as a Thermobarometer for
Mantle-Derived Magmas. Petrology, 11(2): 101-113.
Handbook of physical constants of rocks. S. Clark (ed.)., 1969. Moscow, Mir: 520 p. (Russian
translation)
Hauri, E.H., Whitehead, J.A. & Hart, S.R., 1994. Fluid dynamic and geochemical aspects of
entrainment in mantle plumes. J. Geophys. Res., 99(B12): 24275-24300.
Hékinian, R., Juteau, T., Gracia, E. et al., 2000. Submersible observations of Equatorial Atlantic
Mantle: The St. Paul Fracture Zone region. Marine Geophysical Research, 21: 529-560.
Hess, H.H., 1962. The history of the ocean basins. Geol. Soc. Am. Buddington Vol., 599-620.
Ildefonse, B., Blackman, D.K., John, B.E. et al., 2007. Oceanic core complexes and crustal
accretion at slow-spreading ridges. Geology, 35(7): 623-626.
Cyclic Development of Axial Parts of Slow-Spreading Ridges:
Evidence from Sierra Leone Area, the Mid-Atlantic Ridge, 5-7°N 35

Karason, H. & van der Hilst, R.D., 2000.Constraints on mantle convection from seismic
tomography. The History and Dynamics of Global Plate Motions. Eds. M. Richards, R. G.
Gordon, and R.D. van der Hilst. Geophys. Monogr. 121: 277–289.
Knipper, A.L., Savelieva, A.L., Sharaskin, A.Ya., 2001. Problems of ophiolite classification.
Fundamental problems of general geotectonics. Yu.M. Puscharovsky (ed.). Moscow,
Nauchny mir: 250-281 (in Russian).
Krassivskaya I.S., Sharkov E.V., Bortnikov N.S., Chistyakov A.V., Trubkin N.V., Golovanova
T.I.., 2010. Variolitic lavas in the axial rift of the Mid-Atlantic rift and its origin.
Petrology, 18: P.263-277.
Langmuir, C.H., Klein, E.M. & Plank, T., 1992. Petrological systematics of mid-ocean ridge
basalts: constraints on melt generation beneath ocean ridges. Mantle flow and melt
generation at mid-ocean ridges. Eds.: Morgan J.P., Blackman D.K., Sinton J.M. Geophys.
Monogr. Am. Geophys. Union, 71: 183-280.
MacLeod, C.J., Searle, R.C., Murton, B.J. et al., 2009. Life cycle of oceanic core complexes.
Earth Planet. Sci. Lett., 287(3-4): 333-344.
Ohara, Y., Yoshida, T., Kasuga, S., 2001. Giant megamullion in the Parece Vela Backarc
basin. Mar. Geophys. Res., 22: 47-61.
Pearce, J., 2002. The oceanic lithosphere. Achievements and Opportunities of Scientific Ocean
Drilling. Spec. Issue of the JOIDES Journal, 28(1): 61-66.
Puscharovsky, Yu.M., Skolotnev, S.G., Peyve, A.A., Bortnikov, N.S., Bazilevskaya, E.S. &
Mazarovich, A.O., 2004. Geology and metallogeny of the Mid-Atlantic ridge: 5-7oN. Moscow,
GEOS, 151 p. (in Russian)
Quick, J.E. & Denlinger, R.P., 1993. Ductile deformation and the origin of layered gabbro in
ophiolites. J. Geophys. Res., 98: 14015–14027
Ritsema, J. & Allen, R.M., 2003. The elusive mantle plume. Earth Planet. Sci. Lett., 207: 1-12.
Roden M.F., Hart S.R., Frey F.A., Melson W.G., 1984. Sr, Nd and Pb isotopic and REE
geochemistry of St. Paul’s Rocks: its metamorphic and metasomatic development of an
alkali basalt mantle source. Contrib. Mineral. Petrol. 85(8):
Rudnick, R., 1990. Growing from below. Nature, 347(6295): 711-712.
Savelieva, G.N. 1987. Gabbro-ultrabasite complexes of the Urals ophiolites and their analogs in the
modern oceanic crust. Moscow, Nauka Publ., 246 p. (in Russian)
Savelieva, G.N., Bortnikov, N.S., Peyve, A.A. & Skolotnev, S.G., 2006. Utrabasite rocks of the
Mrkov Deep, rift valley of the Mid-Atlantic Ridge. Geochem. Intern. 44(11): 1192-1208.
Scheidegger, A.E., 1982. Principle of Geodynamics. 3rd edition. Berlin-Heidelberg-Bew York,.
Schilling, J.-G., Hanan, B.B., McCully, B. & Kingsley, R.H., 1994. Influence of the Sierra
Leone mantle plume on the equatorial MAR: a Nd–Sr–Pb isotopic study. J. Geophys.
Res., 99: 12005–12028.
Sharkov, E.V., 2006. Formation of layered intrusions and their ore mineralization. Moscow:
Scientific World, 364 p. (in Russian),
Sharkov, E.V., Abramov, S.S., Simonov, V.A., Krinov, D.I., Skolotnev, S.G., Bel’tenev, V.E. &
Bortnikov, N.S., 2007. Hydrothermal alteration and sulfide mineraliztion in gabbroids
of the Markov Deep (Mid-Atlantic Ridge, 6o N). Geology of Ore Deposits, 49(6): 467-486.
Sharkov, E.V., Bortnikov, N.S., Bogatikov, O.A., Zinger, T.F., Bel’tenev V.E., & Chistyakov,
A.V., 2005.,Third layer of the oceanic crust in the axial part of the Mid-Atlantic Ridge
(Sierra Leone MAR segment, 6oN). Petrology, 13(6): 540-570.
36 Tectonics – Recent Advances

Sharkov, E.V., Chistyakov, A.V., & Laz’ko, E.E., 2001. The Structure of the Layered Complex
of the Voikar Ophiolite Association (Polar Urals) as an Indicator of Mantle Processes
beneath a Back-Arc Sea . Geochemistry Intern., 39(9): 831-847.
Sharkov, E.V., Shatagin, K.N., Krassivskaya, I.S., et al., 2008. Pillow Lavas of the Sierra
Leone Test Site, Mid-Atlantic Ridge, 5o-7o N: Sr-Nd Isotope Systematics, Geochemistry,
and Petrology. Petrology, 16(4): 335-352.
Sharma, M., Hofmann, A.W. & Wasserburg, G.J., 1998. Melt generation beneath ocean
ridges: Re-Os isotopic evidence from the Polar Ural ophiolite. Miner. Mag., V.M.
Goldschmidt Conference. Toulouse Abstracts. 62A: 1375-1376.
Sharma, M., Wasserburg, G.J., Papanastassiou, D.A., Sharkov, E.V. & Laz’ko, E.E., 1995. High
143Nd/144Nd in extremely depleted mantle rocks. Earth and Planet Sci. Letters, 35: 101-114.

Silantyev, S.A., Dmitriev, L.V., Bazylev, B.A. et al., 1995. An examination of genetic
conformity between со-existing basalts, gabbro, and residual peridotites from 15°20'
Fracture Zone, Central Atlantic: evidence from isotope composition of Sr, Nd, and Pb.
InterRidge News, 4: 18-21.
Silantyev, S.A., Krasnova, E.V., Cannat, M. et al. 2011. Peridotite-gabbro-trondhjemite
association of rocks of Mid-Atlantic Ridge in 12о58’ – 14o45’N: hydrothermal fields
Ashadze and Logachev. Geochem. Intern., 49(4): 339-372.
Simonov, V.A., Sharkov, E.V. & Kovyasin, S.V. 2009. Petrogenesis of the Fe-Ti intrusive
complexes in the Sierra-Leone test site, Central Atlantic. Petrology, 17(5): 488-502.
Skolotnev S.G., Beltenev V.E., Lepekhina E.N. & Ipatiev I.S., 2010. Young and ancient
zircons from rocks of oceanic lithosphere of the Central Atlantic, geotectonical
consequences. Geotectonics, 6: 24-59.
Smith, D.K., Escartín, J., Schouten, H. & Cann, J.R., 2008. Fault rotation and core complex
formation: Significant processes in seafloor formation at low-spreading mid-ocean
ridges (Mid-Atlantic Ridge, 13o-15oN). Geochem, Geophys, Geosystems, 9(3): 1525-2027.
Snow, J.E. & Edmonds, H.N., 2007. Ultraslow-spreading ridges. Rapid paradigm changes.
Oceanography, 20(1): 90-101.
Thy, P., 2003. Igneous petrology of gabbros from Hole 1105A: oceanic magma chamber
processes. Eds. J.F. Casey and D.J. Miller. Proc. ODP, Sci. Results, 179: 1-76.
Tucholke, B.E., 1998. Discovery of “Megamullions” Reveals Gateways Into the Ocean Crust
and Upper Mantle. OCEANUS, 41(1): 15-19.
Tucholke, B.E., Behn, M.D., Buck, W.R. & Lin, J., 2008. Role of melt supply in oceanic
detachment faulting and formation of megamullions. Geology, 36: 455-458.
Tucholke, B.E. & Lin J., 1994.,A geological model for the structure of ridge segments in slow
spreading oceanic crust. J. Geophys. Res., 99: 11937-11958.
Turcotte, D.L. & Schubert, G., 2002. Geodynamics. 2nd еdition. Cambridge: Cambridge Univ.
Press, 847 p.
Widom, E., Carlson, R.W., Gill, J.B. et al., 1997. Th–Sr–Nd–Pb isotope and trace element evidence
for the origin of the Saõ Miguel, Azores, enriched mantle source. Chem. Geol., 140(1): 49-68.
Chapter 2

Features of Caucasian Segment of the


Alpine-Himalayan Convergence Zone:
Geological, Volcanological, Neotectonical,
and Geophysical Data

E.V. Sharkov, V.A. Lebedev, A.G. Rodnikov,


A.V. Chugaev, N.A. Sergeeva and L.P. Zabarinskaya

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/50253

1. Introduction
The Caucasus Mountain System (Fig. 1) is the part of the largest Alpine‐Himalayan collision
zone, stretching out for 16,000 kilometers across Eurasia from the Western Mediterranean to
the Western Pacific. In geological terms, it is represented by huge area of the Trans‐Eurasian
Belt (TEB) of Late Cenozoic activation, which was formed after the closure of the Mesozoic‐
Early Cenozoic ocean Neotetis (Sharkov, 2011). TEB is characterized by powerful modern
processes of mountain building, appearance of rift structures, numerous intraplate basaltic
lava plateaus and chains of intracontinental andesite‐latite volcanic arcs that trace suture
zone of continental plates collision. In that both types of magmatism developed almost sim‐
ultaneously with each other and with the tectonic processes of the entire length of the belt.
Large amagmatic blocks (North‐Eurasian and Indian) are located on its both sides.

Despite of long‐standing geological and geophysical investigations, the Caucasus is still


insufficiently studied region, especially in terms of interrelation of its deep‐seated structure
with neotectonics and manifestations of late Cenozoiv volcanism. However, numerous di‐
verse data obtained during the recent years, allow us to consider this region as a testing
place for a comprehensive study of modern geodynamic and petrologic processes in zone of
continental plates collision.

The aim of this chapter is synthesis of geological, petrological and geophysical information
about the deep‐seated processes in the Caucasian parts of the collision zone and their ex‐
pression in the geological processes in the crust.

© 2012 Sharkov et al., licensee InTech. This is an open access chapter distributed under the terms of the
Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
38 Tectonics – Recent Advances

Figure 1. Location of the Caucasus Mountain System in World map of volcanoes, earthquakes and plate
tectonics (World map of volcanoes, earthquakes and plate tectonics;. Compiled by T. Simkin, R.L. Till‐
ing, J.N. Taggart, W.J. Jones and H. Spall. Smithsonian Institution and US Geological Survey, 1989).

2. Alpine‐Caucasian part of the convergence zone


The most complicated structure within of the Trans‐Eurasian Belt of the late Cenozoic
activization has Alpine‐Caucasian segment, where systems of andesite‐latite volcanic
arcs and back‐arc basins (Alboran, Tyrrhenian, and Aegean seas as well as Pannonian
depression) occur (Fig. 2). Despite the differences in the morphology of these structures,
they have several common features. Along their periphery developed fold‐thrust zone, a
kind of ʺaccretionary prismʺ that form the ridges of the Alps, Carpathians, Apennines,
Gibraltar arc, etc. Among tectonic slices are often observed deep‐water sedimentary
rocks of the Tethys, ophiolite complexes, and sometimes blocks of the lower crust and
upper mantle rocks like Ivrea‐Verbano, Ronda, Beni Bousera and others (Magmatic ...,
1988).

In the rear of these structures, repeating their configuration, andesite‐latite volcanic arc of
three types occur: (i) island (Aegean), (ii) ʺsemicontinentalʺ, located partly on the continent
and in part ‐ directly adjacent to it (Alboranian, South‐Italian, etc.), and (iii) ʺintracontinen‐
talʺ (Carpathian, Anatolian‐Caucasian‐Elburssian, etc.). Behind these arcs newly‐formed
depressions occur with thinned at the expense of lower high‐velocity layers (Gize, Pavlen‐
kova, 1988) crust of intermediate or even oceanic type, within which basaltic volcanism
develops. Oceanic crust in the Western Mediterranean has evolved in the Late Cenozoic due
Features of Caucasian Segment of the Alpine-Himalayan
Convergence Zone: Geological, Volcanological, Neotectonical, and Geophysical Data 39

to continental crust the African plate (Ricou et al., 1986; Ziegler et al., 2006) as a result of
back‐arc spreading (Sharkov, Svalova, 2011).

Figure 2. Distribution of the late Cenozoic igneous rocks within the Alpine Belt
1 – back‐arc seas (A – Alboran, T – Tyrrhenian; Ae – Aegian) and “downfall” seas (B –Black, C – Caspian);
2 – back‐arc sedimentary basins (P‐ Pannonian, Po – Po valley); 3 – late Cenozoic andesite‐latite volcanic
arcs (in circles): 1 – Alboran, 2 – Cabil‐Tell, 3 – Sardinian, 4 – South‐Italian, 5 – Drava‐Insubrian,
6 – Evganey, 7 – Carpatian, 8 – Balkanian, 9 – Aegian, 10 ‐12 – Anatolian‐Caucasian‐Elburssian
(10‐ Anatolia‐Caucasian, 11 – zone of the Modern Caucasus volcanism, 12 – Caucasus‐Elburssian);
4 – areas of flood basaltic volcanism (in square): 1 – South Spain and Portugal, 2 – Atlas, 3 – Eastern
Spain, 4 – Central France massif, 5 – Rhine graben, 6 – Czech‐Silesian, 7 – Pannonian,
8 – Western Turkey, 10 – northern Arabia; 5 – suture zones of major thrust structures

According to geophysical data, lithosphere of the Alpine Belt has very complicated structure
and rather different beneath ridges and basins (Hearn, 1999; Artemieva et al., 2006, Kissling
et al., 2006). M. Artemiev (1971) firstly showed that two types of basins exist here (Fig. 3).
The first type (Tyrrhenian, Aegian, and Alboran seas a well as Pannonian depression) is
characterized by large positive isostatic anomalies, which evidence about excess of mass
beneath them, and presence of basaltic volcanism. Very likely, they represent occurrence of
the present‐day extended mantle plume heads, which continues to receive fresh material
resulting in onwards displacement of volcanic arcs in time. Judging by the magnitude of the
anomalies, the most intense it arrives in the areas beneath the youngest back‐arc sea of the
Mediterranean ‐ Aegean, as well as under the Pannonian basin.
Propagation of the plume heads, judging on geological data, had diversy‐oriented character
‐ under the Carpathians material moved to the east (Royden, 1989; Harangi et al., 2006),
under the Tyrrhenian Sea ‐ in the south‐east (Rehault et al., 1987; Harangi et al., 2006), under
the Aegean Sea ‐ to the south (Harangi et al., 2006), under the Alboranian Sea – to west
(Lonengran, White, 1997). At that crustal material above extended plume heads was trans‐
40 Tectonics – Recent Advances

ported to their front, where involved into descending mantle flows with the formation of
subduction zones with the corresponding magmatism; back‐arc basins with transitional and
oceanic crust were formed in their rear (Bogatikov et al, 2009).
Simultaneously with formation of depressions in the collision zone rift systems formed and
basaltic volcanism develops, represented mainly Fe‐Ti subalkaline basalts typical for
intraplate (plume) magmatism (rifts in Central and Western Europe, Atlas, basalt plateaus of
North Africa and Western Arabia, etc.), emerging in front of the mountain ranges (Grachev,
2003; Wilson, Downes, 2006). Based on isotopic and geochemical data, this anorogenic
Circum‐Mediterranean magmatism has a common source ‐ the so‐called Common Mantle
Reservoir (Common Mantle Reservoir, Hofmann, 1997; Lustrino, Wilson, 2007). This
obviously indicates presence of the modern mantle superplume beneath the whole region;
Alpine orogen with its complex system of ridges and basins located over the central part of
this superplume. It is in a good agreement with geophysical data: according to (Hearn, 1999;
Smewing et al., 1991, etc.), these plumes are merged into a single asthenospheric uplift at a
depth of 200‐250 km.

The second type of depressions presented by the Eastern Mediterranean, including the
Ionian Sea, and the Black and Caspian seas. In contrast to the aforementioned first type of
depressions, there are large isostatic minimums occurred beneath them, indicating a deficit
of mass, probably due to the downward currents in the mantle (ʺcold plumesʺ) between
extended heads of ʺhotʺ plumes. For these seas are typical passive margins, thick (up to 16‐
20 km) covers of Meso‐Cenozoic sediments and oceanic crust; magmatic activity absent in
connection with them. One of the main minima located in the eastern Mediterranean
(Levantine basin), where approximately 3‐3.5 Ma rapid subsidence area below sea level
occurred (Emels et al., 1995). Substantial isostatic anomalies are absent in connection with
the Black Sea, which, probably, indicates a reduced rate of downward movement.

Basins of the Black and Caspian seas look like a large ʺdownfallsʺ which cutoff pre‐Pliocene
structures of the Caucasus and the Kopetdag. The formation of these seas began, apparently,
in the early Cretaceous, but a significant deepening of the basins occurred in the late
Oligocene‐early Miocene, and in Miocene gradual shallowing of the basins took place
(Zonenshain, Le Pichon, 1986). New deepening of the Black Sea and South Caspian basins
occurred in Pliocene‐Quaternary, which happened almost simultaneously with the uplift of
the Caucasus and the Crimea, which in the Oligocene‐early Miocene were not expressed in
the relief (Grachev, 2000).

Judging on geophysical data, along the sides of ʺdownfallʺ basins (for example, under the
northern parts of the Black Sea and Eastern Mediterranean), steeply‐dipping seismically
active zones occur, which stretching in the mantle to depths of 60‐70 km (Zverev, 2002;
Shempelev et al, 2001).

An exception to the general rule, a large positive isostatic anomaly is, situated beneath the
Trans‐Caucasian Transverse Uplift on the Great and Lesser Caucasus (see Fig. 3). It is
located between the ʺdownfallsʺ of the Black and Caspian Seas, and continues to south to
eastern Anatolia and north of the Arabian Peninsula. Obviously, this anomaly is also
Features of Caucasian Segment of the Alpine-Himalayan
Convergence Zone: Geological, Volcanological, Neotectonical, and Geophysical Data 41

associated with ascending of a mantle plume. There is no depression here, but the front of
the Anatolian‐Caucasian‐Elburssian andesite‐latite arc shifted sharply to the north, forming
a zone of young volcanism of the Caucasus.

1 ‐ regional lows of average intensity; 2 ‐ of high intensity; 3 ‐ regional highs of average intensity; 4 ‐ intense;
5 ‐ volcanic rocks: a ‐ calc‐alkaline series, b ‐ basalt series; 6 – boundaries of the Alpine Belt

Figure 3. Distribution of regional isostatic anomalies and areas of Cenozoic volcanism in Alpine Belt.
After M.Artemiev.

3. Features of structure of Caucasian segment


The Caucasus Mountains are located in the eastern part of the actual Alpine zone itselfs in
the Arabian‐Eurasian syntaxis, between the ʺdownfallsʺ of Black and Caspian seas (Fig. 1).
As it was mentioned above, these seas, very likely, represent small relics of the Neotethys
Ocean, which gradually shallowed in the Miocene; their new significant deepening began in
the Pliocene‐Quaternary, along with the rise of the Caucasus and the Crimea.

Main Caucasian Ridge (Greater Caucasus, GC), in essence, is the southern edge of the Eurasian
plate, raised along the Main Caucasus Fault (MCF). The latter is a part of a super‐large deep‐
seated fault, traceable from the Kopetdag across the Caspian Sea, the Caucasus and the
Crimea. Very likely, that its further continuation is the Trans‐European Suture Zone
(Tornquist‐Teisseyre Fault Zone), which separates the East European craton from European
Variscides and Alpides (Artemieva et al., 2006). In essence, this Kopetdag‐Caucasus‐Trans‐
European super‐fault divides the Alpine orogen and stable Eurasian plate sensu stricto (Fig. 4).

There is a consensus now that formation of the Alpine tectonic structure of the Caucasus
occurred under leadership of submeridional horizontal compression. It is generated by a
counter movement of two plates: the Arabian ledge (the indenter) and the East European
42 Tectonics – Recent Advances

Craton. Pressure is transmitted from the Arabian to the area of the Greater Caucasus (Arabi‐
(Arabian‐Caucasian syntaxis). The current ʺinvasionʺ of the Arabian wedge into the
Eurasian plate occurs along the Bitlis‐Zagros toward the Greater Caucasus, and its
displacement relative to Eurasia takes place at a rate of several centimeters per year (Saintot
et al., 2006). Reducing the space between Arabian indenter and Eurasian plate in the Late‐
Alpian time reaches a total of 400 km, but it is unevenly distributed over the area (Leonov,
2007). The bulk of it falls on the territory lying south of the Main Caucasian Fault. To the
north of it, within the Greater Caucasus, some reduction also takes place, but it is small and,
apparently, does not exceed a few tens of kilometers.

TESZ – Trans‐European Suture Zone; MCF – Main Caucasian Fault; KDF – Kopetdag Fault

Figure 4. Kopetdag‐Caucasian‐Trans‐European superfault.

In this context, great importance is the structure of the MCF. According to popular
opinion, it is a large overthrust or underthrust Transcaucasian massif under the Great
Features of Caucasian Segment of the Alpine-Himalayan
Convergence Zone: Geological, Volcanological, Neotectonical, and Geophysical Data 43

Caucasus (Khain, 1984; Saintot et al., 2006 and references therein). However, the direct
and indirect observations of geological data, as well as the most straightforward form of
the fault suggests that the leading role play here steep imbricate structures and reverse
faults; a steep or vertical inclination of the MCF observed from geophysical data to depths
of 70‐80 km (Shempelev et al, 2005). From this position the MCF is a reverse fault with a
large value of vertical displacement, but without a large horizontal component (Leonov,
2007). In this case, the reduction of space to the south of the GC may be due to lateral
ʺdiffluenceʺ of the matter to both sides before the hard ʺstopʺ of the East European Craton
under the pressure of the Arabian indenter, as it can follows from the results of the study
of modern GPS‐deformation in the zone of Africa‐Arabia‐Eurasia continental collision
(Reilinger et al., 2006). In other words, judging on geological and geophysical data,
convergence of Arabia with Eurasia is substantially accommodated by lateral transport of
material within the interior part of the collision zone and lithospheric shortening between
the Caucasus and Zagros mountains.

4. Late Cenozoic volcanism of the Caucasus


An important feature of the area of syntaxis is a large belt of Late Cenozoic (up to practically
present‐day) volcanism, which extends in submeridional (Transcaucasian) direction from
the eastern Anatolia via the Lesser to the Greater Caucasus, where large Elbrus and Kazbek
volcanoes occur. Volcanics of this belt on their petrological and geochemical features are
often close to the suprasubduction calc‐alkaline magmas, and represented mostly by basaltic
andesites, andesites and dacites under subordinate role of low‐Ti basalts and rhyolites
(Koronovsky, Demina, 2007; Keskin et al., 2007). Volcanic structures themselves with a lot of
calderas and acid pyroclastics are also very close to the volcanoes of island arcs and active
continental margins. Along with this type of magmatism, extensive lava plateaus, formed by
eruptions (often fissured) of moderately alkaline basalts of within‐plate type, such as
Javakheti, Geghama, Syunik, Kars, and others, are also observed in the region.

Volcanics of the calc‐alkaline series (ʺsuprasubduction typeʺ) is of a prime interest. In


contrast to the island arcs and active continental margins, Eastanatolian‐Caucasus volcanic
belt stretches across the overall structure of the Greater Caucasus, following the direction of
positive Transcaucasian isostatic anomaly. Just south of the Van Lake, this belt is divided
into two branches: one of them can be traced to the west, to Central Anatolia, and the other ‐
to the east, toward the Elburs and the Zagros. Moreover, although the Arabian‐Caucasian
syntaxis is characterized by high seismicity, the earthquake hypocenters with depth not
more than 50‐60 km dominated, and deep‐focus earthquakes (up to 120 km) are extremely
rare and occurred only in the eastern part of the GC at its border with the Caspian Sea (Fig.
5 and 6). From this it follows that there is no any subduction beneath the eastern Anatolia
(Sandvol et al., 2003) and Caucasus at present.

Isotopic and geochemical data suggest that origin of magmatism of ʺsuprasubduction typeʺ
associated with interaction between a mantle plume head and continental crust material
(Lebedev et al, 2006; 2011; Chugaev et al., 2012). Judging by the fact that the orientation of
44 Tectonics – Recent Advances

East‐Anatolian‐Caucasian zone, where joined Anatolian‐Caucasian and Caucasian‐


Elburssian arcs, practically coincides with the zone of syntaxis, i.e. with the area of
maximum stress, we think that melting of crustal material in processes of deformations at
high pressures played an important role in generation of these magmas. As shown earlier,
crystal lattice of minerals under such conditions is at the stress state, which making them
easier to disintegrate, and the conversion to the liquid phase requires much less heat
(Sharkov, 2004 and references herein).

An essential role in this process of melting can play frictional heat generated during
deformation (Frischbutter, Hanisch, 1991; Molnar, England, 1995) and mantle fluids from
degassing due to decompression of the plume head, which penetrated a deformable crust,
introducing some warmth and some components, how it is determined by isotopic and
geochemical methods. From this standpoint the emergence of the Anatolian‐Caucasian and
Caucasian‐Elburssian arcs may be due to the diffluence of deep‐seated crustal material to
both sides from the Arabian indenter, fixing appearance of foci of melting related to the
tectonic flowage of material. In other words, these volcanic arc trace suture zone, on which
outflow of crustal matter from the “stop” of the Eurasian plate occurred in the process of
continental collision. Likely, the location of these sutures zones was determined by
configuration of the plume head which extends to the north and oncoming ʺstreamlinedʺ
from the both sides by tectonized material of shallow lithosphere.

Figure 5. Distribution of the earthquakes in the region. Shallow focuses sharply predominate.
Features of Caucasian Segment of the Alpine-Himalayan
Convergence Zone: Geological, Volcanological, Neotectonical, and Geophysical Data 45

Figure 6. Distribution of earthquake focuses beneath the volcanic area Eastern Turkey – Caucasus

Similar isotopic and geochemical characteristics were established for the late Cenozoic calc‐
alkaline volcanic rocks in the zone of continental collision in the Alpine‐Mediterranean
region, where these features of magmas are usually explained by the complex composition
of the mantle sources, strongly contaminated by crustal material (Harangi et al., 2006 and
references herein).

5. Discordance between deep‐seated situation and geological structure of


the region
Attention is drawn to discordance of deep‐seated processes and geological situation on
the surface: the Kopetdag‐Caucasian‐Trans‐European superfault sinking beneath level of
the Caspian Sea, where it can be traced only by a strip of earthquakes. This fault is
distorted in the north of the Black Sea, which may be indicative of the continued
deepening of the sea which disturb the subsurface geological structures (Fig. 7). In
addition, geophysical and geological data indicate that the head of the mantle plume from
the late Miocene has extended to the north, crossing at the depth the MCF and resulting in
the appearance of the modern volcanism including the late Quaternary volcanoes Kazbek
and Elbrus with its present‐day shallow magmatic chambers (Masurenkov, Sobisevich,
2010; Gurbanov et al., 2011). It is possible that ʺdivingʺ of the plume head under the edge
of the Eurasian plate, which occurred in the Miocene and continued at present, caused a
regeneration of an older suture zone, leading to the rise of the Greater Caucasus. Thus, the
geological situation in the region continues to develop, mainly due to large‐scale deep‐
seated processes.
46 Tectonics – Recent Advances

Figure 7. Discordances of the Kopetdag‐Caucasian‐Trans‐European superfault in the region

It is pay attention that orientation of the Anatolian‐Caucasian volcanic arc does not coincide
with the largest neotectonic structures on the Turkish territory – North‐Anatolian and East‐
Anatolian fractures zones: it is located between them. There is no clear correlation also
between volcanism and the present‐day motions of crustal material which established in the
zone of Africa‐Arabia‐Eurasia continental collision by GPS constraints (Reilinger et al.,
2006). All of these also evidence that deep‐seated processes in the mantle not always found
their reflection on the relatively shallow crustal level.

What could be further scenario? Most likely, this process of propagation of the mantle
plume head to the north could lead to ʺcut openʺ of the lithosphere of the Eurasian plate, the
separation of the Caucasian mountain system into two parts and the formation here
continental rift zone like the Rhine Graben.

6. Conclusions
1. The Caucasus is a part of huge Late Cenozoic Alpine‐Himalayan convergence zone. The
Greater Caucasus is an edge of the Eurasian plate, raised along the large reverse fault ‐
the Main Caucasian Fault. This fault, in turn, is a part of the super‐fault, stretching from
the Kopetdag to the Trans‐European Suture Zone (zone Tornquist‐Teisseyre).
2. The Caucasus is limited from both sides by large depressions modern Black and
Caspian seas of ʺdownfallʺ type, which ʺcutʺ pre‐Pliocene structures both the Caucasus
and the Kopetdag; origin of these seas is associated with downward currents in the
mantle (ʺcold plumesʺ).
3. The peculiar structure of the region is north‐south Transcaucasian Rise, which is located
in the northern part of Caucasian‐Arabian syntaxis. Large positive isostatic anomaly is
confined with it, apparently indicating presence here of the mantle plume head.
4. Along the zone of syntaxis the belt of Neogene‐Quaternary volcanism occurs which
begins in Eastern Anatolia and traced through the Lesser and Greater Caucasus. Two
Features of Caucasian Segment of the Alpine-Himalayan
Convergence Zone: Geological, Volcanological, Neotectonical, and Geophysical Data 47

types of volcanic rocks are represented here: (1) prevailing volcanics of calc‐alkaline se‐
series, very close in petrological and geochemical characteristics to suprasubduction
type, and (2) extensive lava plateaus formed by basalts of intraplate (plume related)
type.
5. However, subduction zone under the Caucasus region, as well as throughout the
Caucasian‐Arabian syntaxis is absent and shallow earthquakes (50‐60 km) are
dominated here. We considered that origin of calc‐alkaline magmas is associated with
interaction between the mantle plume head and crustal material at relatively shallow
depths under conditions of deformation at high pressures, leading to melting of the
material in the zone of collision. In other words, appearance here of such magmas has
not considered to any subduction zone.
6. Reduction of space in the area of Caucasian‐Arabian syntaxis, which occurred during
the Late Cenozoic, reached 400 km; such shortening in absence of subduction was
apparently achieved mainly due to ʺdiffluenceʺ of the crustal material to both sides
before hard “stop” of the East European Craton under the pressure of the Arabian
indenter.
7. Situation in the region continues to develop now mainly due to deep‐seated mantle
processes, destroying the structure of the pre‐Pliocene collision zone, while the
development of the underlying processes occurs independent of the processes in the
earthʹs crust.

Author details
Evgenii Sharkov, V.A. Lebedev and A.V. Chugaev
Institute of Geology of Ore Deposits, Petrography, Mineralogy and Geochemistry RAS, Moscow,
Russia

A.G. Rodnikov, N.A. Sergeeva and L.P. Zabarinskaya


Geophysical Center RAS, Moscow, Russia

7. References
Artemiev, M.E., 1971. Some peculiarities of deep‐seated structure of depressions of
Mediterranean type: evidence from data on isostatic gravity anomalies. Bull. Soc. of
the Nature Investigators of Moscow. Geol. Dept., 4, 5‐10 (in Russian with English
abstract)
Artemieva, I.M., Thybo, H. & Kaban, M.K., 2006. Deep Europe today: geophysical synthesis
of the upper mantle structure and lithospheric processes over 3.5 Ga. In: European
Lithosphere Dynamics. D.G. Gee and R.A. Stephenson (eds). Geol. Soc. London Mem. 32:
11‐42.
Chugaev, F.V., Chernyshev, I.V., Lebedev, V.A., Eremina, A.V., 2012. Isotopic composition
of plumbum and origin of Quaternaly lavas of Elbrus volcano (Greater Caucasus,
Russia): data of high‐precisious MC‐ICP‐MS method, Petrology, 20, in press.
48 Tectonics – Recent Advances

Emels, K.‐Ch., Robertson, A. & Richer, C., 1995. Mediterranean Sea.1. Science operator report
DSDP Leg.160: 11‐20.
Frischbutter A., Hanisch M., 1991. A model of granitic melt formation by frictional heating
on shear planes. Tectonophysics, 194: 1‐11.
Gautheron, G., Moreira, M. & Allegre, C., 2005. He, Ne and Ar composition of the European
lithospheric mantle. Chem. Geology, 1‐2: 97‐112.
Gize, P. & Pavlenkova, N.I., 1988. Structural maps of the Earth’s crust of Europe. Physics of
the Earth, N 10, 3‐14.
Grachev, A.F. (ed.), 2000. Neotectonics, geodynamics and seismicity of the Northern Eurasian.
PROBEL, Moscow, 487 p. (in Russian with English abstract).
Grachev, A.F., 2003. Final volcanism of Europe and it’s geodynamic nature. Physics of the
Earth, N 5: 11‐46.
Gurbanov, A.G., Bogatikov, O.A., Karamurzov, B.S., Tsukanova, L.E., Lexin, A.V., Gazeev,
V.M., Mokhov, A.V., Gornostaeva,T.A., Zharikov, A.V., Shmonov, V.M., Dokuchaev,
A.Ya., Gorbacheva, S.A., & Shevchenko, A.V., 2011. Unusual type of degassing from
melts of peripheral magmatic chambers of “asleep” Elbrus volcano (Russia):
geochemical and mineralogical features. Volcanology & Seismology, 4: 3‐20.
Harangi, S., Downes, H. & Seghedi, I., 2006. Tertiary‐Quaternary subduction processes and
related nagmatism. In: European Lithosphere Dynamics. D.G. Gee and R.A. Stephenson
(eds). Geol. Soc. London Mem. 32: 167‐190.
Hearn, T.M., 1999. Uppermost mantle velocities and anisotropy beneath Europe. Journ.
Geophys. Res., 104(B7): 15123‐15139.
Hofmann A.W., 1997. Mantle geochemistry: the message from oceanic volcanism. Nature,
385: 219‐229.
Keskin, M., 2007. Eastern Anatolia: A hotspot in a collision zone without a mantle plume. In:
Foulger, G.R. and Jurdy, D.M. (Eds.). Plates, plumes, and planetary processes. Geological
Society of America Special Paper 430, 693‐722.
Khain, V.E., 1984. Regional Geotectonics. Alpian‐Mediterranean Belt. Moscow, Nedra, 344 p. (in
Russian with English abstract).
Kissling, E., Schmid, S.M., Lippitsch, R., Ansorge, J. & Fugenschuh, B., 2006. Lithosphere
structure and tectonic evolution of the Alpine arc: new evidence from high resolution
teleseismic tomography. In: European Lithosphere Dynamics. D.G. Gee & R.A. Stephenson
(eds). Geol. Soc. London Mem. 32: 129‐145.
Lebedev, V.A., Bubnov, S.N., Chernyshev, I.V., and Golʹtsman, Yu.V., 2006. Basic
Magmatism in the Geological History of the Elbrus Neovolcanic Area, Greater
Caucasus: Evidence from K‐Ar and Sr‐Nd Isotope Data. Doklady Earth Sciences, 2006,
Vol. 406, No. 1, pp. 37‐40.
Lebedev, V.A., Chernyshev, I.V., Chugaev, A.V., Golʹtsman, Yu.V. & Bairova, E.D., 2010.
Geochronology of Eruptions and Parental Magma Sources of Elbrus Volcano, the
Greater Caucasus: K‐Ar and Sr‐Nd‐Pb Isotope Data. Geochemistry International, Vol. 48,
No. 1, pp. 41‐67.
Lonergan, L., White, N., 1997. Origin of the Betic‐Rif mountains belt. Tectonics, 16: 504‐522.
Features of Caucasian Segment of the Alpine-Himalayan
Convergence Zone: Geological, Volcanological, Neotectonical, and Geophysical Data 49

Louden, K.E., Chian, D., 1999. The deep structure of non‐volcanic rifted continental margins.
Roy. Soc. of London Phil. Trans, 357: 767‐804.
Lustrino, M. & Wilson, M., 2007. The circum‐Mediterranean Anorogenic Cenozoic Igneous
Province. Earth Sci. Rev. 81. P. 1‐65.
Magmatic Rocks. V. 5. Ultramafic Rocks, 1988. E.E. Lazʹko & E.V. Sharkov (Eds.). Moscow,
Nauka Publishers, 508 p. (in Russian)
Masurenkov, Yu.P. and Sobisevich, A.L. Fluid‐magmatic system of Pyatigosk volcanic
center. Doklady Earth Sci., 425, 815‐820.
Rehault, Т. G., Moussat, E. & Fabri, A., 1987. Structural evolution of Tyrrhenian back‐arc
basin. Marine Geology, 74, 123‐150.
Reilinger, R., McClusky, S., Vernant, P. et al., 2006. GPS constraints on continental
deformation in the Africa‐Arabia‐Eurasia continental collision zone and implications for
the dynamics of plate interactions. J. Geophys. Res., 2006. 111, B05411, 1 of 26,
doi:10.1029/2005JB004051.
Ricou, L.E., Dercourt, J., Geyssant, J. et al., 1986. Geological constrain on the Alpine
evolution of the Mediterranean Tethys. Tectonophysics, 123, 83‐122.
Royden, L.H., 1989. Late Cenozoic tectonics of the Pannonian Basin System. Tectonics, 8, 51‐
61.
Saintot, A., Brunet, M.‐F., Yakovlev, F., Sebrier, M., Stephenson, R., Ershov, A., Chalot‐Prat,
F. & McCann, T., 2006. The Mesozoic‐Cenozoic tectonic evolution of the Great
Caucasus. In: European Lithosphere Dynamics. D.G. Gee and R.A. Stephenson (eds). Geol.
Soc. London Mem. 32: 129‐145.
Sharkov, E.V., 2004. Role of the energy of interface formation in the melting and retrograde
boiling. Geochemistry Intern., 42: 950‐961.
Sharkov E., 2011, Does the Tethys Begin to Open Again? Late Cenozoic Tectonomagmatic
Activization of the Eurasia from Petrological and Geomechanical Points of View. New
Frontiers in Tectonic Research ‐ General Problems, Sedimentary Basins and Island Arcs, E.V.
Sharkov (Ed.), Rijeka: InTech, 2011, 4‐18.
Sharkov E. & Svalova V., 2011. Geological‐Geomechanical Simulation of the Late Cenozoic
Geodynamics in the Alpine‐Mediterranean Mobile Belt. In: New Frontiers in Tectonic
Research ‐ General Problems, Sedimentary Basins and Island Arcs, E.V. Sharkov (Ed.), Rijeka:
InTech, 19‐38.
Shempelev, A.G., Prutsky, N.I., Feldman, I.S., & Kukhmazov, S.U., 2001. Geologo‐
geophysical model along profile Tuapse‐Armavir. In: Tectonics of the Neogea: general and
regional aspects. Proceedings of XXXIV tectonic meeting. Moscow, GEOS, 316‐320 (in
Russian).
Shempelev, A.G., Prutsky, N.I., Kukhmazov, S.U. et al,, 2005. Materials of geophysical
investigation along Near‐Elbrus profile (volcano Elbrus‐Caucasian Mineral Waters). In
Tectonics of the Earth’s crust and mantle. Proceedings of XXXVIII tectonic meeting. Moscow,
GEOS, 361‐365 (in Russian).
Spakman, W., van der Lee, S. & van der Hilst, R., 1993. Travel‐time tomography of the
European‐Mediterranean mantle down to 1400 km. Phys. Earth Planet Inter., 79, 3‐74.
50 Tectonics – Recent Advances

Wilson, M. & Downes, H., 2006. Tertiary‐Quaternary intra‐plate magmatism in Europe and
its relationship to mantle dynamics. In: European Lithosphere Dynamics. D.G. Gee and
R.A. Stephenson (eds). Geol. Soc. London Mem. 32, 147‐166.
Ziegler, P.A., Schumacher, M.E., Dezes, P., van Wees, J.‐D. & Cloetingh, S., 2006. Post‐
Variscan evolution on the lithosphere in the area of the European Cenozoic Rift System.
In: European Lithosphere Dynamics. D.G. Gee and R.A. Stephenson (eds). Geol. Soc.
London Mem. 32, 97‐112.
Chapter 3

3D Modelling and Basement Tectonics of


the Niger Delta Basin from Aeromagnetic Data

A.A. Okiwelu and I.A. Ude

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/48158

1. Introduction
Basement structure is crucial in determining the origin, deformation and evolution of basin
as well as the influence of the basement in the overlying Phanerozoic rocks and deposition
and migration of hydrocarbon within a basin [1]. In petroleum exploration the structural
surface interpreted from magnetic depth estimates is often the best available approximation
to the true crystalline (metamorphic/igneous) basement configuration and estimate of base‐
ment depth (sedimentary thickness) is a primary exploration risks parameter [2]. Specifical‐
ly, the magnetic basement is very relevant in the application of magnetics to petroleum
exploration. Magnetic basement is the upper surface of igneous or metamorphic rocks
whose magnetization is so much larger than that of sedimentary rocks. Magnetic basement
may or may not coincide with geologic or acoustic basement. In the application of magnetic
method, the source depth is one of the most important parameters. Others are the geometry
of the source and contrast in magnetization. Basement structure determined from magnetic
depth estimates provides insight into the evolution of more recent sedimentary features
(subbasin, localization of reservoir bearing structures) in areas where the inherited base‐
ment fabrics or architecture has affected either continuously or episodically basin evolution
and development [2]. Depths to the magnetic basement are very useful in basin modeling
such as determination of source rock volume and source rock burial depth. The identifica‐
tion and mapping of geometry, scale and nature of basement structures is critical in under‐
standing the influence of basement during rift development, basin evolution and subse‐
quent basin inversion [3], [4]. From regional aeromagnetic data sets, information such as
tectonic frame of the upper crust can be obtained. The patterns and amplitude of anomalies
reflect the depth and magnetic character of crystalline basement, the distribution and vol‐
ume of intrusive and extrusive volcanic rocks and the nature of boundaries between mag‐
netic terrains [5].

© 2012 Okiwelu and Ude, licensee InTech. This is an open access chapter distributed under the terms of the
Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
54 Tectonics – Recent Advances

Magnetic anomalies are a result of two things: a lateral contrast in rock composition
(lithology) or a lateral contrast in rock structure [6], [7]. Where there is no contrast in
magnetization no anomaly is produced. The magnetization could be due to normal
induction in the Earth’s field or due to remanent magnetization. For accurate modeling and
interpretation of magnetic data it is important to recognize and incorporate the remanent
component where they exist.

Magnetic anomaly transformation/enhancement provides the opportunity to unravel the


basement structure and lithology. Such information is not readily available from the total
intensity data sets especially if they are of low resolution. Our objective in this study is to
demonstrate the relationship between basement framework, magnetic expression and
hydrocarbon prospect in the Niger Delta basin using 3‐D modelling and enhancement data
sets. In the Tertiary Niger Delta basin exploration (seismic) for hydrocarbon is confined to
the sedimentary section despite the fact that basement structure analysis has been used in
locating hydrocarbon targets in other sedimentary basins of the world. We show that the
geodynamics of the deep basement are important phenomena to the explorationist and
could be an important factor that can directly lead to the risk assessment of specific prospect
sites in hydrocarbon exploration. Specifically, we demonstrate that basement structure in
the offshore Niger Delta have control on oil and gas discoveries even though the basement
is known to be beyond drillable depths. It is not possible to prove basement control neither
with subsurface mapping, as few wells penetrate basement, nor with seismic, as the
basement reflector is not always mappable; residual aeromagnetics is the principal
technique used in mapping basement and it is generally applied only to outline the
basement fault block pattern [8]. [9] used aeromagnetic data to show that axis of
hydrocarbon pool in Alberta basin is coincident with the strike of the basement sourced
magnetic signals. [10] reported the relationship between tectonic evolution and hydrocarbon
in the foreland of the Longmen Mountains and showed that superimposed orogenic
movement and related migration of sedimentary basins controlled the generation,
migration, accumulation and disappearance of hydrocarbons. [11] reported three‐sets of
traps from geophysical and geological data in offshore United Arab Emirate of which one is
basement related.

2. Location, geologic and tectonic setting


The study area, fig.1 (Niger Delta) is situated in the Gulf of Guinea. From the Eocene to the
present, the delta has prograded southwestward, forming depobelts that represent the most
active portion of the delta at each stage or its development [12]. The ideas expressed on
location, geologic and tectonic setting is from [13]. The depobelts in this basin form one of
the largest regressive deltas in the world with an area of some 300,000km2, sediment volume
of 500,000km3 [14] and a sediment thickness of over 10km in the basin depocenter. The
onshore portion of the Niger Delta Province is delineated by the geology of Southern Nige‐
ria and Southwestern Cameroon. The northern boundary is the Benin Flank – an east‐
northeast trending hinge line south of the West Africa basement massif. The northeastern
boundary is defined by outcrops of the Cretaceous on the Abakaliki High and further east‐
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 55

south‐east by the Calabar Flank – a hinge line bordering the adjacent Precambrian. The off‐
shore boundary of the province is defined by the Cameroon volcanic line to the east, the
eastern boundary of the Dahomey basin (the eastern‐most West African transformed‐fault
passive margin) to the west [13]

The Niger Delta Province contains only one identified petroleum system [15]. This system is
referred to as the Tertiary Niger Delta (Akata‐Agbada) petroleum system. The maximum
extent of the petroleum system coincides with the boundaries of the province. Most of the
petroleum is in the fields that are onshore or on the continental shelf in waters less than 200
meters deep and occurs primarily in large, relatively simple structures. The Tertiary section
of the Niger Delta is divided into three formations (fig.2), representing prograding
depositional facies that are distinguished mostly on the basis of sand‐shale ratio. The type
sections of these formations are described by [16] and [17]. The Akata Formation at the base
of the delta is of marine origin and is composed of thick shale sequences (potential source
rock), turbidite sand (potential reservoirs in deep water) and minor amounts of clay and silt.
Beginning from the Paleocene and through the Recent, the Akata formation formed during
lowstands when terrestrial organic matter and clays were transported to deep water areas
characterized by low energy conditions and oxygen deficiency. Turbidity currents likely
deposited deep sea fan sands within the upper Akata Formation during development of the
delta. Deposition of the overlying Agbada Formation, the major petroleum‐bearing unit,
began in the Eocene and continues into the Recent. The formation consists of paralic
siliciclastics over 3700 meters thick and represents the actual deltaic portion of the sequence.
The clastics accumulated in delta front, delta‐topset and fluvio‐deltaic environments. In the
lower Agbada Formation, shale and sandstone beds were deposited in equal proportions
but the upper portion is mostly sand with only minor shale interbeds. The Agbada
Formation is overlain by the third formation, the Benin Formation, a Continental latest
Eocene to Recent deposit of alluvial and upper coastal plain sands that are up to 2000m
thick [16].

The tectonic framework of the continental margin along the West Coast of equatorial Africa
is controlled by Cretaceous fractures zones expressed as trenches and ridges in the deep
Atlantic. The trough represents a failed arm of a rift triple junction associated with the
opening of the south Atlantic [13].
In the Delta, rifting diminished altogether in the Late Cretaceous. After rifting ceased, gravi‐
ty tectonics became the primary deformational process. Shale mobility induced internal
deformation occurred in response to two processes. First, shale diapirs formed from loading
of poorly compacted, over‐pressured prodelta and delta‐slope clays (Akata Formation) by
the higher density delta‐front sand (Agbada Formation). For any given depobelt, gravity
tectonics were completed before deposition of the Benin Formation and are expressed in
complex structures, including shale diapirs, roll‐over anticlines, collapsed growth fault
crests (fig.3), back‐to‐back features and steeply dipping closed spaced flank faults [18]. Dep‐
osition of the three formations occurred in each of the five off‐lapping Siliciclastic Sedimen‐
tation Cycle that comprises the Niger Delta. The cycles (depobelts) are defined by synsedi‐
mentary faulting that occurred in response to variable rates of subsidence and
56 Tectonics – Recent Advances

STUDY AREA

Figure 1. Generalized geological map of Niger Delta Basin

Figure 2. Schematic dip section of the Niger Delta

sediment supply. The interplay of subsidence and supply rates resulted in deposition of
discrete depobelts. When further crustal subsidence of the basin could no longer be accom‐
modated, the focus of sediment deposition shifted seaward forming a new depobelt. Each
depobelt is separate unit that corresponds to a break in regional dip of the delta and is
bounded landward by growth faults and seaward by large counter‐regional faults or the
growth fault of the next seaward belt [18]. Five major depobelts are generally recognized,
each with its own sedimentation, deformation, and petroleum history. The northern delta
province, which overlies relatively shallow basement, has the oldest growth faults that are
generally rotational, evenly spaced with increased steepness seaward. The central delta
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 57

province has depobelts with well defined structures such as successively deeper roll over
crests that shifts seaward for any given growth fault. Lastly, the distal delta province is the
most structurally complex due to internal gravity tectonics in the modern continental slope.

Figure 3. Principal Types of oilfield structures in the Niger Delta with schematic indications of common
trapping configurations (After Tuttle etal., 1999)

3. Magnetic data
The total intensity magnetic data (fig. 4) was flown at an elevation of 2500ft (762m) above
sea level with flight line spacing of 2km. This is therefore a low resolution data sourced from
geological survey of Nigeria. The magnetic anomalies are sourced overwhelmly from the
basement. The main advantage of this data for this study is that cultural features such as
railroad tracks, power transmission cables, metals from buildings, drill cores, storage tanks,
steel well casings, oil pipelines and other metallic objects are not sources of anomalies in the
data and therefore, cultural editing are not required. Large concentrations of cultural
sources with particularly strong and pervasive magnetic fields such as cathodically protect‐
ed pipelines can seriously mask the geologic information contained in aeromagnetic survey
data [19]. Gridding of the data were done at 1km interval along the flight lines which is
orthogonal to the regional geologic strike. The grid spacing is tight enough to capture the
anomaly details and meet the objective of this study. All the magnetic maps were plotted
with potent software with the colour interval in all the figures being the convention in mag‐
netic studies. The magnetic highs are depicted with yellows, oranges and reds while pur‐
ples, blues and greens represent magnetic minima (lows). Using colours on the aeromagnet‐
ic map further accentuates the effects of visualization of the magnetic fields. The gradient
zones in the total magnetic intensity field data are shear zones. The shear zones are relics of
basement tectonics and are early Precambrian plate boundaries. They trend NE‐SW and are
principal zones of weakness in the basement and reflects edges of basement blocks.
58 Tectonics – Recent Advances

Figure 4. Total magnetic intensity data offshore Niger Delta. The gradient zones and elliptical contours
reflects basement structures

The magnetic anomalies in fig.4 are as a result of total magnetization of rock and represent
the vector sum of the induced and remanent magnetizations. The induced magnetizations
are produced as a result of the interaction of magnetic minerals with the Earth’s magnetic
field. This is contrary to the remanent magnetization which acts independently of the
Earth’s present field. If remanent magnetization is significantly strong and acts in the direc‐
tion opposite to the present field, it can generate isolated magnetic high at low latitude and
produce a magnetic low at high latitude. If the induced magnetization acts in the direction
of the Earth’s field it produces a magnetic low in low latitude. Experimental work on rock
magnetization has made it abundantly clear that contrary to the earlier belief, presence of
permanent magnetization is often the rule than the exception, in the rocks of the Earth crust
and permanent magnetization associates itself with induced magnetization to orient the
polarization vector of the rock mass in some arbitrary direction [20]. The direction of this
polarization vector influences appreciably the size and shape of the associated magnetic
anomaly. The ratio of the strength of remanent magnetization to induced magnetization is
known as Koenigsberger ratio. If the Koenigsberger ratio is greater than one, it suggests that
the remanent magnetization played a dominant role.

The observed data was used to compute, by least squares, the mathematically describable
surface giving the closest fit to the magnetic field that can be obtained within a specific
degree of detail. We exploited the fact that the regional field is a first‐order surface of the
form:

T ( x , y )  ax  by  c (1)

Where a, b and c are the coefficients and are computed so as to minimize the variation of the
residual. This approach of computing the regional is suitable because higher order
polynomials may be amenable to a large area over which the regional has many
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 59

convolutions. The regional field was subtracted from the total intensity data to obtain the
residual field data (fig. 5a).

In the total intensity data (lat. 40 00N ‐ lat. 30 41N and long. 60 00E ‐ 60 18E) an elliptical
magnetic high and low trending E‐W are separated by strong magnetic gradient. The low is
closely flanked by a high trending NE‐SW as shown in the total intensity and residual maps.
The northeast sector is also characterized with elliptical anomaly trending E‐W. The elliptical
disposition is a pointer to dyke‐like intrusives. The predominance of these lineaments striking
NE‐SW and E‐W can be attributed to regional stresses in the basement. There is a high
gradient in the southeast sector of the study area juxtaposed with elliptical anomalies. The
elliptically shaped anomaly in the residual data has three small circularly shaped anomalies
not revealed in the total intensity map. These are plug‐like intrusives within the basement.

(a)

(b)
Figure 5. a. The Residual magnetic anomaly data showing some circular and elliptical contours not
revealed in total magnetic intensity data. The white lines indicate shear zones. b. The location of profile
lines 800E, 900E, 1400E, 2200E, 2400E and 2900E in the residual magnetic field data. The rectangular
wire frames represents the magnetic sources.
60 Tectonics – Recent Advances

4. 3D magnetic modelling and depth determination


For resource exploration purposes one of the most useful inferences that may be derived
from analyses of potential field (magnetic and gravity) data is the depth to crystalline
basement beneath sedimentary cover [21]. Most magnetic anomalies come from only a few
rock types, such as volcanics, intrusives and basement rocks. Magnetic data therefore can be
used to estimate depth to basement‐ a classic use for such data [22]. Generally, there are two
approaches to potential field modeling: inverse and forward modeling. In magnetic
modelling the inverse approach is whereby a 2D or 3D susceptibility or geometric model is
computed to satisfy (invert) a given observed magnetic field. In this case, the input is the
observed data while the output is the geologic model. That is, the observed data is used to
draw conclusion about the physical properties of the system. Physics principle allows the
means for computing the data values given a geological model. This constitutes forward
modeling (problem).This implies that if one has the knowledge of the properties of a system
one can predict the response of that system. Therefore, the input of a forward model is the
geologic model while the output is the computed values. Forward modeling commences by
erecting a model based on geologic knowledge and geophysical intuition, then calculating
the predicted magnetic field and comparing with observations. The next important step is to
iterate the model to fit. The most significant aspect of forward modeling is that it could
show if the postulated geologic model is incompatible or compatible with potential field
data. This reduces ambiguity in interpretation. Thus, in this study, we adopted 3‐D forward
modeling because the geologic setting of the Niger Delta is well known.

The 3D model constitutes a network or grid values which models a geologic surface repre‐
sented as a surface of susceptibility contrast. The residual magnetic field data (fig. 5b) was
used for modeling instead of the filtered/enhanced magnetic field data. It is not appropri‐
ate to model using filtered data, because we do not know if the component of the magnetic
field removed by the filter is also removed in our model [23]. If an interpreter has two to
three depth points, two at the edges and one on the basin floor, these depths are contoured
with knowledge of the expected structural style [6]. To fulfill the above condition profiles
were taken to model the depth to the basement using rectangular wire frame in fig. 5b. In
our approach, we used a complete quantitative approach‐ complete in the sense that the
three types of information about the geologic target (the depth, geometry/dimensions and
the contrast in the relevant physical properties) were estimated. The 3D forward modeling
is based on models that accommodated both induction in the Earth’s field and remanent
magnetization. Magnetics like other geophysical methods are non‐unique. One way we
adopted to reduce the ambiguity in interpretation is by using geometric simple body. In
potential field modeling, popular geometric bodies usually exploited are ellipsoids, plates,
rectangular prisms, polygonal prisms and thin sheets. In this study, we used rectangular
prism model because of its simple shape and because it makes the process of modelling
simple and stable. Thus, simple models were created using rectangular prism that conform
regularly well with the data on the profiles and that are consistent with anomalies on the
image of the observed field. Secondly, ambiguity is reduced because we know the geologic
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 61

setting (rifting) of the Niger Delta. The most important element required for interpreting
magnetic data is a geologic concept or structural model. We are never blind; even if the
only data available in an area is magnetic data, we know the area is in rift setting or fore‐
land basin or along a passive margin. The data is no longer non‐unique [23]. Another ap‐
proach we used to account for non‐uniqueness was to fix susceptibility and vary geometry
until a reasonable fit was achieved. The modelled magnetic anomalies (figs. 6‐9) resulted
from lithologic and structural changes. Lithologic variation (igneous and metamorphic)
usually produces the strongest magnetic signals. Amplitude of hundreds of nanoTesla is
due to lithologic variations in the basement or igneous rocks within the sedimentary sec‐
tion while amplitude of tens of nanoTesla are related to basement structures [23]. The am‐
plitudes of the anomalies modelled have been moderated by two factors. The main factor is
that the basement rocks in the study area are buried by thick sedimentary sequence, thus
their amplitude is moderated. The second factor is that high amplitude anomalies would
be observed where basement structures are not present. In the study area there are suffi‐
cient basement structures (for example, faults, contacts and dykes). Thus, if a small anoma‐
ly caused by a large structure is superimposed upon a large anomaly caused by lithologic
contrast, the two features may be inseparable.

Zones of lithologic contrast are often loci of structural disturbance [24]. Magnetic and gravi‐
ty data have been traditionally thought of as regional screening tools capable of providing
basin edges or basement mapping. In recent years, the application of these data has greatly
expanded to include modelling of prospect‐level targets. If detailed prospect‐level quantifi‐
cation of the basement structure is required, a 3‐D model would be more appropriate [25].
We exploited the algorithm of [20] based on magnetic anomalies due to rectangular prism‐
shaped bodies to determine depth to basement. This algorithm helped to meet our objectives
because it considered both induction in the Earth’s field and remanent magnetization. The
parameters defining the prisms are shown in fig. 10. Six profiles (Line 800E, Line 900E, Line
1400E, Line 2200E, Line 2400E and Line 2900E) in fig.5b were modeled to obtain geometries
(figs. 6‐9) and physical properties of the basement sources. The attitude (orientation) of the
body (sources) is affected by the manner in which the profiles cut the bodies. The shape of
the magnetic anomalies in all the models were affected by the shape, depth of the sources,
inducing and remanent field which varies in intensity and direction of magnetization [26].
Five discrete basement depth values were obtained from the modelled data and these values
provided additional depth control offshore. A depth to basement at the adjacent onshore
(fig. 11) gave a value of 12000m (fig. 12) by modeling body 6. These basement depth values
which are equivalent to the thickness of sedimentary section in the study area contribute to
basin modeling and put an upper limit on the thickness of source rocks, the base of which
may not be well imaged from seismic information [25]

The range in values of magnetic susceptibility and remanent intensity reflects sources of
basaltic and ultrabasic composition which may have utilized the tensional cracks in the fault
system in the study area.
62 Tectonics – Recent Advances

S N

Observed

(a)

S N

Observed

(b)
Figure 6. Modelling of (a) profile line 800E (b) profile line 900E showing dipping sources buried at
depth 8,500m with a length of 28000m.
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 63

S N

Observed

(a)

S N

Observed

(b)

Figure 7. Modelling of (a) profile line 2200E (b) profile line 2400E. Magnetic signatures are due to
remenance in the Earth’s field
64 Tectonics – Recent Advances

S N

Observed

Figure 8. Modelling of Profile line 2900E which revealed a dyke‐like source

S N

Observed

Figure 9. Modelling of profile line 1400E revealing a tabular body of length 16000m.

The magnetic profiles, Line 2200E and Line 2400E over bodies 1 and 2 show strong rema‐
nence (strong magnetic minima in the north flanked by moderate magnetic high in the
south). This is manifested in the intensity of remanence (0.0600‐0.1600Amp/m) and low
susceptibility values of 0.007‐0.008SI. These values point to a body of basaltic composition
and the depth to the geologic body is 9000m. Bodies 3 and 4 modelled with profile Line
2900E and Line 800E/Line 900E respectively show signatures that are entirely due to induc‐
tion in the Earth’s field (strong magnetic lows) which is consistent with results from equato‐
rial belt. The geophysical explanation of this magnetic low is that the susceptibility of the
anomalous body is lower than that of the host rock. That is, a basaltic body intruded into the
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 65

ultrabasic source of magnetic susceptibility, 0.017SI at a depth of 11,000m. Modelling of


profile Line 1400E incorporated both induced and remanent magnetization. The remanent
magnetization of body 5 is ‐0.3700Amp/m while the magnetic susceptibility is 0.008SI. Rela‐
tively strong high to the south and very moderate low to the north in the magnetic signature
suggest remanence.

Figure 10. Rectangular prisms showing the parameters of the model

Figure 11. Location of profile lines 1800E, 2000E and 2600E on the magnetic data in the adjacent on‐
shore

The depth values obtained from the 3‐D modelling were used to prepare magnetic basement
depth map (fig. 13). A reasonable detailed basement structure map is an integral part of any
regional geological or hydrocarbon evaluation process. Such a map identifies critical
66 Tectonics – Recent Advances

structural trends, the locations of the regions prominent structural prospects and location
and geometry of the hydrocarbon deponcenters [27]. A magnetic basement low (thick
sedimentary section) traverses the southwest and northwest sectors of the study area with a
maximum sedimentary thickness of 11,736m. This is deep basement trough. At lat.30 30 ‐30
41N and long.60 16 ‐ 60 28 E there is a basement high indicating structural high with a
maximum thickness of 5,583m. This basement high is flanked either side by structural lows.
In the northeast sector there is a basement high flanked by basement low. Thus, there is
spatial relationship between paleotopographic highs on the Precambrian basement and
structural and thickness anomalies in the overlying Tertiary sediments. Therefore, the depth
to magnetic basement map (fig. 13) has located deep depocenter, high blocks and major
sedimentary fairways in the study area.

S N
Observed
Calculated

(a)

S N

Observed
Calculated

(b)

S N
Observed
Calculated

Figure 12. Modelling of profile line (a) 1800E, (b) 2000E and (c) 2600E revealed a sill‐like body onshore
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 67

Figure 13. Depth to magnetic basement (thickness of sedimentary section), highlighting basement high,
basement flanks and sedimentary fairways

5. Magnetic anomaly enhancement


The most important and accurate information provided by magnetic data is structural fabric
of the basement. Major basement structures can be interpreted from consistent discontinui‐
ties and /or pattern breaks in magnetic fabric [1]. The basement structures manifest as shear
zones, fault (brittle faults and domain fault boundaries) which are usually weak zones. The‐
se

basement structural features are lineaments and in most cases subtle. Subtle potential field
lineaments could be gradient zones, alignment of separate local anomalies of various types
and shapes, aligned breaks or discontinuities on the anomaly pattern. Subtlety of desirable
lineament requires detail processing using a wide range of anomaly enhancement technique
and display parameters [29]. Filtering and image processing of aeromagnetic data are essen‐
tial tools in mineral exploration. Directional horizontal derivatives enhance edges (figs. 14
& 15a) while vertical derivative (fig. 15b) narrows the width of anomalies and so locate the
source bodies more accurately [30].

The most commonly applied techniques include the horizontal gradient and analytic signal.
Other methods for detecting edges of structures and linear features such as faults include tilt
and diagonal derivatives. [31] and [32] gave expression for magnetic field horizontal gradi‐
ent as

2 2
 M   M 
HG( x , y)      (2)
 x   y 
68 Tectonics – Recent Advances

Maxima in the horizontal gradient magnitude of the reduced‐to‐pole magnetic field are
exploited to locate vertical contacts and estimate their strike directions; where M is the
magnetic field. The analytic signal also reveals basement structure and uses its maxima to
locate the outlines of magnetic sources and their edges. [33] defined the analytic signal from
field derivatives as:

2 2 2
 M   M   M 
AS        (3)
 x   y   z 

While the horizontal gradient is less prone to noise because it calculates only the two hori‐
zontal derivates, it is not well suited to analyzing potential field data at low latitudes. This is
because it requires reduction to the pole. Reduction to the pole is very unstable in magnetic
equator (equatorial belt). The width of a maximum or ridge in analytic signal data is an
indicator of depth of the contact as long as the signal arising from a single contact can be
resolved [34]. While the analytic signal could be discontinuous, the enhancement is very
handy at low magnetic latitude because it eliminates the problems inherent with reduction
to pole (RTP) at low latitude.

One technique we find very useful is the directional horizontal gradient. It appears not to be
popular but it is very effective in revealing basement features. This technique is simple and
like the analytic signal (fig. 16), it can reveal N‐S structures which are difficult to identify in
equatorial belt. The directional horizontal derivative does not require reduction to the pole.
[35] showed that the horizontal derivatives of a smoothly varying scalar quantity,  ( x , y )
measured on a horizontal surface can easily be determined using simple finite‐difference
methods. The horizontal derivatives of  ( x , y ) at point i , j are given approximately by

d ( x , y ) i  1, j  i  1, j
 (4)
dx 2 x

d ( x , y ) i , j  1  i , j  1
 (5)
dy 2 y

This can be performed in the Fourier domain. Thus,

 dn 
F  n   (ikx )n F   , (6)
 dx 

 dn 
F  n   (ik y )n F   (7)
 dy 

Where, (ikx )n and (ik y )n are filters that transform a function measured on a horizontal
surface into nth‐order derivatives with respect to x and y respectively. We exploited the
Fourier domain technique. This approach enhances anomalies with specific orientation and
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 69

is very useful where subtle yet important trend need to be revealed but are obscured and
complicated by trends in other directions [36]. This option enabled us to calculate gradient
in the direction of greatest rate of change and the trend. Since the angle of the output grid
tends to zero in equatorial belt then dx is the gradient to the east and dy is the gradient to
the north. In fig. 14, N‐S striking structures are clearly defined. E‐W striking features in fig.
15 are not surprising at the magnetic equator. Generally, the interpretation of magnetic
anomalies near the equatorial belt is difficult because the ambient (local) field is weak and
horizontal. N‐S striking structures are difficult to detect at the equatorial belt. Magnetic
anomalies are generated when the flux density cuts the boundary of structures and if the
structure strikes parallel with the field then in equatorial belt the flux stays within the
structure and no anomaly is generated [34]. This effect can also be generated when magnetic
field reduced to the equator (RTE) instead of reduced‐to‐the pole is carried out. In this case
the N‐S structures in RTE data are difficult to identify.
The enhancement maps show that the digitized aeromagnetic data is amenable to mathe‐
matical transformation, valuable tools for tectonic interpretation and resource exploration in
the Niger Delta basin. In fig. 14 & fig. 16 the magnetic field defines a more N‐S trending
fabric. Some of the offsets and discontinuities in the gradient maps agree with changes in
the total magnetic intensity and residual maps. This concurrence implies a major structural
contact or faults and represents offsets in the basement which have controlled sedimentation
patterns in the Niger Delta. The directional horizontal derivative maps and the analytic
signal map show clear boundaries of major magnetized zones within the basement. The
internal character and boundaries of the basement blocks and sub‐domains are also re‐
vealed. Thus, the directional horizontal derivative data and the analytic signal map clearly
demonstrate geophysical features and highlight trend directions of magnetic sources even
though the aeromagnetic data is old and is of low resolution. Most of the important geologic
features (faults and contacts) are reflected as lineaments in the magnetic data. A geologic
lineament is a linear zone of weakness in the Earth’s crust that may owe its origin to tectonic
or glacial causes and often represents geologic features such as faults, dykes, lithologic con‐
tact and structural form lines [37]. Large‐scale regional structures are revealed by low pass
filtering. Comparing the low‐passed magnetic data (fig. 17) with the total intensity data
reveals the anomalies that survived the filtering. Principal orientations of magnetic field
anomalies are revealed in the low‐passed data and made the lineaments to be more pro‐
nounced indicating that the lineaments are associated with large scale features. The orienta‐
tion of large scale features is E‐W (the direction of the major domains) while the anomalies
of short wavelength [short scale features] (fig. 18) are discordant with these major trends.

6. Discussion
The depth to magnetic basement map (fig. 13) has revealed a spatial relationship between
the paleotopographic highs and lows in the Precambrian basement and structure and
thickness anomalies in the overlying Tertiary sediments. The basement paleotopography
suggests movement in the shear/wrench fault systems that were active before, during and
70 Tectonics – Recent Advances

after sedimentation. The residual and total intensity data revealed NE‐SW trending
boundaries crossing almost the entire study area. The NE‐SW trending boundaries are shear
zones and are related to primary NE‐SW crustal block faulting that are related to the unique
position of the Niger Delta during the opening of the South Atlantic at the boundary
between the southern area of crustal divergence and the equatorial zones of crustal
translation [18].This trending magnetic anomalies represent ductile healed basement
structure of Early Proterozoic and earlier age. They predominate and obscure the desired
subtle lineaments (brittle faults) trending N‐S which were not revealed in the total intensity
and residual magnetic maps. Appropriate processing using the directional horizontal
derivative (fig. 14) and analytic signal map (fig. 16) clearly revealed the subtle anomalies.
Specifically, the negative analytic signal in fig. 16 reflects zones of low magnetization which
is a pointer to faults/fractures that are associated with possible depletion of magnetite. The
northeast‐southwest basement trends indicate possible extensions within the African
continent of the Charcot and Chain oceanic fracture zones.

Figure 14. Directional horizontal derivative data highlighting subtle N‐S structures and basement fault
blocks. White lines are inferred accommodation zones
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 71

(a)

(b)
Figure 15. Directional derivative maps revealing E‐W structures (a) data obtained by taking gradient
(dy) in north direction (b) First vertical derivative

Figure 16. Analytic signal data highlighting basement block patterns and N‐S structures
72 Tectonics – Recent Advances

Figure 17. Low pass filtered data showing the major trends of the magnetic domains of the deep base‐
ment with discordant small scale structures

Figure 18. Filtered outputs of magnetic anomalies predominantly of short wavelengths which are
discordant to the trend of the major magnetic provinces

The northwest‐southeast trend (Romanche fault zone) equivalent is as a result of block fault‐
ing that occurred along the edge of the African continent during the early stages of diver‐
gence; visible in Calabar flank which is not covered by this magnetic data. [38] recognized
the NE‐SW and ENE‐WSW trends as lineations and interpreted them as fracture zones
trends beneath the Niger Delta. [14] recognized the NE‐SW and NW‐SE trends as the mega‐
tectonic framework of the Niger Delta. A combination of the NE‐SW, E‐W and N‐S struc‐
tures from the residual and enhanced maps resulting from the shear/wrench‐fault tectonics
involving the basement created faulting, fracturing, downwarp and epeirogenic warping
along zones of basement weakness. Both horizontal and vertical movements are involved in
wrench‐fault system but the horizontal movement usually predominates. Wrench‐fault
system often appears as scissor‐type fault. The Faults in this study were recognized from a
combination of offsets and truncations of anomalies and steep gradients in the magnetic
data. The strong shearing in the study area along a wrench fault system has vertical and
horizontal displacements. The vertical displacement could be vividly seen as north‐south
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 73

striking structures in figs. 14 & 16 and horizontal displacement in the E‐W striking struc‐
tures (fig. 15). The N‐S and E‐W bounded fault blocks are secondary faults which must have
influenced stratigraphy and major tectonic elements or as shears which controlled local
features. [39] and [40] mapped family of faults with similar trends that control depositional
history of the sedimentary basin in north‐eastern Morocco and Potiguar rift basin in north‐
east Brazil respectively. The N‐S trending structures were probably induced by a combina‐
tion of differential subsidence across a fault zone and by local uplift due to wrench move‐
ments. These displacements created minibasin and arching of the basement (fig. 13) and
block faults (fig. 14). The basement block boundaries are lineaments which affected deposi‐
tion in the delta. Thus, sediment geometry in the study area is linked to subtle tectonic read‐
justment of basement blocks. These lineaments create conduits which aid the flow of fluids
and may also act as barriers. The N‐S and E‐W structures in the enhancement maps are
relatively weak structures and were created subsequent to the formation of dominant and
stronger NE‐NW and NW‐SE trending anomalies which reflects the shape of the Niger Delta
basin. Thus, the N‐S and E‐W anomalies represent the reactivated structures. These two
trends in addition to the NE‐SW trend form the three potential stress regimes responsible
for the structural architecture of the study area. In individual mega‐tectonic provinces these
three trends are the dominant trends [41].

There are three evidences for reactivation. One of the evidences of reactivation is the arching
up of the basement (fig. 13). During reactivation blocks within the basement may have
moved along faults. The second evidence is that the N‐S and E‐W structures do not correlate
with the basin shape. The third evidence is that when a thick sedimentary cover is forming
pre‐existing structures in the basement have potential to become reactivated. This have been
demonstrated for areas that are evidently tectonically stretched such as shelves or basins on
or adjacent to continental margins and in a slowly subsiding epicontinental basin, where
pre‐existing tectonic structures were reported to have been reactivated at times and
subsidence is enhanced [42]. The N‐S and E‐W structures are bounded by faults. These
faults are brittle in nature and may have developed by shear reactivation of a previously
formed weak surface in a body of rock. In the upper crust of the Earth, roughly 10km in
depth, rocks primarily undergo brittle deformation, creating a myriad of geologic structures
[43].

In this study we opine that the basement structures are identified to play major role in sed‐
iment and hydrocarbon distribution in the Niger Delta in two ways: basement relief (base‐
ment highs and lows) and basement related faults. These two factors are episodic and ap‐
pear to have controlled the trapping and migration of hydrocarbon in the Niger Delta. [8]
identified two basic types of basement control on the overlying sedimentary section in Kan‐
sas: basement topographic control and reactivated basement faults or shear zones. Actual
movement along the shear zones and lineament may be minimal but the minor change in
topographic relief of the overlying sediments is an important control on deposition [44]. The
embryonic faulted margins of the Atlantic are now the continental margins of West Africa
and are prolific oil‐producing regions. The faulted rift systems of Africa developed major
sedimentary basins along its length and generated major oil provinces in Nigeria, Central
74 Tectonics – Recent Advances

Africa and Sudan [45]. The residual map, the enhanced maps and the depth to basement
map show structural characteristics and they are used in this study as evaluation tool in this
hydrocarbon exploration setting. The shear/wrenching and the block faulting in the residual
and enhanced maps represent offsets in the basement and controlled sedimentation pat‐
terns. The development of the delta has been dependent on the balance between the rate of
sedimentation and the resulting sedimentary patterns appears to have been influenced by
the structural configuration and tectonics of the basement [18].

The depth to basement map is characterized by structural highs flanked by structural lows.
The structural low represents syncline/depocenter/subbasin. The structural high anomalies
are interpreted in this study to be the focal points for the migration of oil and gas while the
regional (lows) structural anomalies are the generating depocenters. Thus, structural high
(positive) anomalies near structural low (negative) anomalies are the preferred targets in
hydrocarbon exploration. Thus, the shear/wrench system is reflected as a series of geometri‐
cally arranged downwarp, epeirogenic uplift that may be subjected to continuous adjust‐
ment and compressional stress [44]. The uplifted blocks created the arches while
downdropped ones produced the depocenters and we therefore opine that the flanks of the
basement highs and basement lows are attractive sites for oil and gas accumulation. Oil and
gas generated in such regional lows will migrate updip, where possible onto adjacent struc‐
tural highs. Structural highs located between two adjacent basement lows offer special at‐
tractions for oil and gas migration from both sides [46]. We strongly opine that the basement
structures from the residual map, the enhanced maps and the depth to basement map are as
a result of multiple deep‐seated tensional and shear/wrench faulting within the basement
and that jostling of basement blocks have strongly influenced deposition in the Niger Delta
basin. The aftermath of the basement motion in conjunction with the impact of differences in
topographic relief in the sedimentary section during the Tertiary gave rise to the generation
of the structural lows and structural highs. Subsequent migration of hydrocarbon was aided
by fault induced by basement faulting. The basement blocks jostling beneath the Niger Delta
may have created fracture pattern that may have enhanced or reduced porosity and perme‐
amibility. Basement faults are known to have commonly influenced the distribution of hy‐
drocarbon traps and mineralization zones in sedimentary cover [47]. [48] linked oil pools in
lower productive beds of sedimentary cover to faulted zones in crystalline basement in
known platform hydrocarbon fields.

7. Conclusion
The directional horizontal derivative data, the analytic signal data and filtered maps reveals
the magnetic field lineaments and anomaly fabric that could be related to the basement
faults beneath the Niger Delta basin. E‐W striking structures are brittle faults/fractures
which are usually subtle but are well highlighted even in the total intensity data probably
because they are mineralized or associated with dykes. The N‐S structures in the study area
are due to extensional faulting in the Precambrian crystalline basement giving rise to alter‐
nating system of downwarp and epeirogenic uplift that may have pushed up the Tertiary
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 75

sediments. Hence, the sediment geometry in the Niger Delta can be correlated to subtle
tectonic readjustment of basement blocks beneath the sedimentary section. The downwarp
in this study represents syncline/depocenter/subbasin while the structural high anomalies are
interpreted to be the focal points for the migration oil and gas.

Author details
A.A. Okiwelu and I.A. Ude
Geophysics Research Unit, Department of Physics, University of Calabar, Nigeria

8. References
[1] J. Li and I. B. Morozov, Geophysical structural patterns of the crystalline basement of
the Eastern WCSB, Canadian Society of Geophysicists Convention, 2007, p.672‐675
[2] X. Li, On the use of different methods for estimating magnetic depth. The leading edge:
The meter reader (coordinated by John W. Peirce), vol. 22 (11), pp.1090‐1099, 2003
[3] M. Smith and P. Mosley, Crustal heterogeneity and basement influence on the
development of the Kenya rift, East Africa, Tectonics, vol.12 no.2, pp. 591‐606, 1993
[4] B. L. Crawford, P. G. Betts and L. Ailleres, An aeromagnetic approach to revealing
buried basement structures and their role in the Proterozoic evolution of the Wernecke
Inlie, Ykon Territory, Canada, Tectonophysics, 490, pp.28‐46, 2010.
[5] J. F. Meyer Jr., “The compilation and application of aeromagnetic data for hydrocarbon
exploration in interior Alaska”, in Geologic applications of Gravity and Magnetic: case
Histories” (Eds. Gibson. I. and Millegan. P. S.) (SEG Geophysical Reference No.8 and
AAPG Studies in Geology, No. 43), Tulsa, 1998, pp. 37‐39.
[6] P. S. Millegan and D. E. Bird, “How basement lithology changes affect magnetic
interpretation” in Geologic applications of Gravity and Magnetic: case Histories” (Eds.
Gibson. I. and Millegan. P. S.) (SEG Geophysical Reference No.8 and AAPG Studies in
Geology, No. 43), Tulsa, 1998, pp. 40‐48.
[7] H. V. Lyatsky, Magnetic and Gravity Methods in Mineral Exploration: the Value of
Well‐Rounded Geophysical Skills, Recorder (Canadian Society of Exploration
Geophysics), pp.30‐35, 2010.
[8] S. P. Gay Jr., “Basement control of selected oil and gas fields in Kansas as determined by
detailed residual aeromagnetic data”, in Geophysical Atlas of selected Oil and Gas
fields in Kansas, Kansas Geological Survey Bulletin 237, 1995, pp. 10‐16.
[9] G.E. Leblanc and W.A. Morris, “Aeromagnetics of southern Alberta within areas of
hydrocarbon accumulation”, Bulletin of Canadian Petroleum Geology, vol. 47(4), pp.
439‐454, 1999.
[10] W. Jinqi, Relationship between tectonic evolution and hydrocarbon in the foreland of
the Longmen mountains, Journal of Southeast Asian Eart Sciences, vol.13, Nos.3‐ 5, pp.
327‐336, 1996.
76 Tectonics – Recent Advances

[11] A.S. Alsharhan and M. G. Salah, Tectonic implications of diapirism on hydrocarbon


accumulation in the United Arab Emirates, Bulletin of Canadian petroleum Gelogy, vol.
45, No. 2, pp.279‐296, 1997.
[12] H. Doust and E. Omatsola, , Niger Delta, in Edwards, J. D. and Santogrossi, P.A., eds.,
Divergent/passive Margin Basins, AAPG Memoir 48: Tulsa, American Association of
Petroleum Geologists, pp. 239‐248, 1990
[13] M. L. W. Tuttle, R. R. Charpentier and M. E. Brownfield, The Niger Delta petroleum
system: Niger Delta province, Nigeria, Cameroun and Equatorial Guinea, Africa, U. S.
Geological Survey Open‐file Report‐95‐50H, 31P., 1995.
[14] J. Hospers, , Gravity field and structure of the Niger Delta, Nigeria, West Africa:
Geological Society of American Bulletin, vol. 76, pp. 407‐422, 1965
[15] C. M. Ekweozor and E.M. Daukoru, Northern delta depobelt portion of the Akata‐
Agbada petroleum system, Niger Delta, Nigeria, in, Magoon, L.B., and Dow, W.G. eds.,
The Petroleum System‐‐From Source to Trap, AAPG Memoir 60: Tulsa, American
Association of Petroleum Geologists, pp. 599‐614, 1994
[16] A. A. Avbovbo, Tertiary lithostratigraphy of Niger Delta: American Association of
Petroleum Geologists Bulletin, vol. 62, p. 295‐300, 1978
[17] K. C. Short and A.J. Stäuble, Outline of geology of Niger Delta: American Association of
Petroleum Geologists Bulletin, vol. 51, pp. 761‐779, 1965
[18] B.D. Evamy, J. Haremboure, P. Kamerling, W.A. Knaap, F. A. Molloy and P.H.
Rowlands, Hydrocarbon habitat of Tertiary Niger Delta: American Association of
Petroleum Geologists Bulletin, vol. 62, p. 277‐298, 1978.
[19] J. D. Philips, R. W. Saltus and R. L. Reynolds, Sources of magnetic anomalies over a
sedimentary Basin: Preliminary results from the coastal plain of the Arctic National
wildlife Refuge, Alaska, in Geologic applications of Gravity and Magnetic: case
Histories” (Eds. Gibson. I. and Millengan. P. S.) (SEG Geophysical Reference No.8 and
AAPG Studies in Geology, No.43), Tulsa, 1998,pp.130‐134.
[20] B. K. Bhattacharyya, Magnetic anomalies due to prism‐shaped bodies with arbitrary
polarization, Geophysics, vol. 29 no.4, pp.517‐ 531, 1964.
[21] P. R. Milligan, G. Reed, T. Meixner and D. FitzGerald, Towards automated mapping of
depth to magnetic/gravity basement‐examples using new extensions to an old method,
Australian Society of Exploration Geophysicists 17th Geophysical conference and
Exhibition, Sydney, 2004, 4p.
[22] R. I. Gibson, Magnetic frequency‐depth relationship, in Geologic applications of Gravity
and Magnetic: case Histories” (Eds. Gibson. I. and Millegan. P. S.) (SEG Geophysical
Reference No.8 and AAPG Studies in Geology, No. 43), Tulsa, 1998, pp. 59‐63.
[23] D. E. Bird, Primer: Interpreting magnetic data, American Association of Petroleum
Geologist Explorer, 18 (5), pp. 18‐21, 1997.
[24] R. I. Gibson, Magnetic susceptibility contrast versus structure, in Geologic applications
of Gravity and Magnetic: case Histories” (Eds. Gibson. I. and Millegan. P. S.) (SEG
Geophysical Reference No.8 and AAPG Studies in Geology, No. 43), Tulsa, 1998, pp. 79‐
81.
3D Modelling and Basement Tectonics of the Niger Delta Basin from Aeromagnetic Data 77

[25] J. M. Jacques, M.E. Parsons, A.D. Price and D. M. Swhwartz, Draw upon recent work to
provide evidence as to why gravity and magnetic survey data can still provide vital
geological clues for oil and gas exploration, First break, vol. 21, 2003, pp. 57‐62.
[26] S. Williams, T. D. Fairhead and G. Flanagan, Realistic models of basement topography
for depth to magnetic basement testing, SEG International Exposition and 72 annual
meeting, Utah, 2002, 4p.
[27] M. Alexander, J. C. Pratsch, C. Prieto, Under the northern Gulf basin: basement depths
and trends, Abstract of paper presented at the 1998 SEG 68 annual meeting, New
Oleans, LA, 1998, 3P.
[28] J. Li and I. Morozov, Aeromagnetic mapping of the Williston basin basement, 2006
CSPG‐CSEG‐CWLS Convention, 2006, pp. 65‐68.
[29] H. V. Lyatsky, D.I. Pana and M. Grobe, Basement structure in Central and Southern
Alberta: Insights from Gravity and Magnetic maps, Alberta Energy and Utility
Board/Alberta Gelogical Survey (EUB/AGS) special report72, 83p., 2005.
[30] G.R. J. Cooper and D.R. Cowan, Filtering using variable order vertical derivatives,
Computers and Geosciences, vol. 30, pp.455‐459, 2004.
[31] L. Cordell and V. J. S. Grauch, Mapping basement magnetization zones from
aeromagnetic data in the San Juan basin, New Mexico [Astract]. Program, SEG, 1982
Annual Meeting, 246‐247
[32] L. Cordell and V. J. S. Grauch, Mapping basement magnetization zones from
aeromagnetic data in the San Juan basin, New Mexico, in Hinze, W. J. ed., The utility of
regional gravity and magnetic anomaly maps, Society of Exploration Geophysicists,
pp.181‐197, 1985.
[33] W. R. Roest, J. Verhoef and M. Pilkington, Magnetic interpretation using the 3‐D
analytic signal, Geophysics, vol. 57 pp. 116‐125, 1992.
[34] GETECH GROUP. (2007). Advanced Processing and Interpretation of Gravity and
Magnetic Data. 27. Available:
http://www.getech.com/services/advancedprocessingandinterpretation.pdf
[35] R. I. Blakely, Potential theory in gravity and magnetic applications. Cambridge
University press, Cambidge, 1996, 441p.
[36] R. C. Brodie. (2002). Geophysical and remote sensing methods for regolith exploration,
CECLEME open file report 144, pp.33‐45. Available:
http://www.crcleme.org.au/Pubs/../OfR%20144/06Magnetics.Pdf
[37] M. Lee, W. Morris, J. Harris and G. Leblanc, An automatic network‐extraction algorithm
applied to magnetic survey data for the identification and extraction of geologic
lineaments, The Leading Edge, vol.21, No. 1, pp. 26‐31, 2012.
[38] O. O. Babalola and M. Gipson, Aeromagnetic anomalies and discordant lineations
beneath the Niger Delta: Implications for new fracture zones and multiple sea‐floor
spreading directions in the “Meso‐ Atlantic” Gulf of Guinea Cul‐De‐SAC, vol.18 (6),
pp.1107‐1110, 1991.
[39] R. EI Gout, D. Khattach, M. R. Houari, O. Kaufmann and H. Aqil, Main structural
lineaments of north‐eastern Morocco derived from gravity and aeromagnetic data,
Journal of African Earth Sciences, vol.58, pp. 255‐271, 2010.
78 Tectonics – Recent Advances

[40] D. L. De Castro, Gravity and magnetic joint modeling of the Potiguar rift Basin (NE
Brazil): Basement control during Neocomian extension and deformation, Journal of
South American Earth Sciences, vol.31 (2‐3), pp. 186‐198, 2011.
[41] J. Affleck, Magnetic anomaly trend and spacing patterns, Geophysics, vol. 28 (3),
pp.379‐395, 1963
[42] A. Wetzel, R. Alenbach, V. Allia, Reactivated basement structures affecting the
sedimentary facies in a tectonically “quiescent” epicontinental basin: An example from
NW Switzerland, sedimentary Gelogy, 157, pp. 153‐172, 2003.
[43] B. A. V. Pluijm and S. Marshak, Earth structure: An introduction to structural Geology
and tectonics (2n Ed.), W. W. Norton and company, New York, 2004, pp. 1‐ 476.
[44] D. L. Brown and D. L. Brown, Wrench‐Style Deformation and Paleostructural influence
on Sedimentation in and around a Cratonic Basin, Rocky Mountain Association of
Geologists Symposium, 1987, p. 58‐70.
[45] J. D. Fairhead, Regional tectonics and basin formation: The role of potential field studies
in Phanerozoic Regional Geology of the world, Elsevier, 2012, pp. 330‐341.
[46] C. Prieto and C. Pratsch, Gulf of Mexico study links deep basement structure to oil
fields: Demonstrating deep structural control, Intergrated Geophysics Corporation
Footnotes series on Interpretation, 2000, 4P.
[47] H.V. Lyatsky, G. M. Friedman and V. B. Lyatsky, Principles of practical tectonic
analysis of cratonic regions in Lecture notes in Earth Sciences, Springer‐Verlag, vol. 84,
1999, 369p.
[48] T. N. Plotnikova, Nonconventional hydrocarbon targets in the crystalline basement and
the problem of the recent replenishment of hydrocarbon reserves, Journal of
Geochemical Exploration, vol. 89, pp. 335‐338, 2006.
Chapter 4

Seismic Paleo-Geomorphic System of


the Extensional Province of the Niger Delta:
An Example of the Okari Field

Muslim B. Aminu and Moses O. Olorunniwo

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/48197

1. Introduction
Hydrocarbon exploration and exploitation requires that the spatial and depth distribution
and interplay of factors favorable to commercial hydrocarbon accumulation are thoroughly
appreciated. These factors include the distribution of source rock, reservoir rock, and migra‐
tion pathways, the fidelity of sealing mechanisms, and timing, the temporal relatioinship
between reservoir rock formation and the expulsion of hydrocarbons from the source rock.
The distribution of these elements of the petroleum system is a result of the tectonic history
and fill processes occuring in a basin.
As hydrocarbon exploration moves into geologically and economically more challenging
environments, such as deeper subsurface locations, deepwater regions, sub‐ice in the Artic,
and into geologically and stratigraphyically more complex environments, the costs of explo‐
ration is bound to be on the rise and the risks associated with field development greater.
Continued success in the hunt for Oil and Gas reserves therefore, depends upon a thorough
understanding of the subsurface geology of exploration fields, the ability to accurately pre‐
dict and delineate the spatial and depth distribution of subsurface geologic facies (source
rock, reservoir rock and seal) and the ability to discriminate the fluids saturating reservoirs
(oil, gas or brine) and possibly quantifying such.
Seismic geomorphology which is the integration of three‐dimensional (3‐D) seismic data
with analytical techniques typically used in the study of Earth land forms, is growing in
importance as a tool in understanding the stratigraphy of both mature fields and underde‐
veloped basins (Dunlap et al, 2010). Seismic geomorphologic analysis allows geoscientists to
correctly identify, interpret, and predict the distribution of possible hydrocarbon‐bearing
deposits. It helps to unravel the major factors which have contributed to and or continue to
control the evolution of a basin or field.

© 2012 Aminu and Olorunniwo, licensee InTech. This is an open access chapter distributed under the terms
of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
80 Tectonics – Recent Advances

In this study, we implement a series of conventional interpretation procedures and


advanced imaging techniques to unravel the paleo‐geomorphology, tectonic history and fill
architecture of a hydrocarbon target within the Okari Oil Field.

2. The Niger delta


The Niger delta is situated in the Gulf of Guinea on the west coast of Central Africa (Figure.
1). It is located in the southern part of Nigeria between latitudes 40 00’N and 60 00’N and
longitudes 30 00’N and 90 00’N. It is bounded in the south by the Gulf of Guinea (or the
4000 m bathymetric contour) and in the North by older (Cretaceous) tectonic elements
which include the Anambra Basin, Abakaliki uplift and the Afikpo syncline. In the east and
west respectivily, the Cameroon volcanic line and the Dahomey Basin mark the bounds of
the Delta, Figure 1. The Cenozoic Niger Delta is situated at the intersection of the Benue
trough and the South Atlantic Ocean where a triple junction developed during the
separation of South America from Africa (Burke, 1972; Whiteman, 1982). The delta is
considered one of the most prolific hydrocarbon provinces in the world, and recent giant oil
discoveries in the deep‐water areas suggest that this region will remain a focus of
exploration activities (Corredor et al, 2005).

Figure 1. Map of Niger Delta showing province outline (maximum petroleum system); bounding
structural features; minimum petroleum system as defined by oil and gas field center points (Modified
from Petroconsultants, 1996, as cited in Tuttle et al, 1999).
Seismic Paleo-Geomorphic System of the Extensional Province
of the Niger Delta: An Example of the Okari Field 81

Furthermore, the Niger delta is one of the world’s largest deltas, with a sub‐aerial acreage of
about 75,000 km2. It is composed of an overall regressive clastic sequence, which reaches a
maximum thickness of 30,000 to 40,000 ft (9000 to 12000 m). The development of the delta
has been dependent on the balance between the rate of sedimentation and the rate of
subsidence (Doust and Omatsola, 1990). This balance and the resulting sedimentary patterns
were constrained by the structural configuration and tectonics of the basement (Evamy et al,
1978). Important influences on sedimentary rates have included eustatic sea‐level changes
and climatic variations in the hinterlands. Subsidence has been controlled largely by initial
basement morphology and differential sediment loading of unstable shales (Doust and
Omatsola, 1990).

Figure 2. Generalized dip section of the Niger Delta showing the structural provinves of the Delta. The
slips resulting from the collapse of shelf sediments in the extentional province is diverted onto thrust
ramps through detachemnt structures in the basal Akata formation and comsumed by contractional
folds in the deep‐water fold and thrust belts (Adapted from Whiteman, 1982).

The Delta has built out over the collapsed continental margin, and its core is located above
the collapsed continental margin at the site of the triple junction formed during the middle
Cretaceous. The main sediment supply has been provided by an extensive drainage system,
which in its lower reaches follows two failed rift arms, the Benue and Bida basins. Sediment
input generally has been continuous since the Late Cretaceous, but the regressive record has
been interrupted by episodic transgressions, some of considerable extent (Weber and
Daukoru, 1975). The Niger delta has prograded into the Gulf of Guinea at a steadily
increasing rate in response to the evolving drainage area, basement subsidence and eustatic
sea level changes. Initially, the delta prograded over extensionally thinned and collapsed
continental crust of the West African margin (Figure 2), as far as the triple junction, filling in
the basement graben‐and‐horst topography (Murat, 1970). During the middle and late
82 Tectonics – Recent Advances

Eocene, sedimentary rocks became increasingly sandy, marking the onset of a general
regression of the deltaic deposition.

Of dominant importance in the development of the delta, as in similar settings elsewhere in


the world has been the influence of synsedimentary listric normal faults. These have been
forming at least since the Paleocene and define sites of locally increased subsidence and
sedimentation. They lie sub‐parallel to the paleo‐coastline and are presumed to sole out at
relatively shallow depths within marine shale sequences. A number of major basin‐wide
growth fault zones define depositional realms at succeeding periods of delta history (Doust
and Omatsola, 1990). In the paralic interval, growth fault associated rollover structures
trapped hydrocarbons. Faults in general play an important role in the hydrocarbon
distribution. Growth faults may even function as hydrocarbon migration paths from the
overpressured marine clays. There is an intimate relationship between structure and
stratigraphy. They both depend on the interplay between rates of sediment supply and
subsidence (Evamy et al., 1978).

2.1. Structure
The Niger Delta is regarded as a classical shale tectonics province (Wu and Bally, 2000). The
instability and constant motion of the shales in response to the weight of the advancing
sediment wedge has resulted in the development of different structural styles in different
belts of the Delta.

The Delta has thus been subdivided into five zones. Corredor et al. (2005) identified these
structural zones (Figure 3) as [1] an extensional province beneath the continental shelf that is
characterized by both basinward‐dipping (Roho‐type) and counter‐regional growth normal
faults and associated rollovers and depocenters; [2] a mud diapir zone located beneath the
upper continental slope, which is characterized by passive, active, and reactive mud diapirs,
including shale ridges and massifs, shale overhangs, vertical mud diapirs that form mud
volcanoes at the seafloor, and interdiapir depocenters; [(3] the inner fold and thrust belt,
which is characterized by basinward verging thrust faults (typically imbricated) and associ‐
ated folds, including some detachment folds; [4] a transitional detachment fold zone beneath
the lower continental slope that is characterized by large areas of little or no deformation
interspersed with large, broad detachment folds above structurally thickened Akata For‐
mation; and [5] the outer fold and thrust belt characterized by both basinward‐ and hinter‐
land‐verging thrust faults and associated folds. The inner and outer fold and thrust belts are
most evident in the bathymetry, where ridges represent the crests of fault‐related folds, and
low regions correspond to piggyback basins formed above the backlimbs of fault imbricates.
The inner fold and thrust belt extends in an arcuate path across the center of the offshore
delta, whereas the outer fold and thrust belt consists of northern and southern sections that
define two outboard lobes of the delta (Corredor et al, 2005). These two lobes, and their
associated fold belts, are separated by a major rise in the basement topography that corre‐
sponds to the northern culmination of the Charcot fracture zone (Figure 3). The break be‐
tween the northern and southern sections of the outer fold and
Seismic Paleo-Geomorphic System of the Extensional Province
of the Niger Delta: An Example of the Okari Field 83

Figure 3. Bathymetric Sea‐floor image of the Niger Delta obtained from a dense grid of two‐
dimensional seismic reflection profiles and the global bathymetric database (Smith and Sandwell, 1997,
as cited in Corredor et al, 2005), showing the main structural provinces.

thrust belt are the results of thrust sheets being stacked in a narrow zone above and behind this
major basement uplift (Connors et al., 1998; Wu and Bally, 2000). Deformation across these struc‐
tural provinces is active today, resulting in pronounced bathymetry expressions of structures that
are not buried by recent sediments, as illustrated in Figure 3 (Corredor et al, 2005).
84 Tectonics – Recent Advances

2.2. Stratigraphy
The lithostratigraphy of the Niger Delta Basin (Figure 4) consists of three main rock‐
stratigraphic units of Cretaceous to Holocene origin (Short and Stauble, 1967; Frankl and
Cordry, 1967; Avbovbo, 1978). These units represent the prograding depositional
environments (Corredor et al, 2005).

Figure 4. Stratigraphic column showing the three Formations of the Niger Delta, the Marine Akata
shale, the paralic Agbada formation and the continental Benin sandstone. (Modified from Doust and
Omatsola, 1990).

At the base of the system is the Akata Formation, a sequence of planktonic foraminifera rich
undercompacted transgressive Paleocene‐to‐Holocene marine shales, clays, and silt. The
Akata ranges in thickness from 2000 m (6600 ft) at the most distal part of the delta to 7000 m
(23,000 ft.) thick beneath the continental shelf (Doust and Omatsola, 1990). In the outer fold
Seismic Paleo-Geomorphic System of the Extensional Province
of the Niger Delta: An Example of the Okari Field 85

and thrust belts, the Akata Formation reaches 5000 m (16,400 ft.) thick as a result of structural
repetitions by thrust ramps and in the core of large detachment anticlines (Bilotti and Shaw,
2005). The Akata shales are typically undercompacted and frequently move either down‐
ward or laterally along the continental shelf, or in an upward diapiric motion along growth
faults, in response to the lithostatic pressure of overlying sediment (Ekweozor and Daukoru,
1984). Lateral movements along the continental shelf result in thrusting in more outer bound
regions of the Delta. The Akata Formation is conformably overlain by a paralic sequence of
alternating Lower Eocene to Pleistocene sandstones and sand bodies with shale intercala‐
tions, which is known as the Agbada Formation (Doust and Omatsola, 1990). The Agbada
Formation is more than 3500 m (11,500 ft.) thick and represents the actual deltaic portion of
the sequence. This clastic sequence was accumulated in delta‐front, delta‐topset, and fluvio‐
deltaic environments. The Agbada formation is highly faulted with assays of roll‐over exten‐
sion induced growth faults, compensation listric faults and high angle thrust faults depend‐
ing on which belt of the Niger Delta you are in. Channel and basin‐floor fan deposits in the
Agbada Formation form the primary reservoirs in the Niger Delta. The Agbada Formation is
in turn covered by the Benin Formation which consists of late Eocene to Holocene massive,
porous, and unconsolidated freshwater bearing continental deposits, including alluvial and
upper coastal‐plain deposits that are up to 2000 m (6600 ft) thick (Avbovbo, 1978).

In the outer thrust belt, the Akata Formation lies upon oceanic crust and is diachronously
overlain by deep‐water channel complexes, debris flows, and shales of the Agbada For‐
mation, which represent the sedimentary overburden of the deltaic succession (Maloney et
al., 2010). The Benin Formation is absent in this belt (Maloney et. al., 2010). The Akata For‐
mation and the shaly intercalations of the Agbada Formation are believed to be the source
rocks of the Niger Delta while the paralic Agbada Formation constitutes the reservoir rock. It
has also been suggested (Corredor et al, 2005) that the Akata Formation could contain some
turbidite sands which could represent potential reservoirs in deep‐water environments.

2.3. The Extensional Province


The development of the Niger Delta was initiated by rifting on the onshore portion of the
West African Shield (Merki, 1972). The rifting is associated with the opening of the South
Atlantic Ocean which began in the Late Jurassic and spanned through the Middle
Cretaceous (Lehner and De Ruiter, 1977, as cited in Tuttle et al, 1999). Within the Delta,
rifting had ceased by the Late Cretaceous. Thereafter, gravity induced tectonics became the
primary deformational process and represent the key factor determining the structural style
of the Extensional Province of the Niger Delta. Shale mobility induced internal deformation
and occurred in response to two processes (Kulke, 1995, as cited in Tuttle et al, 1999). First,
shale diapirs formed from loading of poorly compacted, over‐pressured, prodelta and delta‐
slope shales of the basal Akata Formation by the advancing higher density delta‐front paralic
wedge of the Agbada Formation, and second, slope instability resulting from a lack of lateral,
basinward, support for the under‐compacted delta‐slope Akata Formation. These
movements initiated syn‐sedimentary growth faults which compartmentalized the belts into
86 Tectonics – Recent Advances

fault blocks. As faulting advanced fault blocks rotated in fault planes and subsurface
horizons formed roll‐over anticlines. The resulting seaward extension of subsurface strata
was often accommodated by landward dipping compensational listric faults. Other
structures which formed include shale diapirs, collapsed growth fault crests, back‐to‐back
features, and steeply dipping, closely spaced flank faults (Evamy et al, 1978). The Benin
Formation was deposited only after the cessation of gravity induced tectonics and is
unaffected by faulting.

3. The Okari Field example: A gravitational shale tectonics field


At the time of this study, the Okari Oil Field had six wells, four of the wells are vertical and
the other two deviated. Though the wells had varying suites of logs, they generally
possessed lithologic logs (Gamma Ray), porosity logs (sonic, density and neutron),
resistivity logs (LLD and MSFL) and checkshot surveys. Available seismic data consisted of
401 inlines and 221 crosslines and a total of 88,621 post‐stack seismic traces with a record
length of 5 seconds, sampled at 4 ms interval and covering an area of about 56.5 sq km.

Three of the wells, B1, B4 and B5 had encountered hydrocarbon in a stack of three horizons
in a single roll‐over anticlinal structure created by the rotation of an east‐west trending fault
block into the plane of a major structure building fault. The other three wells had turned out
dry at the target horizon encountering water saturated sands and shales. It was understood
that the variability of subsurface facies was more intricate than initially presumed. This is a
known situation in the Niger Delta, the fast rate of sedimentation along the shelf edge and
strong wave action are believed to have ensured rapid spacial variation in subsurface facies
distribution patterns (Aminu and Olorunniwo, 2011). There was therefore a need to fully
image the distribution of reservoir facies within the field to assist with further development
and production.

3.1. Methodology
Our study of the field involved reservoir evaluation using well‐logs, hydrocarbon saturated
zones were identified and their respective petrophysical parameters were computed. Three
horizons, H1, H2 and H3, were found to be petroliferous. However, only the deepest
horizon, H3, was considered prospective. It is a fairly shaly sand with average porosity of
27% and hydroarbon saturation of 77%. Next we created a tectono‐stratigraphic model of
the field by mapping all three horizons and the field’s cascade of intersecting faults (both
growth and compensational listric faults). In this way we mapped a faulted anticlinal
closure against a major structure building fault in the middle of the survey area. Thereafter,
we used tri‐variate crossplots to study rock‐physics relations within the field and to provide
a basis for calibration and interpretation of subsurface facies prediction. We then computed
several seismic attributes, some of which include the instantanous amplitude, frequency and
phase, 1st and 2nd derivative of the seismic trace, and various geometric seismic attributes.
We experienced considerable difficulty deriving correlation between computed attributes
and well information, and at obtaining geologically meaningful maps of subsurface
Seismic Paleo-Geomorphic System of the Extensional Province
of the Niger Delta: An Example of the Okari Field 87

property distribution. This may be due to one of two factors, (1) the seismic attribute ‐ log
relationships are often non‐linear, and (2) the Niger Delta is a soft clastic basin with under‐
compacted subsurface formations, the seismic amplitude response in such cases are often
weak and not well correlated to geology (Chopra and Marfurt, 2007). This necessitated that
we impliment multi‐attributes transforms to predict subsurface facies from the seismic
response. We elected to use the multi‐linear resgression transform and an artifcial neural
network. The multi‐linear transform attempts to derive a multi‐variate linear relationship
between seismic attributes and log derived petrophysical properties. If for a single attribute,
the relationship between the target log L and the attribute A is written as

L  a  bA (1)

Where the coefficients a and b are obtained by minimizing the mean‐squared prediction
error:

1 N
  Li  a  bAi 
2
E2  (2)
N i 1

According to Hampson et al, 2001, the multivariate case for a single attribute is written as:

L  t   w0  w1 A1  t   w2 A2  t   ...  wm Am  t  (3)

Where L(t) represents the log sample, A(t) are the corresponding seismic attributes and w0 …
wm are the coefficients of the linear transform. If the log consists of n samples then the equa‐
tion is written as:

L1  wo  w1 A11  w2 A21  ...  wm Am1

L2  wo  w1 A12  w2 A22  ...  wm Am 2

Ln  wo  w1 A1n  w2 A2 n  ...  wm Amn (4)

Where Aij is the jth sample of the ith attribute (Hampson et al, 2001).

The neural network on the other hand is able to process multi‐parameter and multi‐
dimensional data sets in a non‐linear fashion via a learning process. The task of the neural
network is to perform a non‐linear mapping k of a set of variables onto a target property,

k : Amn  Lm (5)

In the case of log prediction from seismic data, the network reads a set of input predicting
variables, a matrix Amn of seismic attributes and a set of desired output results, a vector Lm
of log responses, computes its own output a vector Lm‘ of pseudo‐log responses, computes
the error vector Em(Lm ‐ Lm‘) and adjust its weights, a matrix Wij using a gradient descent
algorithm in a direction as to minimize the disparity between the network‘s output and
88 Tectonics – Recent Advances

the desired target output. The neural network approximates a mapping function such
that:

L m  k (W ij , Amn )  E m (6)

where k  k . The error given as:

1 (7)
E ( W ij ) 
2
 m ( L m  L m ) 2

and can be minimized using the back‐propagation algorithm (Callan 1999, Liu and Liu,
1998). In practice a momentum factor is usually added to the gradient decent algorithm to
speed up the search for an optimal solution and to prevent the network from getting caught‐
up in a local minimal. This process proceeds iteratively until the error bound is within pre‐
set acceptable limits. At this point the network is presumed to have trained adequately.
Training is achieved with seismic attribute data and well information at well locations. The
network is next implemented to predict pseudo‐well‐log responses throughout the field by
being presented with seismic attribute data from beyond the wells.

3.2. Discussion
3.2.1. Structure
A total of 12 normal faults were mapped in the study area and are annotated F1 to F12 (Fig‐
ure 5). Seven of the faults dip southward towards the sea while five are listric compensa‐
tions for extentions within the fault blocks and dip landwards. These fault are gravity in‐
duced and are the result of the downwad and lateral response of the over‐pressured mobile
shales of the basal Akata Formation. Two of the faults F1 and F2 are major structure building
growth faults which cut across the entire survey area and compartmentalize it into three
east‐west trending fault blocks which are downthrown and dip in the south‐southwestward
seaward direction (Figure 6). These faults appear to sole out as presumed (Corredor et al,
2005) at relatively shallow depths near the top of basal shale sequences. Fault blocks were
created synchronously with sedimentation, this is seen from the relative thickening of sub‐
surface formations on the downthrown side of the faults. Basal shales displaced by these
fault blocks are believed to flow laterally along the continental shelve and the slips thereof
tranferred through detachment structures in the basal Akata Formation onto thrust ramps
and consumed by contractional folds in the deep‐water fold and thrust belts (Bilotti and
Shaw, 2005).

The back‐to‐back alternations of seaward and landward dipping normal faults resulted in
the formation of horst and graben structures partcularly in the western end of the survey
area (Figure 6). These horst and graben structures possibly resulted in the creation of mini
depo‐centres and are probably key factors controlling drainage and depositional patterns in
the field.
Seismic Paleo-Geomorphic System of the Extensional Province
of the Niger Delta: An Example of the Okari Field 89

Contour closures against the major structure building F1 fault represent a faulted anticlinal
structure juxtaposed against the fault. These structures form the majority of structural traps
in the Niger Delta. Exploration wells in this field targeted this closure. Fault throw on the
eastern arm of the anticlinal structure is much greater than on the western arm (Figure 7).
This could indicate that there has been a relative clockwise rotation of the horizon along a
crestal hinge line over the anticline lying orthogonal to the F1 fault. This rotation appears to
have added a sense of south easterly dip to the orientation of the fault block. This sense of
dip could play a key role in determining drainaige and sediment deposition within the field.

Figure 5. Interpreted seismic lines from the Okari field along with the litho‐section. The target horizon
in this study the H3 is indicated in green. Growth faults affect only the paralic Agbada Fm. and com‐
partmentalize the field into east‐west trending blocks downthrown to the south. Horizon thickening on
the downthrown blocks indicates that faulting was synchronous with sediment deposition. The Akata
shales respond (as indicated by yellow arrow) by downward and seaward flowage due to the litho‐
static pressure of the advancing paralic wedge. The Benin Formation were deposited after the ceasation
of faulting and thus are not affected by faulting.
90 Tectonics – Recent Advances

Figure 6. Horizon time map of the target H3 horizon. Overall dip is in the south westerly direction with
fault blocks trending roughly in the east‐west direction. Seaward tensional stress is accomodated by
landward dipping listric faulting to the west of the field thereby creating a graben structure which
apparently played a key role in defining the paleo‐drainage route on the horizon. Exploration wells
targeted the contour closure against the major F1 fault (red).

Figure 7. Annotated 3D view of the H3 horizon. Fault throw (indicated in blue arrows) on the eastern
arm of the major F1 fault is much greater than on the western arm. This appears to have been due to
possible rotation of the fault block along a crestal line (yellow dashed line) over the anticlinal structure
thereby giving the fault block a south easterly dip. A fault controlled channel is seen to trend in the
south westerly direction and terminates against the F1 fault. The fault induced graben structure has a
fairly sinuous configuration.
Seismic Paleo-Geomorphic System of the Extensional Province
of the Niger Delta: An Example of the Okari Field 91

On the upthrown block of the F1 fault, a channel‐like depression trending in the northeast‐
southwest direction appears to terminate against the F1 fault. This depression could have
been a major path through which continentally derived detrites was supplied to the region
of the anticlinal structure on the downthrown block of the F1 fault.

3.2.2. Cross Plots


In cross‐plots of p‐wave impedance vs. porosity (Figure 8), we observed that p‐wave
impedance decreases with increase in porosity (θ), with lower‐porosity sands generally
having higher impedance values, likely a reflection of tighter grain packing. Porosity also
increases with increase in shale content (Vsh), with clean sands being associated with lower
porosities than shaly units of corresponding densities. The shales of the Niger Delta are
known to be under‐compacted and over‐pressured, and the platy nature of the shale
minerals ensures they have high though ineffective porosities. The litho‐facies are almost
separable in two (along the blue line). Furthermore, high hydrocarbon saturation appears
associated with lower‐porosity dirty sands, with porosity in the range 23 to 27%. The shales
and water‐bearing sands of the Okari Field appear, generally, to have higher porosities than
hydrocarbon‐saturated sands.

Figure 8. P‐wave impedance vs. Porosity crossplots, (A) color coded in shale volume. Impedance de‐
creases with porosity increase and shaly units generally possess high porosities. Between shales and
sand facies, separation is almost possible along the dashed blue line, and (B) color coded in water satu‐
ration. Polygons represent subsuface facies; black for hydrocaron sands, light blue for water saturated
sands and brown for shales. Hydrocarbon saturation is associated with lower porosity shaly sand facies.

These lower‐porosity shaly reservoir sand units possibly benefitted from reduced hydrocar‐
bon mobility due to the shale infilling of their pore spaces, while cleaner sands in the field
may have lost their hydrocarbon saturations to invading brine as a result of their higher
porosities and permeabilities, and consequent greater hydrocarbon mobility. Three distinct
litho‐facies are fairly well discriminated in this field; hydrocarbon sands associated with low
relative porosities (23 ‐ 27%) and high p‐wave impedance, water‐saturated sands associated
with medium to high porosities (25 ‐ 35%) and low impedance values, and shales associated
92 Tectonics – Recent Advances

with generally higher relative though ineffective porosities (30 ‐ 42%) and the entire spread
of possible p‐wave impedance values (from low through medium to high) in the field.

3.2.3. Predicted litho‐facies maps


3.2.3.1. Predicted Shale Volume Maps:
Six seismic attributes; original trace, integral of the trace, average of the trace, average of
instantaneous amplitude, and 1st derivative and 2nd derivative of the trace were multi‐
linearly regressed against the target well logs (shale volume and porosity) in the vicinity
of wells penetrating the H3 horizon. These attributes were selected for two major rea‐
sons; firstly, forward entry statistical regression of attributes against target well logs
(porosity and shale volume) had turned up mainly amplitude‐related attributes. Second‐
ly, reviewed literature suggested that amplitude related seismic attributes have robust
physical relationships to lithology and porosity (Banchs and Michelena, 2002; Dorrington
and Link, 2004; Calderon and Castagna, 2005). The resulting multi‐linear transforms
were used to predict shale volume and porosity distributions for H3 horizon. Employing
the same set of seismic attributes used for the multi‐linear regression, we implemented a
total of 50 simulations for shale volume prediction over the H3 horizon using a Multi‐
Layered Feed‐forward Back‐propagation Neural Network (MFLN). Thereafter, we se‐
lected the network which produced the most geologically valid map and strictly honored
well information.
The multi‐linear regression of well log derived shale volume against the earlier mentioned
six seismic attributes (statistics not shown here) turned up a correlation coefficient R of 0.23.
The low statistical correlation underscores the difficulty in predicting lithology using linear
regression algorithms. Mathematically, the chances of correctly imaging lithologic distribu‐
tion over the horizon using multi‐linear regression of these attributes will result only in a
23% correlation or similarity with actual shale volume distribution at the target horizon.
Figure 9, is the result of this multi‐linear attribute transform for the H3 horizon. Our optimal
neural network achieved 0.59 and 0.67 as validation and training correlation coefficients
respectively. The validation coefficient is a significant improvement over the multivariate
statistical regression coefficients (0.23). As neural networks are able to implement non‐linear
mapping, they are better suited to codifying seismic‐well‐log relationships. The neural net‐
work predicted lithology map of the H3 horizon is shown in Figure 10.
The two predicted shale volume maps have stricking semblance in the horizon lithologic
distribution and reservoir geometry and internal architecture they reveal. This engenders
confidence in the robustness of the distribution so imaged. However, the neural network
predicted shale volume distribution possesses greater detail compared to the multi‐linear
regression map. The maps show intricate details of lithologic variation both in the region of
well control (reservoir region) and beyond. They reveal the subsurface paleo‐drainage
system and geologic situation of the field, one typical of the Niger Delta with several me‐
andering sand‐filled channels and sand bars especially in the southwest, north, and west of
the field. The paleo‐channels generally cross the study area in an east‐west direction and
possess shale volumes in the range of 15–20%. This directional trend is most possibly
Seismic Paleo-Geomorphic System of the Extensional Province
of the Niger Delta: An Example of the Okari Field 93

SSC

D2
D1

Vsh%

Figure 9. Multi‐linear regression predicted shale volume map over the top of H3 horizon. The sediment
supply channel is a North‐South trending sinuous channel SSC which terminates against the F1 fault.In
crossing the fault the sediments fan out while plunging into the rapidly emerging accomadation. D1
was possibly the initial drainage channel for the reservoir region but ceased altogether and the sand
bodies became detached due to the onset of faulting induced channelization which created another path
D2 in the western region.

Figure 10. Predicted shale volume map over the top of H3 horizon. The neural network prediction has
high semblance with the result from regression but it provides greater detail in horizon property distri‐
bution and the reservoir geometry and internal architecture it delineates. The western D2 channel is
seen to possess greater sinuosity than on the regression map. Lateral shale seals are predicted to the
south of the reservoir.
94 Tectonics – Recent Advances

in response to dips on either sides of the anticlinal structure, that on the eastern arm re‐
sulting from possible clockwise rotation of the horizon in the plane of the F1 fault around
an axis along the crest of the anticline and roughly orthogonal to the fault plane, while the
accentuated dip on the western arm is largely the result of a fault induced graben struc‐
ture. Channels in the western region demonstrate good lateral continuity while those in
the eastern region are less continuous and appear separated from the reservoir. This is
due in part to structural controls in form of grabens which guide channel flow in the
western region. Channel sand bodies on the eastern arm were possibly separated from the
reservoir as a result of horizon flexing during the formation of the roll‐over anticline
which hosts the known reservoir. It could also have been the result of the creation of con‐
siderably steeper dips on a portion of the western arm of the anticlinal structure due to
the initiation of faulting controlled structural constraints (graben and half graben struc‐
tures). The channel through the eastern arm D1 was possibly the initial drainage path for
the reservoir region, this was temporal and might have ceased entirely with the onset of
extension induced listric faulting on the western arm of the anticline. The faulting in the
western region apparently defined an alternative drainage D2 path for this anticline. A
high‐quality sand‐filled paleo‐channel SSC, trending roughly north‐south on the up‐
thrown block of the F1 fault and possibly turning northwest above the reservoir region is
imaged on both maps. This channel appears to have been the source of sediment fill for
the reservoir region. In crossing the growth fault from north to south, the sediment load
transported by the channel fanned out while plunging into the transient depo‐center cre‐
ated by the active faulting.

The delineated reservoir is an oblate fan which coincides with the position of the struc‐
tural closure on Figure 6. The sand distribution within the reservoir region indicates that
clean reservoir sands are interspersed with shaly sand units. Initial deposits were possi‐
bly reowrked by strong wave acton. These shaly‐sand units serve to compartmentalize
the reservoir and may act as flow barriers to hydrocarbons present and preclude com‐
munication between the segments of the reservior. Extensive lateral shales are predicted
to the south and southeast of the reservoir. Coupled with the growth fault north of the
reservoir, this possibly indicates reasonable lateral sealing mechanisms. Hydrocarbon‐
bearing wells (B1, B4 and B5) penetrated clean to fairly shaly sands with shale volume in
the range 15–20% at the target horizon while wells B2, B3, and B6, penetrated shaly
sands/shales. This result is in keeping with well information for the H3 horizon. It is
nearly possible to trace out the outline of the major structure building F1 fault from the
maps.

3.2.3.2. Predicted Porosity Maps

The multi‐linear regression of well log derived porosity against the same seismic
attributes set used for shale volume prediction (statistics also not shown) turned up a
correlation coefficient R of 0.43. The correlation coefficient is considerably higher than
that obtained for shale volume. This implies that it is easier to predict porosity from
amplitude based seismic attributes using multi‐linear regression than predicting
lithology. Lithology index parameters such as grain shape index, minerology and
Seismic Paleo-Geomorphic System of the Extensional Province
of the Niger Delta: An Example of the Okari Field 95

platiness, to mention but a few, have more intricate relationships to the seismic amplitude
response than coresponding determinants of porosity such as cementation and
compaction. The predicted porosity map is shown in Figure 11. A neural network with the
same network architecture and configuration as that used for predicting shale volume
was trained using the same set of attributes viz. sample number, original trace, integral of
trace, average of trace, average instantaneous amplitude, 1st derivative and 2nd derivative
to predict porosity over H3 horizon. Twenty‐five (25) simulations of the neural network
were run and the networks satisfying the two earlier mentioned criteria were selected.
The most geologically valid map, which honored well information and correlated well
with log information, was selected.

Figure 12 is the predicted porosity map over the H3 horizon. The validation and training
correlations for this network were 0.60 and 0.78 respectively. Though the training correla‐
tion in much higher than that for shale volume prediction, the validation correlation is hard‐
ly better. The validation correlation coefficient indicates the ability of a trained network to
generalize on fresh data sets and is the true measure of the robustness of the prediction
accuracy of a neural network. The minimal difference between the validation correlations
for shale volume prediction and porosity prediction indicates that unlike the multi‐linear
regression transform, neural networks have almost no difficulty codifying the relationship
between seismic attributes and either of well derived shale volume or porosity. This capabil‐
ity is no doubt due to their ability to perform non‐linear multi‐variate mapping in complex
data spaces.

As with the case of shale volume prediction, there is striking semblance in the imaged
reservoir architecture and horizon facies distribution revealed by the porosity maps
predicted by both the multi‐linear regression and the neural network. However, the
neural network map provides higher resolution and images greater details of poro‐facies
variation both within the reservoir and beyond. The reservoir appears to be a fragmented
lobe system with lower porosity facies (likely hydrocarbon bearing) interspaced with
medium porosity facies (likely water bearing sands and or shales as indicated rock‐
physics cross‐plots). This compartmentalization could be due to the re‐working of the
initial deposits by strong wave action. Key stratigraphic features imaged on the predicted
shale volume maps such as the north‐south trending paleo‐channel and the eastern and
western channels which drained the reservoir region are indicated as sinuous low
porosity facies. The maps also indicate that the hydrocarbon‐bearing wells penetrate
lower porosity (20 ‐ 27% porosity) sands in the reservoir region. Rock‐physics studies
earlier discussed showed that hydrocarbon sands in the study area are associated with
lower porosities compared to water saturated sands and shales. The maps also locate the
non‐hydrocarbon‐wells in regions of relatively higher porosities (porosity greater than
32%) indicating that they penetrate either water saturated sands or shales at the target
horizon. The reservoir is bounded to the south by high porosity lithologies which
represent shales, indicating again that the reservoir is surrounded by sufficient lateral
shale seals to the south and the F1 fault to the north.
96 Tectonics – Recent Advances

Figure 11. Multi‐linear regression predicted porosity map over the top of H3 horizon. Hydrocarbon
wells are indicated to have penetrated lower porosity facies while dry wells encountered higher porosi‐
ty facies (shales) at the target horizon. Earlier imaged channel features (SSC, D1 and D2) possess low to
medium porosities in keeping with earlier crossplot results. The reservoir appears as an oblate fan.

Figure 12. Neural network predicted porosity map over the top of H3 horizon. The reservoir is a frag‐
mented lobe system with low porosity facies (hydrocarbon sands) interspaced with medium porosity
facies (water saturated sands), possibly a result of later re‐working of sediments by strong wave action.

4. Conclusion
In this chapter, we have presented our attempt at unravelling the paleo‐geomorphic settings of
the Okari Oil field, a classical example from the extentional province of the Niger Delta. We
employed conventional seismic interpretation techniques to build a tectono‐stratigraphic
model of the Field and to determine key factors which have contributed to and continue to
Seismic Paleo-Geomorphic System of the Extensional Province
of the Niger Delta: An Example of the Okari Field 97

define the development of this province of the Delta. We have further implemented advanced
characterization techniques to image details of facies distribution within the field and further
appreciate the effects of major factors on field evolution. Key factors affecting field evolution
appear to have been the onset of regional seaward dipping syn‐depositional growth faults and
the accomodation of resultant extensions by landward dipping compensational listric faults.
This resulted in rapid creation of accommodation in the form of mini depo‐belts into which
continentally derived detritus plunged. In traversing these growth faults the deposits fanned
out as mini‐deltaic fans which were possibly latter reworked and fragmented by strong wave
action. Fault block rotation possibly played an important part in defining drainage paths but
appear to have been overridden by the subsequent creation of graben structures by compensa‐
tional faults. Paleo‐geomorphic elements in the field include meandering channels, detached
sand bars and a fragmented lobe system. Favorable reservoir facies are to be found in low‐to‐
medium porosity shaly sands units.

Author details
Muslim B. Aminu
Adekunle Ajasin University, Akungba‐Akoko, Nigeria

Moses O. Olorunniwo
Obafemi Awolowo University, Ile‐Ife, Nigeria

5. References
Aminu, M. B., and Olorunniwo, M. A. (2011). Reservoir characterization and paleo‐
stratigraphic imaging over the Okari Field, Niger Delta, using neural networks.
Leading Edge, v.30, no. 6, pp. 650‐655.
Avbovbo, A.A. (1978). Tertiary lithostratigraphy of Niger delta. American Association of
Petroleum Geologist Bulletin, v. 62, pp. 295‐306.
Banchs, R. E. and Michelena, R, J. (2002). From 3D seismic attributes to pseudo well‐log
volumes using neural networks: Practical considerations. The Leading Edge, v.21, no.
10, pp. 996‐1001.
Bilotti, F., and . Shaw, J. H. (2005). Deep‐water Niger Delta fold and thrust belt modeled as a
critical‐taper wedge: The influence of elevated basal fluid pressure on structural styles.
AAPG Bulletin, v. 89, no. 11, pp. 1475–1491.
Burke, K. (1972). Longshore drift, submarine canyons and submarine fans in development of
Niger delta: American Association of Petroleum Geologist Bulletin, v.56, pp. 1975‐ 1983.
Calderon, J. E. and Castagna, J. (2005). Porosity and Lithologic Estimation Using Rock
Physics and Multiattribute Transforms In the Balcon Field, Columbia‐South America.
Expanded Abstracts, 75th SEG Annual International Meeting, pp. 444‐447.
Callan, R. (1999). The Essence of Neural Networks. Prentice Hall Europe. P. 232.
Chopra, S., and Marfurt, K. (2007). Curvature attribute applications to 3D surface seismic
data. The Leading Edge, v. 26, no. 4, pp. 404‐414.
98 Tectonics – Recent Advances

Corredor, F., Shaw, J. H., and Bilotti, F., (2005). Structural styles in the deepwater fold and
thrust belts of the Niger Delta: American Association of Petroleum Geologist Bulletin,
v. 89, no. 6, pp. 753– 780.
Dorrington, K. P. and Link, C. A. (2004). Genetic‐algorithm/ neural network approach to
seismic attribute selection for well‐log prediction. Geophysics, v. 69, no. 1, pp. 212‐ 221.
Doust, H. and Omatsola, E. (1990). Niger Delta, in J. D. Edwards, and P.A. Santogrossi, eds.
Divergent / passive margins basins: American Association of Petroleum Geologist
Memoir 48, pp. 201‐238.
Dunlap, D. B., Wood, L. J., Weisenberger, C., and Jabour, H. (2010). Seismic geomorphology
of offshore Morocco’s east margin, Safi Haute Mer area. American Association of
Petroleum Geologist Bulletin, v. 94, no. 5, pp. 615–642.
Ekweozor, C. M. and Daukoru, E. M. (1984). Petroleum source bed evaluation of Tertiary
Niger Delta. American Association of Petroleum Geologists Bulletin, v. 70, pp.48‐55.
Evamy, B.D., Harem boure, J., Knaap, P., Molloy, F. A. and Rowlands, P.H. (1978).
Hydrocarbon Habitat of Tertiary Niger Delta, American Association of Petroleum
Geologist Bulletin, v. 62, pp. 1‐39.
Frankl, E. J. and Cordry, E. A. (1967). The Niger Delta oil Province: Recent development,
onshore and offshore. Mexico City. Seventh World Petroleum Congress Proceedings v.
2, pp. 195‐209.
Hampson, D. P., Schuelke, J. S. and Quirein, J. A. (2001). Use of multiattribute transforms to
predict log properties from seismic data. Geophysics, v. 66, no. 1, pp. 220‐236.
Liu, Z. and Liu, J., (1998): Seismic‐controlled nonlinear extrapolation of well parameters
using neural networks. Geophysics, v. 63, no. 6, pp. 2035‐2041.
Maloney, D., Davies, R., Imber, J., Higgins, S., and King, S. (2010). New insights into
deformation mechanisms in the gravitationally driven Niger Delta deep‐water fold and
thrust belt. American Association of Petroleum Geologist Bulletin, v. 94, no. 9. pp. 1401–
1424.
Merki, P. J., (1972). Structural Geology of the Cenozoic Niger Delta. In: Dessauvagie, T. F. J. and
Whiteman, A. J. (eds.), African Geology, University of lbadan Press, Nigeria, pp. 635‐646.
Murat, R.C., (1970). Stratigraphy and Paleogeography of the Cretaceous and lower Tertiary
in southern Nigeria, in 1st Conference on African Geology Proceedings, Ibadan
University Press, pp. 251‐266.
Short, K. C. and Stauble, A. J. (1967). Outline of geology of Niger Delta. American
Association of Petroleum Geologist Bulletin, v.51, pp. 761‐799.
Tuttle, M. L. W., Charpentier, R. R., and Brownfield, M. E. (1999). The Niger Delta
Petroleum System: Niger Delta Province, Nigeria, Cameroon, and Equatorial Guinea,
Africa. United States Geological Survey, Open‐File Report 99‐50‐H, P. 65.
Weber, K. J. and Daukoru, E. (1975). Petroleum Geology of the Niger Delta, Tokyo. 9th
World Petroleum Congress Proceedings, v.2, pp. 209‐211.
Whiteman, A. (1982). Nigeria – Its petroleum geology, resources and potential, London,
Graham and Trotman, P. 394.
Wu, S., and A. W. Bally. (2000). Slope tectonics— Comparisons and contrasts of structural
styles of salt and shale tectonics of the northern Gulf of Mexico with shale tectonics of
offshore Nigeria in Gulf of Guinea, in W. Mohriak and M. Talwani, eds., Atlantic rifts and
continental margins: Washington, D.C., American Geophysical Union, pp. 151– 172.
Chapter 5

Geodynamic and Tectonostratigrafic


Study of a Continental Rift:
The Triassic Cuyana Basin, Argentina

Silvia Patricia Barredo

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/49958

1. Introduction
The stress and strain resultant from geodynamic processes control the uplifting and subsid‐
ence of different portions of the crust so sedimentary basins can be considered pieces of the
earth which have suffered prolonged subsidence related to thermo‐mechanical processes.
According to these latter, a mechanical analysis of subsidence in any basin can be used for
interpreting the regional distribution of depositional sequences; similarly the origin of the
sedimentary sequences can be related back to the tectonic activity which controlled the in‐
sertion and evolution of the sedimentary basin [1]. Depositional sequences result from a
complex interaction of the supply of sediments, the availability of the accommodation space
(both with tectonic components), sea level variations (which also may have a significant
tectonic influence) and climate variations. The first‐order control on basin geometry is the
deformation field resulted from the tectonic activity and is a fundamental control on sedi‐
mentation and the location of their resulting environments. For each basin, the geometry of
the main faults or the lithospheric flexure should provide the final morphology of the
trough and the resulting subsidence and thus, will control the sedimentation rate, grain size,
channel migration, avulsion episodes and the development of flood plains and/or lakes.

The Cuyana Basin corresponds to a passive continental rift, sensu [2] developed during Tri‐
assic times as a consequence of the early Mesozoic breakup of Gondwana (Permian to Late
Triassic‐Early Jurassic) e.g. [3, 4]. It was located in its western southernmost portion over a
wide belt of older accreted Eopaleozoic terranes, known as Gondwanides orogen. This latter
was an extensive belt composed of contemporaneous orogens of Palaeozoic age and their
related basins located along the southern Gondwana margin, it was firstly defined by Keidel
[5] but du Toit [6] referred to as “Samfrau geosyncline”. The Cuyana Basin is composed of

© 2012 Barredo, licensee InTech. This is an open access chapter distributed under the terms of the Creative
Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits unrestricted use,
distribution, and reproduction in any medium, provided the original work is properly cited.
100 Tectonics – Recent Advances

several asymmetric half‐grabens linked by accommodation zones that were partially or


completely disconnected in the early rifting phase [7, 8]. This geometry and the coeval tec‐
tonic activity that was mostly characterized by recurrent extensional pulses are thought to
be one of the major factors in the evolution of the sedimentary sequences and the complex
environmental relationships that this basin exhibits. Its continental deposits can be consid‐
ered as a whole a second‐order depositional sequence divided into three third order‐
sequences related with regional to local processes [9].

Much of the research on this basin has been conducted in the Cacheuta depocenter where an
intense hydrocarbon exploration and production has been held in the past decades [10 to
13]. Instead, the northernmost sections of the Cuyana Cuyana Basin, located in the San Juan
province, have been studied in detail by [9, 14‐21].

Interbasinal correlations along the whole basin were due to [7, 18, 22 to 24] who considered
the separate half‐grabens of San Juan and Mendoza provinces as a single trough. Age
control in the basin traditionally has been based on biostratigraphy e.g. [20, 25 to 30]. The
studied area is part of a fold and thrust belt where complex structures and diverse inversion
tectonics phenomena gave place to inverted subbasins composed of normal faults with
reactivated inverse displacement, new inverse faults, and fault propagation folds [8, 31].
Consequently, lithostratigraphy and biostratigraphy proved to be insufficient to understand
sequences contrast along the basin, see [32] for discussion. By the 1990s modern
stratigraphic tools have also been applied in an attempt to establish a more precise
correlation among depocenters and even between the active and flexural margin of the
Rincón Blanco trough, which are presently disconnected [8, 9, 20] but much uncertainty
remained because of the lack of absolute age data in the basin as a whole. Presently, a more
complete chronostratigraphic control has been achieved using the proposed
cyclostratigraphy scheme of [13] with isotopic dating [29, 32 to 34] which permitted to arrive
to an enhanced evolutionary model.

Classically, the Cuyana Basin has been considered as a passive rift developed as a
consequence of extensional to transtensional forces resulted of the collapse of a Permian
orogen and the beginning of the Gondwana breakup, the stratigraphic features represent the
interplay between tectonics (subsidence‐uplift) and sediment accumulation rates. The
evolution of the Gondwanides along the southern portion of Gondwana was a key process
through which the Triassic basins of Argentina developed. In this dominant geodynamic
framework it is emphasized here that making correlations can highlight similarities and
differences among basins and even within a given one, as it is the case of the Cuyana Basin.
Subsidence analysis of the subsurface Bermejo and outcropping Ischigulasto‐Villa Unión‐
Marayes and Cuyana basins (Figure 1) reveals the existence of notably various episodes of
accelerated subsidence during Middle to Late Triassic, suggesting that all of them share a
common tectonic history.

It is explored here these results in detail integrating geodynamic and tectonostratigraphic


concepts to understand the rift evolution and the history of its infilling. The resulting data,
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 101

permitted to understand the coeval deformation and the tectonic processes acting on the
lithosphere in the western margin of Gondwana.

2. Geological setting of the Cuyana Basin


Related to the Gondwanan Orogeny, that affected the western margin of southern South
America (Permian to the Late Triassic‐Early Jurassic) and the beginning of the Mesozoic
Gondwana breakup, a series of rifts were developed (Figure 1). They correspond to a com‐
plex system of rapidly subsiding, NNW‐SSE trending narrow asymmetric half‐grabens
bounded by predominantly normal faults in one margin that parallels the grain of older
Paleozoic structures e.g. [9, 12, 19, 36 to 38]. The border fault is sometimes segmented with
segment boundaries marked by change in strike or fault overlap, relay ramps, transfer faults
and rider blocks(Figure 2). The notable increase in thickness towards the center border fault
suggests that the structures were syndepositionally active and thus it is proposed here the
architecture and fill were influenced by the geometry and the displacement of the bounding
normal faults.

Among them, the Cuyana Cuyana Basin is the largest Triassic rift basin of western Argenti‐
na and considered to extend over an area of more than 60.000 Km2. It corresponds to a pas‐
sive continental rift developed during differential intraplate stresses derived from a backarc
extension acting on a normal crust and moderate thermal flux [9]. At present, it underlies a
lowland segment of the Argentine foreland but several exposures are along both flanks of
the Precordillera, in Mendoza and San Juan provinces (Figure 1). The rift is composed of
several asymmetric half‐grabens roughly triangular in cross section, and linked by accom‐
modation zones [7, 8] which correspond to Cacheuta, Las Peñas‐Santa Clara, Rincón Blanco,
Puntudo General Alvear, Ñacuñan and Beazley (see Figure 1). Border faults (BF) of the
Cuyana Basin strike obliquely to the maximum extension direction (NNE‐SSW), have a
stepping geometry and opposite dip directions (Cacheuta and Santa Clara‐Rincón Blanco)
causing the faulted and hinged margins sometimes shift from side to side of the rift basin as
it is evident between Cacheuta and Santa Clara subbasins e.g. [7, 9, 11], or keep the same dip
orientation as between Santa Clara, Rincón Blanco and Puntudo depocenters (Figure 2).
Each segment of these master structures shows tips that are marked by a change in strike or
fault overlap, relay ramps and rider blocks. Most of them were controlled by the normal to
oblique reactivation of pre‐existing zones of weaknesses in the crystalline basement, and
thus main border fault exhibits N‐S and NNW‐SSE direction while accommodation and
transfer zones between main Puntudo, Rincón Blanco and Cacheuta depocenters display a
mostly NNE‐SSW direction. Major segments do not become hard linked and the displace‐
ment was partitioned among several segments without being physically connected. In this
way, intrabasinal highs could persist throughout synrift and postrift sedimentation along
the whole basin. The lack of polarity reversals in the northern segment is attributed to the
basement fabric control with fault reactivation of ancient N‐S and NNW‐SSW structures. In
particular, within Rincón Blanco depocenter, the border fault is segmented and its closer
spacing segments hard‐linked with evident transfer of displacement from the southern to
102 Tectonics – Recent Advances

northern segments (Figure 2) [1, 8]. The restricted half‐grabens were separated by left‐
oblique‐slip (left lateral?), WNW (Az 112°) that are oblique to the border fault.

Figure 1. Triassic rift basins of central‐western Argentina and the main subbasins. From [35].

The Cuyana Basin contains, approximately, up to 3700 meters of continental rocks of


predominantly alluvial, fluvial, and lacustrine origin interbedded with tuffs of coeval
volcanism with intraplate affinities e.g. [11, 39]. The notable increase in thickness towards
the border faults suggests that these structures were syndepositionally active. Field studies
and seismostratigraphic analysis show that changes in sequence geometry occurred in‐
phase with intra‐continental elastic stress relaxation, as fault reactivation and probable
basement reworking, during the quasistatic event on the western margin of Gondwana.
These recurrent extensional pulses that controlled the extension along the Cuyana Basin
impinged several different characteristic to the infilling of the Cerro Puntudo, Rincón Blanco
and Cacheuta depocenters and thus they will be treated in separate items.
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 103

Figure 2. Cuyana Basin main depocenters. Border faults (BF) have a stepping geometry and opposite
dip directions (Cacheuta and Santa Clara‐Rincón Blanco) or keep the same dip orientation as between
Santa Clara, Rincón Blanco and Puntudo depocenters. Major segments are related through accommoda‐
tions zones which constitute intrabasinal highs. To the right, the scheme shows a plan view of the
Rincón Blanco trough whose segments became hard linked. From Barredo & Ramos [1].

Extension took place along three phases, named Rift I, II and III and their corresponding
synrift deposits by Barredo [9]. During Synrift I, the deposits were restricted to a series of
partially isolated depressions. Thus, in proximal positions the alluvial coarse clastic facies are
covered and interfingered with medium to fine‐grained sandstones and tuffs interpreted as
deposited in fluvial, and lacustrine/playa‐lake settings e.g. [8, 9, 13, 20, 23, 26]. This first
sequence is separated by a regional unconformity by the second depositional phase or Synrift
II. It is a fining‐upward sequence which consists, in general, of cross‐bedded sandstones, black
shales and tuffs, all related to braided to high sinuosity‐river systems which grade into a
widespread lacustrine setting. Finally, this second rifting stage passes upward to shallow
lacustrine to fluvial sandstones, shales, and tuffs deposited during the early to late post‐rift.
This shallowing‐upward succession is widespread in the basin and displays an onlap
relationship with the underlying beds, directly overlying Paleozoic basement [8, 9, 13]. During
this stage, there was a renewal of the subsidence in the basin. Recent investigations permitted
to identify a the third extensional event during early Late Triassic (to Early Jurassic times?)
probably associated with the opening of the Neuquén Basin, located farther south in northern
Patagonia [1, 32]. Main evidences of this event are not regionally found, the best exposures
outcrop in the Rincón Blanco Subbasin, however to the south, the subsurface Barrancas
104 Tectonics – Recent Advances

Formation (Cacheuta depocenter) shares similar tectonostratigraphic characteristics with that


of the Rincón Blanco but reveals certain uncertainties about its age. By the moment, this unit
has been assigned to the Late Triassic [40] and even to Lower Jurassic [41]. The infill of Synrift
III is represented by an alluvial‐fluvial succession developed under renewed mechanical
subsidence of the basin and marked semiarid conditions.

The final thermal relaxation of the basin seemed to have occurred during Jurassic and
Cretaceous times and aborted during Tertiary times when the lithosphere undergone
flexural subsidence induced by the Andean orogenic overloading and by sediment charge.

3. Puntudo Subbasin
The northernmost exposures of the continental Triassic Cuyana Basin crop out at the Cerro
Puntudo locality on the western flank of the Precordillera in the San Juan province (Figure 3).
These exposures record the sedimentation near the northern end of the basin related to a
fault tip end. An angular unconformity separates de Cerro Puntudo succession from the
underlying Choiyoi Group volcanics and it is tectonically truncated at top [34].

Figure 3. Geological map at the Co. Puntudo area. Modified from Mancuso [34]
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 105

The present structure consists of east verging (Andean) thrusts with wavelengths of ~1 km.
These thrusts are responsible of the truncation and suppression of the west El Relincho
facies (El Puntudo fault) (Figure 3). Triassic basin inversion was positive and topographic
uplifting and gently folding of the ancient trough has been observed. The deposits are
folded in northeast striking asymmetric syncline with its eastern flank well developed and
the western almost truncated. The anticline of the eastern margin is interpreted as a result of
reactivation of ancient rider blocks as a post depositional anticline during reverse motion of
the structure. West‐northwest faults seem to follow the Paleozoic fabric and are parallel to
the Rincón Blanco Sub‐basin transfers and thus, genetically related [9].

The extensional structure consists of a north‐south striking fault (border fault?) located to
the east. This assumption is based on the thick fan‐conglomerates with main paleocurrents
showing a marked southwesternward component of the flow parallel to the basin margin
and towards the fault zone centre, where subsidence favoured the insertion of a shallow
lake. This flow direction can be linked southernward with the axial sediment system of the
Rincón Blanco Sub‐basin. The border fault is presumed to be a west‐dipping structure,
probable lystric in cross section, with highest displacement to the center (displacement
estimates are of 400 meters). On the basis of the maximum sequence thickness the “border
fault” has been 12 km long. A notable reduction of the sedimentary record to the north and
south can be extracted of the onlapping of the synextensional strata over basement rocks,
particularly in the southern part where at the Cerro Colorado the Choiyoi volcaniclastics are
exposed. Paleocurrents in this area points to a northward flow and thus, has been
interpreted as an accommodation zone (transfer) that kept this depocenter probably
disconnected from the Rincón Blanco Sub‐basin.

The Triassic column reaches an exposed thickness of approximately 400 meters and has been
divided into two units, the Cerro Puntudo and the El Relincho formations [42] arranged in
two of the three synrift cycles recognized for the Cuyana Basin. The first tectosedimentary
sequence or Synrift I corresponds to the Cerro Puntudo Formation. It is a fining‐upward
succession which starts with a thick package of 400 m alluvial fan conglomerates and cross‐
bedded coarse sandstones (Figure 4). This succession is mainly composed of an alternation
of massive red clast‐supported conglomerates and subordinated sandstones which passes
upward to a braided fluvial system dominated by red and reddish‐brown conglomerates
and sandstones. Sheet‐floods related to ephemeral fluvial deposits, characterized by red
fine‐grained sandstones, mudstones, and light‐coloured tuffaceous limestones, are interfin‐
gered and cover the braided fluvial beds. The upper one‐third of this first cycle comprises 75
m thick well‐stratified and laterally persistent gray micritic and stromatolitic limestones,
reddish brown mudstones, red fine‐grained sandstones with rippled lamination, and thin
levels of green tuffs interbedded. This part of the section was interpreted as deposited in a
relatively shallow carbonate‐rich lake with marked cyclicity recognized by a succession of
retrograding and prograding events. Coeval volcanism is represented by the interbedded
tuffs. To the southeast, the lacustrine deposits are dominated by siliciclastic facies made of
reddish brown mudstones, red fine‐grained sandstones, and green tuffs which were inter‐
preted as deposited in a mudflat environment e.g. [23, 24, 34]. The climatic conditions dur‐
106 Tectonics – Recent Advances

ing the deposition of the first synrift phase were markedly seasonal, as represented by the
cyclical nature of the lacustrine deposits, not associated to tectonic controls. Palynomorphs
suggest the presence of a relatively diverse flora composed by at least riparian vegetation
(ferns and lycopsids) associated to the lake margins and a forest (araucariaceans) in the
fluvial floodplains [34]. Due to the lack of evidence of evaporitic facies and extensive desic‐
cation features humid climate conditions developed at least seasonally thus a subtropical
seasonally dry climate might be assume for the upper part of the first Synrift phase. A U‐Pb
SHRIMP zircon age of 243.8 ± 1.9 Ma (Anisian) obtained from juvenile magmatic zircons in a
tuff interbedded with the lacustrine beds constrained the deposition of the Synrift I at Cerro
Puntudo to the Early‐lower Middle Triassic [34].This age is consistent with palynological
data obtained from equivalent lacustrine levels.

The second tectonosedimentary sequence or Synrift II, is represented by the El Relincho


Formation. It starts at the point where the lacustrine deposition is abruptly interrupted by
the instauration of a coarse alluvial setting. The fining‐upward El Relincho succession
(approx. 140 m thick) is dominated by green clast‐supported conglomerates exhibiting clasts
of the underlying Cerro Puntudo sequence. The succession passes upward into reddish
brown, cross‐bedded, coarse to medium ‐grained sandstones of braided fluvial origin. The
upper limit of the Triassic succession at Cerro Puntudo area is tectonically truncated

Figure 4. Generalized stratigraphic column of the Puntudo subbasin infilling showing the rifting epi‐
sodes recognized across the whole Cuyana Basin. Modified from Mancuso [34].
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 107

4. Rincón Blanco Subbasin


The Triassic Rincón Blanco subbasin is located in the Precordillera fold and thrust belt at 31°
24’ – 31° 33’ south (Figure 5). It is 40 km wide and stretches approximately ~60 km N‐S. The
subbasin was mainly developed over Ordovician‐Silurian distal platform deposits and Car‐
boniferous glacimarine diamictites and turbidites nevertheless, to the north the column
overlies Permian marine platform clastics. The present geometry of the Rincón Blanco active
margin is a north–south tight asymmetric syncline. The structural analysis indicates that the
east margin has been truncated by a series of west verging back‐thrusts since Cenozoic
times. Basin marginal deposits have in places been removed by erosion or preserved in fault
inliers. The western margin has undergone east verging Andean thrusts (Cenozoic) with
wavelengths of 2 km, presumably older than the back‐thrusts, that has separated de out‐
crops from the flexural margin of the depocenter. Only, the south‐eastern portion of the
active margin remains practically unaffected, although some inversion of earlier extensional
structures has been noted (Figure 6). Flexural margin is represented by isolated outcrops
and subsurface deposits and is presently separated by a basement high from the active mar‐
gin. Inversion was positive and topographic uplifting and gently folding of the ancient
troughs has been observed, in particular in the Cerro Bola‐Cerro Amarillo region (south‐
ernmost outcrops). Classic basin‐inversion structures consist of anticlinal folds and reverse
faults, like the BF segment exposed in Cerro Amarillo. Post depositional anticlines and syn‐
clines formed during reverse motion along the reactivated faults. Minor inverse faults area
associated with the backtrhusts, they are composed of fault bend, fault propagation, and
detachment geometries. In the flexural margin they have also uplifted the ancient intrabasi‐
nal highs where sequences are drastically thinned or basement exposed (Figure 7).

The spatial arrangement of the architectural elements permitted to interpret this Triassic
depocenter as an asymmetric trough with a hinged margin. Basin‐scale morphology could
not be determined precisely because Triassic outcrops are tectonically truncated except for
the western margin that could be modelled in subsurface [31]. Faults have displacement
magnitudes of 1000 m and sole into a subhorizontal detachment surface at about 10000 m
deep, below which deposits are undeformed. Using the architecture and some features of
the sedimentary record it was established that the border fault (BF) was composed of right‐
stepping, west‐dipping, normal to oblique faults, separated by a transfer zones (see Figures
2 and 5) became hard‐linked. In cross section they are lystric and in plan view sinuous with
highest displacement toward the center (displacement estimates are of 2 km). The resulting
half‐grabens are supposed to have gradually grown in depth and length through time sur‐
rounded by uplifted footwalls as a consequence of the absolute upward motion during fault
displacement and isostatic adjustment. They correspond to Cerro Amarillo/Rincón Blanco
and Marachemill troughs, in the active margin; Barreal and Agua de Los Pajaritos, in the
flexural margin.
Kinematic indicators and the orientation of diabase dikes suggest a NNE‐SSW (Az 30°)
extension direction for the depocenter and a N‐S to NNW‐SSE border faults strike [8,9].
Local transfers zones correspond to west‐northwest (Az 112°), left lateral strike‐slip faults
interpreted to have been formed to accommodate differential extension within the depocen‐
108 Tectonics – Recent Advances

ter (Figure 6). These faults can be associated with the Paleozoic fabric which consists largely
of generally northwest‐striking and west‐northwest striking reverse and tranpresional faults
[37, 43]. Intrabasinal highs produced by rotation blocks made up internal elevations relative
to the rift flank and trough and, consequently, the flexural margin contained much thinner
sequence of synrift strata (~1800 meters) [31] than the active margin (3000 meters).

Figure 5. Left, geological map of the Rincón Blanco depocenter with detailed information from the
flexural margin. Right, location map of the Rincón Blanco half‐graben outcrops in the Cuyana Basin.
The satellite image shows the distribution of the main depocenters, subsurface is not included. From
Barredo [9].

The infilling consists of almost 3000 m coarse conglomerates interfingered with sandstones,
shales, tuffs, tuffaceous mudstones and bimodal volcanic rocks composed of rhyolites and
rhyiolitic tuffs and ignimbrites associated with the Choiyoi uppermost effusives [39, 45].
Alkaline basalts are also present but only in the flexural margin [9, 44]. The upper limit of
the Triassic succession is marked by a regional uncomformity with the overlying Cenozoic
foreland deposits. Three packages of genetically linked units bounded by regional extended
unconformities of third order are associated with the three rifting stages (Synrift I, II and III)
of the Cuyana Basin. These rifting stages beeing related with the acceleration of faults
subsidence during Middle to Late Triassic. They comprise the Rincón Blanco and
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 109

Marachemill groups (active margin) and the Sorocayense Group (flexural margin) (Figure
7). Their evolution was mainly controlled by tectonic pulses but sediment supply and
climate was also important [1, 9].

Figure 6. Detailed geological map of the active margin of the Rincón Blanco depocenter with inferred
transfers faults. From Barredo [8,9].
110 Tectonics – Recent Advances

Figure 7. Stratigraphic column of the Rincón Blanco active margin. From Barredo & Ramos [1]. Upper
right, north view of the coarse facies of the Marachemill Unit. Lower right,north view of the deep lacus‐
trine facies of the Carrizalito Formation.

In the active margin, the Rincón Blanco Group is composed of, from base to top: the Ciénaga
Redonda, Cerro Amarillo, Panul, Corral de Piedra, Carrizalito, and Casa de Piedra for‐
mations [15]. The trace of the contact between the synrift sediments and the prerift rocks is
concave and well exposed to the south, in the Cerro Bola region (active margin). The Cié‐
naga Redonda and Cerro Amarillo formations correspond to the Synrift I, when the basin
constituted a simple fault‐bounded through, and comprise three facies association, alluvial
fans, alluvial fans braidplain settings dominated by ephemeral streams, and shallow, mostly
ephemeral lacustrine deposits. A maximum of 1200 m has been measured nearby the Ama‐
rillo and Bola hills (Figure 7) decreasing to the north to reach in outcrops less than 100 me‐
ters. This thickness diminishing is also observed to the west. The Synrift I starts with con‐
glomerates and breccias interpreted as cohesive debris flow of fan deposits and un‐cohesive
debris flow developed in non‐canalized sheetflows (Ciénaga Redonda Formation). Interfin‐
gered with these facies, there are ignimbrites and scarce rhyolitic tuffs and tuffs with an
estimated age of 246 Ma [32]. Overlying, there are laterally discontinuous sheet‐like beds of
massive and horizontally bedded sandstones stacked in few meters with alternating con‐
glomerates of the Cerro Amarillo Formation. This sequence corresponds to short‐lived,
wide, poorly canalized flows or low energy sheetflows in ephemeral streams of fans and
bajadas. More distally, these systems are associated with silty‐sandstones, massive to paral‐
lel mudstones with desiccation cracks and anhydrite lenses alternating with black (lignitic?)
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 111

shales deposited in a shallow lacustrine environment. This succession is covered by sand‐


stones, shales and tuffs of braided fluvial origin with a predominantly axial paleoflow in‐
flux.

The Synrift II is separated from the underlying units by a regional unconformity and is
composed of amalgamated conglomerates with angular basement derived lithologies, sand‐
stones and ryolithes from the underlying Triassic units [9]. This sequence passes upwards to
lenticular bodies of massive to trough‐cross conglomerates and pebbly sandstones with
erosive bases. They correspond to alluvial fan and braided fluvial facies association. The
volcaniclastic content is notoriously high compared with the underlying units and consists
of tuffs, and tuffaceous sandstones and conglomerates (Panul Formation). Fining–upward
tabular sequences of fine conglomerates to sandstones with trough‐cross bedding, lag de‐
posits, and erosive bases follow. They grade to medium fine‐grained, moderately to well
sorted tabular sandstones/tuffaceous sandstones, reworked tuffs, and thick overbank mud‐
stones at top of meandering fluvial origin (Corral de Piedra Formation). The whole succes‐
sion passes vertically and laterally into the lacustrine facies of the Carrizalito Formation
composed of thick massive mudstones/tuffaceous siltstones with volcaniclastics, mostly
chonitic, levels. Upward these beds are covered by thinly stratified silty sandstones and
massive mudstones related to the deeper facies of the lake. They are associated with massive
or horizontally laminated marls (30 cm in thickness) and pale‐grey organic levels which
consist of laminated bituminous shales (oil‐prone coals) with micritic calcite. The lake de‐
posits are covered by clast supported planar to trough‐cross bedded lenticular conglomer‐
ates and sandstones interbedded with laminated mudstones and limestones interpreted as
deltaic mouth bars deposits (basal levels of the Casa de Piedra Formation). Upward these
facies are covered by lenticular conglomerates and sandstones with trough‐cross bedding
and erosive bases of fluvial origin. They grade to fine‐grained sandstones/tuffaceous sand‐
stones and thick overbank mudstones which at top interfinger with ash fall tuffs.
A U‐Pb Shrimp age on tuffaceous beds at the base of the lake levels of the Corral de Piedra
Formation, in the Rincón Blanco depocenter, have yielded an age of 239.5±1.9 Ma [32]. This
age constrains the initial deposits of Synrift II to the late Anisian (Middle Tiassic) which
coincides with the recently date of 239.2 ± 4.5 Ma obtained by Spalletti [29] for the Potrerillos
Formation (initial infilling of the Synrift II at the Cacheuta subbasin).

Finally, to the north‐east margin of the basin a series of almost 900 meters of mainly volcani‐
clastic and clastic rocks were interpreted as the Marachemill Unit and included in the (Syn‐
rift III) [1, 32]. It is in tectonic contact with the Rincón Blanco Group through the Tontal
backthrust (see Figure 6). Three facies associations characterized this unit: proximal to distal
alluvial fans, ephemeral streams and fluvial systems. The first depositional environment
comprises thick massive, mud‐rich matrix‐supported red conglomerates and breccias com‐
posed of volcaniclastic and sedimentary clasts mostly from the underlying Rincón Blanco
Group. These facies interfinger with several ignimbrites and rhyolitic tuffs levels and tuffa‐
ceous sandstones. Poorly canalized flows and sheet floods developed in ephemeral envi‐
ronments follow upwards. They are characterized by coarse massive to horizontal laminated
sandstones with erosive to sharp bases and shale intraclasts and are laterally associated with
112 Tectonics – Recent Advances

flood plain sandstones and shales with oxidized organic matter and desiccation cracks.
Volcaniclastics are abundant and correspond to ash fall tuffs and pyroclastic flows. The
third depositional environment corresponds to a braided fluvial system with conglomerates
to granular sandstones in frequently amalgamated bodies. The conglomerates are poorly
sorted with sub‐angular to sub‐rounded clasts from the Ordovician basement. Overlying
beds correspond to tabular sandstones with planar cross stratification which grades to fine
to medium sandstones with small scale trough‐cross stratification and ripple‐cross lamina‐
tion capped by thin beds of massive or rippled greenish grey siltstones, tuffs, and mud‐
stones with invertebrate burrows and mudcracks. Paleocurrents show an east‐northeast
predominant inflow. Ash‐fall deposits are abundant constituting thick tabular massive,
sometimes laminated, deposits and reworked levels that can be interpreted as tuffaceous
plains with isolated poorly developed fluvial channels. U‐Pb zircon age of 230.3 ± 1.5 Ma
(SHRIMP) has been obtained for the Marachemill Unit suggesting an early Carnian age for
its deposition [32].

The Sorocayense Group represents the infilling of the Agua de los Pajaritos and Barreal
groups which correspond respectively to the north and south separated depocenters of the
flexural margin [18,26]. The first is composed of: Agua de Los Pajaritos, Monina, Hilario, El
Alcázar and Cepeda formations e.g. [46] (see Figures 5 and 8) and the other by: Barreal,
Cortaderita and Cepeda formations [18]. The northern Agua de Los Pajaritos consists of a
basal alluvial‐fluvial sequence with abundant subaerial volcanic tuffs and volcaniclastic
rocks overlain by deep lacustrine facies and fluvio‐deltaic facies (Monina Formation) [46,
47]. Clast‐supported conglomerates and breccias dominate at the base and are interpreted as
non‐cohesive debris with subordinate matrix‐supported paraconglomerates (cohesive debris
flow). Non‐canalized and channelized levels follow, the lower ones correspond to sheet‐
flows and are characterized by massive sandstones with high energy directional structures;
the second and most dominant upward, consist of lenticular thin bedded conglomerates
associated with extended and shallow channels. These facies are followed by finning up‐
ward sequences of cross‐bedded mostly tabular sandstones, primary tuffs, reworked tuffs,
shales and bituminous shales, deposited in a high sinuosity river setting with well devel‐
oped flood plains (frequently obliterated by fallout thepra). These facies laterally interfinger
and are covered by sandstones, marls with algae lamination and massive and bituminous
shales. Sandstone and siltstones with wave and climbing ripple stratification and tangential
cross stratification correspond to deltaic mouth bars. At top and intercalated are white and
green fallout tuffs and chonites, massive or thinly laminated and sometimes indurated with
siliceous replacement. The clastic dikes cut these lacustrine facies. The fluvial‐deltaic and
fluvial strata of Hilario Formation follow. They consist of thickening and coarsening up‐
ward units characterized by laminated mudstone and thin sandstone beds which pass into
amalgamated sandstones beds with small scours and cross‐bedded sets of sandstone and
clast‐supported conglomerates, sandstones and thin coaly layers arranged in a series of
fining upwards units. Finally, the facies are replaced by lenticular shaped fining upward
coarse tabular sandstones and shales of a fluvial high sinuosity environment. This clastic
and volcaniclastic sequence resembles that of the Synrift II of the active margin (Rincón
Blanco Group) and has been preliminary correlated with it [46]. Sandstones with erosive
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 113

bases associated with clayed mudstones, tuffs, marls, limestones, and massive and rippled
laminated sandstones of high sinuosity fluvial and palustrine environments are considered
the base of the El Alcázar Formation. This succession unconformably overlies the Synrit II
and has been assigned to the early post‐rift [32]. Tabular multicolored shales with tuffs,
subordinates sandstones, and massive lenticular conglomerates are interpreted as lacustrine
facies with alternating mouth bar and turbidites deposits. It is gradually replaced by lenticu‐
lar sand and conglomerate bodies associated with mudstones, tuffs of floodplain origin
sourced from outside the rifted basin. Lava flows mostly tholeitic interfingers these facies
[44]. A renew of the extensional regime (Rift III) gave place to the deposition of the fan de‐
rived clastics of the Cepeda Formation lateral equivalent to the Marachemill Unit of Rincón
Blanco depocenter [1].

The Synrift I in the Barreal depocenter encompasses distal fans and fluvial braidplain envi‐
ronments followed by shallow lacustrine shales, mudstones, and tuffaceous mudstones and
deltaic sandstones [48] (Figure 8). Braidplain environments consist of clast‐supported, mas‐
sive, sometimes normally graded lenticular conglomerates and breccias with basal sharp
and irregular contacts associated with sandstones that show parallel or low angle planar
cross stratification. These facies are followed by massive and thin laminated shales, fine
current‐rippled sandstones laterally related to silty to fine‐grained sandstones interbedded
in a heterolithic arrangement with coal fragments, interpreted as lacustrine deposits. This
first synrift sequence is included in the lower half of the Barreal Formation.

The second synrift phase (Synrift II) is marked by the deposition of braided fluvial systems
with lenticular to tabular beds of poorly sorted conglomerates with subordinates sandier
lenses and more tabular planar cross stratified sandstones associated with fining upward
channel fill sequences.

This succession is covered by lacustrine deep facies with bituminous shales and thin subor‐
dinate sandstones of mouth bars and turbidites deposits, and finally lenticular conglomer‐
ates and coarse sandstones of fluvial systems, probably high sinuosity sandy braided system
(upper half of the Barreal Formation). The post‐rift facies are represented by the Cortaderita
Formation beds. It starts with tabular shales, tuffs, and massive and rippled wave laminated
sandstones of lacustrine origin. This facies are covered by lenticular pebbly mouth bar sand‐
stones that pass upward to conglomerates and sandstones arranged in lenticular bodies of
sandy braided fluvial system, sometimes obliterated by cinder fall out. The succession fin‐
ishes with the instauration of a high sinuosity river system with well developed flood
plains. The Synrift III (Cepeda Formation) consists of amalgamated clast and matrix sup‐
ported conglomerates of alluvial proximal fans composed of volcaniclastics and siliclastics
mostly from the underlying Cortaderita Formation; subordinate sandstones to the top of the
conglomerates were interpreted as braided ephemeral plains. Sheet floods and poorly cana‐
lized flows developed in ephemeral environments with coarse massive to horizontal lami‐
nated sandstones with erosive to sharp bases and shale intraclasts. Volcaniclastics are abun‐
dant and correspond to ash fall tuffs. Upward, braided fluvial deposits of frequently amal‐
gamated poorly sorted conglomerate bodies with sub‐angular to sub‐rounded clasts with
subordinate sandstones cover the ephemeral facies. It passes upward to tabular sandstones
114 Tectonics – Recent Advances

with planar cross stratification which grade to fine to medium sandstones with small scale
trough‐cross stratification and ripple‐cross lamination. To the top, massive or rippled green‐
ish grey siltstones, tuffs, and mudstones with invertebrate burrows and mudcracks domi‐
nate; this beds were interpreted as mudflat deposits.

Figure 8. Stratigraphic column of the Agua de los Pajaritos depocenter (flexural margin). From Barredo
[46]. Upper right, a south view from the El Alcázar postrift facies of clastic and pyroclastic materials
deposited in shallow lake and fluvial environments. Lower right, east view the deep lacustrine Monina
Formation.

Paleofloristic analysis based on the macrofloral remains preserved nearly in the whole col‐
umn on both the flexural margin and active depocenters indicate the presence of evergreen
subtropical floras adapted to seasonally dry climatic conditions in the Synrift I and lower
part of Synrift II sequences. Thus, they do not indicate a marked climatic shift in this interval
although floristic information from the uppermost part of the successions in this part of the
Cuyana Basin is very fragmentary e.g. [26, 27].

5. Cacheuta Subbasin
The Cacheuta Sub‐basin outcrops constitute the southernmost exposures of the Cuyana Basin.
The best outcrops are located in the southern flank of the Cerro Cacheuta and in the nearby
Cerro Bayo in the Potrerillos locality, west of Mendoza city (Figure 9). The eastern margin of the
sub‐basin was developed over Precambrian crystalline basement conversely; on the western
margin the Triassic succession overlies Silurian‐Devonian turbidites, Cambrian‐Ordovician
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 115

limestones, and the volcanics of the Permian‐Triassic Choiyoi Group complex. A regional
uncomformity determines the upper limit with the overlying Jurassic?–Cenozoic deposits.

Figure 9. Geological map of the Cacheuta Subbasin outcrops at Potrerillos area. Modified from [49].

The Cacheuta Group represents the infilling of this through and is composed of, from base to
top, by: the Río Mendoza, Cerro de las Cabras, Potrerillos, Cacheuta and Río Blanco
formations. These facies characterizes two episodes of rifting followed by a period of post‐rift
subsidence. Unconformabling overlying these sequences, there is another unit, the Barrancas
Formation of clear affinities with upper extensional events registered in the northern portion
of the Cuyana basin. Up to the moment, it has been assigned a Late Triassic age by Rolleri &
Fernández Garrasino [40] and a Lower Jurassic age by Reigaraz [41].

The structure consists of a series of N‐NW to N‐S trending asymmetric folds, broken locally
by reverse high angle faults displaying either eastward or westward vergence [50] (Figure
9). These structures end against the Precordillera province to the north and extend far
southeast to the San Rafael Block (see Figure 1). Tertiary structures encompass compresional
Andean thrusts and positive inverted Triassic faults composed of folds and faults. Basement
involved faults were mildly inverted and they still preserve extensional features. Minor
antithetic and synthetic faults are release structures formed during bending. Compresional
high angle inverse faults (60º) associated with the reactivation of Palaeozoic structures are
also found.
116 Tectonics – Recent Advances

The subbasin was narrow and asymmetric roughly triangular in cross section e.g. [13]. The
more steeply inclined basin margin is located to the west and consists of a predominantly
normal‐slip fault/s associated with the border fault, which generally trends north–northeast.
It is a steep dipping basement‐involved structure with a hinged margin to the east. It coin‐
cides with an important Ordovician suture, the Valle Fértil Lineament [11, 12]. In cross sec‐
tion, it is lystric and in plan view it has been interpreted as being sinuous with highest dis‐
placement toward the center e.g. [7, 13]. Surface and subsurface information show that faults
have displacement magnitudes of thousand meters and sole into a sub‐horizontal detachment
surface. Synrift sequences display reverse drag or rollover folds in subsurface. Strike‐slip
minor faults with a west‐northwest strike and left‐lateral slip can be observed in surface and
subsurface. They were interpreted as inverted Permo‐Triassic oblique‐slip normal faults with
dextral displacement. Intrabasinal highs produced by rotation blocks made up internal eleva‐
tions relative to the rift flank and trough. In this way several sub‐depocenters with local con‐
trol were developed and limited by oblique structures to the border fault. They correspond to
transtensional structures with north‐northwest to northwest trending sinistral high‐angle
faults (60º and 90º to the northeast); additionally, northwest dextral lateral structures were
also found. The Potrerillos area corresponds to a displacement transfer zone in which sinistral
strike‐slip meso‐scale faults play an important role. It was a topographic high that separated
this sub‐basin from the Las Peñas through to the north. In this case, longer north–northeast‐
striking border fault (BF) of the Cacheuta depocenter is connected by shorter segments to the
northern depocenter border by an opposite dipping fault. It has been interpreted as a soft
linked connection as faults transferred displacement from one segment to another without
being physically connected [9].

The Triassic column in this depocenter is of approximately 3000 m thickness (Figure 10). Dur‐
ing the early depositional phase (Synrift I), the succession (approx. 800‐1000 m thickness) is
characterized by reddish conglomerates of alluvian fan facies (Río Mendoza Formation) relat‐
ed to the active margins of the rift. They laterally interfinger with multicolored mudstones,
fine‐grained sandstones, and tuffs of ephemeral‐fluvial and playa‐lake origin deposited
basinward. In the depressed areas of the subbasin, shallow lake facies were accumulated char‐
acterized by the deposition of relatively thin oolitic grainstone beds and stromatolitic lime‐
stones interbedded with tuffs (Cerro de las Cabras Formation). This first sequence is separated
by a regional unconformity by the second depositional phase (Sinryft II) [9, 13, 40]. The second
synrift sequence (Synrift II) is a fining‐upward succession (Potrerillos and Cacheuta for‐
mations) mainly represented by lower energy facies and fine‐grained deposits than the under‐
lying sequence. It was accumulated on a smoother relief due to the infilling of the depocenter
and the deposits reach up to 1200 m in thickness. The Potrerillos Formation is characterized by
fluvial conglomerates at the base intercalated with light greenish cross‐bedded sandstones,
and light tuffaceous sandstones of perennial braided river origin; this fluvial deposits grade
basinward to greenish‐grey laminated siltstone and sandstones interbeded with black bitumi‐
nous shales and tuffs, related to high‐sinuosity river systems. This facies laterally interfinger
and are covered by the widespread lacustrine black shales of the Cacheuta Formation. In the
maximum transgression of the lake, the lacustrine facies overlaps the fluvial deposits of the
Potrerillos Formation to the basin borders. Finally, the post‐rift phase in the depocenter is
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 117

characterized by the red sandstones, mudstones, and tuffs of the Río Blanco Formation. This
succession (up to 1000 m) has an onlap relationship with the underlying beds and represents
the instauration of a fluvial‐deltaic system over the lacustrine black shales. The Barrancas
Formation (100 m) completely developed in subsurface, is composed of distal alluvial to fluvi‐
al facies stacked in amalgamated bodies developed under a semiarid climate [51]. It uncon‐
formable overlies the Río Blanco Formation and is overlain by Middle ‐ Late Jurassic basalts of
Punta de las Bardas Formation.

Figure 10. Proposed stratigraphic column of the Cacheuta sub‐basin infilling showing the rifting episodes
recognized in this work for the whole Cuyana Basin. Modified from Kokogian [13]. To the right, detail
photograph of the Cacheuta lacustrine facies and the transition to the postrift facies of Rio Blanco Formation.
118 Tectonics – Recent Advances

Paleofloristic analysis on the macrofloras preserved nearly in the whole column indicate the
presence of subtropical floras adapted to seasonally dry climatic conditions. Moreover, the
flora preserved in the Potrerillos and Cacheuta beds are evergreen forests in contrast with
that represented in the ovelying Rio Blanco Formation that is characterized by a deciduous
forest. This change indicated a climatic shift to dryer conditions during the deposition of the
fluvial redbeds of the Rio Blanco unit e.g. [27].

Recently, U‐Pb SHRIMP ages on tuffaceous beds of the top of Cerro de las Cabras and the
base of the Potrerillos formations have constrained the initial infilling of the Cacheuta
subbasin (Synrift I) and the beginning of the Synrift II to the early Anisian and the Anisian‐
Ladinian boundary, respectively [29, 33]. This age is coincident with that recently obtained
from the lacustrine beds at the top of the first synrift stage in the northernmost exposures of
the Cuyana Basin at Puntudo [34].

6. Geodynamic and tectonosedimentary evolution of the Cuyana Basin


The depositional sequences of the Cuyana Basin resulted from the complex interaction of the
supply of sediments, the availability of the accommodation space (both with tectonic com‐
ponents), sea level variations and climate variations. The main control on basin geometry
was the deformation field resulted from the tectonic activity along the margin of Gondwana.
In this sense, fault geometry provided the final morphology of the trough and its resulting
subsidence, which in term controlled the sedimentation rate, grain size, channel migration,
avulsion episodes and the development of lacustrine environments.
The Gondwana continental margin has been created as a result of the accretion and amal‐
gamation of different continental blocks and allochthonous terrains to the South American
proto‐margin since the Proterozoic times up to even the Early Paleozoic e.g. [52 to 55] (Fig‐
ure 11). This tectonic cycle terminated as a consequence of the Sanrafaelic orogeny [56] with
the formation of a magmatic arc in the Late Carboniferous‐Early Permian times [58, 57].
Following this deformational event and during Middle Permian to Early Triassic times, an
important silicic magmatism episode occurred associated with an extensional period. It was
named Choiyoi Group [58] and was interpreted as the result of the collapsed of the Permian
orogen e.g. [36, 4, 38, 59] probably because of the declining and/or cease of the long‐lived
subduction of the Panthalassan margin [59].

Additionally, and on the basis of the strain analysis, several authors suggested that these
extensional forces could have been related to the beginning of the Pangea breakup e.g. [36,
60, 61]. Zerfass [62], proposed a regional uplift due to oblique compression for Lower to
Middle Triassic times which is somehow in agreement with the geodynamic contest
envisioned by Llambías [63] for the basement of the Neuquén Basin, except for the
thermomechanical parameters that could have triggered this upwelling, for these latter
authors suggest the probable influence of a thermal rise and/or slab brake off. According to
Zerfass [62] scheme, the basal Talampaya and Tarjados formations, corresponding to the
Ischigualasto Basin (see Figure 1), would constitute together a second‐order sequence and
could be separated of the tectonic history of the basin. By MiddleTriassic ‐ Early Jurassic
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 119

times however, the inception of a new juvenile magmatic arc associated with the renewal of
the subduction processes, added extra intraplate extensional forces [63 to 65]. Consequently,
rapidly subsiding, fault‐bounded, narrow back arc‐related troughs were formed and
arranged in an échelon pattern like the Bermejo, Ischigualasto‐Villa Unión‐Marayes and
Cuyana, basins (see Figure 1) e.g. [11, 32, 66]. Charrier [67], Criado Roque [68], Spalletti [26]
among others, and more recently, Milani & De Wit [69] have proposed a transtensional
origin for Triassic basins related to a sinistral shear zone along in the western plate margin.
According to the latter authors, this transtensional regime could be related with the
geodynamic processes occurring in a still active Gondwanides orogen. By Early to Middle
Jurassic times subduction was completely restored and a huge extensional regime were
developed thereafter in the backarc [3, 70, 71].

Figure 11. Regional Map of the distribution of the Triassic‐Jurassic basins of Argentina. Arrows indi‐
cate stress direction, in blue main compressional regime during Upper Carboniferous‐Lower Permian
and in red oblique extensional regime during Middle to Upper Triassic times. Right, pre‐break‐up
Gondwana configuration taken from Corti [82]. See red asterisk for basin location.

The Cuyana Basin is a NW trending rift whose Triassic record reaches 3700 meters [9]. Its
origin has been associated with an oblique extensional regime (Az 35°) [9, 72] which reac‐
tivated the northwest‐striking lithosferic weakness related to the ancient Gondwanan su‐
tures e.g. [3, 53, 73] (Figure 11). In this scenario, work hardening of the lithosphere by re‐
peated deformational events during the Palaeozoic along the Gondwanides led to an in‐
crease in the lithospheric strength and together with the obliquity of the extensional regime,
gave place to a north to south trend o basin formation. Accordingly, the Bermejo Basin and
120 Tectonics – Recent Advances

Ischigualasto‐Villa Unión‐Marayes basin opened first in Early‐Late Triassic [74] followed by


the Cuyana Basin, during the Middle Triassic [32] and south‐westward in northern Patago‐
nia, the Neuquén Basin during the Late Triassic–Early Jurassic times e.g. [63]. Similar exten‐
sional events have been recognized along the cratonic areas of southern Brazil by Hack‐
spacher [75] and Zerfass [62].

An oblique subduction along the Panthalassan margin has been proposed by Martin [76] for
the Permian–Triassic times which account for the obliquity of the extension as lateral upper
crust shear. Considering that the Gondwanides basement is composed of a series of different
crustal domains with distinct mechanical characteristics [77] and which together are
different from that of the cratonic region, the stress transmission from the this belt to the
backarc and foreland regions could have had shear components with different magnitude
along strike. In this sense, several authors like Llambías & Sato [59, 38, 78], among others,
proposed that during Permian times continental crust block rotation along the Gondwana
margin took place leading to a reorganization of plate boundaries and plate kinematics,
especially along the Triassic.

In the particular case of the Cuyana Basin, Middle Triassic extension took place under a
relative high thermal flux of about 70 mv/m2 [31, 39] which affected a normal to thinned crust
of 30 km thick. Magmatism was due to decompression and melting of the asthenospheric
mantle driven by intraplate stresses and lithospheric thinning and probable by the influence of
a thermal rise during slab brake off [63, 79]. It constituted an alkaline bimodal suite initially
acidic because of crustal melting, and then basaltic and probably sourced from the base of the
lithosphere while subduction was in progress by Middle to Upper Triassic [39].

Faults reflect the historical sequence of changing stress regime during Triassic times in the
Gondwana margin. Some of these structures reactivated ancient preexisting weakness zones
like the Paleozoic sutures within the Gondwamides. In the beginning of extension the lim‐
ited crustal extension led to the formation of a series of restricted half‐grabens separated by
transfer faults which apparently follow a the Paleozoic fabric e.g. [9, 12, 37] or intrabasinal
highs. Extension persisted up to Jurassic (Cretaceous?) times with at least three remarkable
pulses of fault reactivation which are well preserved in the Rincón Blanco subasin; while in
the Cacheuta depocenter there are still doubts. The Barrancas Formation has been consid‐
ered of Late Triassic age by Rolleri & Garrasino [40] using cyclostratigraphic concepts or
Early to Middle Jurassic age by Regairaz [41] on the basis of field relationships. The exist‐
ence of Late Jurassic to Early Cretaceous basalts (Punta de las Bardas Formation) over un‐
conformable overlying this unit made several authors to considered Barrancas Formation as
Jurassic. In any case, enhanced extension during westward migration of the magmatic arc
[63] and the final insertion of the subduction complex with negative trench roll back and
probable extra thermal flux governed the deposition of this unit and de evolution of Punta
de las Bardas basalts. Further north, the Marachemill Unit could be considered the initial
signs of this renewed extensional event of the Gondwana margin.

The Cuyana Basin was filled with a tectonically induced second‐order thick pile of continen‐
tal deposits arranged into three third‐order sequences. These are composed of alluvial, flu‐
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 121

vial, lacustrine, deltaic and pyroclastic deposits [1, 32]. The switch from major fluvial to
lacustrine environments along the basin is interpreted to reflect a change in tectonic activity.
These third‐order depositional sequences can be correlated with those of the Santa María
Supersequence (Paraná Basin‐Southern Brazil) [62].

Depocenters in the Cuyana Basin are half‐grabens roughly triangular in cross section. The
border fault (BF) consists of a network of mainly normal to oblique‐slip faults which are in
plan view soft linked and separated by strike‐slip minor faults (Cacheuta and Rincon Blan‐
co/Santa Clara) or by breached relay ramps (Rincon Blanco‐Puntudo). Only between the
southern and northern segments are inversion of segment polarities. Fault‐displacement
folds were formed and locally influenced sedimentation, with synrift units thickening in the
synclinal lows and thinning onto the highs in the footwalls. This situation produced the
wedge‐shaped sedimentary units that can be traced through the depocenters in the field and
subsurface. The high dip angle of the border faults controlled the significant amount of
throw, especially of the Cerro Amarillo area (Rincón Blanco depocenter) with an initial infil‐
ling of 1200 meters. The footwall of the border faults were uplifted in response to absolute
upward motion coupled with the isostatic unloading. In the particular case of the Rincón
Blanco depocenter this topographic high, with the slope to the east, gave place to the erosion
of the basal Choiyoi deposits, so common in the Puntudo and Cacheuta depocenters. These
shoulders prevented sediment inflow and streams entered the basin along the hinged mar‐
gin but also axially mostly sourced by the transfers (Agua de Los Pajaritos and Barreal for‐
mations). The rift was not connected with the sea and so eustasy played no role in the basin
infilling evolution, drainage systems were in fact controlled by local (mostly lake) base lev‐
els. In this scenario, the relationships among incremental accommodation space (mostly
associated with tectonic subsidence), sediment+water supply and short‐time climatic influ‐
ences determined which depositional system predominated. When the sediment supply
exceeded the incremental accommodation space (basin overfilled), during first stages of
rifting, alluvial to braidplain deposition predominated with strata progressively onlapping
the hanging‐wall (Cerro Puntudo Formation, Ciénaga Redonda and Río Mendoza for‐
mations). These deposits occupied a narrow band close to the main faults and transfers, and
were coeval with the latest tholeitic magmatism (Choiyoi Group) from a Cisuralian Mag‐
matic arc [33,39].

The gradual faults growth in length and displacement drove depocenters to increase in
depth, length, and width and thus, the incremental accommodation space through time. In
this way, they were underfilled with small drainage systems entering from footwall region
between border faults or from the fault tips where footwall uplift was minimal. The main
fluvial systems were mostly axial or longitudinal (Cerro Amarillo, Cerro de Las Cabras and
Cerro Puntudo formations). When the incremental accommodation space significantly ex‐
ceeded the sediment supply and water input, shallow hydrologically closed lacustrine dep‐
osition (playa‐type) predominated (Rincón Blanco and Cacheuta depocenters) and a car‐
bonate‐rich lake in Puntudo depocenter all of them with interbedded tuffs. These lakes were
located close to the faults and subjected to climatic base‐level fall‐to‐rise turnarounds and
thus show a marked cyclicity. Climate was semiarid and seasonally humid. When the in‐
122 Tectonics – Recent Advances

cremental sedimentation rate equalled the accommodation space rate climatic induced or,
less probable, tectonic induced (early post‐rift?) fluvial systems developed over these lakes.
According to U‐Pb zircon dating this tectono‐sequence can be constrained to the base of the
Anisian (early Middle Triassic) [32].

In Middle Triassic (Rift II) a significant reactivation of the faults gave place to the develop‐
ment of a regional unconformity that can be correlated across the three depocenters. It has
been interpreted as being an intra‐Triassic tectonic event, probably associated with the ac‐
celeration of the subduction rate on the west margin as a consequence of the reinsertion of
the convergence [63]. It was responsible of the high sediment supply that gave place to the
deposition of the alluvial‐fluvial facies of El Relincho, Panul and the base of Potrerillos for‐
mations (Synrift II initiation). Fluvial to deep lacustrine facies developed when the accom‐
modation space rate succeeds the sedimentation rate in a basin that widened significantly as
it is suggested by numerous onlaps observed in all the troughs. Climate was sub‐tropical
and seasonally dry and lakes were balanced and deep (meromictic) with deltaic and fluvial
deposition occurring around the margins. Climatic induced lake‐level cycles are clearly seen
in the stacked parasequences bound by lacustrine flooding surfaces which are superim‐
posed on variable rates of subsidence of the rift border fault zone. Upward, the lacustrine–
fluvial transition resulted from a decrease in incremental accommodation space and/or an
increase in sediment supply once extension slowed in the post‐rift stage (Casa de Piedra,
Cortaderita and El Alcázar formations). The shoaling tendency of the lakes during this stage
was due mainly to the basin growth which forced to spread the water over broader regions,
drowning extended regions with the development of palustrine environments. Important
ash fall tuffs and reworked volcaniclastics interfinger these postrift facies along the whole
basin. This second tectono‐sequence corresponds to the Anisian (upper Middle Triassic [32].

Another extensional event took place in Early Late Triassic (Rift III) [32]. At present, best
exposures associated with this event are found in the Rincón Blanco Subbasin (Cepeda and
Marachemill formations) [1, 32]. In the Cacheuta depocenter a tectosedimentary similar unit
is represented by the Barrancas Formation, firstly assigned to Upper Triassic by [10].
However and on the basis of field relationships, Regairaz [41] proposed a Lower to Middle
Jurassic age. Presently, there is no absolute dating for this unit but several authors like
Rolleri & Fernandez Garrasino [40] using the sequence stratigraphy concepts could
demonstrate its Triassic affinities. On a regional scale, the lack of Jurassic ‐ Cretaceous strata
in the northern portion of the Cuyana Basin (Puntudo and Rincón Blanco depocenters)
likely suggests that it stopped subsiding while the southern portion of the basin were still
active contemporaneously with the opening of the Neuquén basin with the deposition of the
Barrancas and Punta de las Bardas formations [32]. Yet, it is worth mentioning that Jurassic
levels have been described in nearby latitude in the Precordillera by Coughlin [80] and
Milana et al. [81] which suggest that further studies are needed. In Southernmost Brazil and
Uruguay, the Triassic Sanga do Cabral Supersequence was divided in three sequences [62]
that can be also correlated with these stages. Proximal fans and braided ephemeral to
perennial streams of semiarid, seasonally humid (?) climatic regime developed as a
consequence of the lost of accommodation space and a high sediment supply. This tectono‐
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 123

sequence was assigned an early Carnian age (lower Late Triassic) in the Rincón Blanco
trough by Barredo [32] but taking into account the evolution of the Barrancas Formation and
Punta de las Bardas Basalts, of the southern Cacheuta depocenter, it could have spanned up
to Jurassic (Cretaceous?) times. Subduction was definitely restored during upper Late
Triassic‐Lower Jurassic under extensional conditions associated with the westward
migration of the arc e.g. [63]. The new extensional regime would have governed the
evolution of the upper section of the Cuyana Basin whose final thermal relaxation should
have occurred during Middle‐Late Jurassic and Cretaceous times and aborted during
Tertiary times when the lithosphere undergone flexural subsidence induced by the Andean
orogenic overloading and by sediment charge.

7. Conclusion
The Gondwana continental margin has been created as a result of the accretion and amal‐
gamation of different continental blocks and allochthonous terrains to the South American
proto‐margin during Paleozoic. In this scenario, Triassic continental rifts of Argentina set‐
tled sometimes controlled by the inherited fabric. Hence, a deep geodynamic analysis is
necessary if we want to understand the evolutionary history of these basins. The uplifting
and subsidence of different portions of the crust will control the insertion and development
of sedimentary basins and the regional distribution of their infilling. In the Cuyana Basin,
depositional sequences resulted from a complex interaction of the supply of sediments, the
availability of the accommodation space (both with tectonic components), sea level varia‐
tions (which also may have a significant tectonic influence) and climate variations, being the
first‐order control on basin geometry the deformation field resulted from the tectonic activi‐
ty. Its origin has been associated with an oblique extensional regime which in some cases
reactivated the northwest‐striking lithosferic weakness related to the ancient Gondwanan
sutures. The resulting sedimentary environments display significant differences across sub‐
basins which made bio and lithostratigraphic correlations quite imprecise. Still, recent iso‐
topic dating permitted to arrive to an enhanced evolutionary model when combined with
thermo‐mechanical analysis, cyclostratigraphy and tectonostrigraphic tools. A more precise
model of the Cuyana Basin evolution now predicts that it was filled with a tectonically in‐
duced second‐order thick pile of continental deposits arranged into three third‐order se‐
quences. These are composed of alluvial, fluvial, lacustrine, deltaic and pyroclastic deposits
separated by key stratigraphic surfaces or sequence boundaries resulting from lacustrine
flooding and/or forced regressive surfaces. The Synrift I contains volcaniclastic, alluvial‐
fluvial and saline lake deposits when accommodation exceeded sediment and water input.
The Synrift II starts with alluvial–fluvial followed by deep lacustrine and deltaic environ‐
ments under wetter climatic conditions. The Synrift III encompasses volcanic and volcani‐
clastic alluvial and fluvial dominated environments developed under semiarid conditions.
These three stages are associated with the evolution of the Gondwana margin during Per‐
mo‐Triassic times, when lithosphere undergone a wide spread extensional regime due to the
almost cease of subduction, the collapse of the ancient Permian orogen and the reorganiza‐
tion of intraplate stresses during Pangea breackup (Rift I and II). The inception of a new
124 Tectonics – Recent Advances

juvenile magmatic arc associated with the renewal of the subduction processes, added extra
intraplate extensional forces which gave place to a third regional scale event, the Rift III,
which begun in Late Triassic and extended up to Middle Jurassic. The final thermal relaxa‐
tion took place during Cretaceous when lithosphere undergone flexural subsidence induced
by the Andean uplift.

Author details
Silvia Patricia Barredo
Department of Petroleum Engeneering, Instituto Tecnológico de Buenos Aires (ITBA), Buenos Aires,
Argentina

Acknowledgement
The author wants to acknowledge discussions specially with Dr Victor Ramos and Don Dr
Pedro Stipanicic over the development of her PhD Thesis. My views from the Triassic
systems highly benefited from their critical observations.

My particular thanks to Laura Giambiagi and Maisa Tunik for their assistance in the field, to
Claudia Marsicano and Guillermo Ottone for sharing their paleontological knowledge and
for providing great inputs to this study, and Luis Stinco for his critical view about the
manuscript. To all my colleagues at conferences and meetings about extensional tectonics
and hydrocarbon exploration. A great part of this study has been supported by UBACYT
X182: 2008‐2010 (VAR.) and PIP CONICET 5120 (EGO.). Additional financial support was
provided by the Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET).
Finally, I want to highlight the great support from the Instituto Tecnológico de Buenos Aires
(ITBA) to my research and for providing the means to perform it through ʺProyecto ITBA‐
70ʺ Geodinámica de cuencas productivas y de potencial generadorʺ.

8. References
[1] Barredo SP, Ramos VA (2010) Características tectónicas y tectosedimentarias del Hemi‐
graben Rincón Blanco: Una Síntesis. Revista de la Asociación Geológica Argentina 65
(1): 133‐145.
[2] Allen PA, Allen JR (2005) Basin Analysis: principles and applications. Second edition.
Blackwell Scietific Publication, 549 p., Oxford.
[3] Uliana MA, Biddle KT (1988) Mesozoic‐Cenozoic paleogeographic and paleodynamic
evolution of southern South America. Revista Brasileira de Geociências, 18: 172‐190.
[4] Zeil W (1981) Volcanism and geodynamics at the turn of the Paleozoic to Mesozoic in
the Central and Southern andes. Zentralblatt für Geologie und Paläontologie 1(3/4): 298‐
318. Stutgart.
[5] Keidel J (1916) La geología de las Sierras de la Provincia de Buenos Aires y sus relacio‐
nes con las montañas de Sudáfrica y Los Andes. Ministerio de Agricultura de la Nación,
Sección Geología, Mineralogía y Minería, anales, XI (3): 1‐78.
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 125

[6] duToit AL (1937) Our wandering continents. Oliver and Boyd: 336 p. Edinburgh.
[7] Legarreta L, Kokogian DA, Dellape D (1993) Estructuración terciaria de la Cuenca Cu‐
yana: ¿Cuánto de inversión tectónica? Revista de la Asociación Geológica Argentina 47:
83‐86.
[8] Barredo SP (2005) Implicancias estratigráficas de la evolución de las fallas normales del
hemigraben Rincón Blanco, cierre norte de la cuenca Cuyana, provincia de San Juan. In:
Cazau LB, editor. Electronic proceedings of the VI Congreso de Exploración y Desarro‐
llo de Hidrocarburos: 9 p.
[9] Barredo SP (2004). Análisis estructural y tectosedimentario de la subcuenca de Rincón
Blanco, Precordillera Occidental, provincia de San Juan. [PhD Thesis]. Buenos Aires,
Argentina. Unpublished: 325 p.
[10] Rolleri EO, Criado Roque P (1968) La cuenca triásica del norte de Mendoza.
Proceedings of the III Jornadas Gelógicas Argentinas (Comodoro Rivadavia, 1966) 1: 1‐
79.
[11] Ramos V.A., Kay S.M. (1991) Triassic rifting and associated basalts in the Cuyo Basin,
central Argentina. In: Harmon, R.S. Rapela C.W., editors. Andean magmatism and its
tectonic setting. Geological Society of America Special Paper 265, 79‐91.
[12] Ramos VA (1992) Control geotectónico de las cuencas triásicas de Cuyo. Boletín de
Informaciones Petroleras, 5: 2‐9.
[13] Kokogian DA, Seveso FF, Mosquera A (1993) Las secuencias sedimentarias triásicas. In:
Ramos VA, editor. Relatorio XII Congreso Geología Argentina and II Congreso de Ex‐
ploración de Hidrocarburos, Geología y Recursos Naturales de Mendoza (Buenos Ai‐
res): 65‐78.
[14] Groeber P, Stipanicic PN (1953) Triásico. In: Groeber P, editor. Mesozoico. Geografía de
la República Argentina: 1‐141 (Sociedad Argentina de Estudios Geográficos, GAEA, 2.
Buenos Aires.
[15] Borrello A.V., Cuerda A.J. (1965) Grupo Rincón Blanco (Triásico San Juan). Comisión de
Investigaciones Científicas. Provincia Buenos Aires, Notas: 2 (10):3‐20. La Plata.
[16] Quartino BJ, Zardini RA, Amos AJ (1971) Estudio y exploración geológica de la Región
de Barreal‐Calingasta. Provincia de San Juan, República Argentina. Asociación Geológi‐
ca Argentina, Monografía 1: 184 p.
[17] Stipanicic PN. (1972) Cuenca triásica de Barreal. In: Leanza AF, editor. Geología
Regional Argentina, Academia Nacional de Ciencias de Córdoba: 537‐566.
[18] Stipanicic PN (1979) El Triásico del Valle del río Los Patos (Provincia de San Juan).
Segundo Simposio de Geología Regional Argentina, Academia Nacional de Ciencias de
Córdoba, II: 695‐744.
[19] López‐Gamundí OR (1994) Facies distribution in an asymmetric half‐graben: the north‐
ern Cuyo basin (Triassic), western Argentina. 14th International Sedimentological Con‐
gress, Abstracts: 6‐7. Recife
[20] Spalletti LA (1999) Cuencas triásicas del Oeste Argentino: origen y evolución. Acta
Geeológica Hispánica, 32 (1‐2)(1997): 29‐50.
[21] Barredo SP, Stipanicic PN (2002) El Grupo Rincón Blanco. In: Stipanicic PN, Marsicano
CA, editors. Léxico Estratigráfico de la Argentina III: 870 p.
126 Tectonics – Recent Advances

[22] Yrigoyen MR, Stover LW (1969). La palinología como elemento de correlación del Triá‐
sico en la cuenca Cuyana. proceedings of the IV Jornadas Geológicas argentina 2: 427‐
447. Buenos Aires.
[23] Strelkov EE, Alvarez LA (1984). Análisis estratigráfico y evolutivo de la cuenca triásica
mendocina ‐ sanjuanina. Proceedings of the IX Congreso Geología Argentina III: 115‐
130. Bariloche.
[24] Sessarego HL (1988) Estratigrafía de las secuencias epiclásticas devónicas a triásicas,
aflorantes al norte del río San Juan y al oeste de las sierras del Tigre, Provincia de San
Juan. [PhD Thesis]. Buenos Aires, Argentina. unpublished: 330 p.
[25] Stipanicic PN, Bonetti MIR (1969) Consideraciones sobre la cronología de los terrenos
triásicos argentinos. Proceedings of the I Simposio Internacional Estratigrafía y Paleon‐
tología del Gondwana, Mar del Plata. UNESCO, Ciencias de la Tierra 2: 1081‐1120. Pa‐
rís.
[26] Spalletti LA (2001) Modelo de sedimentación fluvial y lacustre en el margen pasivo de
un hemigraben: el Triásico de la Precordillera occidental de San Juan, República Argen‐
tina. Revista de la Asociación Geológica Argentina 56 (2): 189‐210.
[27] Artabe AE, Morel EM, Spalletti LA (2001). Paleoecología de las de las floras triásicas
argentinas. In: Artabe AE, Morel EM, Zamuner AE, editors. El Sistema Triásico en la
Argentina. Fundación Museo de La Plata “Fracisco P. Moreno”: 199–225. Argentina.
[28] Morel EM, Artabe AE (1993) Floras mesozoicas. In: Ramos VA, editor. Geología y Re‐
cursos Naturales de Mendoza. Proceedings of the XII Congreso Geológico Argentino y
II Congreso de Exploración de Hidrocarburos (Mendoza), Relatorio, 2(10): 317‐324. Ar‐
gentina.
[29] Spalletti LA, Fanning CM, Rapela CW (2008) Dating the Triassic continental rift in the
southern Andes: the Potrerillos Formation, Cuyo Basin, Argentina. Geologica Acta 6,
267‐283.
[30] Marsicano CA, Barredo SP (2004). A Triassic tetrapod footprint assemblage from south‐
ern South America: palaeobiogeographical and evolutionary implications. Palaeogeog‐
raphy, Palaeoclimatology, Palaeoecology 203, 313‐335.
[31] Zamora G, Cervera M, Barredo SP (2008) Geología y potencial Petrolero de un bolsón
intermontano: Bloque Tamberías, Provincia de San Juan. In: Schiuma M, editor. Traba‐
jos técnicos, VII Congreso de Exploración y Desarrollo de Hidrocarburos: 397‐407. Ar‐
gentina.
[32] Barredo SP, Chemale F, Ávila JN, Marsicano C, Ottone G, Ramos VA (2012) U‐Pb
SHRIMP ages of the Rincón Blanco northern Cuyo rift, Argentina. Gondwana Research
(21): 624‐636. DOI: 10.1016/J.Gr.2011.05.016.
[33] Ávila JN, Chemale F, Mallmann G, Kawashita K (2006) Combined stratigraphic an
isotopic studies of Triassic strata, Cuyo Basin, Argentine Precordillera. Geologial Socie‐
ty of America Bulletin v. 118: 1088‐1098. Doi: 101130/B25893.1
[34] Mancuso AC, Chemale F, Barredo SP, Ávila JN, Ottone G, Marsicano C (2010) Age
constraints for the northernmost outcrops of the Triassic Cuyana Basin, Argentina.
Journal of South American Earth Sciences, 30 (2010): 97‐103. doi:
10.1016/J.Jsames.2010.03.001
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 127

[35] Stipanicic PN, Marsicano CA (2002). Léxico Estratigráfico de la Argentina, Volumen III:
870 p. Asociación Geológica Argentina. Buenos Aires.
[36] Uliana MA, Biddle KT, Cerdan J (1989) Mesozoic‐Cenozoic paleogeographic and geo‐
dynamic evolution of southern South America. Revista Brasileira de Geociencias 18:
172‐190.
[37] Japas MS, Cortés JM, Pasini M (2008) Tectónica extensional triásica en el sector norte de
la Cuenca Cuyana: primeros datos cinemáticos. Revista de la Asociación Geológica Ar‐
gentina 63 (2): 213‐222.
[38] Kleiman LE, Japas MS (2009) The Choiyoi volcanic province at 34°S ‐ 36° S ( San Rafael,
Mendoza, Argentina): Implications for the Late Paleozoic evolution of the Southwestern
margin of Gondwana. Tectonophysics 473: 283‐299.
[39] Barredo SP, Martínez A (2008) Secuencias piroclásticas triásicas intercaladas en la For‐
mación Ciénaga Redonda, Rincón Blanco, Provincia de San Juan y su vinculación con el
ciclo magmático gondwánico del Grupo Choiyoi. Proceedings of the 12th Reunión de
Sedimentología: 91‐92. Buenos Aires
[40] Rolleri EO, Fernández Garrasino C (1979) Comarca septentrional de Mendoza. Proce‐
edings of the Simposio de Geología Regional Argentina, Academia Nacional de Cien‐
cias de Córdoba I: 771‐809. Argentina. Córdoba.
[41] Regairaz AC (1970) Contribución al conocimiento de las discordancias en el área de las
Huayquerías, Mendoza, Argentina. Proceedings of the IV Jornadas Geológicas
Argentinas, Mendoza, 1969, 2: 243‐254.
[42] Mombrú CA (1973) Observaciones geológicas en el Valle de Calingasta‐Tocota. Provin‐
cia de San Juan. YPF Unpublished: 50 p. Buenos Aires.
[43] Japas MS, Salvarredi, J, Kleiman LE (2005) Self‐similar behaviour of Triassic rifting in
San Rafael, Mendoza, Argentina. Proceedings of the Gondwana 12: 210. Argentina.
[44] Rossa N, Mendoza N (1999) Manifestaciones volcánicas en la cuenca triásica de Barreal
Calingasta, San Juan. Proceedings of the XIV Congreso Geologico Argentino II: 171 ‐
174. Salta, Argentina.
[45] Martínez A, Barredo SP, Giambiagi L (2006) Modelo Geodinámico para la evolución
magmática Permo‐Triásica entre Los 32º Y 34º LS, Cordillera Frontal de Mendoza, Ar‐
gentina. Proceedings of the XIII Reunión de Tectónica ‐ San Luis 4 p. ISBN 978‐9871031‐
49‐8.
[46] Barredo SP, Tunik M, Pettinari G, Giambiagi L, Zamora G (2010) A new stratigraphic
synthesis of the Agua de Los Pajaritos depocenter, flexural margin of the Rincón Blanco
Subbasin. Proceedings of the 18Th International Sedimentological Congress.
[47] Baraldo JA, Guerstein PG (1984) Nuevo ordenamiento estratigráfico para el Triásico de
Hilario (Calingasta, San Juan). Proceedings of the 9th Congreso Geológico Argentino 1:
79‐94. San Juan. Argentina.
[48] Bonati S, Barredo SP, Zamora Balcarce G, Cervera M (2008) Análisis tectosedimentario
preliminar del Grupo Barreal, cierre norte de la Cuenca Cuyana, provincia de San Juan.
In: Schiuma M, editor. Trabajos Técnicos, VII Congreso de Exploración y Desarrollo de
Hidrocarburos: 409‐420. Mar del Plata. Argentina.
128 Tectonics – Recent Advances

[49] Massabie AC (1986) Filón Capa Paramillos de Uspallata, su caracterización geológica y


edad, Paramillos de Uspallata, Mendoza. Primeras Jornadas sobre Geología de Precor‐
dillera. Asociación Geológica Argentina Serie A (2): 71‐76. Buenos Aires.
[50] Dellape DA, Hegedus AG (1993) Inversión estructural de la cuenca Cuyana y su rela‐
ción con las acumulaciones de hidrocarburos. Proceedings of the XII Congreso Geología
Argentina y II Congreso de Exploración de Hidrocarburos III: 211‐218, Mendoza, Ar‐
gentina.
[51] Zencich S, Villar HJ, Boggetti D (2008) Sistema petrolero Cacheuta‐Barrancas de la
Cuenca Cuyana, provincial de Mendoza, Argentina. In: Cruz CE, Rodríguez JF,
Hetchem JJ, Villlar J, editors. Sistemas petroleros de las cuencas andinas. VII Congreso
de Exploración y Desarrollo de Hidrocarburos: 109‐134. Argentina.
[52] Ramos VA (1988) The tectonics of the Central Andes, 30 to 33 S latitude. In: Clark S,
Burchfiel D. editors. Process in Continental Lithospheric Deformation, Geological Socie‐
ty America, Special Paper 218: 31‐54. Boulder.
[53] Ramos VA (2000) The Southern central Andes. In: Cordani UG, Milani EJ, Thomaz Filho
A, Campos DA, editors: Tectonic Evolution of South America. Proceedings of the 31 st
International geological Congress: 561‐604. Brasil.
[54] Dalla Salda, de Barrio R, Echeveste H, Fernández, R (2005) El basamento de las Sierras
de Tandilia. In: de barrio R, Etcheverry R, Caballé M, LLambías, E, editors. Geología y
Recursos Minerales de la Provincia de Buenos Aires. XVI Congreso geológico Argenti‐
no. Relatorio: 31‐50. La Plata, Argentina.
[55] Pankhurst RJ, Rapela CW, Saavedra J, Baldo E, Dahlquist J, Pascua I., Fanning CM
(1998). The Famatinian magmatic arc in the central Sierras Pampeanas: an Early to Mid‐
Ordovician continental arc on the Gondwana margin. In: Pankhurst RJ, Rapela CW, edi‐
tors. The Proto‐Andean Margin of Gondwana. Geological Society, London, Special Pub‐
lications, 142:343–367.
[56] Caminos R, Azcuy CL (1996) Tectonismo y diastrofismo. 3. Fases diastróficas neopaleo‐
zoicas. In: Archangelsky S, editor. El sistema Pérmico en la República Argentina y en la
República Oriental del Uruguay. Academia Nacional de Ciencias: 265‐274. Córdoba,
Argentina.
[57] Relledo S, Charrier S (1994) Evolución del basamento paleozoico en el área Punta Cladi‐
tas, región de Coquimbo, Chile, (31°‐32° S). Revista Geológica de Chile 21 (1): 55‐69.
[58] Stipanicic PN, Rodrigo F, Baulies OL, Martinez C G (1968) Las formaciones preseno‐
nianas en el denominado Macizo Nordpatagónico y regiones adyacentes. Revista de la
Asociación Geológica Argentina 23: 76‐98..
[59] Llambías E, Sato A (1995).El batolito de Colangüil transición entre orogénesis y anoro‐
génesis. Revista de la Asociación Geológica Argentina 50 (1‐4): 111‐131.
[60] Charrier R, Pinto L, Rodríguez MP (2007). Tectonostratigraphic evolution of the Andean
Origen in Chile. In Moreno T, Gibbons W, editors. The Geology of Chile. The Geological
Society: 21‐114, London.
[61] Giambiagi L, Bechis F, García V, Clark A (2009). Temporal and spatial relationship
between thick‐ and thin‐skinned deformation in the thrust front of the Malargüe fold
and thrust belt, Southern Central Antes. Tectonophysics 459: 123‐139.
Geodynamic and Tectonostratigrafic Study of a Continental Rift: The Triassic Cuyana Basin, Argentina 129

[62] Zefrass H, Chemale FJr, Schultz C L, Lavina E (2004). Tectonics and sedimentation in
Southern South America during Triassic. Sedimentary Geology, 166: 265 – 292.
[63] Llambías EJ, Leanza HA, Carbone O (2007) Evolución Tectono‐magmática durante el
Pérmico al Jurásico temprano en la cordillera del Viento (37º05’S – 37º15’S): Nuevas ev‐
idencias geológicas y geoquímicas del inicio de la Cuenca Neuquina. Revista de la Aso‐
ciación Geológica Argentina, 62 (2): 217‐235.
[64] Hervé F, Fanning CF (2001) late Triassic detritical zircons in meta‐turbidites of the
Chonos Metamorphic Complex, Southern Chile. Revista Geológica de Chile 28(1): 98‐
104. Chile.
[65] Rapela CW, Pankhurst RJ, Fanning CM, Greco LE (2003). Basement evolution of the
Sierra de la Ventana Fold Belt: new evidence for Cambrian continental rifting along the
southern margin of Gondwana. Journal of the Geological Society, London, 160: 613‐628.
[66] Rincón MF, Barredo SP, Zunino J, Salinas A, Reinante SME, Manoni R (2011). Síntensis
general de los bolsones intermontanos de San Juan y La Rioja. In: Kowlowsky E, Le‐
garreta L, Boll A, editors. Cuencas Sedimentarias Argentinas. XIII Congreso de Ex‐
ploración y Desarrollo de Hidrocarburos: 321‐406. Mar del Plata.
[67] Charrier R (1979) El Triásico de Chile y regiones adyacentes de Argentina. Una recon‐
strucción paleogeográfica y paleoclimática. Comunicaciones 26: 1‐37, Santiago de Chile.
[68] Criado Roque P, Mombrú CA, Ramos VA (1981). Estructura e interpretación tectónica.
In Yrogoyen M, editor. VIII Congreso Geología Argentina, Geología y Recursos Natura‐
les de la Provincia de San Luis. Relatorio: 155‐192, San Luis.
[69] Milani EJ, De Wit MJ (2008) Correlations between the classic Paraná and Cape Karoo
sequences of South America and southern Africa and their basin infills flanking the
Gondwanides: du Toit revisited. Geological Society, London, Special Publications 294:
319‐342.
[70] Somoza R (1996) Geocinemática de América del Sur durante el Cretácico: Su relación
con la evolución del margen pacífico y la apertura del Atlántico Sur. Proceedings of the
XIII Congreso Geológico Argentino 2: 401–402.Buenos Aires, Argentina.
[71] Mpodozis C, Ramos VA (1990) The Andes of Chile and Argentina. In: Ericksen GE,
Cañas Pinochet MT, Reinemud JA, editors. geology of the Andes and its relation to hy‐
drocarbon and mineral resources. Circumpacic Council for Energy and Mineral re‐
sources, Earth Sciences Series11:59‐90. Houston.
[72] Barredo SP, Stinco LP (2010) Geodinámica de las cuencas sedimentarias: Su Importancia
en la localización de sistemas petroleros en Argentina. Revista Petrotecnia. Instituto Ar‐
gentino del Petróleo y del Gas (IAPG) (2): 48‐68.Argentina.
[73] Ramos VA (2009) Anatomy and global context of the Andes: Main geologic features and
the andean orogenic cycle. In: Ka, SM, Ramos VA, Dickinson W, editors. Backbone of
the Americas: Shallow Subduction, Plateau Uplift, and Ridge and Terrane Collision,
Geological Society of América, Memoir 204, p. 31‐65.
[74] Milana JP, Alcober O (1994) Modelo tectosedimentario de la cuenca triásica de Ischigua‐
lasto (San Juan, Argentina). Revista de la Asociación Geológica Argentina 49: 217‐235.
[75] Hackspacher PC, Ribeiro LFB, Ribeiro MCS, Fetter AH, Hadler Neto JC, Tello CES,
Dantas EL (2004) Consolidation and Break‐up of the South American Platform in
130 Tectonics – Recent Advances

Southeastern Brazil: Tectonothermal and Denudation Histories. Gondwana Research 7


(1), 91–101.
[76] Martin MW, Kato TT, Rodríguez C, Godoy E, Duhart P, Mc Donough M, Campos A
(1999) Evolution of the late Paleozoic accretionary complex and overlying forearc‐
magmatic arc, south central Chile (38°‐41°S): constraints for the tectonic setting along
the southwestern margin of Gondwana. Tectonics, 18 (4): 582‐605.
[77] Vaughan APM, Leat PT, Pankhurst RJ, editors (2005) Terrane Processes at the margins
of Gondwana. Geological Society, London, Special Publications, 246.
[78] Visser JNJ, Praekelt HE (1998). Late Palaeozoic crustal block rotations within the
Gondwana sector of Pangea. Tectonophysics 287, 201–212.
[79] Pankhurst RJ, Rapela C, Fanning CM, Márquez M (2006) Gondwanide continental colli‐
sion and the origin of Patagonia. Earth‐Science Reviews 76, 235–257.
[80] Coughlin TJ (2000). Linked orogen‐oblique fault zones in the Central Argentine Andes:
Implications for Andean orogenesis and metallogenesis. [PhD Thesis]. University of
Queensland, Queensland. Unpublished: 268 p.
[81] Milana JP, Bercowski F, Jordan TE (2003) Paleoambientes y magnetoestratigrafía del
Neógeno de la Sierra de Mogna, y su relación con la Cuenca de Antepaís Andina. Re‐
vista de la Asociación Geológica Argentina, 58 (3): 447‐473.
[82] Corti G, Wijk J, Cloetingh, S, Morley C (2007) Tectonic inheritance and continental rift
architecture: Numerical and analogue models of the East African Rift system. Tectonics
26, TC6006, doi: 10.1029(2006TC002086
Chapter 6

Role of the NE-SW Hercynian Master Fault


Systems and Associated Lineaments on the
Structuring and Evolution of the Mesozoic
and Cenozoic Basins of the Alpine Margin,
Northern Tunisia

Fetheddine Melki, Taher Zouaghi, Mohamed Ben Chelbi,


Mourad Bédir and Fouad Zargouni

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/50145

1. Introduction
The Mesozoic and Cenozoic evolution of the northern edge of the African margin (Fig. 1), and
particularly the northern Tunisia, fossilized successive paleogeographic and tectonic episodes.
In fact, after rifting and extensional periods, which started at the end of the Paleozoic and
continued during the Mesozoic [1‐6], was settled the Alpine orogeny that results from the
convergence movements between the African and Eurasian plates; it is induced by compres‐
sive tectonic stresses, beginning at least since the Tertiary intervals and probably the Late
Cretaceous [7‐24]. This orogeny has induced, on the Mediterranean edges, many mountains
chains extend from the Apennines at the East to the Betic Cordilleras at the West.

The various geological works established in northern Tunisia [25‐42,18,43‐47], north‐eastern


Algeria [48‐50,23] and in the Siculo‐Tunisian strait [51‐57], demonstrated that the NE‐SW inherit‐
ed fault networks have controlled sedimentation during the Tethyan rifting and have also con‐
trolled the structuring of the central and northern Atlas during the successive tectonic events.

This margin of northern Tunisia, including the Tell and the Tunisian furrow domains (Fig.
2), is limited to the East by the Zaghouan master fault, which appears to have effect on the
sedimentation since the Jurassic [58,59,33]. It is inherited from NE‐SW trending Hercynian
master fault networks (Fig. 2) and their conjugate faults [60,61]. These lineaments corre‐
spond, from SE to NE, to the Zaghouan fault (ZF), the Tunis‐Elles fault (TEF), the

© 2012 Melki et al., licensee InTech. This is an open access chapter distributed under the terms of the
Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
132 Tectonics – Recent Advances

Figure 1. Geological sketch of the Maghreb (modified from Piqué et al. [4] and Frizon de Lamotte et al.
[23]).

El Alia‐Teboursouk fault (ETF), the Ras El Korane‐Thibar fault (RKTF) and the Cap Serrat‐
Ghardimaou fault (CSGF). Movements of these master faults have effects on the sedimen‐
tary deposition and distribution since the Triassic rifting phase up to now.
Our study is mainly focused on northern Atlas (Fig. 2) where the structures tend to be well
exposed and on the Tunisian Tell marked by sealed structures. These interpretations will be
supported by seismic sections calibrated by petroleum well.
Indeed, based on the structural and paleogeographic zonations, developed during various
geological phases in the Tunisian margin, we propose in this paper (i) to demonstrate the
implication of the Hercynian deep structures, in the deformation of the Atlas. They have a
role in the distribution of the sedimentary basins along the southern Tethyan margin dur‐
ing the Mesozoic and the Cenozoic times; (ii) attempt to clarify the kinematics and chrono‐
logical relationship between the Tell and the northern Atlas, by proposing a coherent geo‐
dynamic model since the Tethyan rifting until the Cenozoic contractional periods, pointing
out the role of the NE‐SW tectonic lineaments; (iii) to trace a tectonic pattern of the northern
Tunisian that contributes to a better comprehension of deformations affecting the Tellian
and Atlasic domains in relation with the global geotectonic framework.

2. Stratigraphy
2.1. Triassic
The Triassic outcrops have often abnormal contact with Jurassic, Cretaceous, Paleogene and
Neogene series in several localities of northern Tunisia (e.g. Ras El Korane, Jebel Ichkeul,
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 133

Jebel Lansarine‐Sakkak, Jebel Chehid and Thibar). It is presented either as a diapiric


structures [62, 31, 63, 64], or as sole of overthrust folds [32], or also as salt glacier structures
[65‐69]. The Triassic deposits are characterized by a chaotic aspect; it is consisted of
Gypsiferous, argillaceous, carbonated and locally sandy facies, of varied color (Fig. 3).

In the Bizerte area (e.g. Bechateur), the Triassic appears rather stratiform. It is represented
by black dolomitic limestones in metric layers becoming centimetric at the top of the
succession [18]. The abundant fauna of lamellibranches indicates Carnian‐Norian [70, 71].
The subsurface data show carbonated Triassic as well in Utique (W8 well), the Cap Bon (W9
well) and in the Gulf of Tunis (W5 well). The Late Triassic is predominantly evaporitic; its
thickness is about 2500m.

2.2. Jurassic
The Jurassic outcrops largely in northern Tunisia [72, 33, 73‐78]. Its thickness changes from
700m [76] to 900m [79]. It shows a Tethyan deep sea facies different from the Nara platform
Formation defined by Burollet [26] in the Nara outcrop in central Tunisia. It is dominantly
calcareous series with some marly intercalations (Fig. 3). In the Ichkeul and Ammar
outcrops, the Jurassic shows a thickness series varied from 260m [73] to 480m [80].

At the level of the “Tunisian Dorsale”, the lower limit of the Jurassic is not well defined due
to the Late Triassic‐Lias transition which still not paleontologically characterized [76].

2.3. Early Cretaceous


The Early Cretaceous largely outcrops at the Tunisian furrow [81]. The Jebel Oust lithostrat‐
igraphic section is considered as the most complete and richest of microfauna and
macrofauna of the Tethyan area [82]. This section consists of fossiliferous marls and clays
with micritic limestone intercalations and sandy and quartzitic recurrences (Fig. 3). This
succession is approximately 2500m thick and its sedimentation has occurred in a subsiding
marine environment [83, 82].

In subsurface, in the Gulf of Tunis, the north‐eastern extension of the Tunisian through
show Early Cretaceous deposits composed of clays, marls and limestones with some sandy
layers [24].

2.4. Late Cretaceous


The Late Cretaceous outcropping in the Tunis area [103] is composed from the base to the
top by two lithostratigraphic Formations (Fig. 3); the Aleg Formation (Turonian‐Coniacian)
composed of alternating limestones and marls rich in faunas (100m), marls and clays (50 to
180m). The Abiod Formation (Campanian‐Early Maastrichtian), composed by marls and
limestones; an argillaceous limestone bar of the zone with Globotruncana arca rugosa;
limestone and marl alternations; a limestone bar (80m), marls locally gypsiferous and
alternations of marl and limestone.
134 Tectonics – Recent Advances

Figure 2. Study area location map of Geologic outcrops, cross section, seismic lines and petroleum
wells (geologic outcrops are modified from 1/500.000 geological map of Tunisia; Castany, 1951).

In the Bizerte area, the Abiod Formation outcrop is composed by the first limestones bar
attributed to Campanian, the intermediate marl and limestone alternations, the second lime‐
stones bar and the upper marl and limestone of Maastrichtian. This Formation makes more
than 200m of thickness [18,81].

2.5. Late Maastrichtian‐Paleocene


This serie is represented by a monotonous brown argillaceous package, which outcrops at
several localities of northern Tunisia. This argillaceous deposits, corresponding to the El
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 135

Haria Formation, presents a lithological succession which starts with marl and limestone
alternations, then brown to grey clays and finally of marl and grey limestone alternations
(Fig. 3).

Compared to the Tunisian furrow, in the Bizerte area, the Tellian Paleocene is characterized
by the presence of the “yellow balls”. Its thickness is about 250m [18,84]. More to the South,
in the Messeftine basin, the Paleocene was recognized at the W9 well, with a thickness of
800m [84]. Towards the West and the North‐West, these deposits exhibit more significant
thickness.

2.6. Eocene
It rests on the black marls and argillaceous limestones of the El Haria Formation; the Eocene
is represented by two different series in northern Tunisia (Fig. 3): (i) a carbonated series of
Early Eocene (Ypresian) known by two facies: the Bou Dabbous Formation with
Globigerines and the El Garia Formation with Nummulites. In the transition zone, we have
a mixed facies; (ii) an argillaceous to marly series of the Souar Formation (Middle Eocene).

In the Bizerte area, the Bou Dabbous Formation includes limestones containing Globigerines
and schistous marls with glauconies and phosphates at the base. These marls are followed
by relatively massive and bituminous limestones with Globigerines. This Formation is 100m
thick [84].

To the West, in the Teboursouk area, the Djebba cross‐section, to the West of the Gorraâ
outcrop, shows a carbonated series of 70m thickness. It is composed, at the base by a car‐
bonated term (10m) with Globigerines, then an upper carbonated term containing Nummu‐
lites.

To the South and the South‐west of the Globigerines province, the Nummulites province
appears represented, in its typical locality in Kef El Garia, by yellow bio‐micritic limestones
with Nummulites and other great Foraminifera. It makes 50m of thickness.

In the Bizerte area, the Souar Formation is composed by yellow marls and clays containing
some limestones layers and yellow balls, typical of the Tellian series. Its thickness is about
300m [18].

Thickness of the Souar Formation, at Jebel Jebbas is about 500m [43] and becomes 800m
thick at the Cap Bon peninsula [85,86].

2.7. Oligocene ‐ Early Miocene


During this period, the Northern Tunisia is marked by two large basins including: a depo‐
center to the SE filled by the Fortuna Formation (600‐800m) and a depocenter to the NW
corresponding to the deposition of the Béjoua series [87] and the Oligo‐Miocene Numidian
Flysch succession (2500m; [88]). These two basins are separated by a “bald” zone [89, 24, 87]
lengthened according a NE‐SW direction, which seem coincides to the domes and diapirs
136 Tectonics – Recent Advances

zone of the Tunisian northern Atlas. The non depositional or erosional zone is inherited
from the Eocene period where we have low deep depositional area separating the two ba‐
sins (Figs. 2 and 3).

Figure 3. Synthetic column of geological series in northern Tunisia (AG‐ Ain Ghrab fm; H‐ Hakima fm;
OM‐ Oued Melah fm; K‐ Kechabta fm; OBK‐ Oued Bel Khedim fm; RR‐ Raf‐Raf fm; PF‐ Porto‐Farina fm;
OH‐ Oued El Hammam fm; Ma: Mahmoud fm)
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 137

At the West of the domes and diapirs zone (Bejaoua group), the Oligocene and Miocene
successions show a stratigraphic continuity with the Eocene series. It contains a clayey and
sandy lower part, a sandy intermediate part and a carbonated sandy upper part. Its thick‐
ness is approximately 100m [87].

The Oligo‐Miocene Numidian Flysch succession, occupies the northwestern part of northern
Tunisia. It is a thick allochthonous unit of turbidites including clays at the base and sand‐
stones and shales at the top [32,90,89,46]. This succession consists of five turbiditic units
filling the channel complexes with silexites at the top [89,88].

The Fortuna Formation occupies all southeastern basin of northern Tunisia. This Formation
includes sandy, clayey and carbonated facies indicating a shallow marine environment [89].
This Formation is 600m thick in the Gulf of Tunis [24] and 800m in the Cap Bon [85,89].

2.8. Middle to Late Miocene


The Middle to Late Miocene series are well developed in the Gulf of Tunis [91,24]. The Mid‐
dle Miocene starts with the Langhian Aïn Grab lumachellic limestone bars of 19m thick. This
unit is marked by conglomerates at the base indicating the beginning of a major transgres‐
sion, known on the scale of the country [26,85,92‐94].

The following strata are marked by an important change facies and geographical distribu‐
tion. In the Gulf of Tunis offshore area, the Saouaf Formation (Serravallian‐Tortonian) and
the Oued Bel Khedim Formation (Messinian) present facies and thickness variations [24].
They are composed of clays, silts and occasionally of sandy limestones at the base and
marls and salts at the top. The thickness of these two Formations is about 1300m (Figs. 2
and 3).

In eastern Tunisia, the series are represented by the deep marine green clays of the
Mahmoud Formation, rich with microfauna. Above, are deposited regressive continental
series, resulting from the erosional strata [95] corresponding to the Segui Formation [58].
Whereas, to the coastline and offshore areas, the Beglia and Saouaf Formations (Serraval‐
lian) constitute the lateral equivalent of the segui Formation. It is a thick series that reach
1700m and composed of clays, sandstones and lignite alternations characterizing an internal
shelf depositional environment [95].

In the Bizerte and Mateur areas, the upper Miocene deposits occupy the foreland basins of
the Tellian domain. They fossilized marine, lagunal and detrital environment
[26,71,96,18,94] and they are concentrating in five different depocenters corresponding to the
Douimis, Jalta, Messeftine, Kechabta and El Alia basins [47], which are insulated either by
morpho‐structural ridges and emerged high zones. Due to the lack of stratigraphic markers,
these deposits were subdivided in lithostratigraphic Formations [26]: Hakima, Oued El
Melah, Kechabta and Oued Bel Khedim (Figs. 2 and 3). In Jalta, the Miocene deposits are
completely continental.
138 Tectonics – Recent Advances

2.9. Pliocene
The Pliocene is dominantly marine deposits outcropping at the East of the Mateur‐Bizerte
basins and on the southern edge of the Cap Bon Peninsula (Fig. 2). These series settle with
unconformity on the various Miocene and ante‐Miocene substratum. They are subdivided
by Burollet [26] in two Formations: the Raf‐Raf at the base and the Porto Farina at the top
(Fig. 3). These two Formations outcrop at the south‐east of the Douimis basin, with a first
primarily argillaceous series (50m) and second predominantly sandy deposits rich in
fossils (50m). In the El Alia‐Ghar El Melah basin, Pliocene shows more significant
thicknesses.

Towards the East, in the Gulf of Tunis, the Lower Pliocene deposits are represented by clays,
sands and sandstones. They exhibit 300m of thickness with notable variations related to the
structuring of the inhirited substratum and Triassic salt movements. Late Pliocene is
essentially made up of sandy series of the Porto‐Farina Formation. Its thickness is about
670m and it is sealed by the villafranchian series. From the current coastline, all along the
master faults and at piedmont of the reliefs, Pliocene deposits becomes completely
continental. These latter are integrated in a detrital sequence of the Ségui Formation, which
is attributed to Late Miocene‐Pliocene.

2.10. Quaternary
At the level of the depocenters and slopes of the reliefs, the Quaternary deposits are repre‐
sented by continental facies; the marine ones are spread out over the entire eastern and
northern coasts of northern Tunisia. The Early Pleistocene of the north of the Kechabta is
composed of silts, continental sands and clays. This series locally made 200m (Fig. 2). The
upper Quaternary (Late Pleistocene / Tyrrhenian) marine and eolian deposits are well de‐
veloped all along the northern and north‐eastern littoral of Tunisia [97].

3. Structural framework
The current tectonic framework of central and northern Tunisia [28,54,24] was guided at
least by five NE‐SW trending master faults (Figs. 4 and 5) that are associated with Triassic
saliferous outcrops forming quite exposed ridges [71,98,63,33,18,99,43]. From the SE to the
NW, we distinguish:

3.1. Zaghouan fault


It is marked by a relatively irregular layout (Fig. 4) and shows a constant tendency with
overlapping towards the SE [100,28,101]. To the North, this master fault is associated to the
Triassic and Jurassic outcrops [33]. Towards the south‐east, this fault disappears before the
Rouhia‐Kalaâ Jerda graben. During the Mesozoic, this N40 master fault has bordered the
ʺTunisian troughʺ to the SE side, then it evolved to SE overlapping fault during the Cenozoic
compressive phases [59,102]. It corresponds to the T2 transversal identified by Jauzein [28].
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 139

Figure 4. A‐ Atlasic and Alpin domains of Tunisia [101]. B‐ Tectonic map of northern Tunisia.

3.2. Tunis‐Elles fault


This lineament of rectilinear layout (Figs. 4 and 5) shows a reverse movement with local
overlapping and imbrications of the series [28]. This master fault has controlled distribution
and evolution of the associated structures since the Triassic rifting phase [103, 43]. During
the extensional periods, this fault has contributed in the individualization of the large
subsiding basins and the delimitation of the Tunisian trough since the Aptian. During the
end of Cretaceous and Cenozoic contractional periods, this fault has modeled the Tunisian
Atlas following its sinistral overlapping movement. It induced the individualization of NE‐
SW poured south‐eastern Atlasic folds and other transverse NW‐SE folds strongly involved
with sigmoid forms [104]. It corresponds to the T3 transversal [28].

3.3. El Alia‐Téboursouk fault


It begins with the El Alia and Kechabta faults to the NE and continues with the fault
delimiting the Sakkak‐Lansarine diapir then the Teboursouk overlapping (Fig. 4) [31, 63,105‐
140 Tectonics – Recent Advances

107]. It corresponds to a sinistral left relay fault system. Towards the SW, it separates the
Oulad Bou Rhanem graben from the Kalaâ Jerda graben [28] and bounds the Tebessa graben
in Algeria [108]. It corresponds to the T4 transversal cited by Jauzein [28].

3.4. Ras El Korane‐Thibar fault


It extends from the Kef Triassic o alignments to the SW to the Ras El Korane in the NE (Fig.
4), crossing the Thibar, Beja and Bazina structures [18]. The Ras El Korane‐Thibar segment
constitutes the NNE extension of the T5 transversal [28]. This fault, which borders the Bazi‐
na Triassic outcrops on the Eastern side [109] extends to the North and delimits the Ras El
Korane Numidian deposits [71, 110, 111, 18, 112, 47].

This master fault corresponds to the paleogeographic limit between the Kroumirie and
Mogods mountains and that of Hedil and Bizerte [71]. It bounds the Numidian Flysch in
Ras El Korane and the Tellian units in Beja. It is a discontinuous and sinistral left relay
fault system which is sometimes shifted by other later NW‐SE faults. Dubourdieu [113]
evokes a recent horizontal displacement towards the SW of about fifteen km on this line‐
ament.

3.5. Cap Serrat‐Ghardimaou fault


It is located in the Cap Serrat area [32] and continues in Algeria crossing Souk Ahras and
Batna [113,28]. It appears to have an important role in separating the Tunisian and Algerian
blocks that have evolved with some independence. [114]. As the other master faults, it is
associated to the Triassic outcrops (Fig. 4). Moreover, it shows some Neogene volcanic
extrusions [32, 45]. This fault corresponds to T6 transversal of Jauzein [28].

We can follow the extension of the majority of these faults in offshore. They affect the
northern Tunisian plate such is the case of the Ras El Korane‐Thibar and Cap Serrat‐
Ghardimaou faults (Fig. 2). Another lineament has the same direction being extended in
offshore and could be attached to that, which delimits the Calabro‐Peloritano‐Kabyle
zone (CPK) [29,35] to the SE (Fig. 2). The other faults affecting the North of Tunisia have
NW‐SE, E‐W and N‐S directions and they have played a significant role, beside the NE‐
SW master faults, on the distribution and evolution of the Mesozoic and Cenozoic
basins.

These lineaments have subdivided the northern Tunisian margin into six compartments
(Figs. 5 and 6) corresponding to the Enfidha‐Cap Bon, Jebel Oust, Mejez El Bab, Mateur,
Nefza and Tabarka. Within each compartment, the sedimentary floor is organized into
several domains corresponding to grabens, half‐grabens and horsts delimited by NW‐SE, E‐
W and N‐S faults related to the regional deformations.

Tertiary contractions on the northern Tunisian margin have also induced folds that have
NE‐SW Atlasic direction within the compartments. Some folds affecting the Neogene strata
rather show near E‐W directions.
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 141

Figure 5. NW‐SE geological cross‐section crossing orthogonally structures of the Atlasic and Tellian
northern Tunisian. The master faults subdivide this region into six compartments. (ZF: Zaghouan Fault;
TEF: Tunis Ellès Fault; ETF: El Alia‐Teboursouk Fault; RKTF: Ras El Korane‐Thibar Fault; SGF: Cap
Serrat‐Gardimaou Fault. Location in Fig. 2.

Figure 6. Structuring of the northern Tunisian margin into six compartments lengthened along NE‐SW
direction.

4. Geodynamic evolution
4.1. Introduction
Structuring and deformation of the substratum is one of the most significant parameters
which induced the distribution and extension of Mesozoic and Cenozoic latter deposits.

Substratum of the Atlasic and Tello‐Rifaine chain is well identified in Morocco, where it
consists of Paleozoic strata deformed by the Hercynian orogenesis [4]. These strata are
affected by N45°E to N70°E strike slip faults with high dip. In northern Algeria, the
142 Tectonics – Recent Advances

substratum seldom outcrops; it extends from eastern Morocco according to W‐E direction
parallel to the current South‐Atlasic lineament (Fig. 7). In Tunisia, the Paleozoic substratum
is little known, due to the outcrop missing, except the Permian of Jebel Tebaga in southern
Tunisia and those encountered in petroleum wells on the Saharan platform in southern
Tunisian.

In northern Tunisia, the ancient fracturing is distributed according the NE‐SW master zone
of faults, limited to the South by the Zaghouan master fault and to the North by the Cap
Serrat‐Ghardimaou master fault [100,32,33]. These old fractures are marked by Triassic and
Jurassic intrusions and magmatic extrusions (Figs. 2 and 4). The NW‐SE direction appeared
especially in offshore Pelagian block of eastern Tunisia [4,93]. The current sedimentary and
tectonic distribution results from the superposition of tectonic phenomena affecting Tunisia
during Mesozoic and Cenozoic periods.
We present in the following sections the geodynamic evolution of sedimentary basins in the
northern Tunisian margin, since Triassic times. We insist for each period on the role of the
inherited Hercynian structures on basin evolution related to the nature and change of the
regional tectonic constraints.

Figure 7. Atlasic domain of the Maghreb during Early Mesozoic [4].

We note the NE‐SW orientation of the Hercynian deep faults in the Tunisian trough.

4.2. Triassic
At the Triassic period the paleogeography of Tunisia was dominated by an extended
platform between the Saharan continent to the South and the Tethys to the North (Fig. 7).
The Triassic series start with a continental detrital sedimentation, then evolves to marine
carbonates and capped by evaporates characterizing a littoral environment. The
paleogeographic changes were related to an unequally subsidence, which is particularly
active in the central and the northern of Tunisia [45]. The Triassic rifting conducts to the
paleo‐Téthys [63,2]. Magmatic green rocks often accompanied the Triassic deposits of
northern Tunisia.
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 143

Detrital deposits at the base of the Triassic indicate the beginning of a major transgression
on the Hercynian unconformity. In the North of Tunisia, Triassic thickness exceeds 2500m.
Due to the absence of well data reaching the base of the Triassic facies, we can’t identify an
appropriate underlying structure to this interval and thus there is no information on
tectono‐sedimentary control. The diapirs and domes zone, currently located between the
Tellian domain and the “Tunisian Dorsale”, have a NE‐SW direction (Fig. 8). During the
Triassic, this zone occupied an inherited horst of the Hercynian substratum and was
delimited by two normal faults, which could correspond to the two known master faults of
the northern Tunisian margin: the El Alia‐Teboursouk and the Ras El Korane‐Thibar faults.
This high zone separates two subsiding domains; the north‐western domain corresponding
to the future Tellian‐Numidian basin and the south‐eastern domain corresponding to future
ʺTunisian troughʺ. The bordering faults facilitated the migration of saliferous facies
upwards. This can be explained by the particular frequency of the evaporate sediments in
the zone located between these two major faults. Rises of the Triassic evaporates begin with
Late Jurassic [6,13,116]; some authors believed Late Jurassic and Early Cretaceous [70,58].

4.3. Jurassic ‐ Early Cretaceous


During the extensional Jurassic period, the southern Tethyan margin was structured into
horsts, grabens, half‐grabens and tilted blocks [117,118]. Since the Late Liassic, this active
kinematics had amplified differential subsidence fossilized by thick series in depocenters
and thin ones even condensed and/or with gaps on the highs [33,119,76].
Through the Atlasic domain, could exist a deep feature that controlled by the substratum
structuring. The most obvious feature is that of the N‐S Axis, which limits the deformed
Atlasic platform to the West and the stable Sahel platform to the East. This master Axis
extends to the North towards the Zaghouan master fault (Fig. 8) and has a continuous
paleogeographic role from the Jurassic to Quaternary series [62,120,54].
In northern Tunisia, these old faults affecting the ante‐Triassic substratum are not well ex‐
pressed. However, from Jurassic and especially during Early Cretaceous, the N‐S to NNW‐
SSE extension of the Tunisian margin induced genesis of subsiding basin (Tunisian furrow)
delimited by the Zaghouan fault to the SE and that of Tunis‐Elles to the NW
[33,121,4,103,75,122,123,43]. This basin will receive an enormous accumulation of deposits,
which exceeds locally 2000m for Barremian ([82]; Fig. 8). Nevertheless, near the Tunis‐Elles
fault, this stage is represented only by a few tens meters of limestone, marl and massive
limestone.
These ancient listric faults will be reactivated and caused the collapse of the NW compart‐
ments. They generate a structuring into half grabens slightly tilted to the SE. At the same
time, these NE‐SW oriented structures are associated at the level of the sedimentary cover
by N‐S, NE‐SW and NW‐SE trending other fractures, which are guided by ascension of
Triassic salt in extensional regime.

The Bou Kornine outcrop of Hammam‐Lif is placed in a paleogeographic and intermediate


structural position between two distinct paleostructural domains, belonging both to the
144 Tectonics – Recent Advances

Figure 8. NW‐SE and NE‐SW Lithostratigraphic correlations of geological series outcropping in the
different identified compartments.
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 145

Maghrebin margin of western Tethys. These domains are the “Tunisian Dorsale”,
corresponding to a carbonated platform and the ʺTunisian troughʺ, corresponding to a NE‐
SW deep graben with sedimentary pelagic filling [124].

Located on an active flexure zone between these two domains, the Jurassic deposits of Jebel
Bou Kornine will record different stages of the geodynamic evolution of the margin [125].
Since Late Toarcian‐Early Aalenian, synsedimentary tectonics of tilted blocks have a
dominating role on the progressive erosion of the near carbonated platform, on the dynamics
of the gravitation flow and on the installation of the four conglomeratic levels [125].

However, at the Tunisian trough, where differentiated grabens, sedimentation is thick and
turbiditic [36]. The Tunisian furrow, characterized by upper Jurassic radiolarite deposits, has
the main structural features of the Tethyan domain [117]. The Zaghouan lineament would
have separated the Jebel Oust compartment from that of Enfidha‐Cap Bon (Figs. 4, 5 and 6).

The Jurassic deposits remain unknown in the Gulf of Tunis because there are no drilled
wells that reach them. The interpretation of seismic lines crossing this area shows
structuring and geometry of the Jurassic limestones and marls above the Triassic carbonated
strata (Fig. 9). These series are marked by downlap progradational structures on the sides of
flanks. The Jurassic is characterized by condensed surfaces of unconformities, marked by
high amplitude and good continuity reflections (Fig. 9).

Figure 9. Interpreted seismic line L1 of the Gulf of Tunis, showing distribution of Mesozoic and Ceno‐
zoic deposits and its evolution towards the NE‐SW and the associated Triassic ascensions [24]. E‐Pl:
Early Pliocene; L‐Pl: Late Pliocene. Location in Fig. 2.

At the level of the northern Tunisian margin, depocenters, half‐grabens and high zones
lengthen preferentially according NE‐SW direction since the Jurassic and especially the
Early Cretaceous [63,126]. Triassic salt rising has been clearly emphasized since the Early
Cretaceous extensional phase corresponding to an intracontinental rifting [4,3]. This
deformation was very active, with formation of tilted blocks related to activity of
synsedimentary normal faults. These structures have been induced by regional N‐S
transtensional event [112,123,24].

In the Gulf of Tunis, the Early Cretaceous is particularly thick in W6 well (2341m). However,
it only presents 912m and 477m in W2 and W3 wells indicating structuring into low and
raised zones under effects of bordering faults (Fig. 10). Differently to the underlying Jurassic
146 Tectonics – Recent Advances

Figure 10. N‐S and WNW‐ESE Lithostratigraphic correlations of petroleum wells in the Gulf of Tunis,
showing inversion of the subsidence at Top Cretaceous, Top Miocene and Top Early Pliocene [47].

reflectors, the Cretaceous deposits are thicker on the ʺGamartʺ tectonic corridor and are
considerably reduced towards the depocenter. This distribution seems to be related to the
movements of the Triassic diapir, which caused the structural inversion and the tilting of
high edges. Towards the ʺRaouadʺ raised structure, the sedimentary sequences are
associated with retrogradational on laps and top laps (Fig. 10).
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 147

During the Berriasian, the Zaghouan fault delimits a subsiding basin, which occupies all the
northwestern part of an uplifted zone of limestones with organic‐detrital deposition [33].
This high zone constitutes the northwestern edge of the Enfidha compartment.
Locally, in Jebel Bou Rahal of the Mejez El Bab area, the El Alia‐Teboursouk fault delimits
the Triassic outcrops to the NW. It separates a NW basin with thick and continuous sedi‐
mentation, of a SE basin with reduction and gap of sedimentation (Barremian‐Coniacian
gap; [127]) in extensional and transtansional context.
The N‐S extension induced appearance of other fractures orthogonal to the Hercynian strike
slip faults already active since Jurassic [103,43]. Thus, the upper Aptian‐Albian period is
characterized by the occurrence of grabens directed close to NW‐SE following the normal
movement of faults that have the same direction [128].
In the Gulf of Tunis, this N‐S to NNW‐SSE extensional and transtensional direction recorded
during this period has mobilized the NE‐SW Hercynian faults, generating tilted blocks and
opening grabens and horsts along the NW‐SE trending faults (Fig. 10).

4.4. Late Cretaceous


A master change of the African plate movement compared to European plate has empha‐
sized in Albian [102], related to beginning of the northern and southern Atlantic Ocean
expansion. This displacement has caused movement of Africa to the North and stopped
accretion of the oceanic lithosphere at the level of African‐European Rift Zone.
During the Albian and Turonian, the Zaghouan master fault has separated a stable eastern
platform from a deformed western platform represented by the ʺTunisian furrowʺ [129].
This last, having a geosynclinal form, was structured into several compartments. This paleo‐
tectonic zonation had a great influence on the development and distribution of the varied
facies during the Senonian.
In the ʺGrand Tunisʺ area [103], the tectono‐sedimentary analysis of the upper Cretaceous
series shows a significant instability of the sedimentary floor inducing a structuring into
tilted blocks, which are bounded by NW‐SE faults [103].
Differential movements of various faults exhibit the ʺkeys of pianoʺ architecture well ex‐
press neighboring the Zaghouan fault [33] and that of Tunis‐Ellès [43]. This extensional to
transtensional period is accompanied by an intense halocinetic and magmatic activity [130‐
132]. Some E‐W contractional pulsations were highlighted in the Tunisian furrow during the
Late Albian‐Cenomanian [43]. This transtensional event is also evident by the presence of
slumped sandstones and synsedimentary N30‐ 40 trending normal faults, which affect the
Cretaceous deposits of the Mejez el Bab area, at the level of El Alia‐Teboursouk lineament
[105] and in the ʺGrand Tunisʺ area (Jeriffet outcrop; [103]).

4.5. Paleocene
In northern Tunisia, several authors [58,129] distinguished two domains; the first where the
Danian is present, whereas the other where it lacking; the limit between these two domains
148 Tectonics – Recent Advances

is roughly located on the Thala‐Elles‐Tunis line (Fig. 2). Paleogeographically, they


correspond to two different basins where the El Haria Formation, of Late Maastrichtian‐
Paleocene age doesn’t have the same stratigraphic significance [58]. The eastern Tunisia
basin with less thick and locally condensed sedimentation and northwestern Tunisian
furrow basin with very thick sedimentation.

Moreover, the limit between the two basins coincides perfectly with the Tunis‐Elles fault
(Figs. 2 and 8). This confirms the role of this fault on the control of deposition, which is well
justified during the Paleocene (El Haria Formation) [42,43].

Furthermore, the Cap Serrat‐Ghardimaou fault seems continued its impact on sedimenta‐
tion by delimiting the Tellian facies to the NW, which is characterized by marl and lime‐
stone alternations from marls of northern facies of the Tunisian furrow to the SE [129].

The Tunis‐Elles fault has controlled the Paleocene deposition; it clearly separates a low sub‐
siding basin to the South‐East from a subsiding basin to the North‐West (Fig. 8).

The Ras El Korane‐Thibar master fault (RKTF) induces an accumulation of more significant
Paleocene series on the Western northern side than on the Eastern southern side. Thus, the
thickness of this series reached 1200m in Bazina to the West of Mateur [109] and 800m in the
Henchir Haroun (W10) petroleum well to the East of Mateur [47].

The distribution of Paleocene deposits on both sides of the NE‐SW master faults indicates
that has contributed to the installation of a tilted blocks and half‐grabens associated with
condensed series and hiatuses near the location of faults and a thick and argillaceous facies
in the distal subsiding depocenters.

4.6. Early Eocene


During the Early Eocene (Ypresian), a high zone was individualized, which extends from
the Kef area to the SW, towards the Mateur area to the NE. It corresponds to the extension of
the Nummulitic limestone facies, which characterizes low subsiding platform. This high
zone separates to the NW and SE two relatively deep provinces (Fig. 11A) with pelagic sed‐
imentation corresponding to a limestone facies with Globigerines [133,34,134,102,135,136].
This high zone coincides with the domes zone. It is located, therefore, between two master
faults: the El Alia‐Teboursouk fault (ETF) to the SE and Ras El Korane‐Thibar fault (RKTF)
to the NW. The paleogeographic position of this high zone is related to the instability of the
Triassic domes whose have rise of and pierced their covers [31,63,43].

This high zone, lengthened according to an NE‐SW average direction, has an asymmetrical
form with a southeastern margin with weak slope and a northwestern margin with steep
slope. This asymmetry, inherited from previous period, was accentuated and reactivated
again by Lutetian contraction [31,112,84,137].

In addition, the morphostructural ridges or underwater peaks [138] seem to have controlled
the distribution of facies, following the movements of the NE‐SW substratum faults, associ‐
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 149

ated with the Triassic salt rising, which began since the extensional Cretaceous periods.
These Triassic bodies length along the Hercynian lineaments, indicate their successive reju‐
venations [40]. Moreover, in the Beja area, the Nummulitic limestone outcrops of the Early
Eocene show a NE‐SW privileged orientation [134].

Figure 11. Paleogeographic and paleotectonic maps of northern Tunisia at the Ypresian (A)
[129,133,34,134,135,84] and at the Aquitanian (B) [129,84,89,86]. Interpreted NW‐SE cross‐sections are
based on the present work.

Under effect of the tectonic deformation, the Paleocene‐Eocene deposits are unequally dis‐
tributed in the Gulf of Tunis. Thickness of these Formations changes from 434m in W2 well,
to 307m in W3 well. We note the absence of the Bou Dabbous and Souar Formations in W5
and W6 wells (Fig. 10). The seismic reflectors are bordered by an angular unconformity and
are associated on the side of the ʺGamartʺ structure by aggradational/retrogradational
onlaps above marly and carbonated Maastrichtian seismic horizons. Reflectors are moder‐
ately continuous and associated with pinch outs on the “Raouad” uplift (Fig. 9). Deposits
are marked, in W2 well by the Eocene breccias. These structures should indicate a slope of
the sedimentary floor along the faulted zone (Fig. 9). The Paleocene‐Eocene deposits, which
are marked by a development in the center of the depression and the ʺRaouadʺ uplift, are
reduced towards the ʺGamartʺ high zones, where clays of El Haria and limestones of
Boudabous are directly deposited on the Triassic evaporites (Fig. 9).
From the end of Cretaceous (Late Maastrichitian) and until the Middle Eocene, the NE‐SW
preexistent faults continue their effects on sedimentation in a contractional and
transpressional regime.
The El Alia‐Teboursouk and Ras El Korane‐Thibar faults have controlled sedimentation during
the Ypresian. Moreover, the NW‐SE faults appeared above the Triassic bodies during the
previous period, will express and we thus attend notable variations of the facies and thicknesses
on both sides of these faults. The extensional movements testified by the NW‐SE normal faults
are integrated in a NW‐SE regional contractional event [139,112,120,21,47,136]. This compressive
constraint generated principally reverse displacements along the NE‐SW ancient faults.

However, some authors [17,38,42,103,122,140,141] consider that this period constitute the
continuation in time of the Mesozoic extension.
150 Tectonics – Recent Advances

4.7. Middle to Late Eocene


During this period, there was the genesis of the Proto‐Mediterranean following the
movement of microplates towards the North. In the high zone, where underwater peak,
identified during the Early Eocene, appear many gaps in the Middle Eocene and especially
in the Late Eocene (Souar Formation) with the presence of several glauconitic levels (Fig. 8).
Salaj [129] announce the absence of the Late Eocene, in most of this zone. This is related to
the high structural position of this zone following rejuvenation of the Ras El Korane‐Thibar
and El Alia‐Téboursouk faults [63,98] (Fig. 11A). Elsewhere, in the Tunisian furrow and at
the level of the eastern Tunisian platform, the Middle and Late Eocene is very developed
[86].

In addition, the Middle and Late Eocene, represented by marl and limestone alternations in
the Mejez El Bab area, is much reduced [127]. These variations of facies accompanied
sometimes by gaps and unconformities characterize the Lutetian contractional period
[98,48,84,142‐144,69,137,24].

At the outcrop scale, the witnesses of contractional tectonics are showed by the presence of
(i) unconformity of the Oligocene on the Middle Eocene in the Jebel Sebâa outcrop and at
the level of the Bizerte town [84]; (ii) unconformity of the Late Eocene at the level of the Bir
Afou structure, which was formed during Late Maastrichtian‐Early Eocene and the presence
of synsedimentary reverse faults affecting marl and limestone alternations of the El Haria
Formation of Late Maastrichtian‐Paleocene age. These faults have N20, N40‐50 and N70‐80
directions [43], (iii) unconformity of the Oligocene on a folded substratum in the Enfidha
area [144].

4.8. Oligocene ‐ Early Miocene


In many localities of the Mejez El Bab, the marine Oligocene deposits, represented by clays
and bioclastic sandy limestones with Nummulites, unconformably rests on the Triassic. It is
surmounted by the Late Oligocene‐Aquitanian continental deposits [127]. This tendency to
emergence since the Kef area towards the Mateur and Bizerte areas is guided by the El Alia‐
Teboursouk and Ras El Korane‐Thibar two master faults (Figs. 8 and 11B). During the Oli‐
gocene, we also attend to the appearance of a bald zone, which lengthens from the Kef area
to the SW towards the El Alia offshore to the NE, passing by the Lansarine chain
[98,89,24,87]. This bald zone separates two different basins, characterized by a clay‐sandy
deposition; (i) basins of Beja‐Ghardimaou and subsiding and deeper Numidian basin to the
NW and (ii) the less deep but subsiding Fortuna basin to the SE.

We think that this distribution is the result of the installation, in northern Tunisia, of an
extensional event controlling the Oligocene‐Early Miocene deposits as was announced by
Piqué et al. [4].

At the regional scale, the Oligocene‐Aquitanian basin is contemporary with reactivation of


old lineaments, which appear as normal faults controlling the genesis of the half‐grabens. At
the local scale post Lutetian extensions are numerous and are fossilized by synsedimentary
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 151

normal faults [63,37,127,43]. The end of the Oligocene‐Aquitanian is marked by a continental


deposit characterized by coarse sands with cross‐bedding stratifications. These deposits
seem to be related to a total emergence of the majority of the study area.

We think that, in spite of the contractional constraints on the northern Tunisian and
northern Algerian margin [15,145], induced by rotation of the Corso‐Sarde block, we have
always an extensional context. This event has induced the siliciclastic deposition in the
Numidian basins [146,45] and of the middle Mejerda on the one hand, and the sandy
deposits of the Fortuna Formation in central and north‐eastern Tunisia on the other hand
[147,89].
Moreover, several indices of NW‐SE to N‐S extensional deformation were announced (i)
sedimentation is controlled by the activity of NW‐SE and NE‐SW faults, which delimited the
different blocks, all along the Tunis‐Elles zone [43]. The passage from the Eocene to Oligo‐
cene is marked by an inversion of subsidence following the reactivation of the NW‐SE faults;
(ii) several indices of N140 synsedimentary normal faults affecting the Oligocene sandstones
of Korbous [147]; (3) the Ras El Korane‐Thibar fault has moved in extensional mode and was
at the origin of the clayey and sandy deposition on the NW side of the mega half graben; (4)
furthermore, in the Téboursouk area, Perthuisot [63] highlighted an extensional event ac‐
companied by N45 diapirism.

4.9. Middle Miocene ‐ Quaternary


The structures recognized in the northern Tunisian margin (Figs. 2 and 4) result from the
whole of the Eocene (Lutetian), Miocene (Tortonian) and Quaternary (Villafranchian) con‐
tractional phases, which followed the multiple extensional and transtensional episodes.
These contractional phases induced tectonic inversions [24,47].
The evidence of major contractions are fossilized in the sedimentary sequences by angular
unconformities as it is the case of the Neogene (post‐Tortonian) deposits, which settled on
the Oligocene‐Aquitanian folded series at the Rmil outcrop in eastern Gaâfour [43] and on
the Campanian‐Maastrichtian (Abiod Formation) at the Mejez El Bab [127]. Moreover, an
unconformity of the Pliocene marine on the underlying strata has showed in the Bizerte [18],
in Messefftine and Kechabta outcrops (Menzel Bourguiba) [26,18,47] in the Cap Bon [37,93],
overlappings to the SSE in the Lansarine chain [98], Quaternary deformations on the El Alia‐
Téboursouk fault near the Sloughia [148] and unconformity of marine Pliocene in the Gulf of
Tunis [24].
During the Middle Miocene‐Quaternary period, three types of Neogene basins (Fig. 12) in
the Bizerte‐Mateur area have been developed following their position compared to the
raised zones delimited by NE‐SW, N‐S and NW‐SE master faults [47]; they correspond to
the (i) narrow, strongly subsiding synclines (Douimis, Kechabta and El Alia basins), (ii)
lozenge‐shaped basins (Messeftine basin) and (iii) trapezoidal basins (Jalta basin).
The Alpine and Atlasic contractional phases have caused the formation of the Tellian and
Atlasic folds trending NE‐SW as well as the installation of the overthrust folds at the north‐
western end of Tunisia [32,38].
152 Tectonics – Recent Advances

Figure 12. 3D Block diagram of the Neogene basins in the Tellian foreland domain [47]. EATM Fault: El
Alia‐Téboursouk master fault; RKTM Fault: Ras El Korane‐Thibar master fault; Pl, Pliocene; M, Mio‐
cene; O1, Oligocene‐Aquitanian (Numidian unit); O2, Oligocene‐Aquitanian; E, Eocene; P, Paleocene; C,
Cretaceous; J, Jurassic; T, Triassic.

These phases have also induced the reactivation of the old lineaments into reverse faults on
the NE‐SW direction and strike slip faults on the other directions. They also caused uncon‐
formity of the Miocene‐Pliocene series (Ségui Formation) on the Oligocene sandy banks
(Goubellat graben, [43]) and unconformity of the Middle Langhian Aïn Grab Formation on
the folded limestones of the Ypresian during the Alpine phase [149].

From the end of the Messinian and during Early Pliocene, the transtensional events induced
increase of the subsidence in the center of the depocenters, which received clayey sediments
of the Raf‐Raf Formation (Fig. 2). At the scale of the outcrop, this succession shows
synsedimentary normal faults.

The contractional event begins again during Late Pliocene and Quaternary [31,150,15] and
continues until the Actual [38,18,44,24]. Thus, several structures having participated in the
Neogene evolution are reactivated under current tectonics [38]. In the Tellian continental
domain, the Neogene basins are also deformed at the level of its levels.

In northern Tunisia, the distribution of epicenters of the earthquakes is oriented NE‐SW


according to the direction of the master faults [151]. The current movements of the Cap
Serrat‐Ghardimaou fault, for example, are characterized by a seismicity which expressed
principally on the level of its active segment of Ghardimaou [152,153]. It is at the origin of
several earthquakes; the last one dates from 17/09/1986. Its focal mechanism is related to
sinistral strike‐slip movement and the axis of sub horizontal shortening is oriented NW‐SE
to N‐S [151].
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 153

The calculation of the composite mechanisms based on the seismic events recorded in the
neighboring areas of the Ras El Korane‐Thibar fault, to the NW of Garaât Ichkeul, shows
sinistral strike slip movements with nodal plans of N45 direction and dip of about 70°
towards NW. This plan can correspond to the Ras El Korane‐Thibar fault and the axis of
shortening is oriented N170 [151].

Other earthquakes are located along the NE‐SW Zaghouan fault indicating its recent
activity. This locally overlapping master fault seems continued in the gulf of Tunis. The
composite mechanisms of the majority of the seismic events allow us to deduce an axis of
pressure oriented NW‐SE to NNE‐SSW [152,154].

5. Synthesis and conclusions


The geodynamic evolution of the northern African margin during the Late Triassic and
Jurassic was mainly guided by the reactivation of the first order NE‐SW Hercynian faults,
associated to a second order conjugate NW‐SE, E‐W and N‐S faults. These lineaments have
differentiated either lozengy basins in the Saharan Atlas [155] or high and subsiding zones
in northern Algeria with basins lengthened according to a N50 direction [156,48,13] or rather
the « pull apart » basins in the high and Middle Moroccan Atlas and in northern Tunisia
[157,117,158]. Moreover, in the Apennines, the Alps and the Betic [159], the Triassic basins
would have evolved in a strike slip mode, which is controlled by the Hercynian directions
(Fig. 13).

The rejuvenation of the southern Tethyan margin faults is related to the movement of the
African plate against the Eurasian plate along the E‐W sinistral transforming fault
[9,160,119]. The explanation of the subsiding NE‐SW oblique Tunisian furrow, compared to
the global direction of the N70° to W‐E Maghrebin furrow is explained by the presence of N‐
S to NNW‐SSE regional transtensional stresses during the Triassic and the Jurassic‐Early
Cretaceous.

This tectonic framework is explained, at the Mediterranean scale, by dextral to reverse dex‐
tral movement of the N70 master faults, separating the African and Eurasian plates [9,160].

The total closing of the Tethys related to the opening of the western Mediterranean in Early
Miocene [16] has induced genesis and rising of the Atlasic chains, which constituted thereaf‐
ter the structural units of Tunisia. Generally, the complexity of the geological structures
increases towards the North of Tunisia, at the level of the Tellian domain, where are devel‐
oped the overthrust folds [32] and where the Tethyan is closing.

The varied and oriented faults and folds affecting the northern Tunisian margin are the
result of complex changes in geometry and style of movements of the African and Eurasian
plates, which started since Triassic and continued until now.

During the Mesozoic, the northern Tunisian margin is characterized by tectonic instability
highlighted by several variations of facies and thickness of series. The NE‐SW and N70°E
Hercynian lineaments have controlled deposition [4] and caused structuring into tilted
154 Tectonics – Recent Advances

blocks and compartments generally lengthened NE‐SW. Each compartment is affected by


other NW‐SE and E‐W conjugated faults.

Figure 13. Table Summarizing the regional and global tectonic event that affected the northern Tunisi‐
an margin and western Mediterranean during the Mesozoic and Cenozoic [58,32,63,139,1,16,
164,168,33,160,2,169,119,37,165,127,40,18,4,166,84,120,75,103,123,54,21,43,45,125,154,24,47].
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 155

These faults have locally controlled the sedimentation and induced horst and graben
structures lengthened orthogonally compared to the old structures [127,103,43,24]. This
interpretation implies that the Tunisian furrow is an oblique « pull apart » graben,
compared to the northern Maghrebian transform margin; that the bordering faults are very
deep; that the Eocene inversion of the old bordering faults increase in Middle and Late
Miocene, as in old Quaternary, generated reverse movements that controlled the rising of
the central and northern Atlas and the Tellian Atlas.

Evolution of the Atlasic and Tellian domains was controlled by the Atlantic opening. Thus,
at Late Liassic, the beginning of oceanic accretion in the central Atlantic induced drifting of
Africa towards the East compared to North America and Europe [2,4,3] (Fig. 13).

During the Aptian and Albian, the opening of the Tunisian rift is related to the anti‐
clockwise rotation of Africa compared to Europe and the opening of the South Atlantic [2,4].

During the Cenozoic, the first effect of the Africa‐Europe collision is marked by a clear
folding phase in Algeria and Morocco at the Middle to Late Eocene [161,157]. In Tunisia,
synsedimentary reverse faults are highlighted in the central and southern Atlas [37,120,142]
as well as folds sealed by the Oligocene deposits [84,69,144], or by the end Eocene [137] in
the northern Atlas (Fig. 13).

An Oligocene‐Aquitanian extensional phase was highlighted in Tunisia


[63,37,127,147,45,120, 162,24]. This phase was followed by upper Miocene major collision
between Europe and Africa, resulting in the Tunisian Atlas and the installation of the
overthrust folds poured to the South‐East following a NW‐SE contractional event
[31,32,139,47]. The folding was largely amplified at Late Pliocene‐Quaternary following the
persistence of the NNW‐SSE to N‐S contractional regime [163,150,164,37,165‐167,154].
These transcurrent movements have been evolved during various tectonic phases according
to the geodynamic context. They controlled sedimentation in extensional context and they
moved either in reverse or in strike slip of contractional event. Thus, these faults, which
have at the beginning, the rather strong dips, tend to lean towards the West and the North‐
West following the NW‐SE contractions. This is seen clearly for the most northern faults due
to their proximity to the zone of contact between the African and Eurasian plates.
The NE‐SW, NW‐SE, E‐W and N‐S trending faults that have affected the North‐African
margin have evolved during the tectonic periods and controlled deposition in relation with
(i) sinistral displacement of Africa compared to Europe during the Late Jurassic‐Early Creta‐
ceous, following the opening of the southern and central Atlantic [160,2,4]; (ii) dextral to
convergent displacement of Africa compared to Europe during Campanian‐Lutetian
[16,168,160]; (iii) collision of Africa against Europe since the Middle Miocene
[32,137,168,169,54,47].

All the authors agree on the fact that faulting recognized in outcrop has related the effect of
master deep lineaments. Movement of master faults are fossilized in sedimentary series by
thickness and facies changes associated with complex structures and accentuated by salt
tectonics along various orientations.
156 Tectonics – Recent Advances

Author details
Fetheddine Melki and Fouad Zargouni
Department of Earth Sciences, FST, Tunis El Manar University, Tunis, Tunisia

Taher Zouaghi and Mourad Bédir


Georessources Laboratory, CERTE, Borj Cédria Technopole, University of Carthage, Soliman,
Tunisia

Mohamed Ben Chelbi


Water Institute of Gabès, University of Gabès, Tunisia

6. References
[1] Zargouni F. Tectonique de l’Atlas méridional de Tunisie, évolution géométrique et
cinématique des structures en zone de cisaillement. Thèse ès‐Sciences, Université Louis
Pasteur Strasbourg; 1985.
[2] Guiraud R., Maurin J.C. Le rifting en Afrique au Crétacé inférieur : synthèse structurale,
mise en évidence de deux phases dans la genèse des bassins, relations avec les
ouvertures océaniques péri‐africaines. Bulletin de la Société Géologique de France, 1991;
5, 811‐823.
[3] Guiraud R. Mesozoic rifting and basin inversion along the northern African Tethyan
margin : an overview. In : Macgregor, D. S., Moody, R. T. J. & Clark‐Lowes, D. D. (eds)
1998. Petroleum Geology of North Africa. Geological Society, London, Special
Publication, 1998; (132), 217‐229.
[4] Piqué A., Aît Brahim L., Ait Ouali R., Amrhar M., Charroud M., Gourmelen C., Laville E.,
Rekhiss F., Tricart P. Evolution structurale des domaines atlasiques du Maghreb au
Méso‐Cénosoïque ; le rôle des structures héritées dans la déformation du domaine
atlasique de l’Afrique du Nord. Bulletin de la Société Géologique de France 1998 ; 169,
797‐810.
[5] Jallouli C., Mickus K. Regional gravity analysis of the crustal structure of Tunisia. Journal
of African Earth Sciences, 2000; 30 (1), 63‐78.
[6] Zouaghi T., Bédir M., Inoubli M.H. 2D Seismic interpretation of strike‐slip faulting, salt
tectonics, and Cretaceous unconformities, Atlas Mountains, central Tunisia. Journal of
African Earth Sciences 2005; 43, 464‐486.
[7] Auzende J.M. La marge continentale tunisienne: Résultats dʹune étude par sismique
réflexion: Sa place dans le cadre tectonique de la Méditerranée occidentale. Marine
Geology Research 1971; 1, 162‐177.
[8] Auzende J.M., Bonnin J., Olivet J.L. La marge Nord‐africaine considérée comme une
marge active. Bulletin de la Société Géologique de France 1975; (7), 486‐495.
[9] Biju‐Duval B., Dercourt J., Le Pichon X. From Tethys ocean to the Mediterranean Sea. In:
Biju‐Duval, B. et Montadert, L. (Ed.), Structural history of the Mediter. basin. Split 1976,
1977; 143‐164.
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 157

[10] Durand‐Delga M., Fontboté J.M. Le cadre structural de la Méditerranée occidentale.


XXVIè Cong. géol. Int., Colloque C5: Géologie des chaînes alpines issues de la Téthys.
Mémoire du Bureau de Recherche Géologique et Minière, Paris 1980; 115, 65‐85.
[11] Cohen, C.R., Schamel, S., Boyd‐Kaygi, P., 1980. Neogene deformation in Western
Tunisia: origine of the eastern Atlas by microplate‐continental margin collision.
Geological Society of American Bulletin, Part I 91, 225‐237.
[12] Obert, D., 1981. Etude géologique des babors occidentaux (domaine tellien, Algérie).
Thèse es Science, Paris VI, 1‐635.
[13] Wildi, W., 1983. La chaîne tello‐rifaine (Algérie, Maroc, Tunisie): Structure, stratigraphie
et évolution du Trias au Miocène. Revue de Géologie dynamique et de Géographie
physique, Paris 24, (3), 201‐297.
[14] Bouillin, J.P., 1986. Le “ bassin maghrébin ”: une ancienne limite entre l’Europe et
l’Afrique à l’Ouest des Alpes. Bulletin de la Société Géologique de France (8), 547‐558.
[15] Letouzey, J., 1986. Cenozoic paleo‐stress pattern in the Alpine Foreland and Structural
interpretation in a plateform basin. Tectonophysics 132, 215‐235.
[16] Dercourt, J., Zonenshain, L.P., Ricou, L.E., Kazmin, V.G., Le Pichon, X., Knipper, A.M.,
Grandjacquet, C., Sborshikov, J.M., Geyssaut, J., Lepvrier, C. Pechersky, D.H., Boulin, J.,
Sibuet, J.M., Savostin, L.A., Sorokhtin, O., Westphal, M., Bazhenov, M.L., Lauer, J.P.,
Biju‐Duval, B., 1986. The geological evolution of the Téthys belt from Atlantic to Pamir
since Liassic. Tectonophysics 123, 241‐315.
[17] Martinez, C., Truillet, R., 1987. Evolution paléogéographique et structurale de la
Tunisie. Mémoire de la Société Géologique d’Italie 38, 35‐45.
[18] Melki, F., 1997. Tectonique de l’extrémité nord‐est de la Tunisie (Bizerte‐Menzel
Bourguiba‐Mateur). Evolution tectonique de blocs structuraux du Crétacé supérieur au
Quaternaire. Thèse de Doctorat d’Université, Université de Tunis II, Faculté des
Sciences de Tunis: 1‐207.
[19] Tricart, P., Torelli, L., Zitellini, N., Bouhlèl, H., Creuzot, G., De Santis L., Morlotti, E.,
Ouali, J., Peis, D., 1990. La tectonique d’inversion récente dans le canal de Sardaigne :
résultats de la compagne MATS 87. C.R. Acad. Sc. Paris, t. 310, série II, 1083‐1088.
[20] Mascle, G.H., Tricart, P., Torelli, L., Bouillin, J.P., Compagnoni, R., Depardon, S., Mascle,
J., Pecher, A., Peis, D., Rekhiss, F., Rolfo, F., Bellon, H., Brocard, G., Lapierre, H., Monié,
P., Poupeau, G., 2004. Structure du canal de Sardaigne : réamincissement crustal et
extension tardi‐orogénique au sein de la chaîne Apennino‐Maghrébienne ; résultats des
campagnes de plongées Cyana SARCYA et SARTUCYA en Méditerranée occidentale.
Bulletin de la Société Géologique de France, Nov. 2004, 175 (6), 607‐627 ; DOI :
10.2113/175.6.607.
[21] El Euchi, H., Saidi, M., Fourati, L., El Mahersi, C., 2004. Northern Tunisia thrust belt:
Deformation models and hydrocarbon systems. In: Swennen R., Roure F. and Granath,
J. W. (Ed.), Deformation, fluid flow, and reservoir appraisal in foreland fold and thrust
belts. American Association of Petroleum Geologists Hedberg Series (1), 371‐390.
[22] Frizon de Lamotte, D., Michard, A., Saddiqi, O. 2006. Some recent developments on the
geodynamics of the Maghreb. C. R. Geoscience 338, 1–10.
158 Tectonics – Recent Advances

[23] Frizon de Lamotte, D., Leturmy, P., Missenard, Y., Khomsi, S., Ruiz, G., Saddiqi, O.,
Guillocheau, F., Michard, A., 2009. Mesozoic and Cenozoic vertical movements in the
Atlas system (Algeria, Morocco, Tunisia): An overview. Tectonophysics 475, 9‐28.
[24] Melki, F., Zouaghi, T., Ben Chelbi, M., Bédir, M., Zargouni, F., 2010. Tectono‐
sedimentary events and geodynamic evolution of the Mesozoic and Cenozoic basins of
the Alpine Margin, Gulf of Tunis, north‐easternTunisia offshore. C. R. Geoscience 342,
741–753.
[25] Solignac, M., 1927. Étude géologique de la Tunisie septentrionale, Dir. Gen. Trav. Publ.,
Tunis, Thèse d’État, Lyon, 1‐756.
[26] Burollet, P.‐F., 1951. Étude géologique des bassins mio‐pliocènes du Nord‐est de la
Tunisie. Annales des Mines et de Géologie, Tunis (7).
[27] Castany, G., 1954. Les grands traits structuraux de la Tunisie. Bulletin de la Société
géologique de France, 6, t. 4, (1‐3), 151‐173.
[28] Jauzein, A., 1967. Contribution à l’étude géologique des confins de la dorsale tunisienne
(Tunisie Septentrionale). Thèse ès Sciences‐Annales des Mines et de Géologie, Tunis
(22).
[29] Auzende, J.M., Olivet, J.L., Bonnin, J., 1973. Le détroit sardano‐tunisien et la zone de
fracture nord‐tunisienne. Tectonophysics, 21, 357‐274.
[30] Caire, A., 1975. Les règles de la fracturation continentale dans l’évolution de l’écorce
terrestre. Revue de Géographie physique et de Géologie dynamique (2), vol. XVII, Fasc.
4, 319‐354.
[31] Zargouni, F., 1977. Etude structurale de la bande triasique de Baouala‐Aroussia‐El
Mecherket (Zone de diapirs, Atlas tunisien). Bulletin de la Société des Sciences
Naturelles, Tunisie 12, 79‐82.
[32] Rouvier, H., 1977. Géologie de l’extrême Nord tunisien: tectonique et paléogéographie
superposées à l’extrémité orientale de la chaîne nord maghrébine. Thèse ès Sciences,
Université de Pierre et Marie Curie, Paris VI, France.
[33] Turki, M.M., 1988. Polycinématique et contrôle sédimentaire associés sur la cicatrice
Zaghouan‐Nabhana. Editée par le Centre des Sciences de la Terre, Institut National de
Recherche Scientifique, Tunisie. Revue des Sciences de la Terre (7).
[34] Bishop, W.F., 1988. Petroleum Geology of East‐Central Tunisia. The American
Association of Petroleum Geologists Bulletin, 72 (9), 1033‐1058.
[35] Tricart, P., Rorelli, L., Brancolini, G., Croce, M., De Santis, L., Peis, D., Zitellini, N., 1991.
Dérives d’arcs insulaires et dynamique méditerranéenne suivant le transect Sardaigne‐
Afrique. C.R. Acad. Sc. Paris, t. 313, série II, 801‐806.
[36] Alouani, R., Rais, J., Gaya, S., Tlig, S., 1991. Les structures en décrochement au
Jurassique de la Tunisie du Nord : Témoins d’une marge transformante entre Afrique et
Europe. C. R. Acad. Sci., Paris, pp. ???.
[37] Ben Ayed, N., 1993. Évolution tectonique de lʹavant pays de la chaîne alpine de Tunisie
du début Mésozoïque à lʹActuel. Annales des Mines et de Géologie, Tunisie (32).
[38] Dlala, M., 1995. Evolution géodynamique et tectonique superposée en Tunisie ;
implication sur la tectonique récente et la sismicité. Thèse Doctorat ès‐Sciences
Géologique, Université de Tunis II, Faculté des Sciences de Tunis, 1‐200.
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 159

[39] Chihi, L., 1995. Les fossés néogènes à quaternaires de la Tunisie et de la Mer
pélagienne : étude structurale et leur sugnification dans le cadre géodynamique de la
Méditerranée centrale. Thèse de Doctorat ès‐Science, Faculté des Sciences de Tunis,
Université Tunis II, 1‐566.
[40] Boukadi, N., 1996. Un schéma structural nouveau pour le Nord de la Tunisie. The
Tunisian Petroleum Exploration Conference, Tunis, Octobre 15‐16, 91‐100.
[41] Rigo, L., Garde, S., El Euchi, H., Bandt, K., Tiffert, J., 1996. Mesozoic fractured reservoirs
in a compressional structural model for the north‐eastern Tunisian atlasic zone. The
Fifth Tunisian Petroleum Exploration Conference, Tunis, October 15‐18, 233‐255.
[42] Boutib, L., 1998. Tectonique de la région grand Tunis : évolution géométrique et
cinématique des blocs structuraux du Mésozoïque à l’Actuel (Atlas nord oriental de
Tunisie). Thèse d’Université, Université de Tunis II, Faculté des Sciences de Tunis, 1‐
151.
[43] Ben chelbi, M., 2007. Analyse tectonique des structures liées à la faille de Tunis Ellès.
Thèse de Doctorat. Univ. Tunis El Manar, Fac. Sci. Tunis, 1‐235.
[44] Rekhiss, F., 2007. Modèle d’évolution structurale et géodynamique à l’extrémité
orientale de la chaîne alpine d’Afrique du Nord, Thèse es‐sciences en Géol., Uni. de
Tunis El Manar, 2007, 1‐285.
[45] Talbi, F., Melki, F., Ben Ismail‐Latrach, K., Alouani, R., Tlig S., 2008. Le Numidien de la
Tunisie septentrionale: données stratigraphiques et interprétation géodynamique.
Estudios Geológicos 64 (1), 31‐44.
[46] Riahi, S., Soussi, M., Boukhalfa, K., Ben Ismail Lattrache, K., Dorrik, S., Khomsi, S., Bédir
M., 2010. Stratigraphy, sedimentology and structure of the Numidian Flysch thrust belt
in northern Tunisia. Journal of African Earth Sciences 57, 109–126.
[47] Melki, F., Z ouaghi, T., Harrab, S., Casas Sainz, A., Bédir, M., Zargouni, F., 2011.
Structuring and evolution of Neogene transcurrent basins in the Tellian foreland
domain, north‐eastern Tunisia. Journal of Geodynamics 52 (2011) 57–69,
doi:10.1016/j.jog2010.11.009.
[48] Vila, J.M., 1980. La chaîne Alpine dʹAlgérie orientale et des confins algéro‐tunisiens.
Thèse ès Sciences, Université Pierre et Marie Curie, Paris VI.
[49] Bracène, R., Frizon de Lamotte, D., 2002. The origin of intraplate deformation in the
system of western and central Algeria: from Jurassic rifting to Cenozoic‐Quaternary
inversion. Tectonophysics 357, 207‐226.
[50] Marmi, R., Guiraud, R., 2006. End Cretaceous to recent polyphased compressive
tectonics along the “Mole Constantinois” and foreland (NE Algeria). Journal of African
Earth Sciences 45, 123‐136.
[51] Casero, P., Roure, F., 1994. Neogene deformation at the Sicilian‐North African plate
boundary. In Roure, F. (Ed), Peri‐Tethyan platforms. Editions Technip, Paris, 189‐200.
[52] Catalano, R., Franchino, A., Merlini, S., Sulli, A., 2000. Central western Sicily structural
setting interpreted from seismic reflection profiles. Memorie della Società Geol. Italiana
55, 5‐16.
160 Tectonics – Recent Advances

[53] Sartori, R., Carrara, G., Torelli, L., Zitellini, N., 2001. Neogene evolution of the south‐
western Tyrrhenian Sea (Sardinia Basin and western Bathyal plain). Marine Geology
175, 47‐66.
[54] Abbès, C., 2004. Structurations et evolutions tectono‐sédimentaires mésozoïques et
cénozoïques, associées aux accidents réghmatiques, à la jonction des marges téthysienne
et Nord‐africaine (Chaîne Nord‐Sud‐Tunisie central). Thèse Doctorat ès‐Sciences,
Université Tunis El Manar, 1‐437.
[55] Zecchin, M., Massari, F., Mellere, D., Prosser, G., 2004. Anatomy and evolution of a
Mediterranean‐type fault bounded basin: the Lower Pliocene of the northern Crotone
Basin (Southern Italy). Basin Research 16, 117‐143.
[56] Corti, G., Cuffaro, M., Doglioni, C., Innocenti, F., Manetti, P., 2006. Coexisting
geodynamic processes in the Sicily Channel in Dilek, Y., and Pavlides, S., eds.,
Postcollisional tectonics and magmatism in the Mediterranean region and Asia: Geolo.
Soci. of America Special Paper 409, 83–96.
[57] Accaino, F., Catalano, R., Di Marzo, L., Giustiniani, M., Tinivella, U., Nicolich, R., Sulli,
A., Valenti, V., Manetti, P., 2010. A crustal seismic profile across Sicily. Tectonophysics
(2010), doi:10.1016/j.tecto.2010.07.011.
[58] Burollet, P.F., 1956. Contribution à l’étude stratigraphique de la Tunisie centrale. Ann.
Mines et Géol., 18, Tunis, 22 pls., 1‐352.
[59] Turki, M.M., 1980. La « faille de Zaghouan » est la résultante de structures superposées
(Atlas tunisien central). Bulletin de la Société géologique de France (7), XXII, 321‐325.
[60] Caire, A., 1970. Tectonique de la Méditerranée centrale. Annales de la Société
géologique du Nord, 307‐346.
[61] Caire, A., 1977. Interprétation tectonique unitaire de l’Atlas à fossés. Comptes Rendus
de l’Académie des Sciences Paris, série D, t. 284, 403‐406.
[62] Burollet, P.F., 1973. Importance des facteurs salifères dans la tectonique tunisienne.
Livre Jubilaire M. Solignac, Annales des Mines et de la Géologie, Tunis, (26), 111‐120.
[63] Perthuisot, V., 1978. Dynamique et pétrogenèse des extrusions triasiques en Tunisie
septentrionale. Thèse ès Sciences, Ecole Normale Supérieure, ERA 604‐CNRS, 1‐321.
[64] Chikhaoui, M., Jallouli, C., Turki, M.M., Soussi, M., Braham, A., Zaghbib‐Turkib, D.,
2002. L’affleurement triasique du Debadib–Ben Gasseur (Nord‐Ouest de la Tunisie) :
diapir enraciné à épanchements latéraux dans la mer Albienne, replissé au cours des
phases de compression tertiaires. C. R. Geoscience 334, 1129–1133.
[65] Vila, J.M., Ben Youssef, M., Charrière, A., Chikhaoui, M., Ghanmi, M., Kamoun, F.,
Peybernès, B., Saâdi, J., Souquet, P., Zarbout, M., 1994. Découverte en Tunisie, au SW
du Kef, de matériel triasique inter stratifié dans l’Albien : extension du domaine à
« glaciers de sel » sous‐marins des confins algéro‐tunisiens. Comptes Rendus de
l’Académie des Sci. de Paris, 318, II (12), 1535‐1543.
[66] Vila, J.M., Ben Youssef, M., Bouhlel, S., Ghanmi, M., Kassaâ, S., Miâadi, F., 1998.
Tectonique en radeaux cénomano‐turonien, au toit d’un « glacier de sel » albien de
Tunisie du Nord‐ouest : exemple du secteur minier de Gueurn Halfaya. Comptes
Rendus de l’Académie des Sciences de Paris, 327, IIa (8), 563‐570.
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 161

[67] Ghanmi, M., Vila, J.M., Ben Youssef, M., Jouirou M., Kechrid‐Benkhrouf, F., 2000. Le
matériel triasique interstratifié dans l’Albien de l’anticlinal autochtone atlasique du
Jebel Takrouna (Tunisie) : Stratigraphie, arguments gravimétriques, signification dans
la transversale N‐S des confins algéro‐tunisiens (Maghreb du Nord‐est). Bull. de la
Société de l’Histoire Naturelle de Toulouse, 136, 19‐27.
[68] Ghanmi, M., Ben Youssef, M., Jouirou M., Zargouni, F., Vila, J.M., 2001. Halocinèse
crétacé au Jebel Kebbouch (Nord‐ouest tunisien) : mise en place à fleur d’eau et
évolution d’un glacier de sel albien, comparaisons. Eclogea geol. Helv. 94, 153‐160.
[69] Ben Chelbi, M., Melki, F., Zargouni, F., 2006. Mode de mise en place des corps salifères
dans l’Atlas septentrional de Tunisie. Exemple de l’appareil de Bir Afou. C. R.
Geoscience 338, 349‐358.
[70] Bolze, J., 1954. Ascension et percée de diapirs au Crétacé moyen dans les monts de
Teboursouk, C.R. somm. Soc. Géolo. France, 139‐141.
[71] Crampon, N., 1971. Etude géologique de la bordure des Mogods, du Pays de Bizerte et
du Nord des Hédil. Thèse es Sci., Uni. Nancy, 1‐522.
[72] Bonnefous, J., 1972. Contribution à l’étude stratigraphique et micropaléontologique du
Jurassique de Tunisie (Tunisie septentrionale et centrale, Sahel, zone des Chotts). Thèse
d’État, université Paris‐6, 1–397.
[73] Alouani, R., 1991. Le Jurassique du Nord de la Tunisie. Marqueurs géodynamiques
d’une marge transformante. Turbidites, radiolarites, plissement et métamorphisme.
Thèse de spécialité, Université de Tunis II, 1‐200.
[74] Kammoun, F., Peybernès, B., Fauré P., 1999. Evolution paléogéographique de la Tunisie
saharienne et atlasique du Jurassique. C. R. Académie des Sciences, Paris 328, 547‐552.
[75] Soussi, M., 2000. Le Jurassique de la Tunisie atlasique : Stratigraphie, dynamique
sédimentaire, paléogéographie et intérêt pétrolier. Thèse Doctorat ès Sciences,
Université de Tunis‐2, 1–661.
[76] Soussi, M., 2003. Nouvelle nomenclature lithostratigraphique « événementielle » pour le
Jurassique de la Tunisie atlasique, Geobios, 36, 761‐773.
[77] Boughdiri, M., Sallouhi, H., Maâlaoui, K., Soussi, M., Cordey, F., 2006. Calpionellid
zonation of the Jurassic–Cretaceous transition in North‐Atlasic Tunisia. Updated Upper
Jurassic stratigraphy of the ‘Tunisian trough’ and regional correlations. C. R. Geoscience
338,1250–1259.
[78] Sekatni, N., 2009. Thèse d’université, Faculté des Sciences de Tunis.
[79] Rakus, M., Biely, A., 1971. Stratigraphie du Lias dans la Dorsale tunisienne. Notes Serv.
Géol. Tunisie, n° 32, 45‐63.
[80] Pini, S., 1971. Carte géologique de la Tunisie. Echelle 1/50 000. Feuille n◦ 13 : Ariana.
Notice explicative, Dir. Mines et Géol. Tunis, 1‐62.
[81] Ben Haj Ali, N., 2005. Les foraminifères planctoniques du Crétacé (Hautérivien à
Turonien inférieur) de Tunisie : systématique, biozonation et précisions
stratigraphiques. Thése d’Etat Es‐Sciences géologiques, Université de Tunis El Manar,
Faculté des Sciences de Tunis, 1‐340.
162 Tectonics – Recent Advances

[82] Maâmouri, A.L., Salaj, J., Maâmouri, M., Matmati, F., Zargouni, F., 1994. Le Crétacé
inférieur du Jebel Oust (Tunisie nord‐orientale) : microstratigraphie‐biozonation‐aperçu
sédimentaire. Zemny plyn a nafta, 39, 1, 73‐105. Hodonin, cerven 1994.
[83] Memmi, L., 1989. Le Crétacé inférieur Bérriasien‐Aptien) de Tunisie: Biostratigraphie,
Paléogéographie et Paléoenvironnement. Thèse d’Etat Es‐Sciences géologqiues, Univ.
Claude Bernard, Lyon 1, 1‐158.
[84] Melki, F., Boutib, L., Zargouni, F., Alouani, R., 1999. Nouvelles données sur l’évolution
structurale de l’extrémité nord‐est de la Tunisie (région de Bizerte). Africa Geoscience
Rev., 149‐157.
[85] Ben Ismail‐Lattrache, K,. Bobier, C., 1984. Sur l’évolution des paléo‐environnements
marins paléogènes des bordures occidentales du détroit siculo‐tnisien et leurs rapports
avec les fluctuations du paléo‐océan mondial. Marine Geology, 55, 195‐217.
[86] Ben Ismail‐Lattrache, K., 2000. Précisions sur le passage Lutétien‐Bartonien dans les
dépôts éocènes moyens en Tunisie centrale et nord‐orientale. Revue de
Micropaléontologie 43 (1‐2), 3‐16.
[87] Boukhalfa, K., 2011. Sédimentologie et stratigraphie des séries oligo‐miocènes de la
Tunisie septentrionale. Thèse d’Université, Faculté des Sciences de Tunis, 1‐327.
[88] Riahi, S., 2011. Sedimentologie, Stratigraphy, Provenance and Reservoir Potentiel of the
Oligo‐Miocene Numidian Flysch of Northern Tunisia. Thèse d’Université, Fac. Sci. de
Tunis, 1‐389.
[89] Yaïch, Ch., Hooyberghs, H.J.F., Durlet, Ch., Renard, M., 2000. Corrélation
stratigraphique entre les unités oligo‐miocènes de Tunisie centrale et le Numidien.
Comptes Rendus de l’Académie des Sciences de Paris 331, 499‐506.
[90] El Mehressi, C., 1991. Dynamique de dépôt du flysch numidien de la Tunisie (Oligo‐
Miocène). Thèse d’Université, Ecole des Mines de Paris. (15), 1‐244.
[91] El Euch, N., Ferry, S., Suc, J.P., Clauzon, G., Carmen, M., Dobrinescu, M., Gorini, Ch,
Safra, A., Zargouni, F., 2009. Messinian deposits and erosion in northern Tunisia:
inferences on Strait of Sicily during the Mesinian Salinity Crisis. Terra Nova, Vol. 21,
No. 1, 41‐48.
[92] Ben Salem, H., 1992. Contribution à la connaissance de la géologie du Cap Bon:
Stratigraphie, tectonique et sédimentologie. Thèse de 3ème cycle, Université de Tunis 2,
Faculté des Sciences de Tunis.
[93] Bédir, M., Tlig, S., Bobier, Cl., Aissaoui, N., 1996. Sequence Stratigraphy, Basin
Dynamics, and Petroleum Geology of the Miocene Eastern Tunisia. American
Association of Petroleum Geologists Bulletin 80 (1), 63‐81.
[94] Mannaï‐Tayech, B., 2009. The lithostratigraphy of Miocene series from Tunisia,
Revisited. Journal of African Earth Sciences 54, 53–61.
[95] Mannaï‐Tayech, B., 2006. Les series silicoclastiques miocènes du Nord‐Est au Sud‐Ouest
de la Tunisie: une mise au point, Geobios, 39, 71‐84.
[96] Ben Ayed, N., Gueddiche, M., El Ghali, A., Amar, H., Boughdiri, M., 1996. Évolution
géodynamique et néotectonique du bassin molassique de Kechabta. The Fifth Tunisian
Petroleum Exploration Conference, Tunis, October 15‐18, 83‐89.
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 163

[97] Oueslati, A., 1989. Les côtes de la Tunisie. Recherche Géomorphologique. Thèse de
Doctorat dʹEtat en Géographie Physique, Uni. Tunis I, Fac. Sc. Hum. Soc., 1‐682.
[98] Zargouni, F., 1978. Analyse structurale de la chaîne de Lansarine (zone des diapirs,
Atlas tunisien). Bulletin de la Société des Sciences Naturelles, Tunisie, t. 13, 97‐104.
[99] Ben Haj Ali, M., Chihi, L., Rabhi, M., 1998. Le Néogène de Tunisie: Analyse des séries
stratigraphiques et le rôle de la tectonique et de l’halocinèse dans la genèse des bassins.
Proceedings of the 6th Tunisian petroleum exploration and production conference,
ETAP, Memoir (12), 59‐76.
[100] Castany, G., 1951. Etude géologique de l’Atlas tunisien oriental. Annales des Mines et
de Géologie, Tunis, 8.
[101] Caire, A., 1978. The central mediterranean mountains chains in the Alpine orogenic
environment. The ocean basins and margins, Vol. 4B, Edited by Alan E.M. Nairm,
William H. Kanes and Francis G. Stehli (Plenum Publishing Corporation), 201‐256.
[102] Morgan, M.A., Grocott, J., Moody, R.T.J., 1998. The structural evolution of the
Zaghouan‐Ressas Belt, northern Tunisia. Geological Society, London, Special
Publication, 132, 405‐422.
[103] Boutib, L., Melki, F., Zargouni, F., 2000. Tectonique synsédimentaire dʹâge crétacé
supérieur en Tunisie nord orientale: blocs basculés et inversion de subsidence. Bull. Soc.
Géol. France 171 (4), 431‐440.
[104] Ben Chelbi, M., Melki, F., Zargouni, F., 2008. Précision sur lʹévolution structurale de
lʹAtlas septentrional de Tunisie depuis le Crétacé (Bassin de Bir MʹCherga). Echos dʹune
évolution polyphasée de la marge tunisienne dans son cadre méditerranéen. African
Geoscience Review, 15, (3 & 4), 229‐246.
[105] El Ouardi, H., Turki, M.M., 1995. Tectonique salifère polyphasée dans la région de
Mejez el Bab–Testour (zone des dômes, Tunisie septentrionale): contrôle de la
sédimentation méso‐cénozoïque. Geol. Mediterr. 22 (2) (1995) 73–84.
[106] Hamdi‐Nasr, I., Inoubli, M.H., Ben Salem, A., Tlig, S., Mansouri, A., 2009. Gravity
contributions to the understanding of salt tectonics from the Jebel Cheid area (dome
zone, Northern Tunisia). Geophysical Prospecting, 2009, 57, 719–728.
[107] Ben Slama, M.M., Masrouhi, A., Ghanmi, M., Ben Youssef, M., Zargouni, F., 2009.
Albian extrusion evidences of the Triassic salt and clues of the beginning of the Eocene
atlasic phase from the example of the Chitana‐Ed Djebs structure (N.Tunisia):
Implication in the North African Tethyan margin recorded events, comparisons. C. R.
Geoscience 341 (2009) 547–556.
[108] Perthuisot, V., Jauzein, A., 1974. L’accident El Alia‐Tebessa dans la région de
Téboursouk. Notes Service Géologique, Tunis (9), 57‐63.
[109] Marzougui, W., Melki, F., Zargouni, F, 2008. Modèle évolutif de mise en place de la
structure d’Aouana‐Bazina (extrémité orientale de la chaîne alpine de la Tunisie
septentrionale) : apport du filtrage des anomalies gravimétriques. 22nd Colloquium of
African Geology, 13th Conference of the Geological Society of Africa, November 04‐06‐
2008, Abstract, 289‐290.
[110] Ben Ayed, N., 1994. Les décrochements‐chevauchements EW et NS convergents de la
Tunisie septentrionale : géométrie et essai de reconstitution des conditions de
164 Tectonics – Recent Advances

déformations. Proceeding of the 4th Tunisian Petroleum Exploration Conference,


Mémoire n°7, ETAP, 25‐37.
[111] Melki, F., Alouani, R., Talbi, F., Zargouni, F., 1996. Evolution géodynamique de la
région de Bizerte. Entraînement et blocage des structures à partir du Paléogène. The
Fifth Tunisian Petroleum Exploration Conference, Tunis, October 15‐18, 413‐421.
[112] Melki, F., Zargouni, F., 1991. Tectonique cassante post Jurassique de la mine de
Hammam Zriba (Tunisie nord‐orientale). Incidence sur la karstification et les
concentrations de fluorine, barytine et célestine, d’environnement carbonaté. Bull. Soc.
Géol. France 162, (5), 851‐858.
[113] Dubourdieu, G., 1959. La déformation récente de l’Afrique du Nord. Extrait des
comptes rendus de l’Académie des Sciences, t. 249, pp. 2799‐2801, séance du 21
décembre 1959.
[114] Glangeaud, 1951. Interprétation tectono‐physique des caractères structuraux et
paléogéographiques de la Méditerranée occidentale. Bulletin de la Société géologique
de France, t.6, 735‐762.
[115] Chandoul, H., Burollet, P.F., Ben Ferjani, A., Memmi, L., 1993. Recueil des Coupes
Types de Tunisie –I‐ Trias et Jurassique. Entreprise tunisienne d’activités pétrolières,
Mémoire (4), 1‐95.
[116] Zouaghi T., Guellala R., Lazzez M., Bédir M., Ben Youssef M., Inoubli M.H., Zargouni
F., The Chotts Fold Belt of Southern Tunisia, North African Margin: Structural Pattern,
Evolution, and Regional Geodynamic Implications. New Frontiers in Tectonic Research
‐ At the Midst of Plate Convergence. Intech; 2011. p49‐72.
[117] Alouani, R., Tlig, S., Zargouni, F., 1990. Découverte de radiolarites du Jurassique
supérieur dans le « sillon tunisien ». Faciès et structures d’une marge SE de la Téthys
maghrébine. C. R. Acad. Sci. Paris, t. 310, série II, 609‐612.
[118] Soussi, M., Mangold, C., Enay, R., Boughdiri, M., Ben Ismail, M.H., 2000. Le Jurassique
inférieur et moyen de la Tunisie septentrionale ; corrélations avec l’axe Nord‐Sud et
paléogéographie. GEOBIOS, 33, 4: 437‐446.
[119] Alouani, R., Rais, J., Gaya, S., Tlig, S., 1992. Les structures en décrochement au
Jurassique de la Tunisie du Nord : Témoins d’une marge transformante entre Afrique et
Europe. C. R. Acad. Sci. Paris, t. 315, série II, 717‐724.
[120] Rabhi, M., 1999. Contribution à l’étude stratigraphique et analyse de l’évolution
géodynamique de l’axe Nord‐Sud et des structures avoisinantes (Tunisie centrale).
Thèse d’Université, Faculté des Sciences de Tunis, Université de Tunis II, 1‐217.
[121] Chikhaoui, M., Turki, M.M., Delteil, J., 1991. Témoignages de la structurogenèse de la
marge téthysienne en Tunisie, au Jurassique terminal‐Crétacé (Région du Kef, Tunisie
septentrionale). Géologie Méditerranéenne, XVIII (3), 125‐133.
[122] El Ouardi, H., 2002. Origine des variations latérales des dépôts yprésiens dans la zone
des dômes en Tunisie septentrionale, C. R. Geoscience 334 (2002) 141–146.
[123] Bouaziz, S., Barrier, E., Soussi, M., Turki, M.‐M., Zouari, H., 2002. Tectonic evolution of
the northern African margin in Tunisia from paleostress data and sedimentary record.
Tectonophysics 357, 227‐253.
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 165

[124] Rais, J., Gaya, S., Alouani, R., Mouguina, M., Tlig, S., 1991. Le sillon tunisien :
structuration synsédimentaire jurassique. Rift avorté et cicatrisé au Dogger supérieur–
Malm, à la marge SE de la Téthys maghrébide, C. R. Acad. Sci. Paris, Ser. II 312 (1991)
1169–1175.
[125] Sekatni, N., Fauré, P., Alouani, R., Zargouni, F., 2008. Le passage Lias–Dogger de la
Dorsale de Tunisie septentrionale. Nouveaux apports biostratigraphiques. Âge Toarcien
supérieur de la distension téthysienne. C. R. Palevol 7 (2008) 185–194.
[126] Bobier, C., Viguier, C., Chaâri A., Chine, A., 1991. The post‐Triassic sedimentary cover
of Tunisia: seismic sequences and structure. Tectonophysics, 195, 371‐410.
[127] El Ouardi, H., 1996. Halocinèse et rôle des décrochements dans l’évolution
géodynamique de la partie médiane de la zone des dômes, Thèse d’Université, Faculté
des Sciences de Tunis, Université de Tunis II, 1‐ 242.
[128] Martinez, C., Chikhaoui, M., Truillet, R., Ouali, J., Creuzot, G., 1991. Le contexte
géodynamique de la distension albo‐aptienne en Tunisie septentrionale et centrale :
structuration éocrétacée de l’Atlas tunisien. Eclogae geol. Helv., 84, 1, 61‐82.
[129] Salaj, J., 1980. Microbiostratigraphie du Crétacé et du Paléogène de Tunisie
septentrionale et Orientale (hypostratotypes tunisiens). Thèse dʹEtat, Institut de
Dionyzhir, 238 p., Bratislava.
[130] Boukadi, N., Bédir, M., 1996. L’halocinèse en Tunisie : contexte tectonique et
chronologique des évènements. Comptes Rendus de l’Académie des Sciences, Paris,
322, série IIa, 587‐594.
[131] Laâridhi‐Ouazaâ, N., Bédir, M., 2004. Les migrations tectono‐magmatiques du Trias au
Miocène sur la marge orientale de la Tunisie. African Geoscience Review 11 (3), 179‐196.
[132] Talbi, F., Jaafari, M., Tlig, S., 2005. Magmatisme néogène de la Tunisie septentrionale :
Pétrogenèse et évènements géodynamiques. Revista de la Sociedad Geológica de
España, 18(3‐4).
[133] Bishop, W.F., 1980. Eocene and Upper Cretaceous reservoirs in the East Central
Tunisia. Oil and Gas Journal, 137‐142.
[134] Er‐Raioui, L., 1994. Environnements sédimentaires et Géochimie des séries de lʹEocène
du Nord‐Est de la Tunisie. Thèse 3ème Cycle, Faculté des Sciences de Tunis, 1‐244.
[135] Zaïer, A., Beji‐Sassi, A., Sassi, S., Moody, R.T.J., 1998. Basin evolution and deposition
during the Early Paleogene in Tunisia. In: Macgrfgor D.S., Moody R.T.J. & Clarck‐Lowis
D.D. (eds) 1998. Petroleum Geology of North Africa, Geological Society, London,
Special Publication, (132), 375‐393.
[136] Tlig, S., Sahli, S., Er‐Raioui, L., Alouani, R., Mzoughi, M., 2010. Depositional
environment controls on petroleum potential of the Eocene in the North of Tunisia.
Journal of Petroleum Science and Engineering 71 (2010), 91–105.
doi:10.1016/j.petrol.2010.01.009.
[137] Masrouhi, A., Ghanmi, M., Ben Slama, M.M., Ben Youssef, M., Vila, J.M., Zargouni, F.,
2008. New tectono‐sedimentary evidence constraining the timing of the positive
tectonic inversion and the Eocence Atlasic phase in northern Tunisia: Implication for the
North Africa paleo‐margin evolution. C. R. Geoscience 340, 771‐778.
166 Tectonics – Recent Advances

[138] Batik, P., 1977. Faciès de l’Eocène dans la région des Hedil. Notes Service géologique
de Tunisie, (43), 63‐72.
[139] Letouzey, J., Trémolière, 1980. Paleo‐stress around the Mediterranean since the
Mesozoic from microtectonics : comparison with plate tectonics data. Rock Mech., 9,
173‐192.
[140] Gharbi, R.A., Chihi, L., Hammami, M., Soumaya, A., Kadri, A., 2005. Manifestations
tectono‐diapiriques synsédimentaires et polyphasées d’âge Crétacé supérieur‐
Quaternaire dans la région de Zag Et Tir (Tunisie, centre‐nord), C. R., Geoscience 337,
1293–1300.
[141] Ben Mehrez, F., Kacem, J., Dlala. M., 2009. Late Cretaceous synsedimentary diapirism
of Bazina‐Sidi Bou Krime Triassic evaporates (Northern Tunisia): Geodynamic
implication. C. R. Geoscience 341, 78‐84.
[142] El Ghali, A., Ben Ayed, N., Bobier, C., Zargouni, F., Krima, A., 2003. Les manifestations
tectoniques synsédimentaires associées à la compression éocène en Tunisie,
implications paléogéographiques et structurales sur la marge Nord‐Africaine. C. R.
Geoscience 335 (2003) 763–771.
[143] Meulenkamp, J.E., Sissingh, W., 2003. Tertiary palaeogeography and
tectonostratigraphic evolution of the Northern and Southern Peri‐Tethys platforms and
the intermediate domains of the African‐Eurasian convergent plate boundary zone.
Palaeogeography, Palaeoclimatology, Palaeoecology 196 (2003) 209‐228.
[144] Khomsi, S., Bédir, M., Soussi M., Ben Jemia, M.G., Ben Ismail‐Lattrache, K., 2006. Mise
en évidence en subsurface d’événements compressifs Éocène moyen–supérieur en
Tunisie orientale (Sahel) : généralité de la phase atlasique en Afrique du Nord. C. R.
Geoscience 338, 41–49.
[145] Benaouali‐Mebarek, N., Frizon de Lamotte, D., Roca, E., Bracène, R., Faure, J.‐L., Sassi,
W., Roure, F., 2006. Post‐Cretaceous kinematics of the Atlas and Tell systems in central
Algeria: Early foreland folding and subduction‐related deformation. C. R. Geoscience
338, 115‐125
[146] De Jong, K.A., 1975. Gravity tectonics or plate tectonics : example of the Numidian
flysch, Tunisia. Geol. Mag., 112 (4), 373‐381.
[147] Er‐Raioui, L., Alouani, R., Tlig, S., 1995. Les bassins de l’Oligo‐Miocène inférieur de la
Tunisie nord‐orientale : structures et remplissage de fossés intracontinentaux. Notes du
Service Géologique de Tunisie, (61), 43‐61.
[148] Rebaï, S., 1992. Sismotectonique et champ de contrainte dans les chaînes alpines et
dans les plate‐formes de l’Europe, d’Afrique du Nord et du Moyen‐Orient. Thèse de
Doctorat, Académie de Montpellier, Université Montpellier II, Sciences et Techniques
du Languedoc, 1‐210.
[149] Tlig, S., Erraoui, L., Ben Aissa, L., Alouani, R., Tagourti, A., 1991. Tectogenèse alpine
atlasique: Des évènements distincts de l’histoire géologique de la Tunisie. Corrélations
avec les évènements clés de la Méditerranée. Comptes Rendus de l’Académie des
Sciences de Paris 312, 295‐301.
[150] Ben Ayed, N., Viguier, C., Bobier, C., 1983. Les éléments structuraux récents essentiels
de la Tunisie nord‐orientale. Notes Service géologique de Tunisie (47), 5‐19.
Role of the NE-SW Hercynian Master Fault Systems and Associated Lineaments on the Structuring
and Evolution of the Mesozoic and Cenozoic Basins of the Alpine Margin, Northern Tunisia 167

[151] Gueddiche, M., Harjono, H., Ben Ayed, N., Hfaiedh, M., Diament, M., Dubois, J., 1992.
Analyse de la séismicité et mise en évidence d’accidents actifs dans le Nord de la
Tunisie. Bulletin de la Société géologique de France, (4), 415‐425.
[152] Hfaiedh, M., Ben Ayed, N., Dorel, J., 1985. Etude néotectonique et séismotectonique de
la Tunisie nord‐orientale. Notes du Service Géologique de Tunisie, n° 51, 41‐55.
[153] Gueddiche, M., Ben Ayed, N., Mohammadioun, G., El Ghali, A., Chekhma, H.,
Diament, M., Dubois, J., 1998. Etude sismotectonique de la Tunisie nord‐orientale.
Bulletin de la Société géologique de France, t. 169, (6), 789‐796.
[154] Zouaghi, T., Bédir, M., Melki, F., Gabteni, H., Gharsalli, R., Bessioud, A., Zargouni, F.,
2010. Neogene sediment deformations and tectonic features of northeastern Tunisia:
evidence for paleoseismicity. Arabian Journal of Geosciences, DOI 10.1007/s12517‐010‐
0225‐z.
[155] Kazi‐Tani, N., 1986. Evolution géodynamique de la bordure Nord africaine: Le
domaine intraplaque nord‐algérien. Approche mégaséquentielle. Thèse ès Sc. Uni. Pau,
Bordeaux, 1‐871.
[156] Guardia, P., 1975. Géodynamique de la marge alpine du continent africain d’après
l’étude de l’Oranie nord‐occidentale. Thèse ès Sciences, Inst., Polytech. Médit., Nice.
[157] Laville, E., 1985. Evolutions sédimentaire, tectonique et magmatique du bassin
jurassique du Haut Atlas (Maroc). Modèle en relais multiples de décrochement. Thèse
ès Science.
[158] Adil, S., 1993. Dynamique du Trias dans le Nord de la Tunisie: Bassin en relais
multiples de décrochement, magmatisme et implications minières. Thèse de 3ème cycle,
Université de Tunis 2, 1‐248.
[159] Alvaro, M., Capote, R., Vegas, R., 1979. Un modela de evolucion geotectonica para la
cadena celtiberica. Acta. Geol. Hispanica, 172‐177.
[160] Uchupi, E., 1988. The Mesozoic‐Cenozoic evolution of Iberia. A tectonic link between
Africa and Europe. Rev. Soc. Geol. Espana, 1, 257‐294.
[161] Raoult, J.‐F., 1974. Géologie du centre de la chaîne numidienne (Nord du
Constantinois, Algérie). Thèse ès Sciences, Paris, 1‐163.
[162] Melki, F., Zouaghi, T., Ben Chelbi, M., Bédir, M., Zargouni, F., 2009. Mesozoic and
Cenozoic tectono‐sedimentary events and geodynamic evolution of the Alpine margin
in the North‐Eastern Tunisia offshore (Gulf of Tunis) using the subsurface data. 4ème
Congrès Maghrébin de Géophysique Appliquée, Hammamet (Tunisie), 26‐27‐28 mars
2009, Abstract, 79.
[163] Castany, G., 1956. Essai de synthèse géologique du territoire Tunisie‐Sicile. Annales
des Mines et de Géologie (16).
[164] Philip, H., 1987. Plio‐Quaternary evolution of the stress field in Mediterranean zones of
subduction and collision. Annales Geophysicae 5 (3), 301‐319.
[165] Dlala, M., Rebaï, S., 1994. Relation compression‐extension Miocène supérieur à
Quaternaire en Tunisie : implication sismotectonique. C. R. Académie des Sciences,
Paris 319, 945‐950.
[166] Chihi, L., Philip, H., 1998. Les fossés de l’extrémité orientale du Maghreb (Tunisie et
Algérie orientale): tectonique mio‐plio‐quaternaire et implication dans l’évolution
168 Tectonics – Recent Advances

géodynamique récente de la Méditerranée occidentale, Notes du Service Géologique de


Tunisie (64), 103‐116.
[167] Kacem, J., 2004. Étude sismotectonique et évaluation de l’aléa sismique régional du
Nord‐Est de la Tunisie: apport de la sismique réflexion dans l’identification des sources
sismogéniques. Thèse de Doctorat d’Université, Université de Tunis El Manar, Faculté
des Sciences de Tunis.
[168] Le Pichon, X., Bergerat, F., Roulet, M.J., 1988. Plate kinematics and tectonics leading to
the Alpine belt formation; a new analysis. Geol. Soc. Am., sp. Paper, 218, 111‐131.
[169] Bédir, M., Zargouni, F., Tlig, S., Bobier, C., 1992. Subsurface Geodynamics and
Petroleum of Transform Margin Basins in the Sahel of Mahdia and El Jem (Eastern
Tunisia). American Association of Petroleum Geologists Bulletin 76 (9), 1417‐1442.
Chapter 7

An Assessment of the Earthquakes of Ancient


Troy, NW Anatolia, Turkey

Akın Kürçer, Alexandros Chatzipetros, Salih Zeki Tutkun, Spyros Pavlides,


Süha Özden, George Syrides, Kostas Vouvalidis, Emin Ulugergerli,
Özkan Ateş and Yunus Levent Ekinci

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/48471

1. Introduction
Many destructive earthquakes occurred in Northwestern Anatolia during historical and
instrumental periods and as a result of these earthquakes civilities were damaged.
Approximately 30 km southwest of Çanakkale the ancient city of Troy is located, containing
remains belonging to the period between B.C. 3000 and A.D. 400 (Figure 1).

According to the intermittent archeological excavations, carried out from 1871 up to the
present, there exist nine different layers of settlements in Troy. Although there is some
archeological evidence which indicates that some of these layers, especially Troy III (B.C.
2200‐2050) and Troy VI (B.C. 1800‐1275) have been damaged by one or more earthquakes, no
multidisciplinary geoscientific research has been carried out so far on the active faults which
could have caused these earthquakes.

Troy which once controlled the commercial crossing point between Asia and Europe over
Dardanos (the Dardanelles) used to be one of the most important trade centers of that era.
Because of this fact Troy, besides being one of the hundreds of ancient cities situated in
Anatolia, was a city that played an important role in the development of Western Anatolian
and Aegean cultures. As a result of Troy’s dominant position, several European countries
believe that their roots lie in Troy and Trojans. When the architecture of Troy, represented
by 9 layers of settlements spanning the period between B.C. 3000 and A.D. 400, is examined,
it is observed that passages between civilizations are not gradual; instead, there are radical
changes in building styles and materials. This observation can lead to the assumption that
the events causing passages between civilizations are natural phenomena such as earth‐
quakes rather than wars, fires or epidemics. On the other hand, Professor Manfred Korfman,
who meticulously presided over Troy excavations between 1968 and 2005, talks

© 2012 Kürçer et al., licensee InTech. This is an open access chapter distributed under the terms of the
Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
172 Tectonics – Recent Advances

Figure 1. A. The major active tectonic structures of Eastern Mediterranean Region (Modified from
Şengör et al., 1985), B. The locational relationship of Troy with the northern and southern branches of
NAFS and the major active faults in the region (Modified from Şaroğlu et al., 1992).

about archaeological findings indicating that especially Troy VI Layer (B.C. 1800‐1275) was
damaged by one or more earthquakes. In the light of this information, some deformation
structures thought to be of seismic origin have been observed, especially on the walls of
Troy VI Layer. These deformation structures can be classified as systematic cracks, rotations
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 173

and blocks tilting (Figure 2). In this article, the results of the research related to the earth‐
quakes that caused the deformation structures observed in the remains of ancient Troy city
are presented.

Figure 2. Deformation structures of seismic origin within Troy VI Layer a) clockwise rotation b) Tilting
c) systematic cracks in the walls of Troy VI Layer that Prof. Manfred Korfman showed the project team

2. Regional geology and tectonic framework


In the study area, rock units having ages from Lower Cretaceous to present‐day outcrop
(Figure 3). The basement of the study area is constituted by Lower Cretaceous aged Deni‐
zgören ophiolites (Okay, 1987). Denizgören ophiolites are unconformably overlain by Ça‐
nakkale formation (Şentürk and Karaköse, 1987). Çanakkale formation, which outcrops at
relatively high plateaus over a large part of the study area, is composed of a succession of
pebblestones, sandstones, sandy limestones and limestones deposited in lagoonal, coastal
and offshore environments. The age of the unit is Upper Miocene‐Pliocene according to
ostracoda and pelecyopoda fossils contained in it (Şentürk and Karaköse, 1987). All units
outcropping in the study area are unconformably covered by Quaternary deposits. Quater‐
nary deposits are represented by paleoterrace deposits of Dümrek river, flood plain deposits
174 Tectonics – Recent Advances

of Dümrek and Karamenderes rivers and alluvial fan deposits developed at the front of the
Troy Fault.

The study area is located in the peninsula in the NW corner of Anatolia. This area, known as
Biga Peninsula, is being deformed under the effect of the roughly N‐S oriented Western
Anatolian Extension System and the western extensions of the North Anatolian Fault Sys‐
tem (NAFS). The study area, situated at the NW part of Biga Peninsula, is mostly under the
influence of NAFS. NAFS, which is one of the most important active tectonic structures of
the eastern Mediterranean Region, is divided into two branches as the north and south
branches, starting from the east of Adapazarı. The northern branch that reaches the Sea of
Marmara from Hersek delta, after traversing the Marmara Sea in approximately E‐W direc‐
tion is connected to the Ganos‐Saroz Fault (GSF) on which Şarköy‐Mürefte Earthquake
(Mw=7.2) of 1912 occurred. GSF is oriented N700E and cutting through Gallipoli Peninsula
in this orientation from Saroz Gulf, reaches North Aegean Sea. The fault segments constitut‐
ing the south branch of NAFS, by traversing Biga Peninsula in NE‐SW direction reach North
Aegean Sea. The most important ones of these NE‐SW oriented faults are, from north to
south, Edincik fault, Biga fault, Sarıköy‐İnova fault, Yenice‐Gönen fault, Evciler fault, Pa‐
zarköy fault and Edremit fault.

The ancient city of Troy is situated within a right lateral deformation zone bordered by the
northern and southern branches of the North Anatolian Fault System (NAFS). The northern
border of this deformation zone is formed by Gaziköy‐Saroz Fault. And the southern border
of the deformation zone is represented by the south branch of NAFS composed of faults
which are parallel or subparallel to each other and extend between Kapıdağ Peninsula and
Edremit Bay (Figure 1b). Troy Fault System, which is represented by E‐W oriented normal
faults, NE‐SW oriented right lateral and NW‐ SE oriented left lateral strike slip faults, exists
in the area within this deformation zone. When taken into account its geologic fault length
and morphotectonic characteristics, 12‐km long Troy Fault is the most important one of the
E‐W oriented faults within the Troy Fault System (Tutkun and Pavlides, 2005). And the most
important of the NE‐SW oriented right lateral faults is Kumkale Fault which possesses a
total length of 15 km including its continuation in the sea.

Destructive earthquakes occurred on the active faults in Biga Peninsula (Troy is also located
here) and in its vicinity, during historical and instrumental eras (Figure 4 and 5).

3. Material and method


This study is a multidisciplinary research in which geomorphological, geological and
geophysical methods were applied in a specific order.

Within the framework of the geomorphological studies, a 1/25 000 scale digital elevation
model of the study area was constructed and on this model, geomorphological features such
as stream/valley drainage areas, mountain front sinuosity ratio and ratio of valley floor
width to valley height were investigated.
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 175

Figure 3. Geological map of Troy and its vicinity.

Figure 4. Western segments of NAFS in the study area and in its near vicinity (modified after Şaroğlu et
al., 1992) and the historical earthquakes that occurred on these faults between A.D. 32 and A.D. 1900
(modified after Ambraseys and Finkel, 1991)
176 Tectonics – Recent Advances

Figure 5. Extensions of North Anatolian Fault System (NAFS) in Biga Peninsula, instrumental period
earthquakes related to these faults and focal mechanism analyses (Focal mechanism analyses were
taken from 1: Kıyak, 1986; 2: Canıtez and Tokgöz, 1971; 3: McKenzie, 1972; 4: Kalafat, 1978; 5: Kalafat,
1988; 6: Jackson and McKenzie, 1984; 7: McKenzie, 1978; 8: Taymaz, 2001; 9: Aksoy et al., 2010)

Within the framework of the geological studies, a 1/25 000 scale geologic map of the region
was prepared. Afterwards, structural observations were made on the faults within the Troy
Fault System. The data obtained from these observations were evaluated in the kinematical
analysis program developed by Carey (1979) and the stress regimes effective in the region at
the present time were determined. In active tectonic researches, shallow (0‐15meter) drilling
works can be conducted in order to determine and document basin asymmetry originating
from active faults. With the help of shallow core drillings, lateral and vertical variations of
the Quaternary deposits in the basin can be determined. In this study, a total of 5 shallow
drillholes with depths varying between 3 and 9 meters were opened in the area north of the
Troy fault and where Quaternary deposits outcrop. The core samples compiled from the
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 177

drillings were logged and interpreted. In this study, along the Troy fault, using Direct
Current Resistivity method measurements were made along 8 profiles with an average
length of 200 meters and perpendicular to the fault. And applying two‐dimensional inverse
solution, geoelectric cross‐ sections for the shallow depths (0‐15 meters) of the earth were
obtained. And lastly, in the study paleoseismological trench works were conducted in the
areas determined according to geological observations and geophysical data. Within the
scope of this work, a total of 4 trench works were carried out, 3 trenches on Troy fault and 1
on Kumkale fault. All of the trenches were excavated perpendicular to the fault, with an
average length of 10‐15 meters, 3‐4m wide and 4‐6m deep. A total of 20 carbon‐containing
soil samples were compiled from the trenches for C‐14 dating. These samples were analyzed
at Beta Analytical Laboratory in the United States of America.

4. Findings
4.1. Morphotectonic analysis of Troy Fault
The geomorphology of a region can be described based on specific measurements of its
morphological characteristics. This method, defined as morphometry, is realized by digitally
deriving information about geomorphological elements from the elevation values (DEM ‐
Digital Elevation Model) belonging to the region and analyzing them. These values,
obtained with the help of morphometry, can provide consistent and fast information both
on the evolution of the drainage in the study area and on the degree, distribution and
character of the structural/lithological control on this evolution (Keller and Pinter, 1996).

For this purpose, through the digitization of the 1/25 000 scale sheets of Ayvalık I 16 a2 and
Ayvalık I 16 b1 belonging to the study area, the digital elevation model of the region was
created (Figure 6). The geomorphological characteristics of the region were investigated on
this model. The geologic map of the region was draped over the digital elevation model as a
separate layer and a relief geologic map was obtained (Figure 3).

The purpose of the morphotectonic analysis is to digitally reveal the degree of influence of
the erosional and tectonic processes effective in the morphological shaping of a region. In
this study, such geomorphological indices as mountain front sinuosity ratio (Smf index),
stream/valley drainage areas and profiles, and valley floor width to valley height ratio (Vf
index) were calculated.

Mountain front sinuosity ratio (Smf) is an index that reflects the balance between erosional forces
that tend to cut embayments into a mountain front and the tectonic forces that tend to produce a
straight mountain front. Mountain fronts uplifted by active tectonism are straight and have low
Smf values. And mountain fronts that move slowly or have lost their activity display irregular
forms and high values because they are destroyed by erosional forces (for further information,
see Keller and Pinter, 1996). If Smf values are between 1 and 2, the fault in question has high
activity. If this value is greater than 2, the activity of the fault should be considered as doubtful.
However, one must bear in mind that Smf index can also be affected by the strength properties of
the rocks forming the mounting front involved in faulting and by erosional activities.
178 Tectonics – Recent Advances

In the evaluation made by taking into account mountain front sinuosity ratios, Troy Fault is
subdivided into three geometric segments (Figure 6). These are, from east to west, Dümrek,
Halileli and Tevfikiye segments, respectively. According to their mountain front sinuosity
ratios, Tevfikiye and Halileli segments in the west have similar properties and are more
active (Smf=1.15 and 1.48, respectively), while Dümrek segment in the east has relatively
lower activity (Smf=2.52).

Morphological cross‐sections were prepared in five areas along Troy fault (Figure 7). While
there is disharmony between Dümrek and Halileli segments with regard to normal fault
scarp elevation, an agreement is observed between Halileli and Tevfikiye segments. Normal
fault scarp values also indicate that Halileli and Tevfikiye geometric segments should be
evaluated together.

Figure 6. Digital elevation model of the study area and on this model, the segments of Troy fault ac‐
cording to mountain front sinuosity ratio.
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 179

Figure 7. Locations of cross‐sections along Troy fault

When the ratio of valley floor width to valley height (Vf) is calculated, the parameters in the
formula are calculated at a certain distance from the mountain front for each valley. Higher
Vf values show lower uplift speed and lower Vf values indicates actively uplifted areas.

Vf values showed that Tevfikiye and Halileli segments show similar properties and are
more active; on the other hand, Dümrek segment has higher Vf values and hence is less
active.

As a result of the morphotectonic analysis studies, since Tevfikiye and Halileli segments
display similar morphotectonic characteristics, they were evaluated together and named the
Troy segment. Since Troy segment is more active compared to Dümrek segment, the
shallow geophysical and paleoseismological investigations related to the Troy fault were
mostly conducted on this segment.
180 Tectonics – Recent Advances

On the other hand, the expected maximum earthquake magnitude for Troy fault was
calculated using the empirical equations proposed by Pavlides and Caputo (2004). The total
length of the Troy fault is about 11‐12km, hence, the expected maximum earthquake
magnitude for the Troy fault was computed from the formula

Ms= 0.9‐log (SRL) + 5.48 (Pavlides and Caputo, 2004)

as Ms= 6.2‐6.5

4.2. Kinematic analysis studies along the Troy fault


Troy fault system, located within the right lateral deformation zone between the northern
and southern branches of NAFS, is represented by approximately E‐W oriented normal
faults, NE‐SW oriented right lateral and NW‐ SE oriented left lateral strike slip faults. When
taken into account its geologic fault length and morphotectonic characteristics, the Troy
fault is the most important one of the approximately E‐W oriented normal faults within the
Troy Fault System.
Troy fault was defined for the first time by Tutkun and Pavlides (2005). It is a normal fault
which is 11‐12 km long, approximately E‐W oriented, dipping 600 to the north (see Figure 3).
Troy fault has highly pronounced normal fault morphology and is composed of two seg‐
ments: Troy and Dümrek segments (See Figure 6).The Troy fault cuts Lower Cretaceous
aged Denizgören ophiolites in the NE parts of the study area. To the east of Dümrek village,
the fault makes a bend towards the west and continues in E‐W direction, and it brings Neo‐
gene aged Çanakkale Formation and Quaternary deposits face to face in this area (Kürçer et
al., 2006).
The fault enters the alluviums of Karamenderes Stream starting from the area where ancient
Troy settlement is located, but its continuation from here towards the west is not clear.
However, when its direction is followed, it can be seen that the fault trace again becomes
evident between Ballıburun and Kesiktepe (See Figure 3). That’s why, the sector of the Troy
fault between ancient Troy settlement and Ballıburun was interpreted as probable active
fault. In the drilling works we conducted in the Quaternary sediments in the hanging wall
of the Troy fault, the top of the Çanakkale Formation was cut at 7‐8 meter depth in the drill‐
ings B‐4 and B‐5 (See Figure 3) just north of the Troy fault. In the morphological cross‐
section (See Profile B in Figure 7) taken from an area between B‐4 and B‐5 drill locations, a
normal dip slip of about 50 meter was measured. When the top of Çanakkale formation is
taken as reference plane, cumulative dip slip in the Troy fault is around 60 meters including
7‐8 meters obtained from the drillings.

In the study area, apart from the Troy fault, strike slip reverse and normal faults are
observed, as well. In this context, within the Troy fault system, from 4 sites (for site
locations, see Figure 3, Table 1) where outcrop conditions permit, a total of 67 fault planes
were measured and calculated using numerical analysis method (Table 2). Site 1 was
measured in Neogene aged Çanakkale formation. Thanks to this site, the kinematic
condition of the final tectonic regime in the Troy region was determined.
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 181

Site Longitude (N) Latitude (E) Age Lithology


1 35 s 0435107 4423466 Neogene Sandstone
2 35 s 0446204 4426875 Lower Cretaceous Serpantinite
3 35 s 0446695 4427003 Lower Cretaceous Serpantinite
4 35 s 0445035 4426292 Lower Cretaceous Serpantinite
Table 1. Coordinates of the sites where fault planes or fault assemblages outcropping in the vicinity of
Troy were measured; lithology and age of the measured geological units.

σ1 σ2 σ3
Site N Az / dip Az / dip Az /dip R M.D. S. D.
1 11 80 / 26 256 / 64 349 / 1 0.29 9.4 13.1
2 27 127 / 59 287 / 29 22 / 9 0.25 7.7 12.0
3 11 279 / 18 100 / 72 9/0 0.43 10.4 9.4
4 18 85 / 11 213 / 73 352 / 13 0.69 8.9 11.0
S1= 285  22° / 4° S3 = 15  21° / 6° Rm= 0.37
Table 2. The conditions of the principal stress axes [(σ1), (σ2), (σ3)] computed as a result of the evalua‐
tion of fault assemblages measured in the vicinity of Troy using Carey (1979) numerical analysis meth‐
od; R ratio, number of measurements (N), fault measurement sites, mean (M.D.) and standard deviation
(S.D.) values

The results of the faults (measured in the vicinity of Troy), obtained employing the method
developed by Carey (1979) for kinematical analysis of fault assemblages are as follows: Alt‐
hough the sites where fault assemblages were measured (Figure 8) are composed of geologi‐
cal units of various ages, one site (Site 1) was measured in Neogene aged units. Accordingly,
the tectonic regime effective at the present time implies transtensional strike slip faulting. In
harmony with this strike slip faulting (compression) normal faulting (tension) regime devel‐
oped, which represent the Troy fault as well.

Figure 8. Presentation of the kinematical analyses carried out at locations belonging to faultings given
in Table 2 on equal angle lower hemisphere (Wulf) (the distribution of the angle of deviation between
the predicted slip vector (τ) and computed slip vector (s) is given in histograms).
182 Tectonics – Recent Advances

For the purpose of correlating the results of the kinematical analyses of the fault assemblag‐
es with the focal mechanism inverse solutions of the recent earthquakes, focal mechanism
inverse solution of the earthquake (Figure 3, Table 3) with a magnitude of 3.1 that occurred
on 03.02.2008 was carried out (Figure 9). Focal mechanism inverse solution presents a fault‐
ing with strike slip normal component, developed under transtensional stress regime that is
characterized by WNW‐ESE oriented compression (P) and NNE‐SSW oriented tension (T).
This data is compatible with the stress conditions obtained from the kinematical analyses of
the fault assemblages (Figure 10).

Figure 9. Focal mechanism inverse solution of the earthquake that occurred on 03.02.2008 (Özden et al.,
2008)

Explanations Values
Date 03.02.2008
Time 15 09
Latitude 39.94 N
Longitude 26.22 E
Magnitude (Mw) 3.1
Depth (km) 12
Plane 1 55°/88°/‐165°
Plane 2 325°/75°/‐3°
P axis 302°/10°
T axis 210°/10°
Source BOUN‐KOERI
Focal solution and reference Özden et al., 2008
Table 3. Numerical values of the earthquake that occurred on 03.02.2008 (Özden et al., 2008)
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 183

Figure 10. Comparative presentation of the regional stress conditions obtained from the kinematical
analyses of the fault assemblages outcropping in the vicinity of Troy and the earthquake (black arrows
represent effective regimes and blue ones represent non‐effective regimes)

As a result; the Troy Fault System and stress regimes effective at the present time in the Troy
region show that a NNW‐SSE trending comppression and a NNE‐SSW oriented tension is
effective in this region and as a result of it, strike slip and normal faults developed under
transtensional tectonic regime.

4.3. Shallow geophysical surveys along the Troy fault


Structural geology studies conducted along Troy fault aroused the suspicion that the fault
scarp might have been degraded in the course of time. In order to determine the fault trace
within Quaternary sediments, shallow geophysical researches employing direct current
resistivity method were carried out. Within this context, shallow depths (0‐20m) were inves‐
tigated along a total of 8 profiles determined perpendicular to the fault trace along the Troy
fault (for profile locations, see Figure 3). Two‐dimensional inverse solutions of the obtained
profiles were made and geoelectric cross‐sections were constructed (Figure 11). According
to the geoelectric cross‐sections obtained by means of shallow geophysical studies, some
discontinuities were encountered (Figure 11).

According to the Direct Current Resistivity studies conducted on the profiles perpendicular
to the fault trace along Troy fault, it passes 30‐100 meters north of morphologic escarpment
bordering the Troy rise. This condition can be explained by the degradation of the Troy fault
scarp in the course of time.
184 Tectonics – Recent Advances

Figure 11. Geoelectric cross‐sections measured by Direct Current Resistivity method along Troy fault
and obtained employing two‐dimensional inverse solutions. A: Profile 1a‐b, B: Profile 1c‐d, C: Profile 2,
D: Profile 3, E: Profile 4, F: Profile 5, G: Profile 7, I: Profile 8 (for profile locations, see Figure 3).
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 185

4.4. Shallow drilling works on the hanging wall of the Troy fault
In active tectonic research, in order to determine and corroborate basin asymmetry, shallow
(0‐15 m.) drilling works can be carried out. Lateral and vertical variations of the Quaternary
sediments within the basin can be determined with the help of shallow drillings. In this
study, a total of 5 shallow drillholes having depths of 3‐9m were opened in the area where
Quaternary sediments outcrop, north of the Troy fault (for drill locations, see Figure 3).

The phases of the shallow drilling works are presented in Figure 12. In the drilling works,
“Percussion drilling set for soils with gasoline powered percussion hammer‐Cobra MK1”
core drill machines were used. In the first phase, Quaternary sediments were drilled with
the core drill machine (Figure 12A). Then with the help of a simple lever of the drill
machine, the drill set was hoisted (Figure 12B). After labeling of the hoisted drill cores
(Figure 12C), the drill cores were split into two along their long axes (Figure 12D). After
these split cores were photographed (Figure 12E) and defined, the drill work was completed
(Figure 12F). The well logs of the drillholes are given in Figure 13. According these well
logs, Quaternary sediments become thicker towards the Troy fault.

When the information obtained from the shallow drillings were transferred onto the three‐
dimensional geologic cross‐section of the region, it was observed that the plane constituting
the boundary between Quaternary deposits and Neogene aged units are tilted towards
south (towards the Troy fault) and consequently Quaternary sediments became thicker
(Figure 14). In other words, the area north of the Troy fault is tilted back. This corroborates
basin asymmetry and points to Quaternary activity of the Troy fault as well.

4.5. Paleoseismological trench works on Troy fault system


The tectonic regime type effective in Troy region at the present time is strike slip faulting
with normal component (transtensional) developed under WNW‐ESE oriented compression
regime. The geological and geophysical investigations we conducted in Troy region showed
that a NNE‐SSW oriented tensional regime depended on this compression was effective in
this region. The products of the transtensional tectonic regime effective in the region are E‐
W oriented normal faults, NE‐SW oriented right lateral and NW‐SE oriented left lateral
strike slip conjugate faults. In this study, paleoseismological trench studies were conducted
on the Troy normal fault, which is prominent by its geological fault length and morphotec‐
tonic properties, and on the right lateral strike slip Kumkale fault.

Within the scope of this work, a total of 4 trench works were carried out, 3 trenches along
the Troy fault and 1 along Kumkale fault (for trench locations, see Figure 3). The trenches
were excavated 15 meters long, 4 meters wide and 4 meters deep.

The existence of a fault, E‐W oriented and dipping to the north, just north of the Troy rise
was stated by various researchers (for example; Kayan, 2000; Tutkun and Pavlides, 2005;
Kürçer et al., 2006). The fault named the Troy Fault (Tutkun and Pavlides, 2005) brings in
contact the Neogene Çanakkale formation and Quaternary deposits outcropping in Dümrek
plain (see Figure 3). There is an elevation difference of 50 meters on average between the
186 Tectonics – Recent Advances

Figure 12. Phases of shallow drill works A: Drilling, B: Hoisting of drill set, C: Numeration, D: Splitting
of drill cores, E: Photographing, F: Logging
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 187

Figure 13. Well logs of shallow drillings conducted on the hanging wall of Troy fault

Figure 14. Three‐dimensional schematic geologic cross‐section of the Troy region constructed making
use of surface geology and drilling data

floor of the plain and the Troy rise. In the shallow drilling works conducted by us, it was
determined that Quaternary‐Neogene boundary in sectors near the fault was at 8 meters. In
the light of this information the cumulative dip slip on the Troy fault is thought to be ap‐
proximately 60 meters.
When all the information about Troy fault, known and obtained by this study, is evaluated
altogether, it is assumed that Troy fault was initially an E‐W oriented normal fault devel‐
oped within Western Anatolia Stress System (WASS). It gained some right lateral strike slip
character owing to the western extensions of NAFS that affected the region as of Late Plio‐
cene. The Troy fault that bears the traces of both WASS and NAFS has become one of the
most significant morphologic elements in the region at the present time.

On the Troy fault, paleoseismological trench studies were conducted in three locations. These
locations are: Tevfikiye (Figure 3, T1), Ballıburun (Figure 3, T2) and Kesiktepe (Figure 3, T3).

Tevfikiye Trench (T1)

Tevfikiye trench is located in an area east of the village road that connects Tevfikiye and
Yenikumkale villages, very near to the drillhole B‐5 and on the geophysical profile line
188 Tectonics – Recent Advances

number 2 (P‐2) (see Figure 3, T‐1). In this area, an excavation work was conducted that is 5‐
5.5 meters deep, 8 meters wide, 48 meters long and multibenched (Figure 15).
Tevfikiye trench was planned in S15E‐N15W direction starting from Neogene aged deposits
as to include the discontinuity indications on the geophysical profile line number 2 (see
Figure 11‐C). In figure 16, the composite trench cross‐section belonging to the Tevfikiye
trench location is presented. The composite trench cross‐section was constructed making use
of the excavation site belonging to Tevfikiye trench, the drilling information of drillhole B‐5
and the observation of the small Tevfikiye trench excavated in this region.
As it can be seen from the trench section, Tevfikiye trench was excavated, at first, 48 meters
long towards the north starting from the Neogene deposits. In the excavated 48‐ meter sec‐
tion, it was observed that Quaternary deposits unconformably overlie Neogene units with
an erosional contact. The contact between Quaternary and Neogene can be distinctly traced
over the 48‐meter section. No trace of deformation that can be attributed to active tectonism
was encountered in this section. However, numerous normal faults having dips not exceed‐
ing several centimeters and some paleoliquefaction structures in the type of fire structure
were observed within the Neogene aged deposits at the basement, between 14th and 15th
meters of the trench (Figure 17).
It was impossible to continue at the targeted depth towards north after 48th meter due to
conditions of groundwater level. Thereupon, an additional trench with an average depth of
6 meters was excavated between 55th and 58th meters. No trace of deformation within Qua‐
ternary deposits at the upper parts of this trench was encountered, either. The Quaternary‐
Neogene contact was encountered at the depth of ‐6 meters (see Figure 16).
B‐5 drillhole is located at the 74th meter of Tevfikiye excavation area (see Figure 16). In B‐5
hole, Quaternary‐ Neogene contact was encountered at the depth of ‐8.10 meters (see Fig‐
ures 13 and 16). When additional trench information and B‐5 drilling data were transferred
onto the trench cross‐section belonging to Tevfikiye excavation area, it was seen that, at the
66th meter of the trench, Quaternary‐Neogene contact is at different depths. This point corre‐
sponds to the discontinuity sign encountered at the 96th meter of the geophysical profile
number 2 (see Figure 11‐C). Geological observations in the Tevfikiye excavation area and
shallow geophysical data point to the presence of a fault the south block of which was up‐
thrown at the 66th meter of the excavation area. As can be seen from the composite trench
cross‐section prepared according to geological information, if a comparison is made taking
Quaternary‐Neogene contact as a basis, it is seen that the south block is uplifted about 2
meters compared to the north block. When evaluated under the regional tectonic frame‐
work, this fault is thought to be a normal fault dipping to the north (the Troy fault). No data,
related to the continuation of the fault in the Quaternary deposits, was obtained from Tevf‐
ikiye trench. The age B.C.1190‐1140 was obtained from the sample numbered TEV‐E‐04
compiled from Tevfikiye trench. TEV‐E‐04 sample was taken from the greenish yellow col‐
ored silty clay number 2. No trace of deformation was encountered within the unit number
2 in the Tevfikiye excavation area. This indicates that no earthquake occurred resulting from
the Troy fault during the period from the deposition of the unit number 2 up to the present
(B.C. 1190‐Present‐day), at least in Tevfikiye trench area.
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 189

Figure 15. Look at Tevfikiye Trench from the North

Figure 16. Composite section of the East Wall of Tevfikiye Trench (T1)
190 Tectonics – Recent Advances

Figure 17. Small scale normal faults and paleoliquefaction structures observed between 14th and 15th
meters of Tevfikiye Trench (Look from West to East)

Ballıburun Trench (T‐2)


Ballıburun Trench is located between the electric resistivity profiles P‐7 and P‐8 at the
Ballıburun location west of Tevfikiye (see Figure 3, T‐2). The trench was excavated 22 me‐
ters long, 3 meters wide and 4 meters deep on average. Ballıburun trench was planned,
starting from Neogene sediments, in S10W‐N10E direction as to involve the discontinuity
signs on the geophysical profile line number 8 (see Figure 11‐I). In Figure 18, the trench
cross section and the photograph belonging to the west wall of Ballıburun Trench are pre‐
sented. As can be understood from the trench section, no sign of faulting was observed in
Ballıburun trench. However, some paleoliquefaction structures that can be secondary proof
of an earthquake were observed within the mud which is black colored and rich in organic
matter, at a depth of ‐2 meters from the surface, between 12th and 14th meters of the trench
(Figure 18B).

Three samples were collected and dated from Ballıburun trench. Two of these samples were
collected from Unit 1 and the third one was taken from Unit 3 in which paleoliquefaction
structures were observed. Since the age of the sample taken from Unit 3 is greater than the
other two samples, it was considered as an allochthonous sample and was not counted in
the interpretation. The ages of the remaining two samples support each other. As there is no
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 191

usable age information from Unit 3, in which paleoliquefaction structures were observed, it
can be said that the age of the earthquake, that might have caused these paleoliquefaction
structures, is younger than the ages obtained from the samples BAL‐5‐13C and BAL‐5‐4C.
Since the youngest age obtained from Unit 1 in the Ballıburun Trench is B.C. 760, the age of
the earthquake that might have caused these paleoliquefaction structures is the time interval
B.C. 760‐Present‐day.

Kesiktepe Trench (T‐3)

The Troy fault enters the alluvium deposits of Karamenderes brook starting from the west
of Tevfikiye village and loses its morphologic trace. It becomes again morphologically
distinct in the sector between Ballıburun and the Aegean Sea (see Figure 3). In the highland
that is located west of Ballıburun and composed of Neogene aged Çanakkale formation is
known as “Kesik Tepe” (Truncated Hill) and this area is characterized by its “V” shaped
morphology. It is thought that this morphologic anomaly was initially shaped by Troy fault
and later modified by human activity. For the purpose of checking the anomaly, electrical
resistivity measurements were carried out along P‐5 profile from Kesiktepe location and
geoelectric cross‐section was obtained (see Figure 11‐F). In the geoelectric cross‐section, a
discontinuity was determined in the region where Kesik Tepe is located and for the purpose
of testing this discontinuity, Kesiktepe Trench was excavated, which is 6 meters long, 2
meters wide and 4.5 meters deep on average (Figure 3, T‐3).The trench cross section of
Kesiktepe trench is presented in Figure 19.
No sign of faulting was encountered in Kesiktepe trench. However, Kayan (2000) mentions
that Neogene deposits were cut at different depths in two drillholes very near to each other,
drilled in the south and north of Kesiktepe. According to the researcher, while the Neogene
deposits were cut at ‐2 meter in the south, they were cut at ‐8meter just in the north. This
indicates that the north block was relatively downthrown. As to support this, some signs of
discontinuity were encountered at the 96th meter on the geophysical profile number 5 (see
Figure 11‐F). On the other hand, it is known that, during Troy era, in Kumkale plain some
drainage canals were excavated in order to dry the swamp that caused malaria (personal
communication with Dr. Rüstem Aslan. 2005; Figure 20). And one of these drainage canals is
situated just on the Troy fault trace in Kesiktepe location. By means of the discontinuity in
the geoelectric cross‐section, archeological information and geomorphological approach, it is
thought that the Troy fault reaches the Aegean Sea over Kesiktepe. However, since the ca‐
nal, excavated as drying canal, was filled with recent sediments during the period from its
last activity to the present time, no trace of active faulting was encountered at the first 4
meters from the surface (from B.C. 1500 up to the present).

Kumkale Trench (T‐4)

In the Troy region, there are also NE‐SW oriented, right lateral strike slip faults that were
developed within NAFS. The best examples for these faults are Kumkale and Yenimahalle
faults (see figure 3). Kumkale fault is a N250E oriented, dipping 750 to NW right lateral strike
slip fault that borders from the north the ridge on which Kumkale village is situated.
Yaltırak et al. (2000), in their shallow sea seismic studies conducted in Dardanelles and its
192 Tectonics – Recent Advances

Figure 18. A: West wall cross‐section of Ballıburun trench (T‐2) B: Paleoliquefaction structures
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 193

Figure 19. A: West wall cross section of Kesiktepe Trench (T‐3), B: Photograph of West Wall of Kesik‐
tepe Trench
194 Tectonics – Recent Advances

Figure 20. Relief geomorphological map of Troy region (Forchammer, 1850) Red lines show drainage
canals opened in order to dry the swamp

vicinity, documented NW‐dipping faults as the continuation of Kumkale fault. These faults
are the continuation of Kumkale fault in the sea. The total length of Kumkale fault is 15 km
including its continuation in the sea. Kumkale fault is thought to be one of the primary syn‐
thetic strike slip faults developed within NAFS.

Kumkale fault brings Neogene units and Quaternary deposits face to face (Figure 21). This
relationship could be observed in the open outcrop in Figure 21 B. Afterwards; this
observation point was leveled with the help of a working machine and transformed into a
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 195

trench (see Figure 3, T‐5). In Figure 22, the photomosaic and trench cross section are
presented.

Four samples were taken from T‐5 Trench and ages could be obtained from three of them.

Through the evaluation of trench microstratigraphy and C‐14 age data together, two
earthquakes were defined in T‐5 Trench. The first of them is associated with the faulting that
cuts Unit C and is covered by Unit B. Consequently, the event horizon for this fault (Event II)
is between Units C and B. The age obtained from Unit C is between B.C. 40 and A.D. 130, and
the age obtained from Unit B is between A.D. 780 and 1000 (see Figure 24). According to
these age data, the first earthquake occurred between A.D. 130 and 780. The second fault
(Event I) is associated with the faulting that cuts Unit B and is covered by Unit A. The event
horizon for this earthquake is between Units B and A. The age obtained from Unit B is A.D.
780‐1000. And the age obtained from Unit A is between A.D. 1300 and 1430 (see Figure 22).
According to these age data, the second fault (Event I) occurred between A.D. 1000 and 1300.

Figure 21. A: Panoramic sight of Kumkale fault, B: Sight of Kumkale fault in the outcrop
196 Tectonics – Recent Advances

Figure 22. Photomosaic (A) and cross‐section (B) of the south wall of Kumkale Trench (T‐5)

5. Results and discussion


As a result of this study;
According to the morphotectonic studies conducted on the Troy fault, the Troy fault is di‐
vided into three segments, from west to east, as Tevfikiye, Halileli and Dümrek segments.
Since their mountain front sinuosity ratio (Smf) and valley floor width to valley height ratio
(Vf) are similar, Tevfikiye and Halileli segments were evaluated together and given the
name “the Troy segment”.
The mountain front sinuosity ratio of the Troy segment was computed as Smf= 1.28 and that
of Dümrek segment was computed as Smf=2.52. Taking into account these values, the Troy
segment was determined to be more active and therefore, subsequent geological and geo‐
physical studies were concentrated on the Troy segment.

In the calculations carried out using the empirical equations suggested in the studies that
investigate the relation between the geological fault length and the greatest earthquake (for
example, Pavlides and Caputo, 2004), the greatest earthquake magnitude that can result
from 11‐12 km long Troy fault was computed as M= 6.2‐6.5.

Both the results of the kinematic analysis studies of the fault assemblages and the focal
mechanism inverse solution of the earthquake having the magnitude of 3.1 that occurred on
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 197

the Troy fault on 03.02.2008, indicate a strike slip with normal component faulting
developed under transtensional stress regime characterized by WNW‐ESE oriented
compression (P) and NNE‐SSW oriented tension (T) in the Troy region at the present time.
The fault systems in the Troy region and the stress regimes which are effective at the present
time show that strike‐slip faults and concordant normal faulting in active in the Troy region.

According to the Direct Current Resistivity profile studies conducted on profile lines
perpendicular to the fault trace along the Troy fault, it was understood that the Troy fault
passes 30‐100 meters north of the morphological escarpment which borders the Troy rise
from the north. This situation can be explained by means of the degradation of the Troy
fault scarp in the course of time. According to the geoelectric cross‐sections obtained from
the shallow geophysical studies, some discontinuity signs were encountered. However, it
was observed that these discontinuities do not reach the surface but come to an end at the
depth of about ‐10 meters from the surface.

In the Quaternary deposition area north of the Troy fault, in order to examine the lateral
thickness variation of the Quaternary deposits towards the fault, 5 shallow drilling works
with depths varying between 3 and 9 meters were conducted at 5 points along a direction
perpendicular to the Troy fault. When the information obtained from these drilling works
were transformed onto the three dimensional schematic geological cross‐section, it was seen
that the plane which constitutes the boundary between the Quaternary deposits and the
Neogene aged units gets thicker towards the south (towards the Troy fault). In other words,
this boundary is tilted towards the Troy fault. This corroborates the asymmetry of the basin
and at the same time points to the Quaternary activity of the Troy fault. From the paleo‐
seismological trench studies it was understood that no earthquake resulting from the Troy
fault occurred in the last 3000 years from B.C. 1190 up to the present, in Tevfikiye trench.
Since the thickness of the sediments accumulated on the fault during the period from the
last earthquake which resulted from the Troy fault up to the present was greater than the
depth that can be reached in paleoseismological trench works, it was not possible to reach
traces of faulting in the trench excavations.

In Ballıburun trench, some paleoliquefaction structures, developed during the time from
B.C. 760 up to the present as a result of an earthquake that was originated from the nearby
faults, were observed.

Two earthquakes were dated in the paleoseismological studies conducted on the Kumkale
fault. The first of these faults occurred between A.D. 130 and 780 and the second one
occurred between A.D. 1000 and 1300.

The paleoliquefaction structures observed in Ballıburun trench might have resulted from the
Kumkale fault.

In this study, it was understood that the earthquakes which destroyed Troy III and Troy VI
layers had not developed depending on the Troy fault system. Although their source is not
clearly known, these earthquakes might have originated from the active faults in Biga
Peninsula or from Gaziköy‐Saroz fault which represents the north branch of NAFS. The
198 Tectonics – Recent Advances

isoseist maps of the large earthquakes that caused surface faulting on Gaziköy‐Saroz,
Yenice‐Gönen and Biga faults in this region, in the last century are presented in Figure 23.

Figure 23. Isoseist maps of Mürefte‐Şarköy (1912, Yenice‐Gönen (1953), and Biga earthquakes (1983)
(Kalafat et al., 2007)

As can be clearly seen in Figure 24, the region where Troy is located was affected at an in‐
tensity of VIII‐IX by all three earthquakes. When such properties as local soil conditions and
building style are taken into account, it is possible to state that historical earthquakes result‐
ed from these faults affected Troy region. Rockwell et al., (2001) conducted paleoseismologi‐
cal studies on Gaziköy‐Saroz segment of NAF in the vicinity of Kavakköy and documented
5 earthquakes from the present‐day backwards. Kürçer et al., (2008), in their paleoseismo‐
logical studies carried out on Yenice‐Gönen fault, documented two earthquakes that they
dated to A.D. 620 and A.D. 1440. In the light of all this information, it is the possible that the
earthquakes that affected the Troy region might have resulted from the other active faults in
the region.

Author details
Akın Kürçer
General Directorate of Mineral Research and Exploration,
Department of Geology, Çankaya, Ankara, Turkey
An Assessment of the Earthquakes of Ancient Troy, NW Anatolia, Turkey 199

Alexandros Chatzipetros, Spyros Pavlides,


George Syrides, Kostas Vouvalidis and Özkan Ateş
Aristotle University of Thessaloniki, Faculty of Sciences,
Department of Geology, Thessaloniki, Greece
Salih Zeki Tutkun and Süha Özden
Çanakkale Onsekiz Mart University, Department of Geological Engineering, Çanakkale, Turkey
Emin Ulugergerli and Yunus Levent Ekinci
Çanakkale Onsekiz Mart University, Department of Geophysical Engineering, Çanakkale, Turkey

Acknowledgement
This study was supported by a bilateral cooperation project between TUBİTAK (The Scientific
and Technological Research Council of Turkey) and GSRT (General Secretariat for Research and
Tecnology). Grant: TUBİTAK ‐ 105Y360. We would like to warmly thank the following col‐
leagues that assisted in the fieldwork during various field periods: A. Michailidou, S. Sboras, S.
Valkaniotis, A. Zervopoulou (from Aristotle Universityof Thessaloniki, Department of Geology)
and Gündoğdu, Y (from Ankara University, Department of Geophysical Engineering).

6. References
Aksoy, M., E., Meghroui, M., Vallee, M. and Çakır, Z., 2010. Rupture characteristics of the is
A.D. 1912 Mürefte (Ganos) earthquake segment of the North Anatolian fault (western
Turkey). Geology, November 2010; v. 38; no. 11; p. 991‐994; doi: 10.1130/G31447.1.
Ambraseys, N.,N. and Finkel, C.,F., 1991. Long‐term seismicity of İstanbul and of the
Marmara region, Engin. Seis. Earthq. Engin. Report, 91/8, İmperial College.
Canıtez, N. and Toksöz, M.N. 1971. Focal mechanism and source depth of earthquakes from
body and surfacewave data. Bulletin of Seismological Society of America 61, 1369–79.
Carey, E.; 1976. Analyse numérique dʹun modéle mécanique élémentaire appliqué a lʹétude
dʹune population de failles : calcul dʹun tenseur moyen des contraintes a partir des
stries de glissement. Thése de 3º cycle, Université de Paris‐Sud, Orsay, 138 p.
Carey, E., 1979. Recherche des directions principales de contraintes associées au jeu dʹune
population de failles, Revue Geological Dynamic and Géography physic., 21, 57‐66
Jackson J., McKenzie D., 1984. Active tectonics of the Alpine–Himalayan Belt between
Western Turkey and Pakistan, Geophy. J. Royal Astr. Soc. 7 (1984) (1984) 185–264.
Kalafat, D. 1988. Active tectonics and seismicity of SW Anatolia and its vicinity. Bulletin of
Earthquake Research, 63, 5–98. Ankara (in Turkish with English abstract).
Kalafat, D., Güneş, Y., Kara, M., Deniz, P., Kekovalı, K., Kuleli, S., Gülen, L., Yılmazer, M., Özel,
N., 2007. A revised and extended earthquake cataloque for Turkey since 1900 (M≥ 4,0).
Boğaziçi University, Kandilli Observatory and Earthquake Research Institute, İstanbul.
Keller, A. E., and Pinter, N., 1996; Active Tectonics, Earthquakes, Uplift and Landscape.
Prentice Hall (ISBN 0‐02‐304601‐5) N. Jersey ‐pp. 377. (Second Edition 2002).
Kürçer, A. and Tutkun, S.Z., 2008. The role of the geological and tectonic processes on
geomorphological evolotuion of the Biga Peninsula, NW Turkey. National
Geomorphology Symposium, Abstracts, 213‐214. Çanakkale, 2008.
200 Tectonics – Recent Advances

Kürçer, A., 2006. Neotectonic Features of the vicinity of Yenice – Gönen and
Paleoseismology of March 18, 1953 (Mw:7,2) Yenice‐Gönen Earthquake Fault (NW
Turkey) (Çanakkale Onsekiz Mart University Natural and Applied Sciences Instıtue ‐
2006). p. 170 (in Turkish with English abstract).
Kürçer, A., ,Tutkun, S. Z., Pavlides, S., Chatzipetros, A., Ateş, Ö., Özden, S., , Ulugergerli, E.,
Gündoğdu, Y., Bekler, T., Syrides, G., Vouvalidis, K., Valkaniotis, S., Zervopoulou, A.,
Şengül, E., Ekinci, Y. L.,Köse K., Demirci, A. and Elbek, Ş., 2006, Morphotectonical
Features of Troia Fault and Preliminary Paleoseismological Studies, NW Turkey. 10th
Meeting of the Active Tectonic Research Group 10th. Proceedings; pp: 60., 2‐4
November 2006, Dokuz Eylül University, Department of Geology, İzmir.
Kürçer, A., Chatzipetros, A., Tutkun, S.Z., Pavlides, S., Ateş, Ö. and Valkaniotis, S. 2008. The
Yenice‐Gönen Active Fault (NW Turkey): Active Tectonics and Palaeoseismology,
Tectonophysics, 453, 263‐275.
McKenzie, D.P., 1972. Active tectonics of the Mediterranean region, Geophys. J. R. Astr. Soc.,
30 (2), 109‐185.
McKenzie, D.P., 1978. Active tectonics of the Alpine‐Himalayan belt: The Aegean sea and
surrounding regions (tectonics of aegean region), Geophys. J. R. Astr. Soc., 55, 217‐254.
Okay, I.A, (1987); Geology and tectonics of the western part of Biga peninsula. TPAO,
Report Number: 2374.
Özden, S.,Bekler, T., Tutkun, S. Z., Kürçer, A., Ateş, Ö., Bekler, F., Kalafat, D.,Gündoğdu, E.,
Bircan, F., Çınar, S., Çağlayan,Ö., Gürgen, M., İşler, H. and Yalçınöz, A., 2008,
Seismotectonics of Biga Peninsula and South of Marmara Sea. Active Tectonic Research
Group 12th. Abstracts; pp: 48‐49, Akçakoca, MTA, Turkey.
Pavlides, S. and Caputo, R., 2004. Magnitude versus faults’ surface parameters:quantitative
relationships from the Aegean. Tectonophysics, 380 (3–4), 159–188.
Rockwell, T., Barka, A., Dawson, T., Akyüz, S. and Thorup K. 2001. Paleoseismology of the
Gazikoy‐Saros segment of the North Anatolia fault, northwestern Turkey: Comparison
of the historical and paleoseismic records, implications of regional seismic hazard, and
models of earthquake recurrence. Journal of Seismology, 5: 433‐448.
Şaroğlu,F., Emre, Ö. and Kuşcu, İ., (1992), Active fault map of Turkey, MTA, Ankara.
Şengör, A.M.C., Görür, N. and Şaroğlu, F., 1985. Strike‐slip faulting and related basin
formation in zones tectonic escape: Turkey as a case study, in strike‐slip deformation,
Basin formation and Sedimentation, edited by Biddle, K.T. and Christie‐Blick, N., Soc.
Econ. Paleontol. Mineral. Spec. Publ., 37, 227‐264.
Şentürk, K. and Karaköse, C, 1987, Geology of Çanakkale strait and its vicinity, MTA Report
Number: 9333 (unpublished).
Taymaz, T. 2001. Active tectonics of the North and Central Aegean Sea. Symposia on
Seismotectonics of the North‐Western Anatolia‐Aegean and Recent Turkish
Earthquakes, XI‐XIX. Istanbul: ITU, 113 pp.
Tutkun, S.Z. and Pavlides, S.B. 2005. The Troy Fault, Bulletin of the Geological Society of
Greece, Vol. XXXVII, 194‐200.
Yaltırak, C., Alpar, B., Sakinc¸, M., Yüce, H., 2000. Origin of the Strait of Çanakkale
(Dardanelles): regional tectonics and the Mediterranean–Marmara incursion. Mar. Geol.
164, 139– 156 (with Erratum 167, 189–190).
Chapter 8

Paleoseismological Three Dimensional Virtual


Photography Method; A Case Study:
Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone,
Central Anatolia, Turkey

Akın Kürçer and Yaşar Ergun Gökten

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/48194

1. Introduction
In order to make earthquake risk analysis in reliable manner in the regions carrying seismic
risks, earthquake behavior of the active faults present in that region must be well deter‐
mined. An active fault can be defined as a fault that produced earthquakes in Quaternary
time associated with surface rupturing or deformation and has the potential to produce
earthquakes in the future. The most important method which is accepted and commonly
used at the present time to reveal earthquake behavior of active faults is paleoseismology.
Paleoseismology is a method that tries to obtain information on the location, nature and date
of historical earthquakes making use of geological and geomorphological data (McCalpin
and Nelson, 2009).

Paleoseismic trenching is one of the methods that is frequently applied and provides
noteworthy data in paleoseismology (Pantosti and Yeats, 1993; Demirtaş, 1997). In this
method, geological evaluations are made according to the principles of sedimentology,
stratigraphy and structural geology within trenches excavated perpendicular or parallel
to the active fault trace depending on the faulting type. Appropriate samples collected
from sediments are enable to date the historical earthquakes by suitable radiometric da‐
ting techniques.

Excavation types and dimensions of paleoseismological trenches can show variations


according to the properties of the studied fault (faulting type, annual slip rate, earthquake
recurrence interval, amount of displacement occurred in each earthquake, the elapse time
from the last known historical earthquake until today, etc.), and the physical parameters of

© 2012 Kürçer and Gökten, licensee InTech. This is an open access chapter distributed under the terms of
the Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
202 Tectonics – Recent Advances

the trench site (groundwater condition, stability of trench sediments, etc.). The
paleoseismological excavations which are conducted in faults having relatively lower slip
rate ( 1 mm / year) and in which dip slip is dominant, are made deeper and wider. In areas
where the groundwater level is shallow or the trench walls are not stable, trenches are
excavated benched and/or sloped for the sake of safety.

After the trench walls are cleaned, gridding is performed taking into consideration detailed
geological features within the trench. In gridding, a grid interval of 1 m2 is generally accepted
as standard (McCalpin, 2009). After gridding the opposite walls in same scale the stratigraph‐
ic levels and displacement figures on the walls are visually recorded and photographed.
Logging can be conducted based on two principles: subjective and objective (Hatheway and
Leighton, 1979). In subjective logging, the logger reflects his personal point of view in the
trench log. As for the objective logging, in that, all sorts of details observed on trench walls
are tried to be reflected in the trench log rather than the personal point of view. Although
objective logging is possible in theory, in practice its application is difficult and most of the
time not fit for purpose. For this reason, the logger must transfer paleoseismological details
on the trench walls with an objective point of view adding some subjective interpretation.
Three methods are used in trench logging. These are: manual trench logging, electronic
trench logging and photomosaic trench logging (McCalpin, 2009). These methods have nu‐
merous advantages and disadvantages compared to each other (for further detail, see
McCalpin, 2009).

Trench logs are, after all, two dimensional figures that reflect the personal interpretation of
the logger about that trench. In recent years, photomosaic trench logging technique started
to be frequently applied in paleoseismology. Photomosaic trench logging technique is a fast
and inexpensive method which minimizes the possible errors may be rised from drawing.
However, this method remains inefficient in deep and benched trench excavations since it
causes some image losts on photographs during stitching them together.

In this article, a new photographic method is suggested to be used for paleoseismological


trench works and this method is named ‘Paleoseismological Three Dimensional Virtual
Photography Method’. This method has been applied for the first time in the paleoseismo‐
logical study conducted on the Tuz Gölü Fault Zone (Central Anatolia, Turkey).

2. Tuz Gölü Fault Zone – Akhisar‐Kılıç segment


The Tuz Gölü Fault Zone (TGFZ) is one of the most important intracontinental active fault
zones of the Central Anatolian Region (Şaroğlu et al., 1992; Dirik and Göncüoğlu, 1996;
Çemen et al., 1999; Koçyiğit, 2000) (Figure 1). When taken into account its morphotectonic
features and the distribution of the epicenters of the earthquakes with magnitudes reaching
up to 5, it can be said that TGFZ is still active (Dirik and Erol, 2000). TGFZ is a NW‐SE
trending, approximately 200 km long, oblique‐slip normal fault zone with a right lateral
strike‐slip component. It extends from Tuz Gölü to the northwest and Kemerhisar (Niğde)
to the southeast. This fault zone is at the same time a transition zone that separates the
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 203

region of normal component strike slip neotectonic regime to the east from the region of
extensional neotectonic regime to the west (Koçyiğit and Özacar, 2003).
TGFZ is composed of fault segments parallel or subparallel to each other and with lengths
varying between 4 and 33 km. Because of its geological fault length and morphotectonic
features, the Akhisar‐Kılıç Segment (AKS) constitutes one of the most important fault seg‐
ments of TGFZ. AKS is a 25 km long, N250 ‐ 300W trending fault. It is located in the middle
section of TGFZ and it extends between Akhisar Village and Hasandağ (Figure 2).

In Figure 3, a geologic map illustrates the region along the Akhisar‐Kılıç Segment. According to
this map, the Akhisar‐Kılıç Segment cuts Lower Pliocene aged Kızılkaya ignimbrites in the
vicinity of Akhisar village to the northwest. Between Akhisar and Yuva villages, it forms the
boundary between Oligocene aged Yassıpur formation and Quaternary aged alluvial fan
deposits and in places it cuts these alluvial fan deposits. The Akhisar‐Kılıç Segment cuts
Hasandağ volcanites starting from the northwest of Yuva village. The fault segment which
borders from the northeast the NW‐SE trending area of depression with an ellipsoidal geometry,
at the same time cuts Late Pleistocene‐Holocene deposits in places. It again cuts Hasandağ
volcanites in the section starting from the southeast of Helvadere to Kılıç Ridge and comes to an
end at Kılıç Ridge locality. The Akhisar‐Kılıç Segment is characterized by the alluvial fans
arranged parallel to each other and the linear fault scarps between Akhisar and Yuva villages. A
number of cold and brackish water springs are present along the fault trace in this area.

Figure 1. Main neotectonic elements of Eastern Mediterranean Region and Location of Tuz Gölü Fault
Zone (Modified from Okay et al., 2000). Black arrovs and corresponding number show GPS‐derived
plate velocities (mm‐year) relative to Eurasia (Reilinger at al., 2006).
204 Tectonics – Recent Advances

Figure 2. Segments of Tuz Gölü Fault Zone and setting of Akhisar‐Kılıç Segment and Bağlarkayası‐2010
Trench on the Digital Elevation Model of the region (Shuttle Radar Topography Mission‐SRTM data
were used for the Digital Elevation Model)

3. Bağlarkayası‐2010 trench
Bağlarkayası‐2010 Trench excavation is the first paleoseismological study carried out on the
Tuz Gölü Fault Zone. The Trench is located in the middle part of the Akhisar‐Kılıç Segment
(See Figure 3, GPS Coordinates: 34.1785050 E – 38.2355610 N). The trench site is situated with‐
in the fault‐controlled Quaternary area of depression to the southeast of Yuva village. In this
depression area, Hasandağ volcanites were cut by the fault and obliquely displaced. Since
the normal component of the fault was dominant, the hanging wall was tilted towards the
fault and owing to this back‐tilting a topographical saddle came into existence.
Bağlarkayası‐2010 Trench was excavated on this topographical saddle (Figure 4).
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 205

Figure 3. Geological map of Akhisar‐Kılıç Segment and its vicinity (Modified from Dönmez et al., 2005).
206 Tectonics – Recent Advances

Figure 4. Panoramic view of Bağlarkayası‐2010 Trench area (view from NE to SE)

In the selection of trench site, the fault plane outcrops existing in the near north and south of
the trench area were made use of (See, points 1 and 2 in Figure 3 and Figure 5). In addition
to this, 8 vertical electric soundings with intervals of 250 meters, along a line perpendicular
to the fault trace were carried out in the area which was considered for the trench site. In
these vertical electric soundings, data were gathered from a depth of approximately 2000
meters. A geoelectric cross section was generated making use of these data. On the
generated geoelectric cross section, a fault that reached up to the surface was detected
(Figure 6). The site of the trench was selected by integrating the fault data determined from
the geoelectric cross section and structural observations with the geomorphology. The 1/500
scale microtopographical map of the Bağlarkayası‐2010 Trench area is presented in Figure 7.
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 207

Figure 5. Fault planes belonging to Akhisar‐Kılıç Segment that cuts Hasandağ volcanites A. SE of Yuva
Village, B. NE of Koçpınar Village (see Figure 3 for Locations)
208 Tectonics – Recent Advances

Figure 6. A. Apparent iso‐resistivity cross section, B. Geoelectric cross section, generated from Vertical
Electric Soundings (see Figure 3 for Profile Location)
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 209

Figure 7. Microtopographic map of the Bağlarkayası 2010 Trench site

The Bağlarkayası‐2010 Trench was excavated perpendicular to the fault. It has a length of 94
meters, a width of 5 meters and an average depth of 5 meters (maximum depth: 8.5 meters).
While the northern 40‐meter section of the trench was excavated as a single slot trench, the
southern part of it was excavated as multi‐bench trench (Figure 8). A gridding of 1 m2 was
applied in the trench. Manual trench logging method was applied for the entire trench, at a
scale of 1/20. In Figure 9, the stages of the trench work are presented.
210 Tectonics – Recent Advances

Figure 8. Aerial panoramic view of Bağlarkayası‐2010 Trench, Photograph was taken from an altitude
of 25 meters at inclined angle, from the fire tower (view from SW to NE)
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 211

Figure 9. The stages of the trench work A. Excavation and cleaning, B. Gridding, C. Manual logging D.
Photographing

In Figures 10 and 11, wall logs belonging to Bağlarbaşı‐2010 Trench are presented. Seven
different microstratigraphic levels were identified within Bağlarbaşı‐2010 Trench. The first
two of these levels that are relatively older were interpreted as ash and block flows of
Hasandağ volcanism. And the relatively younger units are deposits associated with plinian
activity and fluvial processes. Three deformation zones were defined within the trench.
These zones are main fault zone, synthetic faulting zone and antithetic faulting region. The
photomosaic of the main fault zone on the NW wall of the trench is seen in Figure 12.
212 Tectonics – Recent Advances

Figure 10. NW Wall Log of Bağlarkayası‐2010 Trench (close shot of general log and main fault zone)
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 213

Figure 11. SE Wall Log of Bağlarkayası‐2010 Trench (close shot of general log and main fault zone)
214 Tectonics – Recent Advances

Figure 12. The photomosaic of the main fault zone on the NW wall of the trench; A : Uninterpreted B :
Interpreted
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 215

Since carbon‐rich material is limited within Bağlarkayası‐2010 Trench, an supplementary


trench was excavated approximately 100 meters SW of the trench (see Figure 4 and 7). In
this trench, at a depth of 2 meters from the surface, some bones were encountered within a
level that can be correlated with the unit 5 of Bağlarkayası‐2010 Trench (Figure 13).
According to the determination of Anthropologist Dr. Gerçek Saraç, these bones belong to a
human.

Figure 13. Human bones detected within the supplementary trench

18 samples from Bağlarkayası‐2010 Trench and 3 samples (2 of which were bones) from the
Supplementary Trench were collected. The radiometric age determination (14 C ‐AMS) of the
total 21 samples was carried out at BETA Analytical Lab (Table 1). The 14C ages determined
by BETA Analytical Laboratory were afterwards evaluated making use of Oxcal v3.10 cali‐
bration program developed by Ramsey (2001) (Figure 14).
216 Tectonics – Recent Advances

Table 1. Results of 14C Age Determination for the samples collected from Bağlarkayası‐2010 Trench and
Bağlarkayası Observation Trench
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 217

Figure 14. Radiocarbon ages calibrated making use of Oxcal v3.10 calibration program (Ramsey, 2001)
of the samples collected from Bağlarkayası‐2010 Trench and Bağlarkayası Observation Trench

As a result of the evaluation carried out making use of paleoseismological criteria such as
trench microstratigraphy, geometry of fault colluvial wedge, upward termination of fault
strands, and 14C age data, two paleoseismic events were described within the Bağlarkayası‐
2010 Trench, which occurred during the last 10 500 years. The first of these earthquakes
(penultimate event) occurred after the deposition of unit number 2 and as a result of it the
colluvial wedge number 3 was developed (see Figures 10, 11 and 12). The youngest age
obtained from the unit number 2 is 12 420  60 B.P. years. And the base age obtained from
the unit number 4 that covers the colluvial wedge number 3 is 9650  50 B.P. years. In addi‐
tion to these, the ages obtained from the colluvial wedge number 3 are 12 270  60 B.P. and
10 410  50 B.P. years, respectively. Colluvial wedges can be directly used in paleoseismo‐
logical studies as event horizons in dating of earthquakes. In the light of these data, the age
of the first earthquake (penultimate event) defined in the Bağlarkayası‐2010 Trench was
determined between the years 10 410  50 B.P. and B.P. 9650 410  50 B.P.
The second earthquake (Event 1) defined in the trench occurred after the deposition of the
unit number 5. Faulting traces belonging to this earthquake can be traced within the unit
number 5, but the faults do not cut the unit number 6 (see Figures 10, 11 and 12). The
youngest age obtained from the unit number 5 of the trench is 6790  40 B.P. years. In addi‐
tion, the human bones discovered within the observation trench were collected from the
unit number 5. The ages obtained from these bones were determined as 5590  40 B.P. and
5560  40 B.P. years, respectively. And the base age determined from the unit number 6 is
1370  40 B.P. years. According to these age data the last earthquake took place between the
years 5560  40 B.P. and B.P. 1370  40 B.P.

The amount of displacement occurred during the last earthquake (Event 1) was measured as
25 cm by taking the top of the unit 4 as reference plane (see Figure 12). Along the fault, the
amount of displacement in the older units is 50 cm. This shows that these two earthquakes
are of similar magnitude and an average displacement of 25 cm occurred in each
earthquake. This data demonstrates that the Akhisar‐Kılıç segment displayed characteristic
earthquake behavior in Holocene.
218 Tectonics – Recent Advances

On the other hand, it was determined that Kızılkaya Ignimbrites was displaced 268 meters
in the vertical direction by Tuz Gölü Fault Zone in the vicinity of Akhisar village (Figure 15).
The age of Kızılkaya Ignimbrites was determined by various researchers using different
methods. For example, employing K/ AR method, while Innocenti et al. (1975) obtained 5.4 
1.1 million years, Besang et al. (1977) 5.4  1.1 million years, and Schumacher and
Schumacher (1996) 4.3  0.2 and 4.5  0.2 million years of age for Kızılkaya Ignimbrites. On
the other hand, Aydar et al. (2012) obtained 5.19  0.07 and 5.11  0.37 million years of age
employing the method of Ar/Ar. According to these age data, the age of the Kızılkaya
Ignimbrites is approximately 5 million years. Taking into account the radiometric age (5
million years) determined in previous studies and the measured total displacement (268
meter), the annual slip rate of Akhisar‐Kılıç Segment of Tuz Gölü Fault Zone was calculated
as 0.0536 mm starting from Early Pliocene.

Through the evaluation of the annual slip rate (0.0536 mm/ year) of the Akhisar‐Kılıç
Segment together with the amount of vertical displacement (25 cm), the average earthquake
recurrence interval of the Akhisar‐Kılıç Segment has been found as 4664 years. The age of
the first earthquake (penultimate event) that occurred within the Bağlarkayası‐2010 Trench
was determined between 10 410  50 B.P. and 9650 410  50 B.P. years which can be accepted
as 10000 years BP in average. After 4664 years from this earthquake, which is the average
earthquake recurrence interval, the last earthquake (Event 1) must have been occurred. The
age of this earthquake is approximately 5336 B.P. years. Consequently, the elapse time from
the last earthquake to the present time is 5336 years. Since this value is greater than the
average earthquake recurrence interval that is 4664 years, the Akhisar‐Kılıç Segment could
produce an earthquake at any moment. By taking into consider the annual slip rate and the
elapse time from the last earthquake to the present time, the amount of average vertical
displacement accumulated on the Akhisar‐Kılıç Segment was calculated as 28.76 cm.

Figure 15. Google earth image of Kızılkaya Ignimbrites displaced by Tuz Gölü Fault, in the vicinity of
Akhisar Village; the amount of vertical displacement between the points A and B is 268 meters.
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 219

Supposing that the whole of the Akhisar‐Kılıç Segment, which has a length of 27 km, of Tuz
Gölü Fault Zone was broken, the maximum earthquake magnitude that this segment can
produce was calculated making use of the empirical equations proposed by Wells and
Coppersmith (1994).

According to these equations, the magnitude of the maximum earthquake that the Akhisar‐
Kılıç segment could produce (M), the amount of maximum displacement (MD) and the
amount of average displacement (AD) were computed as follows:

M = a + b x log (SRL)
a = 4.86
b = 1.32
SRL = 27 km
M = 4.86 + 1.32 x log (27)
and M was computed as M = 6.74 .
log (MD) = a  b x M
M = 6.74
a = 5.90
b= 0.89
log (MD) = 0.0986
and MD was computed as MD = 1.25 meters.
log (AD) = a  b x M
M = 6.74
a = 4.45
b = 0.63
log (AD) =  0.2038
and MD was computed as MD = 0.62 meters.
The amount of the displacement measured for an earthquake in the Bağlarkayası‐ 2010
Trench is 25 cm. This value is smaller than the amount of average displacement computed
employing the empirical equations proposed by Wells and Coppersmith (1994). There might
be several reasons for this: 1) Since Akhisar‐ Kılıç Segment has a minor right lateral compo‐
nent, the amount of the measured normal displacement is smaller than the computed one 2)
or the segment that has a length of 27 km was not entirely broken when the earthquake(s),
during which a displacement of 25 cm occurred, took place.

By means of the evaluation made taking into account the second possibility, the magnitude
(M) of the earthquake that should occur in order that an average displacement of 25 cm
could take place and the length of the surface rupture (SRL) required for an earthquake with
that magnitude to take place were computed as follows making use of the empirical
equations proposed by Wells and Coppersmith (1994) :

log (AD) = a  b x M
AD = 0.25 cm
a = 4.45
b = 0.63
220 Tectonics – Recent Advances

0.60205991 = 4.45  0.63 M


and M was computed as M = 6.10
log (SRL) = a  b x M
M = 6.10
a = 2.01
b = 0.50
log (SRL) = 1.04
and SRL was computed as SRL = 10.96 km.

4. Paleoseismological three dimensional virtual photography method;


Bağlarkayası‐2010 trench application
In recent years there have been significant developments in the applications related to the
cameras that provide panoramic viewing, especially depending on the developments of
image processing programs. Such cameras started to be widely used in various fields such
as security, teleconference, publicity, virtual tour and robot navigation (Baştanlar and
Yardımcı, 2005; Ergün and Şahin, 2009). Since the angle of sight of cameras is always smaller
than that of human beings and the image of large objects cannot be pictured within a single
photograph, a demand for the creation of a panorama arose just at the beginning of the
photography (Parian, 2007). The efforts to obtain a photographic panorama were realized at
the end of 1800s with joining together several photographs taken from different directions to
form a full panorama (Baştanlar, 2005). ‘Panorama’ is a word that was formed by joining
two Greek words together. ‘Pan’ means ‘all’, and ‘horama’ means ‘sight’ (Baştanlar, 2005).
The difference between traditional photograph and panoramic photograph is similar to the
difference between looking at a city from the window of a small office and from the roof of
it (Kwiatek, 2005). One of the main objectives of Photogrammetry is, by carrying the real
image into the virtual environment, to arouse the impression of ‘really being there’ in view‐
ers and readers. In recent years, with the development of Internet and multimedia technolo‐
gies, important developments were recorded in photogammetry and the Three Dimensional
Virtual Photography Technique started to be used in various fields. Although the method is
frequently used in different sectors such as architecture, tourism and museum publicity; it
has no application either in earth sciences or in paleoseismology. Paleoseismological Three
Dimensional Virtual Photography Method has been applied for the first time in this study,
in the paleoseismological studies conducted on the Tuz Gölü Fault Zone (Central Anatolia,
Turkey).

The application area of the Earth Sciences is the Nature. The presentation of the studies,
which are carried out in the Nature, to the readers without any loss of data and with high
resolution is important. Paleoseismological trench works are comprehensive and high cost
works. Trench works should be completed within a specified period of time and with specif‐
ic standards. The specific physical conditions of the trench area (the state of groundwater,
instability of trench sediments, etc.) in some cases complicate the trench works may some‐
times and cause loss of life and property due to collapse and cave in. In order to eliminate
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 221

such drawbacks, trenches should be excavated as deep and wide as possible and the work
should be completed in a manner as fast and qualified as possible.

Paleoseismological Three Dimensional Virtual Photography Method makes it possible the


transfer of all details within the trench to the reader without any loss of data and with high
resolution. In addition, as the reader can directly see the trench wall image, the method
gives the reader the right to interpret. In the section that follows, the application of this
method to a paleoseismological trench study (Bağlarkayası‐2010 Trench) conducted in the
middle part of the Tuz Gölü Fault Zone is explained.

Paleoseismological Three Dimensional Virtual Photography Method is mainly composed of


four stages. These are :

1. Planning
2. Photographing
3. Stitching the photographs and forming scenes
4. Connecting the scenes and forming virtual tour

4.1. Planning
In order a trench work to be photographed in three dimensions, first of all, the number and
locations of the points at which photographs are to be taken should be planned. In planning
the number of points,

- Paleoseismological detail within the trench and


- Dimensions and excavation type of the trench
should be taken into consideration. In the determination of the locations of the point at
which photographs are to be taken, care should be taken not to leave any blind spots. A
virtual tour of the entire trench can be obtained only under these conditions.
Bağlarkayası‐2010 Trench was excavated as 94 meters long, 5 meters wide and 5 meters
deep on the average (Maximum depth: 8.5 meters). While the northern 40‐meter section of
the trench was excavated as a single slot trench, the southern part of it was excavated as
multi‐bench trench (Figure 8). A gridding of 1 m2 was applied on the walls of the trench.
Three deformation zones are present within the trench, which are main fault zone, synthetic
faulting zone and antithetic faulting zone.
Before starting paleoseismological three dimensional virtual photographing, a general pho‐
tograph should be taken which covers the entire trench and shows the points where the
photographs are to be taken. This photograph can be taken, if possible, from above, at an
inclined angle. The general trench photograph is to be used as the index photograph in the
formation of the virtual tour in the final stage. In this study, the general photograph of the
Bağlarkayası‐2010 Trench was taken from the fire tower, at an elevation of 25 meters and at
an inclined angle (Figure 16).

As a result of the evaluation made taking into account the form of the trench and the
detailed paleoseismological features on the trench walls, it was decided to take photographs
222 Tectonics – Recent Advances

at 15 points in the Bağlarkayası‐2010 Trench (Figure 16). The points where photographs
were taken were selected near the parts of the trench walls that contained fault. In order to
trace the continuation of the fault on one wall of the trench, on the other wall, the floor of
the trench should also be cleared and the fault traces should be made distinct at the floor as
well.

Figure 16. The general image of the Bağlarkayası‐2010 Trench taken from the air at an inclined angle.
Yellow circles represent the points where photographs were taken (look from SW to NE)
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 223

In the virtual tour to be obtained as a result of this method, the image should be as clear as
possible. That’s why, one day before photographing, the time and the angle at which
sunlight comes to the trench walls should be controlled and recorded. Thus, it would be
planned at which point, at which hour photographs be taken. Since 8 photographs are to be
taken, an average of 15 minutes can be envisaged for taking photographs at each point.

4.2. Photographing
The hardware and software required for Paleoseismological Three Dimensional Virtual
Photography are presented in Table 2.

The aim of
Software The aim of use Hardware
use
Adope Photoshop
Correction of color and light Tripod Photographing
CS5
Digital
PTGui Pro 9.0 Stitching Photographing
camera
Cubic Converter 2.2.1 Stitching and creating of stages Pano Head Photographing
PTLens 1.5.3 Correction of distortion Fisheye Lens Photographing
Panotour Pro 1.6.0 Cretating of Virtual tour Computer Office work
Internet Explorer, Google Chrome, Mozilla
Web Browser
Firefox
Table 2. The hardware and software required for Paleoseismological Three Dimensional Virtual Pho‐
tography

In this study, Manfrotto 055x probe tripod, Nikon D200 Model and 10.2 megapixel Digital
photo camera, Manfrotto sph 303 panoramic head and Nikkor fish‐eye objective having 1800
angle of sight were used.
At the stage of photographing, 8 photographs are taken at each point. Six of these photo‐
graphs are taken at the horizontal plane using an angle of 600 and thus a horizontal image of
3600 of that point is obtained. Afterwards, the sky and floor image of the same point are
taken as two separate photographs and the image of that point at the vertical plane is ob‐
tained. In this manner, a total of 8 photographs, 6 of which at horizontal plane and 2 at ver‐
tical plane, of each point are taken.

4.3. Stitching photographs together and forming scenes


Paleoseismological Three Dimensional Virtual Photography Method is a method which is
most of the time applied employing some computer programs. However, from time to time
it requires some manual intervention. In order that the virtual tour to be produced from the
photographs can be as realistic as possible, some corrections are needed to be made on the
raw photographs.

At this stage, first of all, color and light corrections of 8 photographs taken at each point are
made using Adobe Photoshop CS5 program. Following this, employing PTGui Pro 9.0
224 Tectonics – Recent Advances

program, photographs are connected by stitching. At the first stage of the connection, 6
photographs taken at horizontal plane and the sky image taken at the vertical plane are
used. During this process, from time to time manual corrections may be required when
common parts of the photographs are overlapped. At the end of the first stage of the
connection, all photographs, except the floor image, become stitched together.

These 7 stitched photographs are then transferred to Cubic Converter 2.2.1 program. Cubic
Converter 2.2.1 program creates partial photographs by dividing the image composed of 7
stitched photographs into 6 parts in the form of cubes. One of these partial photographs is
the floor photograph yet to be completed.

To complete the floor photograph, of the raw photographs taken at first, the one belonging
to the floor is opened in the program PT lens 1.5.3. The image, which is obtained after its
distortion correction is made in this program, is assembled with the partial photograph
belonging to the floor. This assembled photograph is again transferred to the program Cubic
Converter 2.2.1 and thus the floor photograph becomes completed. Lastly, the image of the
whole scene is exported in TIFF format from the Cubic Converter 2.2.1 program. In Figure
17, the stitched and connected scene image belonging to the point 8 (see SE second bench in
Figure 16) of Bağlarkayası‐2010 Trench is presented.

By repeating these processes for all of the scenes, a total of 15 files were obtained.

Figure 17. The scene image obtained as a result of stitching and connecting the photographs belonging
to the point 8 (see SE second bench in Figure 16) of Bağlarkayası‐2010 Trench.

4.4. Connecting the scenes and forming the virtual tour


All of the scenes that are connected and recorded are opened in Panotour 1.6.0 program.
Transition links are formed for the transition of the scenes with each other to be made.
Afterwards, index photograph is imported into the program and transition links are added
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 225

to the index photograph as well. In this manner, the access to a desired scene is made
possible both over the virtual tour and over the index photograph.

The completed virtual tour is recorded as MTML file. This file not only makes it possible the
visual presentation of a trench study to the reader but also makes significant contributions
especially at the stage of interpretation of the trench data

You can reach the virtual tour produced employing Paleoseismological Three Dimensional
Virtual Photography Method developed in this study over the following link:

https://hotfile.com/dl/155062161/c15656d/webquality.rar.html

5. Results
In this study, a new photography method for paleoseismological studies was developed and
named ‘Paleoseismological Three Dimensional Virtual Photography Method’.
It was observed that this method could be successfully applied especially in deep and
benched paleoseismological trench excavations.
The most important advantage of this method is that all the geologic detail within the trench
can be transferred to the reader without any loss of data and with high resolution. In this
manner, the reader is given the right to interpret.
Paleoseismological Three Dimensional Virtual Photography Method is mainly composed of
four stages which are planning, photographing, forming scenes by stitching the photo‐
graphs together and forming the virtual tour by connecting the scenes.
The virtual tour obtained as a result of this method enables the reader to make a tour, feel‐
ing as if he were really there.
The method has been applied for the first time in this study on the Akhisar‐Kılıç Segment of
the Tuz Gölü Fault zone which is one of the most important active fault zones of the Central
Anatolia Region.
Tuz Gölü Fault Zone (TGFZ) is a NW‐SE trending, dipping towards SW, active, right lateral
strike slip component normal fault zone which extends between north of Tuz Gölü and
Kemerhisar (Niğde) and has a length of 200 km.
TGFZ is composed of fault segments that have lengths varying between 4 and 33 km. The
Akhisar‐Kılıç Segment is one of the most important segments of this zone owing to its geo‐
logical fault length and morphotectonic features.
The total amount of vertical displacement of the Akhisar‐Kılıç Segment is 268 meters accord‐
ing to Lower Pliocene aged Kızılkaya Ignimbrites.

According to the calculation made taking into account the total amount of vertical
displacement (268 meters) and the age of Kızılkaya Ignimbrites (5 million years), the slip
rate of the Akhisar‐Kılıç Segment during the period from Lower Pliocene to the present day
is 0.0536 mm/year.
226 Tectonics – Recent Advances

In this study, a paleoseismological trench study has been conducted on the Akhisar‐Kılıç
Segment of TGFZ for the first time and the trench has been named ‘Bağlarkayası‐2010 Trench’.

According to the paleoseismological findings obtained from Bağlarkayası‐2010 Trench, two


earthquakes that resulted in surface faulting occurred on the Akhisar‐Kılıç Segment during
the last 10 500 years. The first of these earthquakes occurred approximately in the year 10
000 B.P. and the second one took place in the year 5336 B.P. Consequently, the elapse time
from the last earthquake up to the present is approximately 5336 years.

The amount of the average vertical displacement accumulated on the Akhisar‐Kılıç Segment
from the last earthquake to the present time was calculated making use of empirical
equations (Wells and Coppersmith, 1994) as 28.76 cm.

An average displacement of 25 cm was measured in the earthquakes that were defined and
dated within Bağlarkayası‐2010 Trench. According to the evaluation made taking into con‐
sideration this value together with the slip rate (0.0536 mm/year), the earthquake recurrence
interval for the Akhisar‐Kılıç Segment was calculated as 4664 years.

The fact that the average earthquake recurrence interval (4664 years) is smaller than the
elapse time from the last earthquake to the present day indicates that the Akhisar‐Kılıç Seg‐
ment has completed its average earthquake recurrence interval and could produce earth‐
quake(s) at any moment.

The geologic fault length of the Akhisar‐Kılıç Segment is 27 km. The moment magnitude of
the maximum earthquake that this segment could produce in case the whole of it were to be
broken was calculated as M = 6.74 making use of the empirical equations proposed by Wells
and Coppersmith (1994).

According to the computations made using the other equations of the same study, the
amount of the maximum displacement which could occur in an earthquake with the magni‐
tude M = 6.74 is MD = 1.25 meters and the amount of the average displacement is AD = 0.62
meters.
A displacement of 25 cm was measured in both of the earthquakes defined by means of the
paleoseismological excavation studies conducted in the trench. This value is smaller than
the amount of the average displacement that a pure normal fault with a length of 27 km can
produce. The reason for this difference may be the fact that the fault has a minor strike slip
component. Or the whole of the segment that is 27 km long is not broken. In this case, in
order for a displacement of 25 cm to occur, the moment magnitude of the fault is expected to
be M = 6.10 and the required surface rupture length for an earthquake with the magnitude
M =6.10 to occur is expected to be SRL = 10.96 km.

Author details
Akın Kürçer
General Directorate of Mineral Research and Exploration, Department of Geology, Çankaya, Ankara,
Turkey
Paleoseismological Three Dimensional Virtual Photography Method;
A Case Study: Bağlarkayası-2010 Trench, Tuz Gölü Fault Zone, Central Anatolia, Turkey 227

Yaşar Ergun Gökten


Ankara University, Faculty of Engineering, Department of Geological Engineering,
Tectonic Research Group, Ankara, Turkey

Acknowledgement
This study has been carried out within the scope of the project (2010‐30‐14‐02‐3) titled ‘The
Neotectonic Properties and Paleoseismology of the Tuz Gölü Fault Zone, Central Anatolia,
Turkey’conducted by the Department of Geological Researches of the General Directorate of
Mineral Research and Exploration (MTA). We would like to express our thanks to the
General Directorate of Mineral Research and Exploration (MTA). In addition, our thanks
extend to Dr. Ömer Emre (MTA), Dr. Tamer Yiğit Duman (MTA) and Dr. Selim Özalp
(MTA) for their useful critiques of paleoseismological interpretations, Geophysical Engineer
Hayrettin Karzaoğlu (MTA) for his contributions to the geophysical studies and Dr. Gerçek
Saraç for his contributions to the determination of the human bones.

6. References
Aydar, E, Schmitt, A.K.., Çubukçu, H.E., Akin, L., Ersoy, O., Sen, E., Duncan, R.A., ve Atici,
G., 2012. Correlation of ignimbrites in the central Anatolian volcanic province using
zircon and plagioclase ages and zircon compositions. Journal of Volcanology and
Geothermal Research 213‐214 (2012) 83–97.
Baştanlar, Y., 2005. Parameter extraction and image enhancement for catadioptric
omnidirectional cameras. Msc. Thesis, METU Informatic Institute, Ankara
Baştanlar, Y. and Yardımcı, Y., 2005. Hiperbolik Aynalı Katadioptrik Tüm Yönlü Kameralar
için Parametre Çıkarımı, IEEE Xplore,
http://ieeexplore.ieee.org/stamp/stamp.jsp?arnum ber=01567669 (2009.04.20).
Besang, C., Eckhardth, F.J., Harre, W., Kreuzer, H. and Müller, P., 1977. Radiometrische
Altersbestimmungen an neogenen
Bronk Ramsey, C., 2001. Development of the radiocarbon program OxCal: Radiocarbon, v.
43, no. 2A, p. 355–363.
Çemen, İ., Göncüoğlu, M.C. and Dirik, K., 1999. Structural evolution of the Tuz Gölü basin
in Central Anatolia, Turkey. Journal of Geology, 107, 693‐706.
Demirtaş, R., 1997. Paleoseismology, General Directorate of Disaster of Turkey, Ankara,
Turkey.
Dirik, K. and Göncüoğlu, M.C., 1996. Neotectonic characteristics of Central Anatolia, Int.
Geology Review, 38, 807‐817.
Dirik, K. and Erol, O., 2000. Tuz Gölü ve civarının tektonomorfolojik evrimi Orta Anadolu,
Türkiye, Haymana‐Tuz Gölü‐Ulukışla Basenleri Uygulamalı Çalışma (Workshop),
T.P.J.D. Bülteni, Özel sayı 5
Dönmez, M., Akçay, A.E., Kara, H., Türkecan, A., Yergök, A.F. and Esentürk, K., 2005.
Geological maps of Turkey in scale of 1/100 000 Aksaray L 32 sheet. MTA publications,
52, Ankara.
228 Tectonics – Recent Advances

Ergün, B. and Şahin, C. 2009. Digital Spherical Photogrammetry Techniques Recently in Use.
Harita Dergisi, 142, p. 40‐50.
Hatheway, A. W., and Leighton, F. B., 197). Exploratory trenching. Geol. Soc. Am., Rev. Eng.
Geol. 4, 169–195.
Innocenti, F., Mazzuoli, G., Pasquare, F., Radicati Di Brozola, F. and Villari, L., 1975. The
Neogen calc‐alcalin volcanism of Central Anatolia: geochronological data on Kayseri‐
Niğde area: Geol. Mag., 112 (4), 349‐360.
Koçyiğit, 2000. General neotectonic characteristics and seismicity of Central Anatolia,
Haymana‐Tuz Gölü‐Ulukışla basenlerinin uygulamalı çalışması (workshop). Abstracts,
1‐26, Aksaray.
Koçyiğit, A. and Özacar, A., 2003. Extensional neotectonic regime through the NE edge of
the Outer Isparta Angle, SW Turkey: New Field an Seismic Data. Turkish Journal of
Earth Sciences, 12, 67‐90.
Kwiatek, K., 2005. Generation of a virtual tour in the 3D space applying panoramas,
exercised on the sites of Dresden and Cracow. Lisans Tezi, AGH University ofScience
and Technology, Dresden.
McCalpin, J., P., 2009. Field Tecniques in Paleoseismology – Terrestrial Environments,
Paleoseismology, edited by James P. McCalpin, second edition, International
Geophysics Series v. 95, p. 29‐118.
McCalpin, J., P. and Nelson, A., R., 2009. Introduction to Paleoseismology. Paleoseismology,
edited by James P. McCalpin, second edition, International Geophysics Series v. 95, p.
1‐25.
Okay, A.I., Tüysüz, O., Satır, M., Özkan‐Altıner, S., Altıner, D., Sherlock, S. and Eren, R.H.,
2006. Cretaceous and Triassic subduction‐accretion, HP/LT metamorphism and
continental growth in the Central Pontides, Turkey. Geological Society of America
Bulletin, 118: 1247‐1269.
Pantosti, D. and Yeats, R.S., 1993. Paleoseismology of great earthquakes of the late Holocene,
Analidi Geofisica, v. XXXVI, n:3‐4, p. 237‐257.
Parian, J. A., 2007. Sensor Modeling, Calibration and Point positining with Terrestrial
panoramic Cameras. Doctora Thesis, E.T.H. Zurich, Swisszerland.
Reilinger, R.E., Mcclusky, S.C., Vernant, P., Lawrence, S., Ergintav, S., C¸ akmak, R.,
Nadariya, M., Hahubia, G., Mahmoud, S., Sakr, K., Arrajehi, A., Paradissis, D., Al‐
Aydrus, A., Prilepin, M., Guseva, T., Evren, E., Dmitritsa, A., Filikov, S.V., Gomes, F.,
Al‐Ghazzi, R., and Karam, G., 2006. GPS constraints on continental deformation in the
Africa‐Arabia‐Eurasia continental collision zone and implications for the dynamics of
plate interactions: Journal of Geophysical Research, v. 111, p. V05411, doi: 10.1029/
2005JB004051.
Schumacher, R., Mues‐Schumacher, U., 1996. The Kizilkaya ignimbrite—an unusual low‐
aspect‐ratio ignimbrite from Cappadocia, Central Turkey. Journal of Volcanology and
Geothermal Research 70, 107–121.
Şaroğlu, F., Emre, Ö. and Kuşçu, İ., 1992. Active fault map of Turkey, MTA, Ankara.
Wells, D.L. and Coppersmith, K.J., 1994. New Empirical Relationships among Magnitude,
Rupture Length, Rupture Width, Rupture Area, and Surface Displacement. Bulletin of
the Seismological Society of America, Vol. 84, No. 4, pp. 974‐1002.
Chapter 9

Structural Geological Analysis


of the High Atlas (Morocco):
Evidences of a Transpressional Fold-Thrust Belt

Alessandro Ellero, Giuseppe Ottria, Marco G. Malusà and Hassan Ouanaimi

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/50071

1. Introduction
The High Atlas of Morocco, representing the southernmost element of the Perimediterrane‐
an Alpine belts, is a typical example of intracontinental belt (Mattauer et al., 1977). It was
formed within the North African plate during convergence of the African and European
plates during the Cenozoic (Dewey et al., 1989; Gomez et al., 2000). Like other intracontinen‐
tal mountain belts, the High Atlas shows a double sense of vergence and a complex evolu‐
tion of timing and sequence of thrusting.

Many studies have emphasized the role of the inversion tectonics in the evolution of the
High Atlas system (Proust, 1973; Jacobshagen et al., 1988; Giese and Jacobshagen, 1992;
Laville and Piquè, 1992; Beauchamp et al., 1996, 1999; Mustaphi et al., 1997; Hafid, 2000).
Thrust and fold structures would resulted from the reactivation (inversion), caused by the
Cenozoic compressional events, of the preexisting extensional faults associated with the
Triassic‐Liassic Atlasic rifting.

In this context several authors considered strike‐slip faulting as an important component of


the Alpine evolution of the High Atlas belt (Mattauer et al., 1977; Fraissinet et al., 1988;
Froitzheim et al., 1988; Laville and Piquè, 1991, 1992; Morel et al., 2000; Piquè et al., 2002).
However, more recent studies (Frizon de Lamotte et al., 2000; Teixell et al., 2003, 2005; Arbo‐
leya et al., 2004; Ayarza et al., 2005; Missenard et al., 2006) have been aimed to the definition
of the geodynamic model for the Alpine evolution of the High Atlas belt. As a consequence
structural studies have been performed within a regional scale geodynamic pure compres‐
sional framework, neglecting the kinematic meaning of data. Therefore, despite the fact that
the geometries of deformation structures are locally well known, the details of Alpine struc‐

© 2012 Ottria et al., licensee InTech. This is an open access chapter distributed under the terms of the
Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
230 Tectonics – Recent Advances

tural evolution of the High Atlas belt, particularly kinematic information, are still poorly
known.

In this contribution we present the results of a structural and kinematic study we have
carried out in Morocco along a transect crossing the Western High Atlas and along the
southern margin of the Central High Atlas in the Tinerhir‐Boumalne area, two key sectors of
the High Atlas belt. A main objective of this paper is to establish the possible relationships
between the described structures to better understand the general processes of
intracontinental mountain building that have constructed the High Atlas belt.

Our structural and kinematic study, suggesting main transpressional imbricate systems,
indicates that strike‐slip movements played an important role in the deformation evolution
of the High Atlas fold and thrust belt. We propose that the main Alpine deformation of
High Atlas belt was transpressional in character, with efficient kinematic strain partitioning
focussing the strike‐slip component along pre‐existing major faults. The potential influence
of pre‐existing tectonic boundaries on weakness and evolving tectonic fabrics on kinematic
strain partitioning can be therefore emphasized.

2. Regional geology
The High Atlas represents the highest mountain belt of Morocco, with peaks of over 4000 m
a.s.l. (Mt. Toubkal, 4165 m), crossing the country along the SW‐NE direction from the
Atlantic Ocean to Algeria for a length of about 2000 km (800 km in Morocco) and a width
ranging from about 50 km to 100 km, framed between the Meseta domains (Morocco Meseta
and Oran Meseta) to the north, and the northern boundary of the West African Craton (Anti
Atlas belt) to the south (Figure 1).

The High Atlas fold and thrust belt is formed by a Precambrian and Paleozoic basement and
a Mesozoic‐Cenozoic succession. The pre‐Mesozoic basement is exposed in several inliers of
the High Atlas, forming the most elevated areas of the Western High Atlas. The Mesozoic
successions, mostly Jurassic in age, crop out almost exclusively in the Central High Atlas
and the Atlantic basin south of Essaouira (Figure 1). The Cenozoic deposits, substantially
absent along the belt axis, characterize the High Atlas boundaries with the plains where the
Neogene formations deposited.

In the Western High Atlas, the Precambrian basement crops out mostly in the Ouzellarh
Block, being composed by metamorphic rocks and granitoids topped by late Precambrian
volcanics. The Paleozoic succession ranges from Lower Cambrian to Carboniferous and it is
mostly characterized by clastic rocks deformed during the Variscan orogeny. In particular,
the Late Visean‐Early Westphalian tectonic event (Main Variscan Phase; Michard et al. 2008
and references) produced tight folds associated with metamorphism and granite intrusions.
These folds show NE‐SW to N‐S axes and sub‐vertical to generally E‐dipping axial planes
(western vergence) and developed a pervasive axial plane cleavage. The P‐T conditions of
metamorphism did not exceeded low‐grade greenschist‐facies conditions during the main
Variscan folding, except in the regions close to the granite intrusions (Michard et al., 2008).
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 231

However the Variscan structural pattern is dominated by ENE‐WSW to NE‐SW major dex‐
tral fault zones (e.g. Tizi nʹTest, Medinat, Erdouz) which broke up the Western High Atlas
into several structural blocks (Proust et al., 1977; Ouanaimi and Petit, 1992; Houari and
Hoepffner, 2003; Hoepffner et al., 2005). A recent paper (Dias et al., 2011) emphasises the
occurrence of WNW‐ESE sinistral shear zones, already documented by Fraissinet et al.
(1988), in the framework of conjugated fault systems with the ENE‐WSW dextral shear
zones developed during the main Variscan phase.

Figure 1. Schematic structural map of Morocco with location of Figures 2 and 6 (modified after Hafid et
al., 2006 and Michard et al., 2008). SAF: South Atlas Fault; NAF: North Atlas Front; TnTFs: Tizi n’Test
Fault system; JeT: Jebilet Thrust; A: Agadir; C: Casablanca; F: Fes; G: Gibraltar; M: Marrakech; O: Oujda;
R: Rabat; T: Tinerhir.
232 Tectonics – Recent Advances

The Mesozoic succession of the High Atlas belt started with the Late Permian‐Triassic red
beds (conglomerates, sandstones, siltstones and mudstones), unconformably resting on the
Lower Paleozoic rocks or on the Precambrian basement. These continental deposits (Fabuel‐
Perez et al., 2009) represent the detrital infilling of basins developed during the Late
Permian‐Triassic Atlasic pre‐rifting phase when the Variscan shear zones were reactivated
as normal and strike‐slip faults. The pre‐rift deposits are capped by tholeiitic basalt flows of
the CAMP (Central Atlantic Magmatic Province) that provide absolute ages of about 200 Ma
corresponding to the Triassic‐Jurassic transition (Fiechtner et al., 1992; Knight et al, 2004;
Marzoli et al, 2004).

The overlying limestones and dolomites represent the transgressive Lower Liassic platform.
Within the Liassic the transition from massive carbonates to a layered sequence of marls and
limestones indicates a platform‐basin boundary documenting the progressive disruption
and drowning of the Liassic platform (Jossen and Filali‐Moutei, 1992). The Upper Liassic‐
Lower Dogger (from Toarcian to Bajocian) are varicolored marls and reefal limestones
underlying Bathonian red sandstones and silty shales pointing to a continental
sedimentation (e.g. Ellouz et al., 2003). The Cretaceous is characterized by red sandstones
and conglomerates (ʺInfracenomanianʺ; Gauthier, 1957) evolving to platform white
limestones of Cenomanian‐Turonian age which mark a global transgression to the scale of
the entire Atlas domain (Ettachfini and Andreu, 2004). During the Upper Cretaceous‐
Paleogene the sedimentation is mainly continental and lacustrine with minor marine
bioclastic limestones of Eocene age (Marzoqi and Pascal, 2000).

The Neogene continental deposits, occurring above a regional unconformity, resulted


essentially from the syndeformation erosion of the mountain belt (Miocene‐Pliocene
molasses).

The post‐Jurassic deposits which probably formed a rather continuous cover overlying the
High Atlas and the surrounding areas (Meseta and Anti Atlas domains) are now exposed
mainly in the areas bordering northward and southward the High Atlas (Haouz, Souss and
Ouarzazate basins), while they have been largely eroded into the High Atlas being pre‐
served only in few limited outcrops.

The Alpine (Atlasic) tectonic evolution of the High Atlas fold and thrust belt has been gen‐
erally considered to be characterized by at least two main deformation steps spanning in age
from Late Eocene to Oligocene‐Miocene and from Pliocene to Early Quaternary, respectively
(Gorler, 1988; Jacobshagen et al., 1988; Giese and Jacobshagen, 1992; Frizon de Lamotte et al.,
2000; El Harfi et al., 2001, 2006; Missenard et al., 2007). Teson and Teixell (2006) documented
a rather continous thrusting at the southern border of the High Atlas (Boumalne area) active
from the Oligocene to the Pliocene. Nevertheless Quaternary tectonics, deforming alluvial
terraces, has been also documented (e.g. Morel et al., 2000; Sebrier et al., 2006; Cerrina Feroni
et al., 2007; Delcaillau et al., 2010).

The present‐day seismicity along the marginal zones of the High Atlas testifies that the
orogenic movements are still active.
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 233

The tectonic style characterizing the High Atlas is mainly thick‐skinned, as the basement
was involved in the compressional deformation (e.g. Frizon de Lamotte et al., 2000; Teixell
et al., 2003; El Harfi et al., 2006). However, the structures of the southern border have been
interpreted as evolved within a thin‐skinned style of deformation (e.g. Beauchamp et al.,
1999; Bennami et al., 2001; Teixell et al., 2003).

The limits that bound, to the north and to the south, the High Atlas orogenic system are
represented by two tectonic lineaments of regional importance.

In particular, the northern boundary of the Western High Atlas is represented by a complex
system of thrusts and high angle faults that divides the main range from the Haouz Plain.

The northern border of the Central High Atlas belt is marked by a N‐verging thrust,
associated with strike‐slip faults, that finds its westernward extension with the thrust that
subtends and bounds northward the Jebilet range (Figure 1). On the contrary, the southern
boundary of the Central High Atlas matches the South Atlas Fault Zone (Russo and Russo,
1934), a major tectonic feature extending from Morocco, where the High Atlas belt is
juxtaposed to the Anti Atlas, to Tunisia (Bracene et al., 1998). In the Western High Atlas the
South Atlas Fault Zone corresponds to the western termination of the Tizi n’Test Fault
system.

3. The studied sectors


We have focused our study in two selected sectors corresponding to the southern boundary
of the Central High Atlas in the Tinerhir‐Boumalne area and to the Western High Atlas
transect between Imi n’Tanoute and Taroudant region (Figure 1).

In these sectors, the occurrence of Mesozoic‐Cenozoic deposits allowed to describe the


geometry and kinematics of the main Alpine structures and define the ages of the tectonic
events that characterize the polyphase deformation of the High Atlas borders. The
availability of industry seismic data, discussed in earlier papers (e.g. Mustaphi et al., 1997;
Hafid et al., 2006; Sebrier et al., 2006), provided further useful information.

In particular, the structural study of the Tinerhir‐Boumalne area is important in the under‐
standing of the kinematics induced by the South Atlas Fault Zone. The Western High Atlas
is a privileged sector where it is possible the study of the entire mountain belt along a rela‐
tively short transect (about 50 km) between its northern and southern boundaries, that is to
say from the Imi nʹTanoute Fault to the Tizi nʹTest Fault Zone.

Actually, whereas the Imi nʹTanoute Fault is located at the northern boundary of the
Western High Atlas at the margin of the Haouz Plain, from a structural point of view it does
not represent the northern belt front as the Jebilet range, north of Marrakech (Figure 1), must
be considered as part of the High Atlas itself (Hafid et al., 2000, 2006; Michard et al., 2008).
Nevertheless, along the Western High Atlas transect it is possible to acquire new data about
the characterization of fault and thrust‐fold systems evolving from northern to southern
vergences crossing the belt from north to south.
234 Tectonics – Recent Advances

3.1. Central High Atlas (Tinerhir‐Boumalne area)


3.1.1. Geological setting
The study area corresponds to the NE‐SW trending zone, from Boumalne and Tinerhir,
between the Central High Atlas to the north, and the Eastern Anti Atlas (Saharan Craton) to
the south (Figures 1 and 2). The geological setting is characterized by the eastern
termination of the Ouarzazate basin, a topographic low where most of the Miocene‐Pliocene
and Quaternary deposits sedimented. It represents the physiographic separation between
Central High Atlas and Eastern Anti Atlas belts. In the Tinerhir area, where the Ouarzazate
basin ends, the two belts face each other, being separated only by the Cretaceous‐Eocene
succession.

The structural pattern of the southern boundary of the Central High Atlas between
Boumalne and Tinerhir is dominated by the high angle faults belonging to the South Atlas
Fault system showing an overall direction of about N70E. In particular the South Atlas Fault
main zone corresponds to a strike‐slip fault that juxtaposes a northern block represented by
the Central High Atlas belt s.s. and a southern block characterized by the Mesozoic‐
Cenozoic succession unconformably overlapping the Paleozoic basement outcropping north
the Proterozoic rocks of the Jebel Saghro. The Mesozoic‐Cenozoic succession of the southern
block is deformed by a south‐verging fold‐thrust system that affects also the underlying
Paleozoic basement. The latter is formed by terrigenous clastics with subordinate carbonate
sediments ranging in age from Lower Cambrian to Carboniferous (Visean‐Namurian). The
Paleozoic succession is characterized by a polyphase deformation consistent with the Late
Carboniferous Variscan evolution (Michard et al., 1982; Hoepffner et al., 2006; Cerrina
Feroni et al., 2010). Nevertheless, a post‐Variscan tectonics, connected with the Cenozoic
Atlasic orogeny, has been documented in the Eastern Anti Atlas (Malusà et al., 2007).

In the Tinerhir area, the Mesozoic‐Cenozoic succession which unconformably overlies the
Paleozoic basement is characterized by a basal unit formed by Upper Cretaceous continental
conglomerates and sandstones (ʺInfracenomanianʺ) which directly overlaps the Paleozoic
and Proterozoic basement (Figure 2). The pebbles of the conglomerates are formed by
Precambrian and Paleozoic rocks of the Anti Atlas domain; the basal unconformity sealed
the structures generated by the Variscan deformation.

Above the Infracenomanian deposits the succession evolves upward to Upper Cretaceous‐
Eocene mainly marine deposits which can be subdivided in many lithostratigraphic units
(Figure 2B). The overlying deposits consist of red pelites and fine sandstones with gypsum
lenses and intercalations of sandstones and conglomerates (e6‐e7); this lagoonal‐continental
succession corresponds to the Hadida‐Ait Ouglif formation which has been referred to the
Upper Eocene‐Early Oligocene (El Harfi et al., 2001).

The uppermost part of the Mesozoic‐Cenozoic succession is characterized by a very thick


(up to more than 700 m) Miocene‐Pliocene formation constituted by polygenic
conglomerates with pebbles of Precambrian to Paleogene rocks. It is deposited in an alluvial
fan environment evolving to shales and lacustrine limestones and to alluvial fan
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 235

conglomerates mainly composed by Jurassic limestone clasts. The Miocene‐Pliocene


deposits lies unconformably over all the previous formations comprising the

Figure 2. A: Geological map of the Tinerhir‐Boumalne area (simplified after Hindermayer et al., 1977,
and Milhi, 1997.) indicating the geological cross‐sections of Figure 3. SAF: South Atlas Fault. B:
Lithostratigraphy of geological cross‐sections in Figure 3.

Jebel Saghro Precambrian rocks (outside the area represented in the geological map of Figure
2A). The Miocene‐Pliocene basal contact represents a regional unconformity as the Miocene‐
Pliocene deposits directly overlap also the Middle Jurassic rocks belonging to the stratigraphic
succession of the Central High Atlas, north of the South Atlas Fault zone (Figure 2B).
236 Tectonics – Recent Advances

Figure 3. Geological cross‐sections through the South Atlas Fault zone between Tinerhir and Boumalne
(location in Figure 2). In sections 1‐1’ and 2‐2’ stereonet diagrams (Equal area projection, lower hemi‐
sphere) represent poles to bedding (yellow circles: normal beds; blue circles: overturned beds), and
measured fold axes (yellow triangles); the β axes are indicated (red squares). In sections 3‐3’ and 4‐4’
stereonet diagrams represent fault planes and striae.

The Central High Atlas succession exposed in the study area consists of Triassic red silt‐
stones, sandstones and basalts evolving upward to Liassic‐Dogger mainly massive lime‐
stones and dolostones and varicolored marls, locally with gypsum (Figure 2B).

3.1.2. Structure geometry and kinematics


The geological‐structural data collected in the field allowed to produce four cross‐sections
that describe the geometry and kinematic of the main structures occurring at the southern
margin of the Central High Atlas between Boumalne and Tinerhir villages (Figure 3).

The outstanding structure is the South Atlas Fault zone along which the mainly Lower Ju‐
rassic rocks of the Central High Atlas succession are juxtaposed to the Cretaceous‐Eocene
succession of the Anti Atlas block. The main fault zone is characterized by NE‐SW trending
high angle fault planes displaying horizontal to oblique slickensides that indicate an overall
dextral displacement (Figure 4A). The kinematic analysis allowed to obtain two palaeostress
tensors by inversion of the fault data collected along the South Atlas Fault zone northwest of
Tinehrir (Figure 5). The Imarirene site of measurement evidences an homogeneous popula‐
tion of NE‐SW dextral strike‐slip faults compatible with a strike‐slip palaeostress tensor
with a sub‐horizontal maximum compression σ1 axis directed roughly E‐W. The dextral
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 237

strike‐slip faults collected at Ait Snane are conjugated with NNW‐SSE sinistral strike slip
faults. This fault system is consistent with a strike slip palaeostress regime (sub‐vertical
intermediate compression axis σ2) and a sub‐horizontal σ1 axis showing a NNW‐SSE direc‐
tion (Figure 5). This strike slip tensor is kinematically compatible with the ENE‐WSW trend‐
ing inverse faults occurring in the same site of measurement.

Figure 4. Geometrical and kinematic features along the South Atlas Fault zone in the Tinerhir‐
Boumalne area. A: Main South Atlas Fault plane corresponding to the Lower Jurassic rock wall.
The direction of the red arrow corresponds to the movement of the missing block, indicating a dextral
sense of movement. B: Thrust surface affecting the Quaternary deposits. The thrust plane is character‐
ized by the same direction of the South Atlas Fault zone with a top to the South sense of movement. C:
Hectometre‐scale south‐vergent fold developed in Lower Jurassic rocks
of the High Atlas domain. D: Hectometre‐scale south‐vergent fold developed
in the Upper Cretaceous‐Eocene succession of the Anti Atlas domain.

Thrust folds are structures characterizing also the deformation pattern of the study area,
affecting the Jurassic rocks and the Upper Cretaceous‐Eocene succession, as well as the
Miocene‐Pliocene deposits and the Quaternary terrace gravels (Figures 3 and 4B). The
thrusts are directed about ENE‐WSW (~ N70E), broadly parallel to the South Atlas Fault
zone, dipping toward NNW of 30°‐40°. The fault surfaces display down‐dip slickenside
striations indicating a top to the S sense of movement. The resulting palaeostress tensors are
238 Tectonics – Recent Advances

consistent with pure compressive tectonic regimes with sub‐horizontal σ1 axes directed
NNW‐SSE as shown in the Ait Arbi and Sidi Ali Ou Bourk stations (Figure 5).

Figure 5. Geological sketch map with stress invertion results (TENSOR program; Delvaux, 1993). Stere‐
ograms (Schmidt net, lower hemisphere) with traces of fault planes, observed slip lines and slip senses;
the principal stress axes (S1, S2, S3) and type of stress tensor are reported.

Northwestward this thrust system links to the high angle faults (South Atlas Fault system)
displaying an overall asymmetrical positive flower structure geometry (Figure 3).

The thrust system is associated with a fold system characterized by anticlines and synclines
showing steeply dipping axial planes; the fold axes, generally showing sub‐horizontal
plunging, trend from N70E to about E‐W, again sub‐parallel to the South Atlas Fault trend
(Figure 4C and 4D). The fold asymmetry indicates a southward vergence. The sub‐vertical
limbs of the folds are often affected by thrust faults that cut off the anticline‐syncline hinge
zones (Figure 3).

The analysis of the geological cross‐sections indicates that the deformation is not
homogeneously distributed in the study area: deformation zones constituted by folds and
thrusts are more developed close to the South Atlas Fault zone, while moving to the SE the
deformation decreases generating more spaced open syncline and anticline folds.

The geometric‐kinematic analysis suggests that thrusting and folding can be linked to the
development of contemporaneous strike‐slip faulting in a complex polyphase tectonic
evolution. In fact, the relationships of the Miocene‐Pliocene deposits that unconformably
sealed the fold structures of the Upper Cretaceous‐Eocene succession, and the deformation
affecting the Quaternary deposits indicate two distinct tectonic phases characterizing the
Alpine evolution of the southern boundary of the Central High Atlas belt.
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 239

3.2. Western High Atlas


The studied sector of the Western High Atlas develops between Imi nʹTanoute village and
Menizla village (SE of Argana), at the northern limit of the Souss Plain (Figure 6).

Figure 6. Schematic structural map of the Western High Atlas with location of Figures 7 and 9 (modi‐
fied after Hollard et al, 1985). 1: Quaternary; 2: Neogene; 3: Eocene; 4: Cretaceous; 5: Jurassic; 6: Permi‐
an‐Triassic, 7: Carboniferous; 8: Ordovician, Silurian and Devonian; 9: Cambrian; 10: Variscan Granites.
InTF: Imi n’Tanoute Fault; SekF: Seksaoua Fault; MedF: Medinat Fault; TzMF: Tizi Maachou Fault; IkaF:
Ikakern Fault; TnTFs: Tizi n’Test Fault system; WAFZ: Western Atlasic Fault Zone.

The central part of the belt is formed by a 10 km thick Paleozoic succession, mostly repre‐
sented by the Cambrian metasediments (sandstones, schists and greywackes). East of the
Western Atlasic Fault Zone (Cornée and Destombes, 1991), a major N‐S trending Cambrian
tectonic lineament (Figure 6), the Lower Cambrian succession is characterized by volcano‐
detritic schists with conglomeratic lenses overlying a complex of schists with arkoses, volca‐
no‐detritic and calcareous intercalations. The Paleozoic succession evolves up to Ordovician
sandstones and to Silurian‐Devonian units respectively composed of black/reddish schists
and sandstones and conglomerates with lens of limestones.

The Carboniferous formations are restricted to the Ida Ou Zal basin in the southwestern
sector of the Western High Atlas (Figure 6). It is composed by 1800 m thick succession of
conglomerates followed by sandstones, pelites, coal seams and argillaceous sandstones
alternating with dolomitic calcareous layers (De Koning, 1957). The succession was accumu‐
240 Tectonics – Recent Advances

lated during Stephanian‐Autunian time span in this basin which has been interpreted as a
Late Variscan basin created along a strike‐slip fault system in a transextensional regime
(Saber et al., 2001). The Stephanian‐Autunian deposits of the Ida Ou Zal basin were de‐
formed by a folding phase with E‐W to ESE‐WNW axial direction developed in a transpres‐
sional tectonic regime (Saber et al., 2001). A more complex tectonic evolution has been pro‐
posed by Qarbous et al. (2003) consisting in superimposed folding and faulting defor‐
mations in alternating compressional and extensional regimes. The last deformation stage,
connected to the Alpine tectonics, produced the reactivation of the ENE‐WSW faults as
reverse faults in the context of a roughly N‐S compression.

The western boundary of the Western High Atlas is characterized by the Upper Permian‐
Triassic deposits (sandstones and siltstones) whose outcrops are limited to a NNE‐SSW
trending basin (Argana Corridor). This basin evolves westward to the Agadir‐Essaouira
basin where the Mesozoic‐Cenozoic succession developed. The Argana Corridor deposits
are affected by a network of ENE‐WSW, NE‐SW and WNW‐ESE faults (Tixeront, 1974) that
extend eastward cutting the Paleozoic basement.

The Mesozoic‐Cenozoic succession characterizes also the northern and southern marginal
sectors of the Western High Atlas, respectively north the Imi nʹTanoute Fault at the margin
with the Haouz Plain and south the Tizi nʹTest Fault system at the margin with the Souss
Plain. The Souss Plain, which acted as the High Atlas foreland basin during the Cenozoic,
constitutes an E‐W oriented depression separated from the Ouarzazate basin by the Siroua
high plateau (Figure 1).

The Haouz Plain, on the contrary, is interpreted as an intra‐mountain basin located between
the Jebilet and the High Atlas (Michard et al., 2010) and characterized by Miocene‐Pliocene
molasse deposits.

In the following paragraphs we will discuss separately the structural geology of two sectors
of the Western High Atlas, respectively the Imi nʹTanoute area (northern sector) and the
Menizla area (southern sector).

3.2.1. Northern boundary (Imi nʹTanoute Fault)


The geological setting of the northern sector (Imi nʹTanoute area) is characterized by the
contact between the Paleozoic basement and the Mesozoic‐Cenozoic succession which de‐
velops northward in the Haouz Plain. This contact corresponds to the Imi nʹTanoute Fault, a
major fault zone showing an about N70 direction (Figure 7). However, ESE of Imi nʹTanoute
village the unconformable overlap of the Jurassic succession, starting with purple conglom‐
erates and sandstones, above the Paleozoic rocks is preserved. This unconformity is well
visible also at the map scale for the strong discordance between the sub‐horizontal Jurassic
beds and the Paleozoic succession. The latter displays a sub‐vertical attitude, resulted from
the Variscan tectonics which produced N‐S directed folds associated with a pervasive sub‐
vertical cleavage. Further, this area is characterized by a N‐S trending thrust, connected to
the WAFZ, that duplicated the Paleozoic succession with an east vergence.
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 241

In the Paleozoic succession the Lower‐Middle Cambrian schists and greywackes evolves to
schists and schists with quartzite bars of Ordovician age. The Paleozoic succession ends
with Silurian black/reddish schists and Devonian sandstones and conglomerates with lens
of limestones.

Figure 7. Geological map of the northern boundary of the High Atlas in the Imi n’Tanoute area (modi‐
fied after Duffaud, 1981). 1: Quaternary; 2: Miocene; 3: Eocene; 4: Paleocene; 5: Upper Cretaceous; 6:
Cenomanian‐Turonian; 7: Lower Cretaceous; 8: Jurassic; 9: Devonian; 10: Ordovician; 11: Middle Cam‐
brian; 12: Lower Cambrian; 13: normal beds, 14: overturned beds; 15: vertical beds; 16: fold axes, ver‐
gence is indicated; 17: main foliation in Paleozoic rocks; 18: main faults; 19: secondary faults;
20: tectonic boundaries. InTF: Imi n’Tanoute Fault; SekF: Seksaoua Fault. At the top left of the figure
stereogram (Schmidt net, lower hemisphere) with traces of fault planes, observed slip lines and slip
senses. The principal stress axes (S1, S2, S3) and type of stress tensor are indicated.

The post‐Jurassic Mesozoic succession of the Imi nʹTanoute area begins with Lower
Cretaceous marine deposits (yellow and reddish marls, limestones and sandstones with
242 Tectonics – Recent Advances

gypsum of Barremian FIGURE 8 age) followed by a thick Cenomanian‐Turonian sequence


formed by grey and red marls with anhydrite and by Senonian red and white sandstones
with lumachellic limestones and white marlstones. The overlying Maastrichtian
conglomerates and phosphatic sandy marls and limestones were deposited above an
unconformity and evolved up to the Paleocene‐Eocene reddish sandstones and brown
marls. Slightly north of the Imi nʹTanoute village, up to 20 m thick white conglomerates with
limestone and chert pebbles, considered to be Oligocene in age, are also documented
unconformably overlapping the Eocene succession (Zuhlke et al., 2004). Nevertheless the
main unconformity inside the Cenozoic succession corresponds to the basal contact of the
Miocene conglomerates and sandstones (molassic deposits) that rest directly above different
levels of the Upper Cretaceous‐Eocene succession, sealing the thrusting and folding
deformation of the first tectonic phase of the Western High Atlas belt.

Figure 8. Geometrical and kinematic features along the northern boundary of High Atlas in the Imi
n’Tanoute area. A: Panoramic view of the north‐vergent folding in the Cretaceous succession associated
to the dextral sense of movement of the Imi n’Tanoute Fault. B: Detail of a fault plane (Middle Cambri‐
an) bearing oblique slickenlines with a dextral strike‐slip movement. The direction of the red arrow
corresponds to the movement of the missing block. C) Shear zone developed in the Lower Cretaceous
rocks along the Imi n’Tanoute Fault. D: Kilometer‐scale north‐vergent fold developed in Cretaceous‐
Paleogene successions, characterized by a secondary fold with rabbit‐ear geometry. The deformed strata
are unconformably overlain by the Miocene clastic deposits.

In the study area, the geological structures at the Western High Atlas northern border can be
observed along a natural cross‐section directed roughly N‐S (Figure 8A). The Cambrian
metamorphic rocks are juxtaposed to the Cretaceous sedimentary sequences along the Imi
nʹTanoute Fault. A slice of Jurassic rocks is also isolated inside the fault zone. The foliation
attitude of the Cambrian metasandstones and metapelites, that normally shows a sub‐
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 243

vertical N‐S direction, close to the fault zone suffered a virgation produced by the fault
activity becoming sub‐parallel to the fault itself and showing a northward high angle dip‐
ping (Figure 7). North the Imi nʹTanoute Fault, the Lower Cretaceous red marlstones and
sandstones belong to the southern vertical limb of a syncline‐anticline‐syncline system
showing northward vergence. This folding system involved the entire Mesozoic‐Cenozoic
succession of the Imi nʹTanoute area up to the Paleocene‐Eocene deposits, as can be ob‐
served northwest the Houdjanene village (Figure 8D) where the Turonian limestones evi‐
dence a spectacular secondary fold that can be interpreted as a rabbit‐ear fold (Narr &
Suppe 1994; Missenard et al., 2007). In the Houdjanene cross‐section the sub‐horizontal
(slightly dipping) Miocene conglomerates are unconformable above the sub‐vertical Paleo‐
cene‐Eocene beds (limestones and sandstones).

The overall axial direction of the described fold system is directed sub‐parallel to the Imi
nʹTanoute Fault, i.e. N70E as evidenced also by the spatial arrangement of the bedding data
collected in the Cretaceous‐Eocene succession (Figure 7).
The kinematic data collected along the main fault zone of the Imi nʹTanoute Fault indicate a
main dextral strike‐slip movement (Figure 8B and 8C) consistent with a strike‐slip tectonic
regime displaying a sub‐horizontal WNW‐ESE directed maximum compression axis, where
the associated roughly E‐W directed normal faults are compatible too (Figure 7).
In the Cambrian metasandstones a population of NNE‐SSW directed dextral strike‐slip
faults have been also collected, being compatible with a different palaeostress tensor where
the sub‐horizontal axis is directed ENE‐WSW.

3.2.2. Southern boundary (Tizi nʹTest Fault system)


The Paleozoic succession of the Menizla area is characterized by the occurrence of the Upper
Carboniferous deposits of the Ida ou Zal basin. Unlike the northern boundary of the West‐
ern High Atlas, the Mesozoic‐Cenozoic succession starts with the Lower Cretaceous red
sandstones directly overlying the Ordovician rocks outcropping in two small inliers north of
Addouz (Figure 9). The upper part of the succession is characterized by the Maastrichtian‐
Ypresian phosphate series and by the Miocene‐Pliocene continental deposits.
The Mesozoic‐Cenozoic succession at the southern boundary of the Western High Atlas was
deformed by southward verging fold systems comprehensively formed by two wide anti‐
clines separated by a wider syncline. The sub‐vertical, locally reversed, southernmost limb
of the southern anticline represents the margin with the Quaternary deposits of the Souss
Plain.
For the main object of our study, we analyzed the relationships between the folding struc‐
tures and the Tizi nʹTest Fault system in the area of Tafrawtane (Figures 9 and 10). The main
fault that juxtaposes the Cambrian rocks with sub‐vertical principal foliation to the Upper
Cretaceous deposits is a sub‐vertical dextral strike‐slip fault directed about E‐W. The rela‐
tive kinematic data allow defining a strike‐slip palaeostress tensor with a sub‐horizontal
NW‐SE trending σ1 axis (Figure 9).
244 Tectonics – Recent Advances

Figure 9. Geological map of the southern boundary of the High Atlas in the Menizla area (modified after
Choubert, 1957 and Tixeront, 1974). 1: Quaternary; 2: Miocene‐Pliocene; 3: Eocene; 4: Paleocene; 5: Upper
Cretaceous; 6: Cenomanian‐Turonian; 7: Jurassic; 8: Upper Triassic; 9: Middle Triassic; 10: Permian‐
Triassic; 11: Carboniferous; 12: Devonian; 13: Silurian; 14: Ordovician; 15: Middle Cambrian; 16: Lower
Cambrian; 17: Variscan Granites; 18: Main faults; 19: Tectonic Boundaries; 20: Direction of tectonic
transport. ArgF: Argana Fault; BigF: Bigoudine Fault; IfeF: Iferd Fault; TirF: Tirkou Fault; TnTFs: Tizi
n’Test Fault system. Stereograms (Schmidt net, lower hemisphere) with traces of fault planes, observed slip
lines and slip senses are reported for the Iferd, Menizla and Tirkou Faults and for Tizi n’Test Fault system
in the Tafrawtane area. The principal stress axes (S1, S2, S3) and type of stress tensor are indicated.

Actually, the main fault zone is characterized by a slice of Permian‐Triassic red sandstones
interposed between the Cambrian and the Cretaceous rocks that are also affected by a south‐
verging thrust that roots into the main sub‐vertical fault plane.
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 245

The Upper Cretaceous marls and limestones are affected by complex, disharmonic, fold
structures showing a general southward vergence. As a result bedding is variable in
orientation but it generally strikes ENE‐WSW, sub‐parallel to the main direction of the Tizi
nʹTest Fault system; bedding dips follow the fold structures, being progressively steeper
approaching the fault zone. The axis of the fold system is directed about N70E, sub‐parallel
to the Tizi nʹTest Fault system direction, showing a slight eastward plunging.

Figure 10. Panoramic view of the south verging folding in the Upper Cretaceous associated to the
dextral sense of movement of the Tizi n’Test Fault system in the Tafrawtane zone.

However, from a structural point of view, the most important feature occurring in the
Menizla area is the Tizi nʹTest Fault system and in particular the western component of this
major tectonic lineament consisting of different anastomizing branches, the main three of
which (Menizla Fault, Tirkou Fault, Iferd Fault; Baudon et al., 2012) have been studied with
more detail (Figure 9).

Starting from the margin with the Souss Plain, the southernmost fault is the Menizla Fault
characterized by strike‐slip movements associated with mostly south‐verging thrusts sub‐
tending sub‐vertical limbs of folds that deformed the Carboniferous rocks. In particular, the
Menizla Fault developed dextral strike‐slip faulting along high‐angle planes directed about
E‐W. The relative palaeostress tensor indicates a maximum compression directed WNW‐
ESE (Figure 9); in this tectonic context oblique‐normal faults can be also compatible and
some of these have been detected.

Moving toward the north, the first main fault is the Tirkou Fault showing a direction
varying from E‐W to NE‐SW. The structural analysis was performed where the Tirkou Fault
trends ENE‐WSW between the Carboniferous deposits and the Devonian dolostones and
shows a very thick (300‐400 m) fault zone cutting off a slice of Permian‐Triassic red
sandstones (Figure 9). The mesoscale observations evidenced that the Tirkou Fault is
characterized by folds linked to the development of double‐verging thrust systems that root
in sub‐vertical fault planes comprehensively describing a positive flower structure (Figure
11A, C). Fold styles vary from structures with rounded hinges to kink‐like or chevron folds
with steeply dipping axial planes. On average, the fold axes trend sub‐parallel to the
direction of the Tirkou Fault zone (~N100E) with shallow plunging. The fold limbs are cut
246 Tectonics – Recent Advances

by interlinked faults producing imbricate zones; the fault array associated with these folds
provided striations indicating inverse‐oblique displacement. As a consequence, the SSE‐
dipping and the NNW‐dipping thrusts display roughly north and south vergences
respectively. The occurring sub‐vertical faults are characterized by nearly horizontal
slickensides showing dextral displacements (Figure 11B). The fault data collected for the
strike‐slip faults of the Tirkou Fault zone indicate a dominant ENE‐WSW fault direction and
allow to obtain a palaeostress tensor characterized by a sub‐vertical σ2 axis and sub‐
horizontal σ1 and σ3 axes oriented N110E and N10E respectively (Figure 9). Within the
obtained palaeostress tensor, ESE‐WNW directed normal faults, such as that observable in
Figure 11A, can be also compatible.

Figure 11. Geometrical and kinematic features along the southern boundary of High Atlas in the
Tirkou Fault zone.
A: Mesoscale positive flower structure developed in the Carboniferous succession. This structure is
characterized by double‐verging thrust systems that root in sub‐vertical fault planes, associated to the
development of double‐verging folds that trend subparallel to the faults.
B. Detail of a fault plane bearing oblique slickenlines with a dextral strike‐slip movement (Carbonifer‐
ous). The direction of the red arrow corresponds to the movement of the missing block.
C. Shear zone developed in the Carboniferous rocks with top to the south sense of movement.

Likewise, the Iferd Fault is outlined by double‐verging structures within an overall dextral
strike‐slip fault zone, locally juxtaposing Silurian rocks and Permian‐Triassic deposits
(Figure 9). The flower structures that characterize the Iferd Fault zone are well developed in
the Silurian schists consisting of a series of anastomosing convex‐upward reverse faults
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 247

which steepen progressively at depth into sub‐vertical strike‐slip faults. The reverse faults
are directed about E‐W steeply dipping toward N and S and display opposite senses of
shear, that is southward and northward respectively (Figure 12A). The inversion of strike‐
slip fault data has resulted in a palaeostress tensor similar to that obtained for the Tirkou
Fault (Figure 9). In the Ifern Fault zone the possible development of normal faults, due to
the permutation of σ1 and σ2 axes, has been documented by the occurrence of curved
striations on a single fault plane suggesting a dextral‐oblique to oblique‐normal movement
(Figure 12B).

Figure 12. Geometrical and kinematic features along the southern boundary of High Atlas in the Iferd
Fault zone. A. Anastomosing reverse faults root progressively at depth into sub‐vertical strike‐slip
faults in the Silurian schists (positive flower structure). The reverse faults are characterized by opposite
senses of shear. B. Fault plane bearing curved slickenlines suggesting the transition from dextral‐
oblique to oblique‐normal movement.

3.2.3. The major faults of the inner belt sectors


In the Western High Atlas, between the Tizi nʹTest Fault system and the Imi nʹTanoute Fault,
other two major faults, cutting both the Paleozoic basement and the overlying Permian‐
Triassic succession, occur: the Ikakern Fault and the Tizi Machou Fault (Figure 13A).

In the studied outcrops the Ikakern Fault zone is formed by nearly E‐W dextral strike‐
slip/oblique faults (Figure 13C) consistent with a palaeostress tensor with a sub‐vertical σ2
axis and sub‐horizontal σ1 and σ3 axes directed WNW‐ESE and NNE‐SSW respectively
(Figure 13A). Several thrust faults associated to the main fault zone have been observed,
showing north and south vergences.

The Tizi Machou Fault shows a dextral map‐scale offset evidenced by the displacement of
the Western Atlasic Fault Zone. However the sense of shear along individual faults can be
rarely deduced at the outcrop scale since kinematic indicators are only sporadically pre‐
served. The kinematically defined structures suggest a predominance of high‐angle dextral‐
oblique faults clustered along two trends: NNE‐SSW and NE‐SW (Figure 13A and 13B). In
addition, few about E‐W normal faults occur, showing moderate south‐dipping.
248 Tectonics – Recent Advances

Figure 13. Kinematic data from the inner belt sector. A. Schematic structural map of Figure 6. The two
principal fault zones, the Tizi Maachou and Ikakern Fault, are evidenced, for which stereograms
(Schmidt net, lower hemisphere) with traces of fault planes, observed slip lines and slip senses are
reported. The principal stress axes (S1, S2, S3) and type of stress tensor are indicated for the Ikakern
Fault only. B. Fault plane bearing oblique slickenlines with a dextral strike‐slip movement developed in
the Triassic rocks along Tizi Maachou Fault. The direction of the red arrow corresponds to the move‐
ment of the missing block. C. Fault plane bearing oblique slickenlines with a dextral strike‐slip move‐
ment developed in the Cambrian rocks along Ikakern Fault. The direction of the red arrow corresponds
to the movement of the missing block.
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 249

4. Discussion
The results of our field study highlight widespread Cenozoic deformation on both the
southern boundary of the Central High Atlas and on the northern and southern boundaries
of the Western High Atlas. The main deformation structures are represented by NE‐SW
trending high angle dextral strike‐slip faults and sub‐parallel thrust faults linked together
forming asymmetric positive flower structures. The overall deformation of these structures
is completed by fold systems associated with the fault and thrust systems and involving
mainly the Mesozoic‐Cenozoic successions. The flower structures are typical structures
developed under a transpressional tectonic regime where the deformation is partitioned
between high‐angle strike‐slip faults and lower angle reverse faults (Wilcox et al, 1973;
Sanderson and Marchini, 1984; Tikoff and Teyssier, 1994). Large scale geometries and
mesoscale data indicate that reverse faults merge into the main NE‐SW oriented strike‐slip
faults. Strike‐slip faults are not offset by the thrusts and vice versa, this supporting that
thrusts are genetically related to the sub‐parallel strike‐slip faults. The observed structure
relationships support therefore that the analysed sectors of the High Atlas belt were affected
by a transpressional evolution during the Alpine tectonics.

Along the southern boundary of West and Central High Atlas the fault‐thrust‐fold systems,
belonging respectively to the Tizi nʹTest Fault system and the South Atlas Fault, show clear
southward vergences. On the contrary, the structures of the northern margin of the Western
High Atlas are connected with the Imi nʹTanoute Fault activity, being characterized by
northward vergences.

The kinematic inversion of the collected fault‐slip data in the Western High Atlas indicates
that deformation is controlled by sub‐horizontal maximum and minimum stress axes,
within a strike‐slip tectonic setting with a WNW‐ESE directed σ1 axes (Figure 7, 9 and 13).
The lack of pure compressive tensors is probably due to the general scarce preservation of
kinematic indicators on the thrust planes of the analysed outcrops. Pure compressional
stresses of NW‐SE direction have been sometime documented in the northern part of the
Western High Atlas (Amrhar, 2002).

The transpressional character of deformation is confirmed by the relevant occurrence of


oblique striations on the fault planes, thus that pure strike‐slip and/or pure inverse faults are
relatively few.

The WNW‐ESE trending normal faults, which have been collected within some transpressional
fault zones, should be also inserted in the documented tectonic pattern of the Western High
Atlas attesting a NNE‐SSW extension in the late stages of the Cenozoic Alpine evolution.

On the contrary, palaeostress determinations from the Boumalne‐Tinehrir area provide both
strike‐slip and compressional tensors, with a quite steady sub‐horizontal σ1 axis trending
NNW‐SSE. The maximum compression σ1 axes obtained from the palaeostress tensors from the
Central High Atlas are consistent with the palaeostress fields reconstructed in the same area for
the Pliocene‐Quaternary stage of the High Atlas tectonic evolution by Ait Brahim et al. (2002).
250 Tectonics – Recent Advances

Our observations documented a WNW‐ESE compression that was never detected before as
individual phase for the Cenozoic deformation of the Western High Atlas and comes to
enriche palaeostress evolution of the High Atlas. In fact, analogous WNW‐ESE directions of
compression have been also evidenced by Qarbous et al. (2003) but referred to the Carbonif‐
erous (Namurian‐Westphalian) phase and therefore to the Variscan orogeny, and to the
Middle Permian tectonics of the Tizi nʹTest Fault system. As the WNW‐ESE compression
derived from the fault‐slip data collected along fault zones clearly affecting also the Mesozo‐
ic‐Cenozoic successions at the High Atlas belt boundaries (Tizi nʹTest Fault system at
Tafrawtane; Imi nʹTanoute Fault) we can consider this compressional direction referable to
the Alpine orogeny.

In the inner belt sectors, the fault zones that do not cut the Cretaceous‐Eocene deposits but
only the Paleozoic basement and the Permian‐Triassic rocks show geometric‐kinematic
features similar to those of the bordering fault zones. Therefore they have been interpreted as
Cenozoic Alpine faults, admitting the possible reactivation of older high‐angle shear zones.

Regional data and mesostructural analyses suggest the superposition of younger Cenozoic
deformation on older structural trends producing reactivation of previous major fault zones.
As a consequence, it is a generally shared opinion that Tizi nʹTest Fault system and Imi
nʹTanoute Fault were active since the Early Paleozoic and in turn reactivated more times
during the successive tectonic events up to the Alpine Cenozoic phases.

Following Baudon et al. (2012), also the other about E‐W oriented faults of the Western High
Atlas (from south to north Tirkou, Iferd and Argana faults; Figure 9) can be interpreted as
faults reactivated during the Alpine transpressional phase, after deposition of the Late
Triassic deposits.

About the timing of deformation, the results of our study define two main episodes of de‐
formation separated by the basal unconformity of the Miocene‐Pliocene molassic deposits.
The first episode occurred post‐Eocene time, as the Upper Cretaceous‐Eocene successions
were deformed by fold‐thrust systems that were sealed by the unconformable Miocene‐
Pliocene deposits.

The second tectonic event deformed the Miocene‐Pliocene deposits as well as the
Quaternary deposits and therefore can be assigned to a Quaternary age.

We documented the complete polyphase evolution in the Boumalne‐Tinehrir area (Central


High Atlas) whereas we have not new data about the Quaternary deformation in the
Western High Atlas that has been already documented by Sebrier et al. (2006) in the Souss
Plain. Likewise, Quaternary reactivations consisting of thrusting associated with strike‐slip
faulting characterize also the northern boundaries of Western and Central High Atlas and
the Houaz Plain (Morel et al., 2000).

Seismic data suggest that several brittle structures along northern and southern margins of
the High Atlas belt are still active. The few available focal solutions show that the southern
boundary of the High Atlas is characterized mostly by strike slip faulting and subordinate
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 251

transpressional mechanisms, with the direction of the maximum compression P axes


ranging from NW‐SE to N‐S (Serpelloni et al., 2007). In particular the focal mechanism
solutions obtained for the earthquakes of magnitude Mb=5.2 occurred on October 23 and 30,
1992, in the Rissani region (Eastern Anti Atlas) indicate for both events a pure strike slip
faulting (Hanou et al., 2003; Bensaid et al., 2009). The seismogenic zone has been interpreted
to be an E‐W trending structure that could corresponds to the South Atlas Fault and/or
associated structures activated as a dextral strike slip fault. However, the largest earthquake
ever recorded within the Atlas system corresponds to the event of M=5.7 occurred on
February 29, 1960, which destroyed the Agadir city causing 12.000 victims.

Also the GPS data indicate that the deformation along the High Atlas is still active and
accommodates about 1.5 mm/year of NW‐SE compression related to the convergence
between Africa and Iberia plates (Serpelloni et al., 2007).

In addition, as more general result, we propose two regional geological cross‐sections


representing at the belt scale the structural patterns of Western and Central High Atlas,
performed on the basis of our field studies integrated with literature data (Figure 14)
(Hollard et al, 1985; Froitzheim et al, 1988; Teixell et al., 2003).

Figure 14. Simplified geological cross‐sections through Western (A‐A’) and Central (B‐B’) High Atlas.
1: Neogene and Quaternary; 2: Mesozoic and Paleogene, 3: Precambrian and Paleozoic. SAF: South
Atlas Fault; NAF: North Atlas Front; TnTFs: Tizi n’Test Fault system; JeT: Jebilet Thrust; InTF: Imi
n’Tanoute Fault; TzMF: Tizi Maachou Fault; IkaF: Ikakern Fault.
252 Tectonics – Recent Advances

Comprehensively, these schematic cross‐sections represent a possible model of Alpine


transpressional evolution for the whole High Atlas belt. The main structures are
characterized by high‐angle geometries and dextral strike‐slip kinematics along an ENE‐
WSW direction. These fault zones generated also thrust planes and folds that in the inner
belt sectors show double vergences at the mesoscale whereas along the bounding areas they
are more developed characterizing the entire High Atlas belt by a double vergence, as at the
northern margin the vergence is toward north (North Atlas Fault and Imi nʹTanoute Fault
zones in A‐Aʹ and B‐Bʹ cross‐sections, respectively) while at the southern margin it is toward
south (South Atlas Fault zone and Tizi nʹTest Fault system in A‐Aʹ and B‐Bʹ cross‐sections,
respectively).

The proposed geological cross‐sections evidence also that the Alpine deformation was/is not
limited to the High Atlas mountain range but involved/involves wider sectors. In the B‐Bʹ
cross‐section the Jebilet Thrust has been interpreted as a low‐angle surface that roots into the
high‐angle transpressional zone of the Imi nʹTanoute Fault (Figure 14) and therefore
Paleozoic basement and Mesozoic‐Cenozoic deposits of the Haouz Plain are faulted and
folded by the Cenozoic tectonic events. Along the southern boundary of the Western High
Atlas we considered the Alpine tectonics deforming the Paleozoic rocks of the Anti Atlas
belt as well as the Souss and Ouarzazate deposits of which they represent the basement, as
already documented (Sebrier et al., 2006; Malusà et al., 2007).

5. Conclusion
The field study, mainly consisting in detailed mesostructural analyses, from the Western
High Atlas transect and the Boumalne‐Tinerhir region in the Central High Atlas, indicates a
major role for transpressional tectonics in the Alpine structural evolution of the High Atlas
belt. In particular, the deformation in the studied regions is controlled by two regional right‐
lateral fault systems (Tizi nʹTest‐South Atlas and Imi nʹTanoute‐North Atlas) and their
associated structures that involved, at south, the Anti Atlas belt and, at north, the Western
Meseta domain (Jebilet range). Between these two major tectonic lineaments, a 50 to 100 km
wide region is characterized by a complex tectonic framework, dominated by strike‐slip
faulting, in which strong uplift and exhumation occurred.

Kinematic measurements on major fault planes and mesoscale structural analysis reveal that
the prevailing structural associations correspond to ENE‐WSW trending dextral strike‐slip
faults and sub‐parallel thrust faults, describing typical positive flower structures.

In the proposed transpressional model (Figure 14) the High Atlas belt appears to have a
flower structure cross‐sectional geometry with greater amount of thrust displacement along
its northern and southern boundaries. This pattern showing double‐verging structures re‐
quires a downward extrapolation of surface thrusts rooting into high‐angle fault zones.

Fault analyses and palaeostress reconstructions suggest that flower structures and fold sys‐
tems evolved into a right‐lateral transpression which is related to a direction of maximum
compression varying from about E‐W to about N‐S.
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 253

The High Atlas is a significant example of transpressional belt dominated by strain


partitioning between distinct strike‐slip and thrust faults which result from reactivation of
pre‐existing structures inherited from the pre‐Alpine complex evolution.

The Cenozoic reactivation occurred during two main deformation events: Late Eocene‐
Oligocene and Pliocene‐Pleistocene, the latter being still active.

The High Atlas can be therefore considered an example of active transpressional belt as
defined by Cunningham (2005).

This study, although incomplete, furnished interesting results that indicate as structural and
kinematic analyses are important methodologies and that their future development can
provide new data for the better understanding crustal architecture, history of structural
reactivation, partitioning of strain and distribution of present‐day tectonic activity within
the orogens. Particularly, in this case, the definition of the orogenic mechanisms represents
the main step for interpreting the origin of topographic elevation, the principal still debated
topic of the High Atlas geology.

Author details
Alessandro Ellero and Giuseppe Ottria
CNR Institute of Geosciences and Earth Resources, Pisa, Italy

Marco G. Malusà
Department of Geological Sciences and Geotechnology, University of Milano Bicocca, Milano, Italy

Hassan Ouanaimi
Université Cadi Ayyad Ecole Normale Supérieure Département de Géologie, Marrakech, Morocco

Acknowledgement
This paper is dedicated to Piero Elter, outstanding geologist and our professor who died
when we were preparing the final draft of the manuscript.

This work was partly carried out within the CNR Short Term Mobility Project 2009 “Study
of the structural‐geological evolution of the High Atlas (Morocco)”.

6. References
Aït Brahim L, Chotin P, Hinaj S, Abdelouafi A, El Adraoui A, Nakcha C, Dhont D, Charroud
M, Sossey Alaoui F., Amrhar M, Bouaza A, Tabyaoui H, Chaouni A (2002) Paleostress
evolution in the Moroccan African margin from Triassic to Present. Tectonophysics 357:
87‐205.
Amrhar M (2002) Paléocontraintes et déformations syn‐ et post‐collision Afrique–Europe
identifiées dans la couverture mésozoïque et cénozoïque du Haut Atlas occidental
(Maroc). C. R. Géoscience 334: 279–285.
254 Tectonics – Recent Advances

Arboleya ML, Teixell A, Charroud M, Julivert MA (2004) Structural transect through the
High and Middle Atlas of Morocco. Journal of African Earth Sciences 39: 319‐327.
Ayarza P, Alvarez‐Lobato F, Teixell A, Arboleya ML, Teson E, Julivert M, Charroud M
(2005) Crustal structure under the central High Atlas Mountains (Morocco) from
geological and gravity data. Tectonophysics 400: 67‐84.
Barbero L, Teixell A, Arboleya ML, Rio PD, Reiners PW, Bougadir B (2006) Jurassic‐to‐
present thermal history of the central High Atlas (Morocco) assessed by low‐
temperature thermochronology. Terra Nova 19: 58‐64.
Baudon C, Redfern J, Van Den Driessche J (2012) Permo‐Triassic structural evolution of the
Argana Valley, impact of the Atlantic rifting in the High Atlas, Morocco. Journal of
African Earth Sciences 65: 91‐104.
BeauchampW, Barazangi M, Demnati A, El Alji M (1996) Intracontinental rifting and
inversion: Missour Basin and Atlas Mountains, Morocco. AAPG Bull. 80: 1459‐1482.
Beauchamp W, Allmendinger RW, Barazangi M, Demnati A, El Alji M, Dahmani M (1999)
Inversion tectonics and the evolution of the High Atlas mountains, Morocco, based on a
geological geophysical transect. Tectonics 18: 163‐184.
Benammi M, Toto EA, Chakiri S (2001) Les chevauchements frontaux du Haut Atlas central
marocain: styles structuraux et taux de raccourcissement diff´erentiel entre les versants
nord et sud C R Acad Sci Paris 333: 241‐247.
Bensaid I, Medina F, Cherkaoui TE, Buforn E, Hahou Y (2009) New P‐wave first motion
solutions for the focal mechanisms of the Rissani (Morocco) earthquakes of October 23d
and 30th, 1992. Bulletin de l’Institut Scientifique, Rabat, section Sciences de la Terre.31:
57‐61.
Bracène R, Bellahcene A, Bekkouche D, Mercier E, Frizon de Lamotte D (1998) The thin‐
skinned style of the South Atlas Front in central Algeria. In: Macgregor DS, Moody RTJ,
Clark‐Lowes DD, editors. Petroleum Geology of North Africa. Geol. Soc. Spec. Publ.
133: 395‐404.
Delcaillau B, Laville E, Amhrar M, Namous M, Dugué O, Pedoja K (2010) Quaternary
evolution of the Marrakech High Atlas and morphotectonic evidence of activity along
the Tizi NʹTest Fault, Morocco. Geomorphology 118: 262‐279.
Cerrina Feroni A, Ellero A, Ottria G, Malusà M, Polino R, Musumeci G, Pertusati PC (2007)
Kinematic analysis of the High Atlas in the Tinerhir Area (Southern Morocco): Evidence
of a transpressional fold‐thrust belt. The First MAPG International Convention
Conference & Exhibition, Marrakech, October 28‐31, 2007. Abstract Book, pp. 165.
Cerrina Feroni A., Ellero A., Malusà M.G., Musumeci G., Ottria G., Polino R., Leoni L. (2010)
‐ Transpressional tectonics and nappe stacking along the Southern Variscan Front of
Morocco. International Journal of Earth Sciences, (Geol Rundsch) 99: 1111‐1122.
Choubert G (1957), Carte Géologique du Maroc au 1/500.000, Feuille Marrakech. Notes et
Mém. Serv. géol. Maroc 70.
Cornée JJ, Destombes J (1991) Lʹordovicien De La Partie W Du Massif Ancien Du Haut‐Atlas
Occidental (maroc Hercynien). Geobios 24: 403‐415.
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 255

Cunningham D (2005) Active intracontinental transpressional mountain building in the


Mongolian Altai: Defining a new class of orogen. Earth and Planetary Science Letters
240: 436–444.
De Koning G (1957) Geologic des Ida ou Zal (Maroc). Stratigraphie, pétrographie et
tectonique de la partie Sud‐Ouest du bloc occidental du massif hercynien du Haut‐Atlas
(Maroc). Leidese Geology Meded 23: 209p.
Delvaux D (1993) The TENSOR program for reconstruction: examples from the East Africa
and the Baikal Rift System. Terra Abstract 5: 216.
Dewey JF, Helman MNL, Turco E, Hutton DHW, Knott SD (1989) Kinematics of the western
Meditettanean. In: Coward M, editor. Alpine Tectonics. Geological Society London
Special Publication 45. pp. 265‐283.
Dias R, Hadani M, Leal Machado I, Adnane N, Hendaq Y, Madih K, Matos C (2011)
Variscan structural evolution of the western High Atlas and the Haouz plain (Morocco).
Journal of African Earth Sciences 61: 331‐342.
Duffaud F (1981) Carte géologique du Maroc au 1/100000, feuille Imi n’Tanout. Notes et
Mém. Serv. géol. Maroc 203.
El Harfi A, Lang J, Salomon J, Chellai EH (2001) Cenozoic sedimentary dynamics of the
Ouarzazate foreland basin (Central High Atlas Mountains, Morocco). Int. J. Earth Sci.
90: 393‐411.
El Harfi A, Guiraud M, Lang J (2006) Deep‐rooted ‘‘thick skinned’’ model for the High Atlas
Mountains (Morocco). Implications for the structural inheritance of the southern Tethys
passive margin. Journal of Structural Geology 28: 1958‐1976.
Ellouz N, Patriat M, Gaulier JM, Bouatmani R, Saboundji S (2003) From rifting to Alpine
inversion: Mesozoic and Cenozoic subsidence history of some Moroccan basins. Sedim.
Geol. 156: 185–212.
Ettachfini E, Andreu B (2004) Le Cénomanien et le Turonien de la Plate‐forme Préafricaine
du Maroc. Cretaceous Research 25: 277‐302.
Fabuel‐Perez I, Redfern J, Hodgetts D (2009) Sedimentology of an intra‐montane rift‐
controlled fluvial dominated succession:The Upper Triassic Oukaimeden Sandstone
Formation, Central High Atlas, Morocco. Sedimentary Geol. 218: 103‐140.
Fiechtner L, Friedrichsen H, Hammerschmidt K (1992) Geochemistry and geochronology of
early Mesozoic tholeiites from Central Morocco. Geol. Rund. 81: 45‐62.
Fraissinet C, Zouine EM, Morel JL, Poisson A, Andrieux J, Faure‐Muret A (1988) Structural
evolution of the southern and northern Central High Atlas in Paleogene and Mio‐
Pliocene times. In: Jacobshagen V, editor. The Atlas system of Morocco. Lect. Notes
Earth Sci. pp. 273‐291.
Frizon de Lamotte D, Saint Bézar B, Bracène R, Mercier E (2000) The two main steps of the
Atlas building and geodynamics of the western Mediterranean. Tectonics 19: 740‐761.
Froitzheim N, Stets J, Wurster P (1988) Aspects of Western High Atlas tectonics. In:
Jacobshagen V, editor. The Atlas system of Morocco. Lect. Notes Earth Sci. pp. 219‐244.
Gauthier H (1960) Contribution à l’étude géologique des formations post‐liasiques des
basins du Dadès et du Haut Todra (Maroc méridional). Notes et Mém. Serv. Géol.
Maroc 119.
256 Tectonics – Recent Advances

Giese P, Jacobshagen V (1992) Inversion tectonics of intracontinental ranges: High and


Middle Atlas, Morocco. Geol. Rundsch. 81: 249‐259.
Gomez F, Allmendinger R, Barazangi M, Beauchamp W (2000) Role of the Atlas Mountains
(northwest Africa) within the African‐Eurasian plate‐boundary zone. Geology 28: 769‐
864.
Görler K, Helmdach FF, Gaemers P, Heissig K, Hinsch W, Mädler K, Schwarzhans W, Zucht
M (1988) The uplift of the Central High Atlas as deduced from Neogene continental
sediments of the Ouarzazate province, Morocco. In: Jacobshagen V, editor. The Atlas
system of Morocco. Lect. Notes Earth Sci. pp. 361‐404.
Hafid M (2000) Triassic‐early Liassic extensional systems and their Tertiary inversion,
Essaouira Basin (Morocco). Marine Petro. Geol. 17: 409‐429.
Hafid M, Ait Salem A, Bally AW (2000) The western termination of the Jbilet ‐High Atlas
system (Offshore Essaouira Basin, Morocco). Marine Petrol. Geol. 17: 431‐443.
Hafid M, Zizi M, Bally AW, Ait Salem A (2006) Structural styles of the western onshore and
offshore termination of the High Atlas, Morocco. C. R. Géoscience 338: 50‐64.
Hahou Y, Jabour N, Oukemeni D, El Wartiti M (2003) The October 23; 30, 1992 Rissani
earthquakes in Morocco: Seismological, macroseismic data. Bull. Int Inst. Seismol.
Earthq. Eng. Special edition: 85‐94.
Hindermeyer J, Gauthier H, Destombes J, Choubert G., Faure‐Muret A. (1977) Carte
géologique du Maroc, Jbel Saghro‐Dadès (Haut Atlas central, sillon sud‐atlasique et
Anti‐Atlas oriental) – Echelle 1/200000. Notes et Mém. Serv. géol. Maroc 161.
Hoepffner C, Soulaimani A, Piqué A (2005) Moroccan Hercynides. Journal of African Earth
Science 43: 144–165.
Hoepffner C, Houari MR, Bouabdelli M (2006) Tectonics of the North African Variscides
(Morocco, Western Algeria), an outline. In: Frizon de Lamotte D, Saddiqi O, Michard A,
editors. Recent Developments on the Maghreb Geodynamics. C. R. Géoscience 338: pp.
25‐40.
Hollard H, Choubert G, Bronner G, Marchand J, Sougy J (1985) Carte géologique du Maroc –
Echelle 1/1000000. Notes et Mém. Serv. géol. Maroc 260.
Houari MR, Hoepffner C (2003) Late Carboniferous dextral wrench‐dominated
transpression along the North African craton margin (Eastern High Atlas, Morocco).
Journal of African Earth Sciences 37: 11‐24.
Jacobshagen V, Brede R, Hauptmann M, Heinitz W, Zylka R (1988) Structure and post‐
Paleozoic evolution of the Central High Atlas. In: Jacobshagen V, editor. The Atlas
system of Morocco. Lect. Notes Earth Sci. pp. 245‐271.
Jossen JA, Filali‐Moutei J (1992). A new look at the structural geology of the southern side of
the central and eastern High Atlas Mountains. Geol. Rundsch. 81: 143‐156.
Knight B, Nomade S, Renne PR, Marzoli A, Bertrand H, Youbi N (2004) The Central Atlantic
Magmatic Province at the Triassic–Jurassic boundary : paleomagnetic and 40Ar/39Ar
evidence from Morocco for brief, episodic volcanism. Earth Planet. Sci. Let. 228: 143‐160.
Laville E, Lesage JL, Séguret M (1977) Géométrie, cinématique, dynamique de la tectonique
atlasique sur le versant sud du Haut Atlas marocain : aperçu sur les tectoniques
hercyniennes et tardi‐hercyniennes. Bull. Soc. geol. Fr. 19: 527‐539.
Structural Geological Analysis of the High Atlas (Morocco): Evidences of a Transpressional Fold-Thrust Belt 257

Laville E, Petit JP (1984) Role of synsedimentary strike‐slip faults in the formation of the
Moroccan Triassic basins. Geology 12: 424‐427.
Laville E, Piqué A (1991) La distension crustale atlantique et atlasique au Maroc au début du
Mésozoique: le rejeu des structures hercyniennes. Bull. Soc. geol. France 162: 1161–1171.
Laville E, Piqué A (1992) Jurassic penetrative deformation and Cenozoic uplift in the central
High Atlas (Morocco). A tectonic model Structural and orogenic inversions. Geol.
Rundsch. 81: 157‐170.
Laville E, Piqué A, Amrhar M, Charroud M (2004) A restatement of the Mesozoic Atlasic
rifting (Morocco). Journal of African Earth Sciences 38: 145‐153.
Malusà M, Polino R, Cerrina Feroni A, Ellero A, Ottria G, Baidder L, Musumeci G (2007)
Post‐Variscan tectonics in eastern Anti‐Atlas (Morocco). Terra Nova 19: 481–489.
Marzoli A, Bertrand H, Knight KB, Cirilli S, Buratti N, Verati C, Nomade S, Renne PR, Youbi
N, Martini R, Allenbakh K, Neuwerth R, Rapaille C, Zaninetti L, Bellieni G (2004)
Synchronism of the Central Atlantic magmatic province and the Triassic‐Jurassic
boundary climatic and biotic crisis. Geology 32: 973‐976.
Marzoqi M, Pascal A (2000) Séquences de dépots et tectono‐eustatisme à la limite
Crétacé/Tertiaire sur la marge sud‐téthysienne (Atlas de Marrakech et bassin de
Ouarzazate, Maroc). Newslett. Stratigr., 38: 57‐80.
Mattauer M, Tapponier P, Proust F (1977) Sur les mécanismes de formation des chaines
intracontinentales L’exemple des chaines atlasiques du Maroc. Bull. Soc. geol. France 19:
521‐526.
Michard A, Yazidi A, Benziane F, Hollard H, Willefert S (1982) Foreland thrusts and
olistostromes on the presaharian margin of the variscan orogen, Morocco. Geology
10:253‐256.
Michard A, Saddiqi O, Chalouan A, Frizon de Lamotte D (2008) Continental Evolution: The
Geology of Morocco. Structure, Stratigraphy, and Tectonics of the Africa‐Atlantic‐
Mediterranean Triple Junction. Springer‐Verlag, Berlin Heidelberg 116: 404 p.
Milhi A. (1997) Carte géologique du Maroc, Tinerhir – Echelle 1/100000. Notes et Mém. Serv.
géol. Maroc 377.
Missenard Y, Zeyen H, Frizon de Lamotte D, Leturmy P, Petit C, Sébrier M, Saddiqi O (2006)
Crustal versus asthenospheric origin of the relief of the Atlas mountains of Morocco. J.
Geophys. Res. 111 (B03401) doi:101029/2005JB003708.
Missenard Y, Taki Z, Frizon de Lamotte D, Benammi M, Hafid M, Leturmy P, Sebrier M
(2007) Tectonic styles in the Marrakesh High Atlas (Morocco): the role of heritage and
mechanical stratigraphy. Journal of African Earth Sciences 48: 247‐266.
Morel JL, Zouine EM, Andrieux J, Faure‐Muret A (2000) Déformations néogènes et
quaternaires de la bordure nord‐haut‐ atlasique (Maroc); rôle du socle et conséquences
structurales. Journal of African Earth Sciences 30: 119‐131.
Mustaphi H, Medina F, Jabour H, Hoepffner C (1997) Le bassin du Souss (Zone de faille du
Tizi n’Test, Haut Atlas occidental, Maroc): résultat d’une inversion tectonique controlée
par une faille de d´etachement profonde. Journal of African Earth Sciences 24: 153‐168.
Narr W, Suppe J (1994) Kinematics of basement‐involved compressive structures. American
Journal of Science 294: 802‐860.
258 Tectonics – Recent Advances

Ouanaimi H, Petit JP (1992) The southern limit of the Hercynian belt in the High Atlas
(Morocco): reconstitution of an undeformed projecting block. Bull. Soc. Geol. France
163: 63‐72.
Piqué A, Tricart P, Guiraud R, Laville E, Bouaziz S, Amrhar M, Aït Ouali R (2002) The
Mesozoic‐Cenozoic Atlas belt (North Africa): An overview. Geodinamica Acta 15: 185‐208.
Proust F (1973) Etude stratigraphique, petrographique et structurale du bloc oriental du
Massif Ancien du Haut Atlas (Maroc). Notes et Mém. Serv. géol. Maroc 34, 254: 15‐54.
Proust F, Petit JP, Tapponnier P (1977) Lʹaccident de Tizi nʹTest et le rôle des décrochements
dans la tectonique du Haut Atlas occidental (Maroc). Bulletin de la Societé Géologique
de France 7: 541‐551.
Qarbous A, Medina F, Hoepffner C, Ahmamou, M, Errami A, Bensahal A (2003) Apport de
l’étude des bassins stéphano‐autuniens et permo‐triasiques du Haut Atlas occidental
(Maroc) à la chronologie du fonctionnement de la zone de failles de Tizi n’Test. Bulletin
de l’Institut Scientifique (Rabat), Sciences de la Terre. 25: 43‐53.
Russo P, Russo L (1934) Le grand accident sud ‐ atlasien . Bull. Soc. géol. France 5: 375‐384.
Saber H, El Wartiti M, Broutin J (2001) Dynamique sédimentaire comparative dans les
bassins stéphano‐permiens des Ida Ou Zal et Ida ou Ziki (Haut Atlas occidental,
Maroc). Journal of African Earth Sciences 32:573–594.
Sanderson DJ, Marchini WRD (1984) Transpression. Journal of Structural Geology 6: 449–458.
Saint‐Bézar B, Frizon de Lamotte D, Morel JL, Mercier E (1998) Kinematics of large scale tip
line folds from the High Atlas thrust belt, Morocco. J. Struct. Geol. 20: 999‐1011.
Sébrier M, Siame L, El Mostafa Z, Winter T, Missenard Y, Leturmy P (2006) Active teconics
in the Moroccan High Atlas. C. R. Géoscience 338: 65‐79.
Serpelloni E, Vannucci G, Pondrelli S, Argnani A, Casula G, Anzidei M, Baldi P, Gasperini P
(2007) Kinematics of theWestern Africa‐Eurasia plate boundary from focal mechanisms
and GPS data. Geophys. J. Int. 169: 1180‐1200.
Teixell A, Arboleya ML, Julivert M, Charroud M (2003) Tectonic shortening and topography
of the central High Atlas (Morocco). Tectonics 22: 1051, doi:10.1029/2002TC001460.
Teixell A, Ayarza P, Zeyen H, Fernandez M, Arboleya ML (2005) Effects of mantle
upwelling in a compressional setting: the Atlas Mountains of Morocco. Terra Nova 17:
456‐461.
Teson E, Teixell A (2008) Sequence of thrusting and syntectonic sedimentation in the eastern
thrust belt (Dadès and Mgoun Valleys, Morocco). Int. J. Earth Sci. 97: 103‐113.
Tikoff B, Teyssier C (1994) Strain modeling of displacement‐field partitioning in
transpressional orogens. Journal of Structural Geology 16: 1575–1588.
Tixeront M (1974) Carte géologique et minéralisations de couloir d’Argana (Haut Atlas
occidental) au 1/100 000. Notes et Mém. Serv. géol. Maroc 205.
Wilcox RE, Harding TP, Seely DR (1973) Basic wrench tectonics. American Association of
Petroleum Geologists Bulletin 57: 74–96.
Zühlke R, Bouaouda MS, Ouajhain B, Bechstädt T, Reinfelder R (2004) QuantitativeMeso‐
Cenozoic development of the eastern Central Atlantic Continental shelf, western High
Atlas, Morocco. Marine and Petroleum Geology 21: 225‐276.
Chapter 10

Plate Tectonic Evolution of the Southern Margin


of Laurussia in the Paleozoic

Jan Golonka and Aleksandra Gawęda

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/50009

1. Introduction
The role of an active margin of Eurasia during Mesozoic and Cenozoic times was well
defined (Golonka, 2004). The trench-pulling effect of the north dipping subduction, which
developed along the new continental margin caused rifting, creating the back-arc basin as
well as transfer of plates from Gondwana to Laurasia. The present authors applied this
model to the southern margin of Laurussia during Paleozoic times. The preliminary results
of their work were presented during the Central European Tectonic Group (CETEG) in
2011 in Czech Republic. The supercontinent of Laurussia, defined by Ziegler (1989), in-
cluded large parts of Europe and North America. The southern margin of this superconti-
nent stretched out between Mexico and the Caspian Sea area. The present authors attempt-
ed to characterize the entire margin, paying the special attention to Central and Eastern
Europe.

2. Methods
The present authors were using a plate tectonic model, which describes the relative motions
plates and terranes during Paleozoic times. This model is based on PLATES, GPLATES and
PALEOMAP software (see Golonka et al. 1994, 2003, 2006a,b, Golonka 2000, 2002, 2007a,b,c,
2009a,b). The plate tectonic reconstruction programs generated palaeocontinental base
maps. It takes tectonic features in the form of digitised data files, assembles those features in
accordance with user specified rotation criteria (Golonka et al. 2006a).

The rigid, outer part of the earth divided into many pieces known as lithospheric plates,
comprising both the continental landmasses and oceanic basins These plates are in motion
relative to each other and to the earth itself. Assuming the earth is a sphere, the motion of a
plate across the earth's surface can be described as motion about the axis of a pole of rotation
that goes through the centre of the earth. The intersection of the pole's axis with the earth's

© 2012 Golonka and Gawęda, licensee InTech. This is an open access chapter distributed under the terms
of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
262 Tectonics – Recent Advances

surface is referred to by its latitude/longitude coordinates. The distance the plate travels
about the pole is an angular distance and is recorded in degrees. A stage pole of rotation
describes the distance a plate moved from one time to the next time (i.e. from 20 Ma to 10
Ma). A finite pole of rotation describes the total distance a plate moved from some time in
the past to the present day (i.e. from 20 Ma to 0 Ma). A rotation file contains a list of poles of
rotation for various plates. The rotation files used by applied software contain finite poles of
rotation. Thus for each plate, there are several finite poles of rotation for different times in
the past. Plate models that use rotation files that describe the motion of plates relative to
other plates are called relative framework models. Among, the data that show the relative
motions between plates are fracture zones. Fracture zones are essentially flowlines between
plates. For example, in the South Atlantic, the fracture zones show the motion of South
America relative to (or away from) Africa. The finite pole of rotation describing this motion
is a relative pole. South America is referred to as the moving plate and Africa as the fixed
plate (Golonka et al. 2006a). The rotation file contains a list of finite rotations between pairs
of tectonic elements, at different episodes of time, with brief bibliographic notes or general
comments for each individual rotation.

3. Paleozoic major continental plates and ocean related to the Laurussia


supercontinent
The break-of the supercontinent Pannotia (Dalziel at al., 1994) during the latest Precambrian
times (Golonka, 2000, 2002) lead to the formation of the new continents. Fig 1. depicts the
position of these continents at the beginning of the Paleozoic.

Figure 1. Plate tectonic global map of Early Cambrian (plates position as of 544 Ma). Mollweide projec-
tion. Modified from Golonka (2012). 1 - oceanic spreading center and transform faults, 2 - subduction zone,
3 - thrust fault, 4 - normal fault, 5 - transform fault..
Plate Tectonic Evolution of the Southern Margin of Laurussia in the Paleozoic 263

3.1. Baltica
The continent Baltica was named after the Baltic Sea. It consisted of a major parts of north-
ern and eastern Europe. It was bounded by the north by the border of the shelf of Norway,
southern Barents Sea and Novaya Zemlya, on the east by the Ural suture, and on the south-
west by a suture located close, but not quite along the Teisseyre-Tornquist line (Scotese &
McKerrow, 1990; Golonka et al., 1994). The southern boundary is more controversial. The
Ukrainian shield certainly belonged to Baltica (Zonenshain et al, 1990). Perhaps also frag-
ments of the North European platform like the Malopolska block, Bruno-Vistulicum, Moesia
and other small blocks located now around the Baltic Sea belonged to Baltica (Kalvoda &
Bábek 2010, Żelaźniewicz et al., 2009, Besutiu, 2001). In Central and southwestern Europe
the possible boundary of Baltica is marked by the extent of North European plate below the
Carpathian and Balkan nappes (Figs. 1-3)

3.2. Gondwana
The supercontinent Gondwana, also known as Gondwanaland (Vevers, 2004) was named after
the ancient Indian tribes Gonds; in Sanskrit Gondwana means ”the forest of Gonds”. The con-
tinents forming the core of Gondwana include South America, Africa, Madagascar, India,
Antarctica and Australia. The location of numerous smaller continental blocks that bordered
Gondwana is less certain. The following were adjacent to Gondwana at some time during the
Paleozoic: Yucatan, Florida, Avalonia, central European (Cadomian) terranes between the
Armorica and Bohemian Massif, Moesia, Iberia, Apulia and the smaller, southern European
terranes, central Asian terranes (Karakum and others), China (several separate blocks), and the
Cimmerian terranes of Turkey, Iran, Afghanistan, Tibet and Southeast Asia (Figs, 1, 2).

Figure 2. Plate tectonic map of Gondwana. Late Vendian - Early Cambrian (plates position as of 544
Ma). Stereographic polar projection. Modified from Golonka (2012). Legend as in Fig. 1
264 Tectonics – Recent Advances

3.3. Laurentia
The continent Laurentia was named after the Laurentia Shield, in turn after St. Lawrence
(Laurentius in Latin) River. North America was a major component of Laurentian plate. This
plate also included Greenland, Chukotka peninsula, Svalbard and large part of the Barents
Sea (Barentsia) fragment of Alaska (North Slope), northwest Ireland, and Scotland,
(Golonka, 2000, 2002, Ford & Golonka, 2003, Golonka et al., 2003). (Figs 1-3). Its southern
(present day eastern boundary is located within Appalachians, its northern (present day
western) boundary is located within Rocky Mountains. The relationship between Laurentia
and Siberia remains speculative (Figs 1-5).

Figure 3. Plate tectonic map of Middle Cambrian (plates position as of 510 Ma). Mollweide projection.
Modified from Golonka (2012). 1 - oceanic spreading center and transform faults, 2 - subduction zone, 3
- thrust fault, 4 - normal fault, 5 - transform fault.

Figure 4. Plate tectonic map of Late Cambrian (plates position as of 498 Ma). Mollweide projection.
Modified from Golonka (2012). 1 - oceanic spreading center and transform faults, 2 - subduction zone, 3
- thrust fault, 4 - normal fault, 5 - transform fault.
Plate Tectonic Evolution of the Southern Margin of Laurussia in the Paleozoic 265

3.4. Avalonia
The name Avalonia is derived from the Avalon peninsula, Newfoundland, eastern Canada.
The names ‘Avalon Composite Terrane’ (Keppie, 1985), or Superterrane were also used.
Western Avalonia included terranes in northern Germany, the Ardennes in Belgium and
northern France, England, Wales, southeastern Ireland, eastern Newfoundland, much of
Nova Scotia, southern New Brunswick and some coastal parts of New England (Golonka
2009, McKerrow et al. 1991). The inclusion of terranes in the eastern part of Avalonia is more
speculative (Fig. 5).

Figure 5. Plate tectonic map of Early Ordovician (plates position as of 485 Ma). Mollweide projection.
Modified from Golonka (2012). 1 - oceanic spreading center and transform faults, 2 - subduction zone, 3
- thrust fault, 4 - normal fault, 5 - transform fault.

Figure 6. Plate tectonic map of Middle Ordovician (plates position as of 472 Ma). Mollweide projection.
Modified from Golonka (2012). 1 - oceanic spreading center and transform faults, 2 - subduction zone, 3
- thrust fault, 4 - normal fault, 5 - transform fault.
266 Tectonics – Recent Advances

Perhaps eastern Avalonia consisted of northwestern Poland fragments now included into
Sudety Mountains and Bohemian Massif, terranes accreted in Carpathian-Balkan and
Pannonian regions, containing fragments consolidated during Cadomian orogeny, and also
Serbo-Macedonian massif, Rhodopes and Pontides in eastern Europe and adjacent part of
Asia (Golonka, 2012). Avalonia originated after Gondwana break-up during Early-Middle
Ordovician times (Fig.6).

3.5. Cadomia
Cadomia is named after city of Caen, Roman/Latin Cadomia in northern France also known
as Armorica or Armorica Group Terrane ( e.g. Lewandowski 2003 and references therein) or
Gothic terranes (Stampfli, 2001). It consisted of the fragments of western and Central
Europe between northwestern France (Brittany) and Czech Republic consolidated during
Cadomian orogeny (Figs.1-6). It includes Saxoturingian zone/terrane in Germany and
northwestern Czech Republic. Cadomia belonged to Perigondwana during Early Paleozoic
times. Perhaps it was detached from Gondwana during Silurian times (Golonka et al. 2006a,
Lewandowski 2003), the existence and nature of this detachment as well as extension of
Cadomia eastward and westward remain quite speculative.

3.6. Iapetus Ocean


The Iapetus Ocean is named after Titan, son of Uranus, the sky, and Gaia, the Earth, father
of Atlas. It preceded Atlantic off the American coast, therefore Atlantic was derived from
Atlas and proto-Atlantic was named after Atlas’ father. The Iapetus Ocean was located be-
tween Gondwana and Laurentia (Figs 1-6), later also between Avalonia and Laurentia (Fig.
6). It started to open as a rift between Laurentia and Rodinia in late Neoproterozoic, about
760-700 Ma, and the maximum obtained at about 600-520 Ma (Kamo et al. 1989; Cawood et
al. 2001). Its closure is connected with the rotation and collision of Baltica, Laurentia and the
fragment of Gondwana – Avalonia (Hartz & Torsvik 2002). The Iapetus suture was formed
Caledonian Orogeny was formed in the time interval 480-440 Ma in the west while the east-
ern part (in Europe) closed at 440-420 Ma. The eastern extension of Iapetus is not so certain
(Golonka, 2002, 2006a, b). Perhaps it was located between Baltica and Gondwana. Part of
the Iapetus Ocean is known as Tornquist Sea, the oceanic basin located between the south-
west and southern margin of Baltica and Gondwana, and later also Avalonia plate. The
name is derived from the Tornquist zone in central Europe, already mentioned border of
Baltica. The position of Tornquist Sea is speculative, because Baltica rotated during Early
Paleozoic times.

3.7. Rheic Ocean


The Rheic Ocean was named after Rhea, Titaness-Goddess, daughter of Uranos, the sky, and
Gaia, the Earth, wife of Cronos and mother of Zeus. Rheic Ocean originated between
Gondwana and Avalonia (Fig. 6), later between Gondwana and Laurussia, during Early
Paleozoic times (Nance et al. 2010 and references therein). It was opened during Late
Plate Tectonic Evolution of the Southern Margin of Laurussia in the Paleozoic 267

Cambrian – Early Ordovician times, around 500 Ma. System of back-arc basins developed
along the northern branch of Rheic Ocean after formation of Laurussia during Devonian
times (von Raumer & Stampfli, 2008). These basins are considered either as a part of the
Rheic Ocean (e.g. Nance et al. 2010, Golonka, 2007b, McKerrow et al., 1991) or as a separate
entity known as Rheno-Hercynian or Moldanubian basin in Central Europe (e.g. Golonka et
al. 2006b, Golonka 2002, Schulmann et al., 2009). Rheic Ocean narrowed during Devonian
and was closed during Carboniferous times as a result of collision of Laurussia with
Gondwana.

4. Global Early Paleozoic plate tectonic evolution leading to the assembly


of Laurussia
Gondwana supercontinent was located around the South Pole at the beginning of Paleozoic
(Figs 1-2). Baltica and Laurentia were located at the high latitude in the southern
hemisphere, their southern margins close to the South Pole. They drifted apart from
Gondwana during the Late Vendian times (Torsvik et al., 1996, Golonka et al, 2002, 2009b,
2012). Their breakup led to the formation of new oceans, including Iapetus (Figs. 1-2).
Continued seafloor spreading occurred in this ocean during Cambrian times (Figs. 2-4). The
fragmentation of northern margin of Gondwana was also marked by magmatic activity in
the 550-500 Ma (Dörr et al., 1998, Turniak et al., 2000, Tichomirowa, 2002, Burda and Klötzli,
2011). Laurentia drifted rapidly northward and rotated counter-clockwise, reaching low
latitudes (Golonka, 2000, 2002). Seafloor spreading also occurred within the Pleionic Ocean
between East Siberia and Baltica. The relationship between Laurentia and Siberia remains
quite speculative. Latest Cambrian – earliest Ordovician was the time of maximum
dispersion of continents during the Paleozoic. Baltica, Laurentia and Siberia drifted further
northward (Fig. 5). The subduction along the central margin of Gondwana caused the onset
of rifting of the Avalonian terranes (Golonka et al., 1994, Golonka, 2000, 2002). The
subduction along the northern margin of Baltica was perhaps related to the Ordovician
rotation of this plate (see Golonka et al., 1994, 2006b, Torsvik et al., 1996, McKerrow et al.,
1991, Golonka, 2000, 2002, cocks & Torsvik, 2011). The distance between Gondwana and
Laurentia, which was situated on equator reached 5000 km (Golonka, 2002, Golonka et al.
2006).

Early-Middle Ordovician were the times of a major plate reorganization (Golonka, 2000,
2002, 2009b, 2012, Golonka et al. 2006b). Avalonia probably started to drift from Gondwana
and move northward toward Baltica and Laurentia (Golonka, 2000, 2002, 2009b, 2012). This
movement was related to the origin of Rheic Ocean. (Fig. 6). The Iapetus Ocean had begun
to narrow.

During Late Ordovician times (Fig. 7 the Rheic Ocean between Gondwana and Avalonia
widened significantly (Golonka, 2000, 2002, 2009b, 2012). The position of Cadomia remains
uncertain. On the presented reconstruction, the Cadomian blocks are positioned relatively
close to Gondwana. This is mainly based on the paleobiogeographical data (Scotese and
McKerrow, 1990, Robardet et al., 1993, Golonka, 2002, 2009b, 2012). The alternative
268 Tectonics – Recent Advances

Figure 7. Plate tectonic map of Late Ordovician (plates position as of 452 Ma). Mollweide projection.
Modified from Golonka (2012). 1 - oceanic spreading center and transform faults, 2 - subduction zone,
3 - thrust fault, 4 - normal fault, 5 - transform fault.

reconstructions assumed that these terranes were rifted away and formed separate Cado-
mia plate floating within the Rheic Ocean (Lewandowski, 1993, see also Golonka et al.,
2006a). Latest Ordovician- Early Silurian were the times of collision between Avalonia and
Baltica (Fig. 8).

Figure 8. Plate tectonic map of Early Silurian (plates position as of 435 Ma). Mollweide projection.
Modified from Golonka (2012). 1 - oceanic spreading center and transform faults, 2 - subduction zone,
3 - thrust fault, 4 - normal fault, 5 - transform fault.
Plate Tectonic Evolution of the Southern Margin of Laurussia in the Paleozoic 269

This convergence was dominated by a strike-slip suturing of the two continents, rather than
by full-scale continent-continent collision (Golonka, 2000, 2002). Northwestern Poland and
adjacent part of Germany was joined with Baltica, along a strike-slip fault zone known as
the Tornquist-Teisseyre line forming new continent - Balonia. Perhaps the Brunovistulicum
and Malopolska terranes of southern Poland also belonged to Avalonia and joined Baltica
(Moczydłowska, 1997, Bełka et al., 2000, Golonka, 2002, 2009b). In the Central Western
Carpathians Late Ordovician – Early Silurian tonalitic gneisses of calc-alkaline character,
associated with meta-gabbros, revealed the presence of magmatic episodes at 470-435 Ma
(Kohut et al. 2008, Janák et al. 2002, Gaab et al. 2003, Gawęda & Golonka, 2011). These rocks
and granites present results of docking of Avalonia to Baltica. Also in the East Carpathians
in Romania the intrusion of 459-470 Ma granitoids (Munteanu & Tatu, 2003, Pana et al.,
2002, Ballintoni et al., 2010) document the collision-related tectono-magmatic effects of
docking of eastern prolongation of Avalonia to Baltica. The Scandian Orogeny was the
result of the collision between Balonia and Laurentia. the onset of the orogeny occurred
during the Early Silurian times and by late Silurian the orogeny was concluded (Golonka et
al., 1994, Golonka, 2000, 2002, 2009b). The main phase of the Scandian orogeny is marked by
nappes in Norway and Greenland as well as large crustal thickening (Dewey & Burke 1973,
Torsvik et al., 1996, Golonka, 2000, 2002, 2009b). During the Mid-Silurian Avalonia collided
with Laurentia (Fig. 9). This collision is known as the Caledonian orogeny, named after
Roman name of Scotland – Caledonia. This name is also extending into other related
orogenic events in Scanidinavia, Greenland and Ventral Europe. After the complete closure
of the Iapetus Ocean, the continents of Baltic, Avalonia, and Laurentia formed the continent
of Laurussia (Ziegler, 1989).

Figure 9. Plate tectonic map of Late Silurian (plates position as of 425 Ma). Mollweide projection. Modi-
fied from Golonka (2012). 1 - oceanic spreading center and transform faults, 2 - subduction zone, 3 -
thrust fault, 4 - normal fault, 5 - transform fault.
270 Tectonics – Recent Advances

5. Laurussia during Devonian times


Supercontinent Laurussia existed during Late Silurian and Devonian times. Late Silurian
Global paleogeography depicted on Fig. 9 id showing separation of Gondwana and Laurus-
sia as well as subduction along the southern margin of Laurussia. Fig 10 depicts the global
paleogeography during Early Devonian times. It is showing a possibility of Early Devonian
collision between South and North America according to Golonka (2002, 2007b, 2012 see
also Keppie 1989, McKerrow et al. 1991, Dalziel et al. 1994, Keppie et al. 1996). This collision
was marked by orogenic events in Venezuela, Columbia, Peru, and northern Argentina
(Gallagher & Tauvers 1992, Williams 1995). Paleomagnetic data (Kent & Van der Voo 1990,
Van der Voo 1993, Lewandowski 1998, 2003) and paleobiogeography (Young 1990) also
support the hypothesis about the proximity of South and North America (Golonka, 2007,
20012 b). This proximity is also shown by recent maps of Cocks & Torsvik (2010), however
without the collision. The Carolina region (Rast & Skehan 1993) also contains an element of
collision and transpression between South and North America, acting as an indenter, with a
dextral strike-slip component (Golonka, 2012).

Figure 10. Paleogeographic c map of Early Devonian (plates position as of 401 Ma). Mollweide projec-
tion. Modified from Golonka 2007b, 2012) Ophiolites – 1 Lizard ophiolite, 2 – central and eastern Euro-
pean ophiolites (depicted in details on Fig. 11).
Plate Tectonic Evolution of the Southern Margin of Laurussia in the Paleozoic 271

The accretion of Avalonia was followed by the development of north dipping subduction,
along the new continental margin of Laurussia. That subduction caused rifting and creat-
ing the new back-arc basin. That basin was first recognized in central Europe as Rheno-
Hercynian zone (Ziegler, 1989, Franke, 1992, Franke et al., 1995, Golonka, 2007b and refer-
ences therein). Its paleogeography resembled present-day marginal Seas of East Asia.
Fragment of Rheno-Hercynian zone, with oceanic crust, was also named Lizard-Giessen
Ocean (Zeh & Gerdes, 2010). The name Lizard is derived from the well known Lizard
Ophiolitic Complex in Cornwall, United Kingdom (Bromley, 1975, 1976, Kirby, 1976),
marked as number 1 on Fig.10. This Devonian complex contain peridotites, serpentinites
dolerite dikes as well as amphibolites (Bromley, 1975, 1976, Kirby, 1976, Cook et al., 2002,
Clark et al., 2003). Probably El Castillo volcanic rocks from Spain are related to the Lizard
ophiolites (Gutiérez- Alonso et al., 2008). They yield Devonian, 394,7±1,4 Ma age of mag-
matism.

Figure 11. Paleogeography of Laurussia margin in central-eastern Europe during Early Devonian).
Mollweide projection. Legend as on Fig. 10. Abbreviations: T – L - Tepla Barrandien – Lugia terranes ,
Carp. – Central Carpathian and Balkan terranes. Ophiolites: 1- Central-Sudetic ophiolite, 2 - Western
Carpathian ophiolite, 3 - Balkan-South Carpathian ophiolite.

The eastern extension of the Rheno-Hercynian Basin is marked by ophiolites in Sudety area
in Poland and in the Carpathian-Balkan area (Figs. 10, 11). During Devonian times the
Central-Sudetic ophiolites were located between Avalonian terrenes, sutured to Laurussia,
272 Tectonics – Recent Advances

and several microplates/terranes now included in the mosaic structure of Bohemian Massif
and Sudety Mountains. These terranes are marked on the Fig. 11 as T – L – Tepla-Barrandien
– Lugia. They include several small blocks like Tepla-Barrandien, Góry Sowie and others,
with uncertain, speculative position between Cadomia- Saxoturungia and Laurussia (see.
e.g. Aleksandrowski et. al., 2000, Franke & Żelaźniewicz, 2000, Winchester et al., 2002,
Mazur, et al., 2006, Schulmann et al., 2009, Kryza & Pin, 2010, Nanece et al., 2010).
According to Kryza & Pin (2000) the SHRIMP zircon data supplied evidence for Devonian
age of the Central-Sudetic ophiolites, around 400 Ma.

The age of the ophiolitic remnants within the Carpathians from Tatric and Gemeric units
Unit yield various ages: 371 Ma - 385-383 Ma (Putiš et al. 2009), 391 Ma (Gawęda, 2008), and
394 Ma (Kohut et al., 2006). The recorded events are related to the development of a back-arc
basin with ocean crust as a result of the slab roll-back and derivation of ribbon-like Proto-
Carpathian Terrane from Laurussia (see Gawęda & Golonka, 2011). The associated granitoid
magmatism post-dated that event (387 Ma; Burda & Klötzli 2011, Putiš et al., 2008). The
Balkan-South Carpathian ophiolite belt yield the isotopic age of 406-399 Ma (Zakariadze et
al, 2007), comparable with the Central-Sudetic ophiolites age being around 400 Ma (Kryza,
Pin, 2010).

Gondwana drifted northward and rotated clockwise during Devonian times (Scotese and
McKerrow, 1990, Scotese and Barret, 1990, Golonka, 2000, 2002, 2007b). At the same time,
Laurussia was rotating clockwise (Torsvik et al., 1996) at a somewhat faster rate. Figure 12
depict global paleogeography during the Late Devonian times.

Figure 12. Paleogeographic c map of Late Devonian (plates position as of 370 Ma). Mollweide projec-
tion. Legend as on Fig. 10. Modified from Golonka 2007b, 2012) .Ophiolites – 1 Lizard ophiolite, 2 – cen-
tral and eastern European ophiolites (depicted in details on Fig.13).
Plate Tectonic Evolution of the Southern Margin of Laurussia in the Paleozoic 273

The first contact between Laurussia and the Cadomian promontory of Gondwana occurred
in Central Europe This contact marks the onset of Hercynian orogeny. The Saxoturingian
part of the Cadomian plate collided with small terranes in Germany, Czech Republic and
Poland (Fig. 13). Collisional events were marked also in Carpathians and Balkans. The
Rheno-Hercynian Basin changed its character from extensional into compressional one. By
latest Devonian – Early Carboniferous it displayed synorogenic features, filling with
turbiditic flysch and culm facies (Golonka, 2007b).

Figure 13. Paleogeography of Laurussia margin in central-eastern Europe during Late Devonian).
Mollweide projection. Legend as on Fig. 10. Abbreviations: T – L - Tepla Barrandien – Lugia terranes ,
Carp. – Central Carpathian and Balkan terranes. Ophiolites: 1- Central-Sudetic ophiolite, 2 - Western
Carpathian ophiolite, 3 - Balkan-South Carpathian ophiolite, 4 – Lizard ophiolite.

The new subduction zone on the southern margin Laurussia and the formation of a back-arc
resulted in the intensive partial melting and magmatic activity from 370 to 340 Ma (Burda &
Gawęda 2009, Schulmann et al. 2009). The continuous subduction along both old and new
subduction zones resulted in docking of the Cadomia- Saxoturingia Terrane to Laurussia
and significant consumption of the Rheic Ocean (Fig. 10). Finally the subduction caused
formation of the Variscan Orogenic Suture, extending from Turkey to Mexico. That event
resulted in voluminous granitoid magmatism and amphibolite-granulite facies metamor-
phism – the most prominent features of all crystalline cores, present in all the Variscan com-
plexes (Stipska et al., 1998; Schulmann et al. 2009). These Variscan granitoid magmas were
274 Tectonics – Recent Advances

formed and intruded in the interval of 370-340 Ma, contemporaneously with Variscan
nappes formation (Dallmayer et al., 1996; Burda et al., 2011) and synchronous with uplift of
the continental blocks during prolonged collision (Janak et al., 1999, Gawęda et al., 2000).
The resulting collision was possibly associated with basaltic underplating or slab break-off
(Broska & Uher, 2001; Finger et al., 2009). Most of the granitoid bodies show hybrid charac-
ter, with both mantle and crustal components involved (e.g. Słaby & Martin, 2005, Burda et
al. 2011). Their VAG and CAG affinities suggest that melted metasediments, representing
crustal component of the magma, were originally deposited both in the volcanic arc and
intracontinental basins during subduction at the active, Andean-type continental margin
(Schulmann et al., 2009). The melted material represented mainly recycled Proterozoic com-
ponents, with addition of Paleozoic ones and influence of DM (depleted mantle) component
as a source of heat (Poller et al., 2000). The subsequent uplift caused the exhumation of eclo-
gitic remnants (Janak et al., 1996) and caused their retrograde metamorphism in granulite –
amphibolite facies regime.

Figure 14. Plate tectonic map of Early Carboniferous (late Visean – Serpukhovian), plates position as of
328 Ma). Mollweide projection. Modified from Golonka (2012). 1 – oceanic spreading center and trans-
form faults, 2 – subduction zone, 3 – thrust fault, 4 – normal fault, 5 – transform fault.

The ongoing Hercynian convergence in Europe led to large scale dextral shortening,
overthrusting and emplacement of parts of the accretionary complexes (Edel & Weber,
1995). The amount of convergence was modified by large, dextral and sinistral transfer
faults. The thrusting took place in the Tatra Mts. in the Carpathians (Gawęda et al., 2000,
Golonka, 2000, 2002, 2007b). The collision between Gondwana and Laurussia continued to
develop during Carboniferous times (Figs. 14, 15). The intercontinental collision began to
affect the northwestern part of Africa, developing Mauretinides, Bassarides, Rokelides
orogens is. The Alleghenian orogeny in North America continued (Hatcher et al., 1989, Rast
& Skehan, 1993), prograding westwards to the Ouachita fold belt in Arkansas, Oklahoma,
Plate Tectonic Evolution of the Southern Margin of Laurussia in the Paleozoic 275

Texas and adjacent part of Mexico (Arbenz, 1990, Golonka, 2000, 2002). The clockwise
rotation of Gondwana resulted in the involvement of the deformation. This Gondwanian
influence resulted in the convoluted shape of the Hercynian orogen, strike-slip zones
(Franke et al., 1995) and Hercynian deformation at the eastern end in Poland. The European
foreland basin was elevated or changed its sedimentation regime from flysch to molasse.
The central Pangean mountain range was formed, which extended from Mexico to Poland
(Golonka, 2000, 2002, 2007b). The Laurussia continent ceased its independent existence
becoming a part of the supercontinent Pangea.

Figure 15. Plate tectonic map of Late Carboniferous (plates position as of 302 Ma). Mollweide projec-
tion. Modified from Golonka (2012). 1 – oceanic spreading center and transform faults, 2 – subduction
zone, 3 – thrust fault, 4 – normal fault, 5 – transform fault.

6. Conclusions
1. The Late Precambrian (Vendian) to present plate tectonic processes contributed to the
complex structure of Western and Central Europe.
2. The supercontinent of Laurussia, originated as a result of a closure of Iapetus Ocean
and collision of Baltica, Avalonia and Laurentia.
3. Laurussia originated during Late Silurian times
4. The accretion of Avalonia was followed by the development of north dipping sub-
duction, forming the new Rheno-Hercynian Basin back-arc basin during Devonian
times
5. The oceanic crust of the Rheno-Hercynian Basin is recognised by ophiolites in Western,
Central and Eastern Europe. The Lizard ophiolite in U.K. and Central-Sudetic ophiolite
in Poland represent the best developed Devonian oceanic complexes.
6. The new geological research (including dating) in the Carpathian-Balkan area supports
the idea about the prolongation of the Rheic Suture to the east.
276 Tectonics – Recent Advances

7. The new geological research (including dating) in the Carpathian-Balkan area supports
the idea about the prolongation of the Rheic Suture to the east.
8. The tectonic events and associated magmatism point out the development of the back-
arc at the southern margin of Laurussia in the time interval 406-371 Ma.

Laurussia cease to exist during Hercynian orogeny in Carboniferous times and was included
into supercontinent Pangea.

Author details
Jan Golonka
AGH University of Science and Technology, Kraków, Poland

Aleksandra Gawęda
Faculty of Earth Sciences, University of Silesia, Sosnowiec, Poland

Acknowledgement
This research has been partially financially supported by Ministry of Science and Higher
Education grants (Statutory Activities) through the AGH University of Science and
Technology in Krakow grants no. 11.11.140.447 (J. Golonka) and Ministry of Education and
Science grant No N307 027837 (A. Gawęda).

7. References
Aleksandrowski, P., Kryza, R., Mazu,r S., Pin, C. & Zalasiewicz, J.A,. (2000). The Polish
Sudetes: Caledonian orVariscan? Transactions of the Royal Society, Edinburgh, Earth
Sciences (1999), Vol. 90, 127–146.
Arbenz, J. K., 1990, The Ouachita system, The Geology of North America, Bally, A. W. &
Palmer, A. R., (Eds.,) Vol. . A., 371-396, Boulder, Geological Society of America,
Ballintoni, I., Balica, C., Seghedi, A. & Ducea, M.N. (2010). Avalonian and Cadomian
terranes in North Dobrogea, Romania. Precambrian Research, Vol. 182, No. 3, 217-229.
Bełka, Z., Ahrendt, H., Franke, W. & Wemmer, K. (2000). The Baltica-Gondwana suture in
central Europe: Evidence from K-Ar ages of detrital muscovites and biogeographical
data, Geological Society, London, Special Publications, Vol. 179, 87-102.
Besutiu L. (2001). Moesia – a Baltica derived terrane? Pancardi 2001 II. Abstracts, pp. CP29.
Ádám, A., Szarka, L., Szendröi, J. (Eds.) Hungarian Academy of Sciences, Sopron,
Humgary.
Bromley, A.V. (1975). Is the Lizard Complex, South Cornwall, a fragment of Hercynian
oceanic crust?. The Lizard: A Magazine of Field Studies, Vol., No. 3, 2-11.
Bromley, A.V. (1976). A new interpretation of the Lizard Complex, S. Cornwall, in the light
of the ocean crust model. Proceedings. Geological. Society London, Vol. 132, 114.
Broska, I, & Uher, P. (2001). Whole-rock chemistry and genetic typology of the West-
Carpathian Variscan granites. Geologica Carpathica, Vol. 52 no. 2, 79-90.
Plate Tectonic Evolution of the Southern Margin of Laurussia in the Paleozoic 277

Burda, J. & Gawęda, A. (2009). Shear-influenced partial melting in the Western Tatra
metamorphic complex: geochemistry and geochronology. Lithos, Vol. 110, . 373-385.
Clark, A. H., Sandeman, H. A. I., Nutman, A.P., Green, D.H & Cook, A. C.. (2003).
Discussion on SHRIMP U–Pb zircon dating of the exhumation of the Lizard Peridotite
and its emplacement over crustal rocks: constraints for tectonic models. Journal of the
Geological Society, London, Vol. 160, 331–335.
Cocks, L.R.M. & Torsvik T.H. (2011). The Palaeozoic geography of Laurentia and western
Laurussia: a stable craton with mobile margins. Earth Science Reviews, Vol. 106, 1-51.
Cook, C. A., Holdsworth, R.E. & Styles, M.T. (2002). The emplacement of peridotites and
associated oceanic rocks from the Lizard Complex, southwest England, Geological
Magazine, Vol. 139, 27-45.
Dallmeyer, D., Neubauer, F., Handler, R., Fritz, H., Muller, W., Pana, D., & Putis, M. (1996).
Tectonothermal evolution of the internal Alps and Carpathians: evidence from 40Ar-39Ar
mineral and whole-rock data. In: Schmid S.M., Frey M., Froitzheim N., Heilbronner R.,
Stuenitz, H. (eds.): Alpine geology, proceedings of the second workshop: 2nd workshop
on Alpine geology, Eclogae Geologica Helvetica, Vol. 89, pp. 203-227.
Dalziel, I.W.D., Dalla Salda, L.H & Gahagan, L.M. (1994). Paleozoic Laurentia-Gondwana
interaction and the origin of the Appalachian-Andean mountain system, Geological
Society of America Bulletin, Vol. 106, 243-252.
Dewey, J. F. & Burke, V. B. S., (1973). Tibetan, Variscan, and Precambrian basement
reactivation, products of continental collision, Journal of Geology, Vol. 81, 683-692.
Dörr, W., Fiala, J., Vejnar, Z., & Zulauf, D. (1998). U-Pb zircon ages and structural
development of metagranitoids of the Tepla crystalline complex: evidence for pervasive
Cambrian plutonism within the Bohemian Massif (Czech Republiv). International Journal
of Earth Sciences, Vol. 87, 135-149.
Edel, J. B., Weber, K., 1995. Cadomian terranes, wrench faulting and thrusting in the central
Europe Variscides: geophysical and geological evidence, Geologische Rundschau, Vol. 84,
412-432.
Finger, F., Gerdes, A., Rene, M., & Riegler G. (2009). The Saxo-Danubian Granite Belt:
magmatic response topost-collisional delamination of mantle lithosphere below the
south-west sector of the Bohemian Massif (Variscan Orogen). Geologica Carpathica, Vol.
60, No. 3, 205-212.
Ford, D. & Golonka J. (2003). Phanerozoic paleogeography, paleoenvironment and
lithofacies maps of the circum-Atlantic margins, Thematic set on paleogeographic
reconstruction and hydrocarbon basins: Atlantic, Caribbean, South America, Middle East,
Russian Far East, Arctic. Golonka J. (ed.), Marine and Petroleum Geology, Vol. 20, 249-285.
Franke, W. (1992), Phanerozoic structures and events in central Europe, The European
Geotraverse, A continent revealed, Blundell, D., Freeman, R. & Mueller, S. (Eds.), 164-180,
University of Cambridge, Cambridge.
Franke, W. & Żelaźniewicz, A.,. (2000).The eastern termination of the Variscides: terrane
correlation and kinematic evolution, Orogenic processes:quantification and modelling in the
Variscan belt, Geological Society Special. Publication, Vol. 179, 63–86.
278 Tectonics – Recent Advances

Franke, W., Dallmeyer, R. D. & Weber, K., 1995, Geodynamic Evolution, Pre-Permian geology
of Central and Eastern Europe, IGCP 233 international conference: Gottingen, Federal Republic
of Germany: Berlin, Dallmeyer, R. D.., Franke, W. & Weber, K., (Eds.), 579-593, Springer-
Verlag.
Gaab, A., Poller, U., Todt, W. & Janák M. (2003). Geochemical and isotopic characteristics of
the Murán Gneiss Complex, Veporic Unit (Slovakia), Journal of the Czech Geological
Society, Vol., 48, No. 1-252.
Gallagher, J. J. & Tauver,s P. R. (1992). Tectonic evolution of northwestern South America,
Basement tectonics. Mason, R. (Ed), 123-137.Kluver Academic Publishers..
Gawęda, A. (2008). An apatite-rich enclave in the High Tatra granite (Western Carpathians):
petrological and geochronological study. Geologica Carpathica, Vol. 59, No. 4, 295-306.
Gawęda, A., Golonka J., (2011). Variscan plate dynamics in the Circum-Carpathian area.
Travaux Géophysiques XL (2011), Abstracts of the 9th Central European Tectonic
Groups meeting, Hotel Skalský Dvůr, Czech Republic, 13 - 17. April 2011, 19. Prague.
Institute of Geophysics Academy of Sciences of the Czech Republic,.
Gawęda, A., Kozłowski, K., & Piotrowska, K. (2000). Early-Variscan collision and generation
of leucogranite melts in the Western Tatra Mountains (S-Poland, W-Carpathians).
Journal of the Czech Geological Society, Vol. 45, 230.
Golonka, J. (2000). Cambrian-Neogene Plate Tectonic Maps. 1-125, Wydawnictwa Uniwersytetu
Jagiellońskiego, Kraków.
Golonka, J. (2002). Plate-tectonic maps of the Phanerozoic. Phanerozoic reef pattern, Kiessling
W., Flügel E. & Golonka J. (eds.), SEPM (Society for Sedimentary Geology) Special
Publication, Vol. 72, 21-75.
Golonka, J. (2004). Plate tectonic evolution of the southern margin of Eurasia in the Mesozoic
and Cenozoic, Tectonophysics, Vol. 381, 235-273.
Golonka, J. (2007a). Late Triassic and Early Jurassic paleogeography of the world.
Palaeogeography, Palaeoclimatology, Palaeoecology, Vol. 244, 297-307.
Golonka, J. (2007b). Phanerozoic Paleoenvironment and Paleolithofacies Maps. Late
Paleozoic. Mapy paleośrodowiska i paleolitofacje fanerozoiku. Późny paleozoik.
Kwartalnik AGH. Geologia, Vol .33, No. 2,: 145-209.
Golonka, J. (2007c). Phanerozoic Paleoenvironment and Paleolithofacies Maps. Mesozoic.
Mapy paleośrodowiska i paleolitofacje fanerozoiku. Późny mezozoik. Kwartalnik AGH.
Geologia, Vol. 33, No. 2, 211-264.
Golonka, J. (2009a). Phanerozoic Paleoenvironment and Paleolithofacies Maps. Cenozoic.
Mapy paleośrodowiska i paleolitofacje fanerozoiku. Kenozoik. Kwartalnik AGH.
Geologia, , Vol. 35 No. 4, 507-587.
Golonka, J. (2009b). Phanerozoic Paleoenvironment and Paleolithofacies Maps. Early
Paleozoic. Mapy paleośrodowiska i paleolitofacje fanerozoiku. Wczesny paleozoik.
Kwartalnik AGH. Geologia, , Vol. 35, No. 4, 589-654.
Golonka, J. (2012). Paleozoic paleoenvironment and paleolithofacies maps of Gondwana,
Wydawnictwa AGH Publishing House. Kraków. In press
Plate Tectonic Evolution of the Southern Margin of Laurussia in the Paleozoic 279

Golonka, J., Ross M. I. & Scotese C. R., 1994. Phanerozoic paleogeographic and paleoclimatic
modeling maps. Pangea: Global environment and resources, Embry A. F., Beauchamp, B. &
Glass, D. J., (Eds.),. Canadian Society of Petroleum Geologists Memoir, 17, 1-47.
Golonka, J., Krobicki, M.,Oszczypko, N., Ślączka, A. & Słomka, T. (2003). Geodynamic
evolution and palaeogeography of the Polish Carpathians and adjacent areas during
Neo-Cimmerian and preceding events (latest Triassic - earliest Cretaceous), Tracing
tectonic deformation using the sedimentary record, McCan, T. & Saintot, A. (Eds.),Geological
Society Special Publications. 208, 138-158.
Golonka J. Gahagan L., Krobicki M., Marko F., Oszczypko N. & Slaczka A. 2006a. Plate
Tectonic Evolution and Paleogeography of the Circum-Carpathian Region. In: Golonka,
J. & Picha, F. (Eds.) The Carpathians and their foreland: Geology and hydrocarbon
resources. American Association of Petroleum Geologists Memoir, 84, 11-46.
Golonka, J., Krobicki, M., Pająk, J, Nguyen Van Giang & Zuchiewicz, W. (2006b). Global plate
tectonics and paleogeography of Southeast Asia. Faculty of Geology, Geophysics and
Environmental Protection, AGH University of Science and Technology, Arkadia,
Kraków.
Gutiérrez-Alonso G., Murphy J.B., Fernández-Suárez J. & Hamilton M.A. (2008). Rifting
along the northern Gondwana margin and the evolution of the Rheic Ocean: A
Devonian age for the El Castillo volcanic rocks (Salamanca, Central Iberian Zone).
Tectonophysics, Vol. 461, 157-165.
Hartz E.H., & Torsvik T.H. (2002). Baltica upside down: a new plate tectonic model for
Rodinia and Iapetus Ocean. Geology, Vol. 30, No. 3, 255-258.
Hatcher R. D., Jr, Thomas. W. A., Geiser, P. A., Snoke, A. W., Mosher S. & Wiltschko D. V.,
(1989). Alleghenian orogen, The Appalachian-Ouachita Orogen in the United States,
Hatcher R. D., Jr, Thomas, W. A. & Viele, G. W., (Eds.),V. F, . 233-318, Boulder,
Geological Society of America, The Geology of North America,
Janák, M., Hurai, V., Ludhova, L., O`Brien, P.J., & Horn, E.E. (1999). Dehydration melting
and devolatilization during exhumation of high-grade metapelites: the Tatra
Mountains, Western Carpathians. Journal of Metamorphic Geology, Vol. 17, 379-395.
Janák, M., Finger, F., Plašienka, D., Petrik, I., Humer, B., Meres, S. & Luptak, B., (2002).
Variscan high P-T recrystallization of Ordovician granitoids in Veporic Unit (Nizke
Tatry Mountains, Western Carpathians): new petrological and geochronological data.
Geolines, Vol. 14, 38-39.
Kalvoda, J. & Bábek O. (2010) .The Margins of Laurussia in Central and Southeast Europe
and Southwest Asia. Gondwana Research, Vol. 17, 526-545.
Kamo S.L., Gowar C.F., & Krogh T.E. (1989). Birthdate for the Iapetus Ocean ? A precise U-
Pb zircon and baddeleyite age for Long Range dikes, southeast Labrador. Geology, Vol.
17, No. 7, 602-605.
Kent, D. V. & Van der Voo, R., (1990). Palaeozoic palaeogeography from palaeomagnetism
of the Atlantic-bordering continents, Palaeozoic palaeogeography and biogeography,
McKerrow, W. S. & Scotese, C. R., (Eds.), Geological Society Memoir, Vol. 12, 49-56.
Keppie, J.D. (1985). The Appalachian collage, The Caledonide orogen, Scandinavia, and
related areas, 1217-1226, Gee, D.G., & Sturt, B. (Eds)., New York, J. Wiley and Sons.
280 Tectonics – Recent Advances

Keppie, J. D. (1989). Northern Appalachian terranes and their accretionary history, Terranes
in the Circum-Atlantic Paleozoic orogens. Dallmeyer, R. D. (Ed), Geological Society of
America, Special Paper, Vol. 230, 159-192.
Keppie, J. D., Dostal J., Murphy, J. B., & Nance, R. D. (1996). Terrane transfer between
eastern Laurentia and western Gondwana in the Early Paleozoic: Constraints on global
reconstructions, Avalonia and Related Peri-Gondwanan Terranes of the Circum-North
Atlantic, Nance R. D. & Thompson M. D. (Eds), Geological Society of America Special Paper,
Vol. 304, 369-380.
Kirby, G. A. (1979). The Lizard Complex as an ophiolite. Nature, Vol. 282, 58-61.
Kohut M., Konecny P. & Siman P. (2006) The first finding of the iron Lahn-Dill
mineralization in the Tatric Unit of the Western Carpathians. Mineralogia Polonica –
Special Papers, Vol.28, 112-114.
Kohut, M., Poller, U., Gurk, Ch. & Todt W. (2008). Geochemistry and U-Pb detrital zircon
ages of metasedimentary rocks of the Lower Unit, Western Tatra Mountains (Slovakia),
Acta Geologica Polonica, Vol. 58, 371-384.
Kryza, R. & Pin C. (2010). The Central Sudetic ophiolites (SW Poland): petrogenetic issues,
geochronology and paleotectonic implications. Gondwana Research, Vol. 17, 292-305.
Lewandowski, M. (1998). Assembly of Pangea: Combined Paleomagnetic and Paleoclimatic
Approach, Circum-Arctic Palaeozoic Faunas and Facies, Ginter, M. & Wilson, M. H. (Eds),
Ichtyolith Issues Special Publication, Vol. 4, 29-32.
Lewandowski, M. (2003). Assembly of Pangea: Combined Paleomagnetic and Paleoclimatic
Approach. Advances in Geophysics, Vol. 46, 199-236.
Mazur, S., Aleksandrowski, P., Kryza ,R. & Oberc-Dziedzic, T,.( 2006). The Variscan Orogen
in Poland, Geological Quarterly, Vol. 50 89–118.
McKerrow, W. S. , Dewey, J. F. & Scotese, C. R. (1991). The Ordovician and Silurian
development of the Iapetus Ocean The Murchison symposium; proceedings of an
international conference on the Silurian System in Bassett, M. G., Lane, P. D. & Edwards, D.
(Eds.), Special Papers Palaeontology, Vol. 44, 165-178.
Moczydłowska, M. (1997). Proterozoic and Cambrian successions in Upper Silesia: an
Avalonian terrane in southern Poland, Geological Magazine, Vol. 134, 679-689.
Munteanu M. & Tatu M. (2003.) The East-Carpathian Crystalline-Mesozoic Zone (Romania):
Paleozoic amalgamation of Gondwana- and East European Craton- derived terranes.
Gondwana Research, Vol. 6, No. 2, 185-196.
Nance, R.D., Gutiérrez-Alonso, G., Keppie, J.D., Linnemann, U., Murphy, J.B., Quesada, C.,
Strachan, R.A. & Woodcock, N.H., (2010). Evolution of the Rheic Ocean. Gondwana
Research, Vol. 17, 194-222.
Pana, D., Ballintoni, I., Heaman, L. & Creaser R. (2002). The U-Pb and Sm-Nd dating of the
main lithotectonic assemblages of the East Carpathians, Romania. Geologica Carpathica,
Vol. 53, 177-180.
Putiš, M, Ivan, P, Kohút, M, Spišiak, J, Siman, P, Radvanec, M, Uher, P, Sergeev, S,
Larionov, A, Méres, Š, Demko, R. & Ondrejka M. (2009). Meta-igneous rocks of the
West-Carpathian basement, Slovakia: indicators of Early Paleozoic extension and
shortening events. Bulletin Societe Géolologique France, Vol. 180, No.6, 461-471
Plate Tectonic Evolution of the Southern Margin of Laurussia in the Paleozoic 281

Rast, N. & Skehan, J. W., 1993. Mid-Paleozoic orogenesis in the North Atlantic, the Acadian
orogeny. The Acadian Orogeny, recent studies in New England, Maritime Canada, and the
autochtohonous foreland. , Roy C. & Skehan J. W. (Eds), Geological Society of America Special
Paper, Vol. 275, 1-25.
Robardet, M., Blaise, J., Bouyx, E., Gourvennec, R., Lardeux, H., Le Hérissé, A., Le Menn, J.,
Melou, M., Paris, F., Plusquellec, Y., Poncet J., Régnault, S., Rioult, M. & Weyant M.
(1993). Paléogeographie de l'Europe occidentale de l'Ordovicien au Dévonien;
Paleogeography of Western Europe from the Ordovician to the Devonian, Bulletin
Societe géologique de France, Vol. 164, 683-695.
Schulmann, K., Konopásek, J., Janoušek, V., Lex, O., Lardeaux, J.-M., Ede,l J.-B., Štípská & P.,
Ulrich S. (2009). An Andean type Palaeozoic convergence in the Bohemian Massif.
Comptes Rendus – Geoscience, Vol. 341, 266-286.
Scotese, C. R. & McKerrow, W. S. (1990). Revised world maps and introduction, Palaeozoic
palaeogeography and biogeography, McKerrow W. S. & Scotese C. R. (Eds), Geological
Society of London Memoir, 12, 1-21.
Scotese, C.R. & Barret, S.F. (1990). Gondwana's movement over the South Pole during the
Paleozoic: evidence from lithologic indicators of climate. In: W.S. McKerrow and C.R.
Scotese (Eds.) Paleozoic Paleogeography and Biogeography, , Geological Society of London,
Memoir Vol. 12, pp. 75-85.
Sears, J.W. (2012). Transforming Siberia along the Laurussian margin. Geology, doi:
10.1130/G32952.1
Słaby, E., & Martin, H. (2005). Mafic and felsic magma interaction in granites: the
karkonosze Hercynian pluton (Sudetes, Bohemian Massif). Journal of Petrology, Vol. 49,
353-391.
Štipská, P., Schulmann, K., Kroener, A. (1998). From Cambro-Ordovician rifting to Variscan
collision at the NE margin of the Bohemian Massif: petrological, geochronological and
structural constraints. Paleozoic orogenesis and crustal evolution of the European lithosphere –
post-conference excursion, pp. 24-31.
Tichomirova, M., (2002). Zircon inheritance in diatexite granodiorites and its consequence
on geochronology – a case study in Lusatia and Erzgebirge (Saxo-Thuringia, eastern
Germany). Chemical Geology, Vol. 191, 209-224.
Torsvik, T. H. , Smethurst, M. A., Meert, J. G., Van der Voo, R., McKerrow, W. S., Brasier, M.
D. & Sturt, B. A., Walderhaug, H. J. (1996). Continental break-up and collision in the
Neoproterozoic and Palaeozoic; a tale of Baltica and Laurentia, Earth-Science Reviews,
Vol. 40, No.3-4, 229-258.
Turniak, K., Mazur, S., & Wysoczański, R. (2000). SHRIMP zircon geochronology and
geochemistry of tyhe Orlica snieznik gneisses (Variscan Belt of Central Europe) and
their tectonic implications. Geodinamica Acta, Vol. 13, 293-312.
Veevers, J.J. (2004). Gondwanaland from 650-500 Ma assembly through 320 Ma merger in
Pangea to 185-100 Ma breakup: Supercontinental tectonics via stratigraphy and
radiometric dating. Earth-Science Reviews, Vol. 68, 1-132.
282 Tectonics – Recent Advances

von Raumer, J.F., & Stampfli, G.M. (2008). The birth of the Rheic Ocean – Early Paleozoic
subsidence patterns and subsequent tectonic plate scenario, Tectonophysics, Vol. 461, 9-
20.
Williams, K. E., (1995). Tectonic Subsidence Analysis and Paleozoic Paleogeography of
Gondwana, Petroleum basins of South America, Tankard, A.J., Suarez, S. & Welsink, H. J.
(Eds),American Association of Petroleum Geologists Memoir, Vol. 62, 79-100.
Winchester J.A., Floyd P.A., Crowley Q.G., Piasecki M.A.J., Lee M.K., Pharaoh T.C.,
Williamson P., Banka D., Verniers J., Samuelsson J., Bayer U., Marotta A.-M., Lamarche
J., Franke W., Dörr W., Valverde-Vaquero P., Giese U., Vecoli M., Thybo H., Laigle M.,
Scheck M., Maluski H., Marheine D., Noble S.R., Paarish R.R., Evans J., Timmerman H.,
Gerdes A., Guterch A., Grad M., Cwojdzinski S., Cymerman Z., Kozdroj W., Kryza R.,
Alexandrowsk, P., Mazur S., Stedrá V., Kotková J., Belka Z., Patoćka F. & Kachlik
V.,(2002). Palaeozoic amalgamation of Central Europe: New results from recent
geological and geophysical investigations. Tectonophysics, Vol. 360, 5-21.
Young, G. C., (1990). Devonian vertebrate distribution patterns and cladistic analysis of
paleogeographic hypothesis, Palaeozoic palaeogeography and biogeography. McKerrow W.
S. & Scotese C. R. (Eds), Geological Society of London Memoir, Vol. 12, 243-255.
Zakariadze, G. S.; Karamata, S. O. & Dilek, Y. (2007). Significance of E. Paleozoic Paleo-
Tethyan Ophiolites in the Balkan Terrane and the Greater Caucasus for the Cadomian-
Hercynian Continental Growth of Southern Europe . American Geophysical Union, Fall
Meeting 2007, Abstract #V43B, 1367, American Geophysical Union, Washigton D.C.
Zeh, A. & Gerdes, A. (2010). Baltica- and Gondwana-derived sediments in the Mid-German
Crystalline Rise (Central Europe): Implications for the closure of the Rheic ocean,
Gondwana Research, Vol. 17, 254-263).
Ziegler, P.A., 1989. Evolution of Laurussia. Kluwer Academic Publishers, Dordrecht.
Zonenshain, L. P., Kuzmin, M. L. & Natapov, L. N., 1990. Geology of the USSR: A Plate-
Tectonic Synthesis. Page, B. M. (Ed), Geodynamics Series, American Geophysical Union,
Vol. 21, 1-242.
Żelaźniewicz, A., Seghedi, A., Żaba, J., Fanning, M. &, Buła Z. (2009) More evidence on
Neoproterozoic terranes in Southern Poland and southeastern Romania, Geological
Quarterly, Vol. 51, 93-124.
Chapter 11

Was the Precambrian Basement of Western


Troms and Lofoten-Vesterålen in Northern
Norway Linked to the Lewisian of Scotland?
A Comparison of Crustal Components, Tectonic
Evolution and Amalgamation History

Steffen G. Bergh, Fernando Corfu, Per Inge Myhre,


Kåre Kullerud, Paul E.B. Armitage, Klaas B. Zwaan, Erling K. Ravna,
Robert E. Holdsworth and Anupam Chattopadhya

Additional information is available at the end of the chapter

http://dx.doi.org/10.5772/48257

1. Introduction
Temporal and spatial linkage of Archaean and Palaeoproterozoic crustal provinces in the
North Atlantic realm requires a well‐established geological and geodynamic framework.
Such a framework is well established for the Fennoscandian Shield of Finland, Sweden and
northwestern Russia [1‐4], for Greenland/Laurentia [5, 6] and for the Lewisian of NW
Scotland [7, 8], but not yet for the Precambrian crystalline rocks within and west of the
Scandinavian Caledonides in North Norway (Figure 1).

In western Troms (Figure 2) and in the Lofoten‐Vesterålen areas of North Norway (Figure 3)
Neoarchaean and Palaeoproterozoic continental crust (2.9‐1.67 Ga) is preserved as an
emerged basement horst bounded to the east by thrust nappes of the Scandinavian
Caledonides (Figure 1b) [9‐11] and to the west by offshore Mesozoic basins [12]. These
basement outliers are believed to be part of the Archean‐Palaeoproterozoic Fennoscandian
Shield [1, 11, 13] that stretches from NW Russia, through Finland and Sweden (Figure 1a).
Similarly, a pronounced magmatic suite in the Lofoten area [9, 10] corresponds in age (1.80‐
1.78 Ga) and structural position with the NNW‐trending Transscandinavian igneous belt of
Sweden [14, 15]. In spite of the internal position relative to the Caledonides, and in great
contrast to the basement inliers in southern Norway where Caledonian high‐grade
metamorphic reworking is widespread, the geotransect in western Troms and Lofoten

© 2012 Bergh et al., licensee InTech. This is an open access chapter distributed under the terms of the
Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
284 Tectonics – Recent Advances

displays only modest Caledonian reworking and, thus, provides a reliable framework for
regional correlation of Neoarchaean and Palaeoproterozoic crust [9,10, 16,17].

Figure 1. (a) Location of Precambrian basement outliers in western Troms (WTBC), Lofoten‐Vesterålen
(LOF) and NW Scotland (LW) and related basement provinces within today’s plate setting of the North
Atlantic Ocean. (b) Geologic map of Fennoscandia with location of the West Troms Basement Complex
and the Lofoten‐Vesterålen province west of the Scandinavian Caledonides [16].

In NW Scotland, the Lewisian Complex is situated to the west of the Caledonian Moine
thrust and covers the Outer Hebrides, the NW Scottish mainland and part of the Inner
Hebrides (Figures 1a and 4). The Lewisian rocks also form inliers in the Caledonian orogenic
belt, possibly continuing under the Moine Group up to the Great Glenn fault in the
southeast, while Mesozoic to early Cenozoic extensional basins offshore Scotland largely
disrupted the continuity of the Lewisian outcrops [18, 19, 20].

A possible linkage of the western Troms and Lofoten‐Vesterålen basement rocks with the
Lewisian basement inliers of the Caledonides in NW Scotland (Figure 1a) and with Lau‐
rentia‐Greenland has been raised [3, 4, 8, 21, 22], but excact correlation of these cratonic‐
marginal provinces in the North Atlantic realm, their role during assembly of Fennoscandia
and Laurentia in the Neoarchaean [23, 24], and the situation prior to Palaeoproterozoic
orogenies [3, 25, 26], still remain enigmatic.

This paper reviews the current knowledge of the crustal components, tectono‐magmatic
evolution and amalgamation history of the basement rocks in western Troms and Lofoten‐
Vesterålen, North Norway, and compares them with the Lewisian of Scotland (Figures 2‐4,
Table 1). New and focused structural and geochronological work in the West Troms
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 285

Basement Complex [16, 17, 27‐32] has sparked off new interest in these provinces. Questions
specifically raised for these basement suites concern the age and nature of supracrustal
units, and the character of crustal‐scale ductile shear zones, either as potential terrane
boundaries between assembled older crustal blocks or just reflecting episodes of basin
formation and later reworking. Such boundaries can, in general, help to restore the outline
and correlation of each craton and the cratonic margin characteristics and to unravel cycles
of tectono‐magmatic events [33].

Figure 2. Regional map of the West Troms Basement Complex, North Norway, that shows the main
crustal components and tectonic features, with a generalized cross‐section. Frame shows location of
Figure 10a. The map is revised after [17, 34].
286 Tectonics – Recent Advances

Figure 3. Geological map of the Lofoten‐Vesterålen province, showing the Neoarchaean‐


Palaeoproterozoic basement rocks and the anorthosite‐mangerite‐charnochite‐granite igneous suite [10,
35, 36]
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 287

Figure 4. Simplified geologic and structural map of the Lewisian Complex in northwest Scotland show‐
ing the main rock units and the overall subdivision of the mainland into multiple regions (or terranes)
separated by main Palaeoproterozoic ductile shear zones. The map is modified from from [20, 37].
Abbreviations: LSZ=Laxford shear zone, CSZ=Canisp shear zone, GaSz=Gairloch shear zone,
DSZ=Diabaig shear zone. Frame shows outline of figure 18.
288 Tectonics – Recent Advances

West Troms Basement Complex Lofoten‐Vesterålen Province Lewisian Complex


Age (Ga) Components and events Age (Ga) Components and events Age (Ga) Components and
events
2.92‐2.8 Ga Neoarchaean cratonization: Neoarchaean 3.145‐2.75 Neoarchaean
‐Tonalite crystallization cratonization: Ga cratonization:
(Dåfjord & Kvalsund gneisses) 2.85‐2.7 Accretion, convergence 2.9‐2.7 ‐Various types of TTG‐
Volcanism and sedimentation: Ga and crustal thickening Ga gneisses, protholiths of
‐ Ringvassøya greenstone belt tonalite (Scourian
2.85‐2.83 Ga Continued Neoarchaean Gneiss), granites
cratonization ‐Tonalite magmatism (Laxfordian Gneiss),
‐Mafic plutonism (Bakkejord 2.75‐2.68 Regional high‐grade
2.75‐2.6 Ga diorite) in the southwest Ga metamorphism, crustal
Neoarchaean deformation and Neoarchaean accretion (?) and
2.75‐2.7 Ga metamorphism: 2.72‐2.66 deformation and thickening
‐Magmatism, migmatization Ga metamorphism: Volcanism and
(Gråtind migmatite ) and ‐ Various orthogneisses 2.8 Ga? sedimentation: ‐
ductile shearing (in Dåfjord (e.g. Bremnes gneiss) Mafic/ultramafic
2.7‐2.67 Ga and Kvalsund gneisses) formed by crustal 2.7‐2.6 volcanics and
‐ Main gneiss foliation shortening Ga supracrustal rocks
(initially flat‐lying), ductile (Eanruig paragneiss,
shear zones, tight folds and ‐ Emplacement of Claisfearn supracrustals).
dip‐slip stretching lineation. 2.64 Ga tonalities, followed by Neoarchaean
‐ Medium/high‐grade high‐grade deformation and
metamorphism, ENE‐WSW metamorphism and metamorphism:
crustal contraction and localized migmatization ‐ Subhorizontal
thickening by accretion and/or and ductile crustal foliation & tight folding
underplating shearing (Sigerfjord and thrusting, regional
migmatite, Ryggedalen granulite facies
granulite) metamorphism. NE‐SW
shortening
2.69‐2.56 Ga ‐High‐grade metamorphism
and resetting ‐Open macro‐folds and
2.49‐2.4 axial‐planar dextral
Ga transpressive shear
zones (Canisp and
Laxford? shear zones).
‐ Reworking and
retrogression to
amphibolite facies
2.40 Ga Crustal extension and ? 2.4‐2.0 Crustal extension and
intrusion of the Ringvassøya Ga intrusion of the Scourie
mafic dyke swarm mafic dyke swarm and
Na‐rich pegmatites.
Dextral transtensional
setting.
2.4‐2.2 Ga Deposition of Vanna group ? Deposition of 2.2‐1.9 Deposition of the Loch
clastic sediments in a marine supracrustal units, Ga Maree Group clastic and
subsiding basin heterogeneous mafic volcanic succession in
gneisses, banded iron marine extensional
formations, quartzite, and/or arc‐settings
marbles, graphite ‐Marine mudstones
schists (Flowerdale schists,
Aundrary amphibolites)
2.22 Ga Intrusion of Vanna diorite sill ? ?
2.2‐1.9 Ga Deposition of Mjelde‐
Skorelvvatn, Torsnes and 1.87‐1.86 Precursory stage 1.9‐1.8 Intrusion of early‐stage
possibly, the Astridal Ga magmatism, AMCG‐ Ga granites
supracrustal belts suite, Lødingen granite
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 289

West Troms Basement Complex Lofoten‐Vesterålen Province Lewisian Complex


1.993 Ga Intrusion/volcanism in the 1.8‐1.79 Main stage intrusion of Kalk‐alkaline
Mjelde‐Skorelvvatn belt Ga AMCG‐suite plutonic 1.9‐1.87 volcanism: Ard kalk‐
1.80 Ga Magmatism/intrusion of rocks and the Tysfjord Ga alkaline gneisses
granites and norite in Senja Granite Magmatism: South
1.79 Ga Magmatism/intrusion of 1.77 Ga Intrusion of granite Harris Igneous Complex
Ersfjord Granite in Kvaløya pegmatite dykes (Hebrides)
Intrusion of late‐stage
granites
1.7‐1.68
Ga
c. 1.9‐1.7 Ga Palaeoproterozoic Palaeoproterozoic 1.97‐1.67 Palaeoproterozoic
deformation: deformation Ga deformation:
1.9‐1.8 Ga? Early: ‐Mylonitic foliation Early stage ‐Deep
(initially flat‐lying), NW‐SE 1.87‐1.79 Strong ductile 1.8‐1.85 thrusting (Gairloch
trending gently‐plunging Ga deformation and high‐ Ga? shear zone) and
isoclinal folds, NE‐directed grade (granulite facies) amphibolite facies
ductile thrusts with dip‐slip metamorphism and metamorphism,
stretching lineation. Prograde reworking accretion of the Loch
medium/high‐grade I.83 Ga Major ductile shear Maree Group onto the
metamorphism in the zone (suture)? Arc‐ continental crust, in
southwest. NE‐SW orthogonal related and/or subduction or arc‐
shortening, NE‐directed collisional setting setting.
thrusting/accretion Mid/main stage ‐
1.78‐1.76 Retrogressive Isoclinal folding and
1.78‐1.77Ga Mid: ‐Regional open upright Ga metamorphism of 1.85 Ga thrusting, upright
folding, NW‐SE trend, flat‐ AMCG‐suite rocks macro‐folding and
lying hinges and steep limbs. transpressive shear
Medium/low grade zones (Canisp, Laxford
metamorphism. and Shieldaig shear
Continued NE‐SW orthogonal zones), Accretionary
crustal shortening tectonic event
Late: ‐Regional steep N‐
c. 1.75 Ga? plunging folds, NW‐SE Main/late stage ‐
striking, steep ductile shear Partitioned
zones (strike‐slip). 1.70‐1.67 deformation, thrust and
‐Retrogressive low grade Ga steep dextral‐oblique
metamorphism. strike‐slip shear zones
on a flat detachment
Latest: ‐NE‐SW trending (on Laxford and Canisp
1.7‐1.67 Ga? upright folds of the Vanna shear zones),
group and SE‐directed thrusts, Amphibolite facies
steep semi‐ductile strike‐slip metamorphism, likely
shear zones. Retrogressive low collisional event
grade metamorphism
Late‐stage NW‐SE
Partitioned NE‐SW shortening
steep transpressive
and orogen‐parallel (NW‐SE)
shear zones,
strike‐slip shearing
retrogression to
greenschist facies,
Intrusion of felsic pegmatites
crustal rejuvenatiion.
and retrogression
1.57 Ga 1.6 Ga? Crustal uplift,
retrograde
metamorphism, cooling

Table 1. Summary and comparison of tectono‐magmatic components and events in the West Troms
Basement Complex, the Lofoten‐ Vesterålen area and the Lewisian of Scotland. The data is based on
references listed and discussed in the text.
290 Tectonics – Recent Advances

2. Geological features of western Troms and Lofoten


Archaean and Palaeoproterozoic crust underlies much of the northeastern part of the Fen‐
noscandian Shield [1, 3], including also basement outliers west of the much younger, Palae‐
ozoic Scandinavian Caledonides (Figure 1). Here, the West Troms Basement Complex [17]
and basement in Lofoten‐Vesterålen [9, 10] emerge as a c. 300 km long horst, separated from
the Caledonian nappes by Mesozoic rift‐normal faults [11, 12]. The West Troms Basement
Complex (Figure 2) is composed of various Mesoarchaean to Palaeoproterozoic plutonic
rocks and orthogneisses (2.9‐1.7 Ga), metasupracrustal rocks, mafic dyke swarms, and net‐
works of ductile shear zones [17, 38]. The basement of Lofoten and Vesterålen (Figure 3)
consists of similar metamorphic Neoarchaean rocks intruded by a very extensive suite of
1.80‐1.78 Ga plutonic rocks of the anorthosite‐mangerite‐charnockite‐granite (AMCG) suite
[9, 10] that appears to link up with the Transscandinavian Igneous Belt of southern Sweden
(Figure 1a). Both areas display dominant NW‐SE structural trends parallel with Archaean
and Palaeoproterozoic orogenic belts of the Fennoscandian Shield that stretch from Russia,
through Finland and northern Sweden into the Bothnian basin of central Scandinavia [1].

2.1. The West Troms Basement Complex


The West Troms Basement Complex (Figure 2) is underlain by Meso to Neoarchaean
gneisses, various generations of metasupracrustal rocks and mafic dyke swarms that were
later intruded by felsic and mafic plutonic suites and variably reworked, deformed and
metamorphosed during the main Palaeoproterozoic (Svecofennian) orogeny [17].

2.1.1. Archaean crust


The Meso‐Neoarchaean rocks of the West Troms Basement Complex (Figure 2) consist of
tonalite‐trondhjemite and anorthositic gneisses with mafic and ultramafic layers and banded
intercalations (Figure 5a) and are overlain by the Neoarchaean Ringvassøya greenstone belt.
These rocks were deformed and metamorphosed up to granulite/migmatite facies prior to
deposition of Palaeoproterozoic cover units and intrusion of a 2.4 Ga mafic dyke swarm [17,
39]. A steep NW‐SE trending transposed gneiss foliation with dip‐slip stretching lineations
(Figure 5b) and tight ENE‐vergent intrafolial folds (Figure 5c) attests for WSW‐ENE
contraction and thrusting during the Meso/Neoarchaean [17]. Prominent high‐grade
migmatite zones interpreted as a ductile shear zone (Figure 5d) separate compositionally
different gneisses [17], e.g. the Kvalsund migmatite zone separating the Dåfjord and
Kvalsund gneisses on Ringvassøya [36] and similar zones within the Senja Shear Belt [38,
40]. Polyphase refolding and thrusting is common and suggests protracted Neoarchaean
deformation [17]. In Ringvassøya, tonalitic orthogneisses and granitoids (Dåfjord gneiss)
reveal U‐Pb zircon crystallization ages of 2.92‐2.77 Ga (Figure. 6) [36, 41], and these rocks are
considered to be related to tonalites on the island of Vanna farther north, where a U‐Pb
crystallization age of 2885 ± 20 Ma has been obtained [33]. This Mesoarchaean basement was
also intruded by the 2695 ±15 Ma Mikkelvik alkaline stock [42]. The overlying Ringvassøya
Greenstone belt [43] comprises arc‐related meta‐volcanic rocks with MORB‐transitional,
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 291

tholeiitic to calk‐alkaline affinity. Two meta‐volcanic rocks of the Ringvassøy greenstone


belt yield ages of c. 2.85 Ga [44].

Figure 5. Outcrop features of Meso‐Neoarchaean tonalitic gneisses in the West Troms Basement Com‐
plex: (a) Foliated tonalitic and mafic gneisses (Dåfjord gneiss) in central part of Ringvassøya. (b) Steep
SW‐dipping foliation in tonalitic Dåfjord gneiss with dip‐slip stretching lineation. (c) Banded tonalitic
gneiss with tightly folded mafic inclusions, cut by granitic pegmatite veins presumed to be related to
the 1.79 Ga Ersfjord granite. (d) Zone of major migmatized Kvalsund gneiss in southwestern part of
Ringvassøya. The zone is cut by a mafic dyke, which is part of the Ringvassøya dyke swarm dated at 2.4
Ga [39].

By contrast, all the dated metaplutonic rocks on the islands of Kvaløya and Senja farther
south (Figure 2) are Neoarchaean. The Bakkejord pluton, neosome in the Kattfjord gneiss,
and granodiorite units bordering the Torsnes belt on Kvaløya, as well as several major in‐
trusive bodies on Senja all yield ages between 2.72 and 2.68 Ga [16, 32]. A somewhat young‐
er element at 2.67 Ga is shown by mafic dykes that cut the Bakkejord pluton on Kvaløya
[32]. The only potentially younger Archaean event is the formation of migmatites in south‐
ern Senja, where zircon in a neosome suggests crystallization at c. 2.6 Ga [32]. The Kvalsund
migmatite zone in southwestern Ringvassøya (Figure 2) appears to represent a boundary
separating the Mesoarchean crust to the north from Neoarchean crust to the south. The time
of deformation is not yet dated, but dynamic melting structures in the migmatite are cut by
mafic dykes considered to belong to the 2.4 Ga swarm, hence indicating an Archean age of
shearing. One sample of neosome has a primary age of about 2.7 Ga indicated by zircons,
which, however, also records an event ≤ 2.55 Ga possibly reflecting the time of deformation
[32].
292 Tectonics – Recent Advances

Figure 6. Summary of the main geochronological features in the Lewisian Complex (a) and the West‐
Troms Basement Complex and Loften province (b). The age compilation is based principally on U‐Pb
data. The references are listed and discussed in the text.

2.1.2. Ringvassøya mafic dyke swarm


The Neoarchaean gneisses and the oldest metasupracrustal belts of the West Troms Base‐
ment Complex have been intruded by a huge mafic, plagioclase phyric and gabbronoritic
dyke swarm (Figure 7), the Ringvassøya dykes [39]. These dykes are widely distributed and
display offset, shearing and reworking, thus providing a good time‐marker for resolving the
subsequent Svecofennian deformation [17, 27]. Zircon and baddeleyite from a dyke swarm
on Ringvassøya provides a crystallization age of 2403 ± 3 Ma, and the dykes are classified as
transitional between MORB and within‐plate basalts with an affinity to continental tholeiites
[39].
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 293

Figure 7. Outcrop features of the 2.4 Ga mafic dyke swarm that intruded Neoarchaean massive to‐
nalites in Ringvassøya [39]. Note irregular and varied dyke orientations (a, b) and that the dykes trun‐
cate the main tonalitic gneiss fabric (c).

2.1.3. Palaeoproterozoic supracrustal rocks


The metasedimentary Vanna group represents unconformable continental deposits [29, 45].
Common rock types include layered meta‐psammites locally exhibiting pronounced cross‐
bedding (Figure 8a, b). The age of deposition is constrained between 2403 Ma, the age of the
underlying Ringvassøya dykes, and 2221 ± 3 Ma, the age of a diorite sill in the supracrustal
rocks [29, 39].

Figure 8. Outcrop features of meta‐supracrustal rocks in the West Troms Basement Complex: (a) Layered
meta‐psammites with interbedded mudstones/mica‐schists of the Vanna group [29, 45]. (b) Subvertical
thinly bedded meta‐sandstone of the Vanna group, with pronounced trough and planar cross‐bedding in
internal lenses. Up is to the right. (c) Basal meta‐conglomerate of the Torsnes supracrustal belt. Note that
the dominant clast‐type is tonalite, tonalitic gneiss and granitoid gneisses (d) Rythmical laminated quartz‐
feldspatic meta‐psammite with enrichment of iron‐hydroxide staining from the Astridal belt, Senja. The
beds are steeply dipping and subvertically folded. The fold hinge is located near person.
294 Tectonics – Recent Advances

Arc‐related meta‐volcanic rocks with MORB‐transitional, tholeiitic to calk‐alkaline affinity


occur in the central and southwestern parts, e.g. in the Astridal and Torsnes belts [17, 27,
31]. The Mjelde‐Skorelvvatn belt (Figure 2) is dominated by metabasaltic rocks together with
ultramafic rocks, meta‐psammites, marble and calc‐silicate gneisses. A differentiated gab‐
broic portion of the meta‐basaltic pile yields an age of 1992 ± 2 Ma [31]. The Torsnes belt
comprises a basal conglomerate (Figure 8c) overlain by meta‐psammite and a thick sequence
of mafic metavolcanic rocks. Detrital zircons indicate a maximum‐age of deposition of 1970
±14 Ma [31]. The Astridal belt is compositionally similar to the Mjelde‐Skorelvvatn belt, but
there are no direct radiometric dates yet. It contains abundant mica‐schists, local graphite
schists, mafic meta‐volcanic rocks and widespread sulphide ore deposits (Figure 8d).

2.1.4. Late palaeoproterozoic igneous suites


The Neoarchaean crust in Kvaløya and Senja was intruded by an extensive suite of felsic
and mafic plutonic rocks [17]. The most prominent are the Ersfjord granite (Figure 9a, b) on
Kvaløya [46] and large granites and mafic plutons (Hamn norite) on Senja (Figure 2). The
Ersfjord granite has a U‐Pb zircon crystallization age of 1792 ± 5 Ma [16] and the Hamn
norite 1802.3 ± 0.7 Ma [28], whereas the granitic masses farther south in Senja give Rb‐Sr [47]
and zircon‐titanite ages of 1805 ± 2.5 Ma [16]. Metamorphic overprints of the Ersfjord granite
are recorded by U‐Pb titanite ages of 1769 ± 3 Ma and 1756 ± 3 Ma [16]. Pronounced and
widespread granite pegmatite dykes (Figure 9b, c) formed syn‐tectonically with shear zones
in the metasupracrustal belts at c. 1768 ± 4 Ma, probably genetically related to the main
intrusive activity [16]. All these ages are within the interval when most known Precambrian
juvenile crust generated by arc‐related magmatism [48].

Figure 9. (a) Aerial view of the Ersfjord granite with its rugged mountains and presence of both steep
and gently dipping planar fabrics. (b) Ersfjord granite pegmatite dykes and veins cutting Neoarchaean
tonalitic gneiss foliation in southwestern ringvassøya, and later boudinaged during Palaeoproterozoic
tectonism. (c) Ersfjord granite pegmatite dykes cutting dioritic gneisses near its western boundary
against the Kattfjord gneiss.

2.1.5. Palaeoproterozoic deformation and metamorphism


Strong deformation and metamorphism at 1.8‐1.76 Ga produced mega‐blocks or segments
delineated by NW‐SE trending, variously mylonitized metasuprascrustal belts and lens‐
shaped ductile shear zones as outlined in Figure 10 [17, 38]. This shear belt deformation was
characterized by high‐strain, complex and multiphase deformation and up to amphibolite
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 295

facies metamorphism and reworking. The most distinct one, the Senja Shear Belt is c. 30 km
wide (Figures 2 and 10) and is delimited by the Svanfjellet belt to the south [38, 49] and the
Torsnes belt in the north [17]. This linear crustal structure is thought to continue beneath the
Caledonian nappes into parallelism with the Bothnian‐Senja shear zone of the Swedish part
of the Fennoscandian shield, but gravity and magnetic anomaly patterns do not uniquely
confirm such a correlation [11, 13, 50].

Figure 10. Tectonic map of the Senja Shear Belt in central part of the West Troms Basement Complex
(for location see figure 2), illustrating the lens‐shaped architecture of the metasupracrustal belts. Note
macro‐scale, steep‐plunging folds of the belt and adjacent tonalitic gneisses, where fold hinges are bent
into parallelism with the trace of the Astridal belt. Note also the major shear zone boundaries with
thrust and sinistral strike‐slip characters. The map is from [17].

The Palaeoproterozoic deformation structures of the West Troms Basement Complex (Table
1) include a main NNW‐SSE striking, mylonitic foliation mostly present in the meta‐
supracrustal belts (Figure 11a), that formed axial‐planar to early‐stage isoclinal folds (Figure
11b) with gently‐plunging axes, at amphibolite facies conditions. The foliation has a steep
296 Tectonics – Recent Advances

WSW dip and exhibits a dip‐slip, west‐plunging stretching lineation. Macroscopic NW‐SE‐
trending and mostly upright antiform‐synform folds (Figure 11c) are widespread, causing
the steep tilt and apparent repetition of most of the supracrustal belts in synformal troughs
[17]. Corresponding upright folds also occur in the adjacent gneisses. Younger (Late Palaeo‐
proterozoic) superimposed structures include tight to vertical folds on all scales (Figure 11d,
e) with axial‐planar cleavages coeval with an anastomosing network of steeply‐dipping,
NW‐SE to N‐S trending sinistral and dextral strike‐slip shear zones (Figure 10). The latter
zones are mylonitic and retrogressed into greenschist facies. These shear zones caused sub‐
vertical drag‐folding of the surrounding gneisses and became boudinaged and masked by
quartz precipitates along the main foliation. Later on the foliation was folded by steeply
north‐plunging shear folds and cut by oblique‐slip crenulation cleavages, sigmoidal clasts
and multiple shear bands, all supportive of strike‐slip displacements (Figure 11d, e). The
youngest set of structures occurs in the northeastern parts of the West Troms Basement
Complex, as gently‐dipping, SE‐directed phyllonitic shear zones (thrusts) with abundant SE‐
verging folds and thrusts [17].

Figure 11. Outcrop illustrations of Palaeoproterozoic deformation structures in the West Troms Base‐
ment Complex: (a) Subvertical high‐strain mylonites along the eastern contact between the Torsnes belt
and the granitoid Kattfjord gneiss (right). (b) Foliation‐parallel felsic vein in garnet‐mica‐schists of the
Astridal bel that is isoclinally folded and sinistrally sheared, causing multiple repeatitions of the vein.
View is to the NE on near‐horizontal surface. (c) Upright asymmetrical NE‐verging fold that refolds
isoclinal folds in meta‐volcanic and siliciclastic rocks of the Svanfjellet belt, Senja. (d) Subvertical phyl‐
lonitic shear fabric with quartz precipitates along the main foliation.Note dextral subvertical folding of
the main fabric and quartz veins. (e) Ersfjord granite pegmatite dyke in mafic Kattfjord gneiss northeast
of the Torsnes belt that is folded by subvertical sinistral folds.

The early and middle stages of deformation were associated with prograde metamorphism
varying from low grade in the northeast to amphibolite and granulite facies in the central
and southern parts and terminated with late stage retrogressive greenschist facies reworking
[27]. A migmatitic shear zone in meta‐psammites southwest of Ringvassøya displays a gar‐
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 297

net granulite facies assemblage succeeded by a two‐pyroxene granulite facies assemblage


and is dated with the zircon U‐Pb method to 1777 ± 12 Ma, the same age obtained for a gra‐
nitic dike (1776.6 ± 1.1 Ma) cutting the lens [51]. Other critical radiometric ages for Palaeo‐
proterozoic high‐strain deformation include metamorphic overprints of Ersfjord granite
pegmatite dykes at c. 1.77‐1.76 Ga, and an interval of 1774, 1768 ± 4 Ma and 1751 ± 8 Ma for a
granitoid pegmatite dyke formed syn‐kinematically with late‐stage strike slip shear zones in
the Astridal belt [16].

2.1.6. Mesoproterozoic reactivation


Late‐tectonically deformed granitic pegmatites in the Astridal belt of the Senja Shear Belt
yield U‐Pb ages of 1725 ±22 Ma and 1562 ± 2 Ma, indicating that formation of the pegmatites
occurred after termination of the main orogenic events (Corfu et al. in review). These
occurrences are attributed to ‘anorogenic’ far‐field effects reflecting intracratonic strain,
possibly caused by the emplacement of A‐type massifs in the core of the Fennoscandian
Shield.

2.2. The Lofoten‐Vesterålen province


This province comprises gneisses and major plutonic suites of Precambrian age [52] that
suffered major tectono‐magmatic events at 2.8 and 1.8 Ga [53]. The basement complex in
western parts of Lofoten and Vesterålen (Figure 3) comprises granulite facies rocks, whose
distribution also coincides with a major magnetic and gravity high caused by the presence
of dense rocks in the crust and an elevated Moho discontinuity [9, 11, 54, 55]. The latter is
the result of differential uplift and extension in the aftermath of the Caledonian orogeny and
the subsequent Late Palaeozoic and Mesozoic processes that led to the opening of the North
Atlantic. The eastern part of the region consists of various amphibolite facies gneisses, mig‐
matites, greenstone belts and granitic plutons. After the initial geochronological studies
more work followed using Pb‐Pb, Rb‐Sr and Sm‐Nd dating [9, 36, 56‐58]. The chronology of
the basement complex and the anorthosite‐mangerite‐charnockite‐granite suite (AMCG) has
now been refined by modern U‐Pb geochronology [10].

2.2.1. Neoarchaean basement gneisses


Neoarchaean crust occupies large parts of the islands of Langøya in Vesterålen and Hinnøya
farther east (Figure 3), and small remnants are also present at the southwestern tip of
Austvågøya [58]. The Neoarchaean rocks of Langøya consist mainly of high‐grade gneisses
interpreted to represent metasupracrustal rocks of intermediate composition, while Neoar‐
chaean gneisses on Hinnøya define an amphibolite facies metamorphic domain that was mig‐
matized in the Neoarchaean and subsequently deformed and metamorphosed at granulite‐
facies conditions in the Palaeoproterozoic [9]. The appearance of orthopyroxene to the west of
the amphibolite‐facies domain on Hinnøya has been interpreted as either a prograde meta‐
morphic transition [9, 54], or an abrupt transition marked by a crustal scale ductile shear zone
of presumed Neoarchaean age. The zone east of the metamorphic boundary (Figure 3) is dom‐
298 Tectonics – Recent Advances

inated by intrusive rocks of tonalitic to granitic composition, migmatitic domains, and local
greenstone belt remnants, considered to be Neoarchaean in age [9, 59].

2.2.2. Palaeoproterozoic supracrustal rocks, deformation and metamorphism


A younger sequence of heterogenous mafic gneisses in Lofoten has been interpreted as a
Palaeoproterozoic supracrustal succession of volcanogenic derivation as deduced from
geochemical compositions [9]. Metasedimentary rocks consist of fine‐grained gneisses,
locally quartzitic (Figure 12a), with subordinate graphite schist, banded iron formation
and marble. These gneisses have been overprinted by the same granulite‐facies metamor‐
phism as the Neoarchaean rocks, and the boundary between the two metamorphic do‐
mains corresponds to the eastern limit of the magnetic high and the first appearance of
orthopyroxene [54] in eastern Langøya (Figure 3). In the south this boundary is consid‐
ered a major thrust that juxtaposes the Eidsfjord anorthosite and deformed intrusive
mangerite [54, 60]. Some studies [9] suggest that the orthopyroxene isograde is folded and
continues southeastward across Hinnøya (Figure 3). The granulite facies gneisses tend to
be rather massive, with faint banding, and they equilibrated at 3 to 4 kb and 750 to 780 °C
[61]. They are considered to be orthogneisses [59] or migmatized supracrustal rocks of
intermediate composition [9].

Figure 12. Outcrop photographs of the AMCG suite and metasupracrustal rocks in western parts of
Lofoten (Figure 3)(a) Foliated paragneisses composed of alternating quartz‐rich and mafic meta‐
volcanic rocks. (b) Detail view of mangerite (hypersthene‐bearing monzonite) which is the dominant
rock type of the Lofoten igneous province. Note phenocrysts of plagioclase and orthoclase. (c) Man‐
gerites with mafic intercalations and dykes aligned parallel to a weak magmatic foliation. From a road
cut in eastern Lofoten (Austvågøya). (d) Panorama view of the Hopen pluton mangerite and its bound‐
ary to migmatized gneisses and altered mangerites. View is toward the north.
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 299

2.2.3. Palaeoproterozoic magmatic rocks, deformation and metamorphism


Plutons of the AMCG suite occupy about 50 % of the Lofoten islands [62]. The suite (Figure
3) is dominated by mangerite and charnockite and their metamorphosed equivalents (Figure
12b‐d), with local but important occurrences of gabbro, anorthosite, and granite and associ‐
ated mafic dykes (Figure 12c). The mafic and felsic phases locally grade into, or mutually
cross‐cut each other indicating that some of the intrusions are genetically related. In addi‐
tion, a widespread network of feldspathic veins is present throughout the granulite facies
domain [9]. Mangerite and charnockites are widespread throughout the region, in the form
of large plutons (Figure 12d) such as in the southwestern Lofoten islands [9, 62, 63]. Anor‐
thosite and gabbro forms plutons in most of the islands[54, 64, 65]. There is one distinctive
body of granite‐syenite on Langøya [54, 66]. To the east, on Hinnøya, Neoarchaean gneisses
are cut by Palaeoproterozoic plutons such as the Lødingen granite with a zircon and titanite
age of 1870 Ma [10, 36], whereas the Tysford granite covers a large area on the mainland and
has given U‐Pb ages between 1.8 and 1.7 Ga [67].

The mangerites and charnockites were shown to have intruded at about 4 kb and >925 to 850
°C, whereas the anorthosite on Flakstadøy records a polybaric crystallization history from 9
kb to about 4 kb at 1180 to 1120 °C. The data imply that anorthosites, mangerites and char‐
nockites in Lofoten were emplaced at the same depth of about 12 km [64]. Later on in the
same time span, the Palaeoproterozoic plutonic rocks and surrounding Neoarchaean
gneisses were deformed and metamorphosed up to granulite facies conditions, and portions
of the plutonic rocks, e.g. anorthosite bodies, were thrusted over mangeritic gneisses (Figure
13a, b) [54]. Structural fabrics include transposed foliations (Figure 13c), intrafolial isoclinal
folds (Figure 13d) and local migmatization structures, and upright folds that refolded the
earlier shear fabrics (Figure 13e). A younger, very extensive but narrow ductile shear zone
network is characterized by greenschist facie retrogressive shear zones that truncate all
other structures (Figure 13f) and display both low‐ and high‐angle attitudes. The exact age
of the latter is unknown, but presumably, late Palaeoproterozoic.

The structural and isotopic data show that the Neoarchaean crust played an important role
in the genesis and deformation of the Lofoten igneous rocks, and both as a source and as a
contaminant [58, 68]). The Pb data define a linear trend that may represent mixing between
Neoarchaean lower crustal components and late Palaeoproterozoic juvenile additions,
whereby the mangerites and charnockites contain more Neoarchaean Pb than the mafic
rocks. A multistage evolution with a basaltic parental melt undergoing polybaric
crystallization and differentiation to form anorthosites as cumulates and ferrodiorites as
residual melts has been proposed [69]. The mangerites and charnockites are inferred to
represent feldspar cumulates and residual liquids, respectively, derived from magmas
broadly syenitic in composition.

The more recent U‐Pb results show that the Lofoten AMCG suite was emplaced during two
quite distinct events (Figure 6), the first one at 1.87‐1.86 Ga followed by a second and domi‐
nant magmatic event at 1.80‐1.79 Ga. A concluding period, lasting some 20‐30 my, was char‐
acterized by local hydration of the dry AMCG rocks, and by the infiltration of pegmatite
300 Tectonics – Recent Advances

melts [10]. Local granitic pegmatites belong to a distinct Palaeoproterozoic (ca. 1.77 Ga)
generation.

Figure 13. Outcrop features illustrating Palaeoproterozoic deformation fabrics of the Lofoten‐
Vesterålen province. (a) View of an anorthosite complex in Langøya which is thrusted over granulite
facies mangeritic gneisses to the southeast. The ductile thrust zone is c. 50 m thick and made up of
mylonitic gneisses [54]. Note the lack of vegetation on the grey‐coloured Eidsfjord anorthosite above the
shear zone contact. View is toward NE. (b) Thrust in Palaeoproterozoic granitic gneisses on Langøya
[54]. (c) Transposed foliation in mangeritic gneiss with granitoid bands and intercalations. Locality:
Austvågøya. (d) Ductile shear zone in mangeritic gneiss with felsic intercalations. (e) Open upright fold
in mangeritic gneiss. The fold axis trends NW‐SE, and view is toward SE. (f) Steep and localized retro‐
gressed ductile shear zone in mangerite in southern part of Vestvågøya.

3. Geological features of the Lewisian Complex of Scotland


The Lewisian Complex in NW Scotland (Figure 1a) is situated to the west of the Caledonian
Moine thrust and covers the Outer Hebrides and the NW Scottish mainland of the Inner
Hebrides in the south (Figure 4). The Lewisian rocks also form inliers in the Caledonian
orogenic belt possibly continuing under the Moine Group up to the Great Glenn Fault in the
southeast [18‐20], while Mesozoic to early Cenozoic extensional basins bound the Lewisian
outcrops offshore Scotland (Figure 1a).
The Lewisian Complex was considered by earlier workers [70, 71] as a continuous crustal
block composed of up to three Neoarchaean gneiss domains overlain by Palaeoproterozoic
metasedimentary, metavolcanic and intrusive rocks separated by NW‐SE trending ductile
shear zones. In the last decade it has been proposed that the region consists of distinct ter‐
ranes [7, 72], although there are disparate views on how these terranes are related [8]. A
classical Lewisian nomenclature (Badcallian, Scourian, Inverian, Laxfordian) evolved pro‐
gressively, originally to designate specific rock forming, deformational or metamorphic
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 301

events, and eventually becoming linked to specific dates obtained from forthcoming geo‐
chronological studies. Some of these terms, however, have now become problematic since
they have been and can be used to designate different tectonic expressions and times. To
maintain clarity in the following we shall therefore avoid their use.

3.1. Main structures


The Neoarchean rocks of the Lewisian were deformed during an early granulite facies
tectonic event involving crustal shearing and isoclinal folding that formed a gently dipping
gneiss foliation (Figure 14a, b). During a later retrogressive event the gently dipping fabric
was macro‐folded into NE and SW‐dipping steep attitudes and subjected to steep strike‐slip
shearing (Figure 14 c, d) in amphibolite facies [73‐75]. A key observation is the local
truncation of these macrofolds by the c. 2.4 Ga Scourie dykes [76], as outlined later (Figure
15), indicating that these early deformation events are likely Archaean in age.

Figure 14. Outcrop features of Neoarchaean tonalitic gneisses in the Lewisian Complex. (a) Tight to
isoclinal intrafolial folds in the TTG gneisses with a presumed 2.7 Ga age, northwest of the Canisp shear
zone in Assynt terrane. (b) Cliff face made up of Neoarchaean TTG‐gneisses with a subhorizontal folia‐
tion. Height of the cliff is ca. 50 m. Locality: in Assynt terrane. (c) Steeply dipping, alternating banded
tonalitic, granitic and mafic orthogneisses. Locality: north tip of Rhiconich terrane. (d) TTG‐ gneiss with
an older foliation cut and transposed into a steep ductile shear zone of presumed 2.7 Ga age. Locality:
north of Canisp shear zone.
302 Tectonics – Recent Advances

Figure 15. The Scourie dyke swarm in the Assynt terrane (see Figure 4). (a) Steep, Scourie mafic dyke
that cuts through Neoarchaean gneiss foliation. Note the very sharp an unaffected intrusive contacts.
Dyke thickness is approximately 5 m. Locality, north of Canisp shear zone. (b) Scourie dyke intruded
into gently dipping TTG‐gneisses. (c) Scourie dyke cutting Neoarchaean gneisses and which is again
strongly sheared along steep, presumed 2.49 Ga Palaeoproterozoic shear zones. Locality, near contact to
Canisp shear zone.

Figure 16. (a) Regional structural map showing major ductile shear zones in the Lewisian Complex,
that suffered Palaeoproterozoic deformation and reworking [37]. (b) Map of the Laxford shear zone, with
Scourie dykes that cut presumed regional 2.49 Ga folds that are becoming tightened toward the Laxford
shear zone [37]. (c) Schematick NNE‐SSW cross‐section of the Canisp shear zone showing major upright
folding of the gneiss foliation and formation of localized, steep, axial‐planar shear zones [77].

The main episodes of Palaeoproterozoic crustal deformation in the Lewisian produced ma‐
jor folds and NW‐SE striking, dextral‐reverse, transpressive shear zones (Figure 16a) that
were superimposed on, and largely obliterated the pre‐Scourie dykes fabrics, except in some
low strain lenses. An early/main stage of deformation involved tight to isoclinal folding of
the flat‐lying Neoarchaean gneiss foliation and subsequent upright folding leaving the limbs
in a steep attitude (Figure 16b, 17a). In addition, localized moderately NE‐dipping ductile
reverse and dextral oblique shear zones developed by strain partitioning, likely due to reac‐
tivation of steep pre‐Scourie dyke shear zones [77], and they affected the Scourie dykes
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 303

(Figure 17b). This event was associated with upper amphibolite, locally granulite facies
metamorphism [37] and intrusion of syn‐tectonic pegmatite sheets and veins in e.g. the
Laxford shear zone (Figure 4). The major ductile shear zones show complex multiphase
strain partitioning, including thrusting and refolding of early‐stage subhorizontal shear
zones developed parallel to the pre‐existing foliation [78] and later development of steep
strike‐slip shear zones [77]. The late stage of deformation coincides with retrogression to
greenschist facies conditions [79] associated with the formation of steep‐plunging asymmet‐
ric folds and retrogressed cleavage and strike‐slip shear zones [37].

From northeast to southwest the main shear zones are the Laxford, Canisp, Gairloch and
Diabaig shear zones (Figure 16a). The ca. 8 km wide, SW‐dipping Laxford shear zone [37] is a
major zone of folding (Figure 16b) that reflects thrusting of the gneisses in the Assynt block
over the Rhiconich gneisses to the north. This presumed terrane‐bounding shear zone [80]
evolved from a pre‐existing, steep fabric that displayed early stages of sinistral and dextral
shearing and subsequent oblique‐thrusting and dextral strike‐slip movement [81], evidenced
by SSE plunging stretching lineations [37]. Numerous granites and pegmatite sheets were
injected within this shear zone.

Figure 17. Outcrop features illustrating Palaeoproterozoic (c. 1.9‐1.67 Ga) deformation fabrics in the
Lewisian Complex. (a) Meso‐scale upright folds within low‐strain domain of Canisp shear zone. Ham‐
mer is parallel to fold axis, ESE‐WNW, and the axial‐surface dips steeply to the ENE. A steep, SSW‐
dipping mylonitic shear zone developed on the fold limb (to the right). (b) Steep shear zones that cut
and displace a Scourie dyke and the Neoarchaean gneiss foliation near Canisp shear zone. (c) Upright
folding of TTG‐gneisses outside the Canisp shear zone (left) and refolding by tight sub‐vertical folds
within the shear zone (central and right). View is toward ESE. (d) Contact between TTG‐ gneiss and the
Canisp shear zone. Note subvertical sinistral drag‐folding of the gneisses into the mylonitic shear zone.
View is toward WNW. (e) High‐strain mylonite in the Canisp shear zone, with retrogressed chlorite‐
mylonitic schist and quartz veins aligned along the main fabric. View is toward the WNW. (f) Detail
from the steep mylonite zone in e, showing asymmetric sinistral folding of the main fabric, including
the quartz veins, seen on a horizontal view .
304 Tectonics – Recent Advances

Farther south, the Canisp and Gairloch shear zones define steep oblique crust‐internal shear
zones of the Assynth and Gruinard terranes (Figure 16a). The Canisp shear zone dips steeply
SSW [37] and typically truncates the Scourie dykes (Figure 17b) and displaces them into
zones of alternating high‐ and low‐strain [77]. Variably dipping high‐strain thrust zones
evolved in the hinge zone of a major ESE‐WNW trending fold (Figure 16c), while steep
high‐strain mylonitic shear zones truncate the limbs of the macro‐fold (Figure 17c, d). Strain
partitioning is observed at all scales, and low strain lenses are typically overprinted and
transposed into high strain zones. Folds in the high strain zones are asymmetric, tight and
have generally steep plunge (Figure 17c). These folds may have been produced by early
deformation which partitioned later into the higher strain domain as tighter and steeply
plunging folds (Figure 16b), leaving a weak signature in the low strain domain [82]. Fabric‐
parallel quartz veins are abundant in the high‐strain zones, and these veins have been fur‐
ther deformed and folded internally (Figure 17e, f). Stretching lineations in high‐strain parts
of the Canisp shear zone suggest dextral‐oblique‐reverse movement (south‐side up thrust)
followed by strike‐slip shearing [77, 79, 82]. The Diabaig shear zone is thought to be an in‐
clined thrust ramp dipping toward NE. Similar shear zones exist in the Outer Hebrides
portion of the Lewisian Complex (see Figures 4 and16a).

3.2. A chronology of events


In the past sixty years the Lewisian Complex has been the subject of very extensive geo‐
chronological studies with the application of many different decay systems. Much of the
initial work, especially that done with whole rock methods, documented the existence of
Archaean and Palaeoproterozoic events, matching the subdivision proposed by [70] based
on the pre‐ and post‐ Scourie dyke position of the rocks [83‐88]. The details of the picture,
however, have remained fuzzy due to the complications introduced by the multistage evo‐
lution of the rocks. More recent dating of zircon and other minerals with U‐Pb has helped to
shed more light into the timing and importance of the events in the different domains. This
evolution included five major events, interspersed by some minor but tectonically important
events (Figure 6): (i) Meso‐ and Neoarchean orogenic activity, mainly between 3.00 and 2.70
Ga, built the bulk of the Lewisian crust; (ii) an earliest Palaeoproterozoic event at 2.50‐2.48
Ga, had a profound influence on the Assynt block, but is not seen elsewhere, except for
anorthosite in South Harris; (iii) emplacement of Scourie dyke during at least two episodes
at 2418 and 1992 Ma; (iv) deposition of clastic sediments at Loch Maree sometime between
2.0 and 1.9 Ga, leading to, or associated with a localized but intense episode of arc magma‐
tism at 1.90‐1.85 Ga at Loch Maree, Laxford Bridge and South Harris; and (v) local migmati‐
zation and emplacement of pegmatite dyke swarms at 1.70‐1.65 Ga, mainly in South Harris
but recorded by sporadic pegmatites and titanite across the entire Lewisian Complex.

i. Meso‐ and Neoarchaean evolution


The Lewisian Archaean crust is dominated by banded, felsic to intermediate TTG‐gneisses
of presumed igneous origin. Meso‐ and Neoarchaean supracrustal rocks composed of semi‐
pelites, calc‐silicate schists, meta‐arkoses and metavolcanic rocks associated with mafic to
ultramafic and anorthositic rocks with a tholeiitic composition are also preserved, and when
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 305

present, may represent the protoliths of some of the gneisses. The metamorphic grade
reached granulite facies in the Assynt block, and varies from granulite to amphibolite facies
elsewhere. The geochronology of these gneisses tends to be very complex because of the
multiple overprints which are recorded in zircon so that in many cases the U‐Pb data are
scattered and do not entirely resolve the sequence of events. However, the overall picture
permits to distinguish a main pattern typical of the Assynt block and distinct from that of
the other segments (Figure 6).

The TTG rocks in the Assynt block preserve some of the oldest zircon ages of 3.04‐2.96 Ga
and it has been suggested that this is the age of formation of the rock [89‐91], but it remains
uncertain whether the oldest grains may just be xenocrystic and the gneisses actually
formed about 2.85 Ga [92]. The early high grade metamorphism occurred at around 2.80‐
2.75 Ga, and was followed by the intrusion of trondhjemite at 2.72 Ga [93‐95] and emplace‐
ment of mafic‐ultramafic rocks [88].

The other Archaean segments of the Lewisian rocks had the most prominent period of
development between 2.9 and 2.0 Ga, with a peak at 2.84‐2.82 Ga [7, 90, 91, 93, 96‐101].
An exception is the Rona segment where ages between 3135 and 2880 Ma have been
reported for tonalitic gneisses [102]. The Gruinard block underwent granitic to
trondhjemitic magmatism and high‐grade metamorphism at about 2.73 Ga [91, 96, 97]
and in the Richonich block there are some indications for further activity as late as ca. 2.6
Ga [90].

ii. Earliest Palaeoproterozoic metamorphic progression: 2.50‐2.48 Ga

The Assynt block was affected by a second high‐grade metamorphic event at 2.50‐2.48 Ga.
This event caused strong resetting of U‐Pb in zircon due to recrystallization and local new‐
growth [90‐94] and is also recorded by Sm‐Nd isochrons obtained from garnet and coexist‐
ing minerals in ultramfic pods [104], as well as by U‐Pb in titanite and monazite [94, 96]. The
dry high temperature metamorphism was concluded by re‐hydration which caused local
formation of granitic pegmatites and leucosome in migmatites [94] and likely led to retro‐
gressive amphibolite facies metamorphism and deformation, termed the Inverian [73, 74].
The only well constrained temporal analogue to these metamorphic events elsewhere in the
Lewision Complex is an anorthosite body present in the younger South Harris igneous
complex [105].

iii. Scourie dyke swarms

The Scourie dykes form part of an extensive dyke swarm present throughout the Lewisian
Complex [37] but most abundant in the southern region. These dykes display many
different geometries and attitudes, mostly steeply dipping, and they cut all Neoarchaean
folds and planar fabrics (Figure 15). The composition of the dykes varies from mafic to
ultramafic [105], and their trace element geochemistry indicate formation in a marginal
continental setting, as island arcs or from a mixtures of crustal and oceanic material [106,
107]. The Scourie dykes intruded during at least two different events at 2418 Ma and 1992
Ma [72, 75, 108, 109]. The youngest of these dykes intruded into hot gneissic and migmatitic
306 Tectonics – Recent Advances

rocks syn‐kinematically with a late stage of shear zone deformation [79, 107], of the same
kind as that shown in Figure 14d.

iv. Deposition of supracrustal rocks (2.00‐1.90 Ga) and arc magmatism (1.90 – 1.85 Ga)
Banded iron‐formation, marble, chlorite schists and meta‐psammites of the Loch Maree
Group were deposited on the Neoarchaean crust and are now arranged in two narrow, NW‐
SE trending synformal belts [110, 111] in the Gairloch area (Figure 18). Deposition of the
Loch Maree Group is constrained between 2.0 Ga, the youngest detrital zircons from a
metagreywacke in the Gairloch area [112], and 1903 Ma, the age of the syntectonic Ard
gneiss intrusion [102, 110]. The latter was emplaced during the early stages of deformation
associated with amphibolite to granulite facies metamorphism and interpreted to be related
to the development of a subhorizontal shear zone net work. The Ard gneiss is considered to
be a product of arc magmatism and the deformation caused by lateral accretion of oceanic
plateaus and primitive island arcs [110].

Figure 18. Simplified map of the Palaeoproterozoic Loch Maree Group (shaded) within the Lewisian
Complex of the Gairloch area, NW Scotland [110]. This group consists of highly deformed amphibolites
and metasedimentary rocks cut by 1.9 Ga granitoid alkaline rocks. Note macro‐scale reverse and sinis‐
tral folds and related off‐sets and lateral shear zone displacement.
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 307

The other important tectonic element with similar age and genesis is the South Harris ig‐
neous complex in the Outer Hebrides (Figure 4) that comprises a magmatic arc sequence of
mafic to felsic intrusive rocks formed at 1.89‐1.88 Ga [98, 100]. Clastic sedimentary rocks in
this region contain abundant detritus of the same age [7, 98]. Crudely coeval high grade
metamorphism strongly overprinted zircon in the 2.5 Ga anorthosite [104] and has also
been dated by Sm‐Nd mineral systems [113]. This event also caused deformation and met‐
amorphic zircon growth in a shear belt at the northern tip of the Tarbert block at Nis (Fig‐
ure 6) [98, 99]. As for the Loch Maree assemblage, the South Harris situation has been in‐
terpreted ‘to represent a magmatic arc, complete with contemporaneously derived clastic
sediments, developed in a collisional orogen, which culminated in granulite facies meta‐
morphism’ [98].

A third important occurrence of intrusive rocks of this age is within the Laxford shear
zone. Granite sheets, with a U‐Pb age of 1854 ±13 Ma [72] occur subparallel to the shear
zone boundaries but cutting the pre‐Scourie dyke fabric (Figure 19a, b). Synchronous shear
deformation affected the dykes and aligned them as lenses into shear zones that also affect‐
ed the surrounding tonalitic and mafic gneisses by mostly contractional deformation (Fig‐
ure 19b‐d). There is some consensus that the Laxford shear zone likely represents a terrane
boundary but the timing of juxtaposition is debated, and the role played by the granites
unclear. Recent work [80] concludes that the terranes were probably brought together
sometime between 2.5 and 2.4 Ga, after the earliest Palaeoproterozoic deformation and
retrogression, but before intrusion of the Scourie dykes. Others, however, suggest that the
juxtaposition probably occurred around the time of emplacement of the granites, consistent
with the timing proposed for the Loch Maree group and South Harris igneous complex [7,
8, 114].

v. Late orogenic events (1.70 – 1.65 Ga)

A metamorphic event at around 1.75 Ga is indicated by a Sm‐Nd age of garnet and coexist‐
ing metamorphic minerals from a mafic dyke of the Assynt block [109,], and by titanite from
rocks near the Laxford shear zone [90, 93]. Because of the localized occurrence of the rocks,
however, the tectonic significance of this age is uncertain.

By contrast a later event at 1.70 to 1.65 Ga had a much stronger impact across all of the Lew‐
isian Complex (Figure 6). The main expression of this event is the granitic and pegmatitic
migmatite complex associated to, and bordering, the South Harris igneous complex [7, 100,
101, 115, 116]. Felsic dykes and pegmatites of this age have also been found in most other
parts of the Lewisian [7, 98, 110] and they also coincide with the ages of rutile and a younger
titanite generation [76, 93, 102]. The exact significance of this event is still uncertain. One
view is that emplacement of pegmatites of this ages pre‐dated, but likely broadly coincided
with the late flexuring, steep shearing and greenschist facies retrogression [110], thus attrib‐
uting these events to contractional processes during late‐stage collision of the Lewisian
Complex with e.g. a southern block [8].
308 Tectonics – Recent Advances

Figure 19. Outcrop features of Palaeoproterozoic intrusive rocks in the Lewisian complex. (a) Granitic
pegmatite intrusion into TTG‐gneisses of the Rhiconich terrane. (b) Syn‐kinematic felsic pegmatite
dykes intruded into mafic gneisses and sheared during the main Palaeoproterozoic tectono‐thermal
events. Note boudinaged and asymmetric lenses indicating top‐NE movement sense (right in photo).
View is toward SE. (c) Syn‐ and post‐kinematic granite pegmatite dykes injected into mafic Neoarchae‐
an gneisses. Note lensoidal sheaped mafic bodies in the sheared gneisses. (d) Revers ductile shear zone
displaying a duplex geometry, cutting obliquely across steeply dipping TTG‐gneisses in the Rhiconich
terrane. View is to the NW.

4. Discussion
4.1. Similarities of terranes and terrane juxtaposition
In order to discuss terrane aspects and hypothetic assembly history of the West Troms
Basement Complex, the Lofoten‐Vesterålen province and the Lewisian Complex as a
framework for further correlation, we focus on critical similarities such as the presence of
crustal segments with contrasting age, tectono‐magmatic and/or metamorphic histories,
crustal‐scale ductile shear zones (sutures), overlapping nature and character of deformation
(convergent and strike‐slip) and major metamorphic breaks that may have juxtaposed dif‐
ferent crustal levels instead of spatial terranes [8, 72]. The first step would be to locate differ‐
ent crustal segments, then to localize the suture(s) formed by the juxtaposition of terranes,
and finally, discussing assembly of different crustal levels in order to explain metamorphic
and petrologic differences known from all the studied regions [8, 17].
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 309

Although metamorphic and structural characteristics of Precambrian continental crust in


general is not uniquely diagnostic for correlation [107], the geotransect in western Troms
and Lofoten is underlain by age‐equivalent TTG‐gneisses and granitoid gneisses of Meso-
and Neoarchaean age, with only modest differences in metamorphic grades and histories
[32]. This indicates that at the end of the Archaean these rocks were likely assembled as a
large single terrane, formed during a prolonged cratonization event (2.92‐2.67 Ga), though
with considerable complexity in detail [32]. In this scenario, the migmatized ductile shear
zones bounding the Dåfjord and Kvalsund gneisses on Ringvassøya [17] and, potentially
also the major tectonic and metamorphic boundary in the Lofoten‐Vesterålen area [10] may
reflect Neoarchaean intra‐cratonic terrane boundaries. An argument for multiple terranes in
the West Troms Basement Complex at the end of the Neoarchaean, however, appears from
the contrasting ages, e.g. 2.8 Ga for the Ringvassøya greenstone belt, 2.4‐2.2 Ga for the Van‐
na group [29] and c. 1.9 Ga for the Torsnes belt [31].

By comparison, the Lewisian Complex was earlier considered to have originated from a
single Neoarchaean continent that split up into multiple terranes in the early Palaeoprotero‐
zoic and later on were juxtaposed during the 1.90‐1.67 Ga events [79]. A revised model of
Palaeaoproterozoic juxtaposition in the Lewisian, however, included ten terranes and one
block [7, 72, 90]. The Laxford shear zone was considered a terrane boundary between the
Rhiconich terrain in the northeast and the Assynth terrane in the southwest (see Figure 4).
These two terranes were likely accreted after intrusion of the pegmatite sheets in the Rhi‐
conich terrane at c. 1.855 Ga, since they are absent in the Assynth terrane, but prior to the
early‐Palaeoproterozoic structures common in both regions. One model suggests that the
accretion occurred at 1.74 Ga, synchronously with an amphibolite facies metamorphism
recorded in the Rhiconich terrane and a metamorphic retrogression associated with the
formation of shear zones in the Assynth terrane [7]. Recently, it was argued that juxtaposi‐
tion of the Assynth and Rhiconich terranes occurred prior to the 1.9‐1.75 Ga period, e.g. in
late Neoarchaean since Scourie dykes are present on both sides of the Laxford shear zone
[80].

A second model [8] involved only two continental plates during the Neoarchaean and Pal‐
aeoproterozoic history of the Lewisian Complex. In this model the Lewisian Complex was
divided into two blocks classified as upper‐plate and lower‐plate blocks (Figure 20) that dif‐
fered considerably with regards to position of the supracrustal rocks relative to the accreted
versus the overriding plates, which is critical for the metamorphic conditions. The upper‐
plate block preserved rare and weak Palaeoproterozoic deformation and amphibo‐
lite/greenschist facies retrogression of granulite facies gneisses, while the lower‐plate portion
involved prograde and peak amphibolite facies metamorphism and high‐strain assemblag‐
es. The upper‐plate blocks displayed weak deformation and could be located on the low‐
strain areas above the terrane boundary shear zone in the crustal model for the mainland [8,
37], while the lower‐plate blocks with strong Palaeoproterozoic structures could correspond
to the mid‐deep level of the shear zone itself, i.e. as during the emplacement of the Loch
Maree Group (Figure 20).
310 Tectonics – Recent Advances

Figure 20. Schematic model of an idealized subduction‐accretion collision sequence [8] leading to
domains of contrasting deformation and metamorphism, e.g. high‐grade/prograde in lower plate and
low‐grade retrogressive metamorphism in upper plate settings.

By analogy, in the West Troms Basement Complex, a significant Palaeoproterozoic tectono‐


metamorphic break is thought to exist southwest of the island of Senja (Figure 2), in a region
that separates dominantly amphibolite facies gneisses from granulite facies AMCG‐suite
rocks of the Lofoten‐Vesterålen province [9, 10]. This major boundary is also inferred by
contrasting gravity and magnetic characters, and could therefore reflect a Palaeoproterozoic
suture [11, 13, 50]. The slightly older ages of the peak Palaeoproterozoic deformation (1.87‐
1.78 Ga) versus 1.78 Ga in the West Troms Basement Complex (Table 1) suggest progressive
southwestward accretion toward an orogenic hinterland near Lofoten (Figure 21), which is
also consistent with the observed increase in metamorphic grade. On the other hand, the
contrast in U‐Pb ages of basement rocks in the southwest (2.7‐2.6 Ga) compared to in the
north (2.92‐2.8 Ga), could indicate a second terrane in the northeast (Figure 21)[10, 17].

Figure 21. Schematic model of Palaeoproterozoic accretion in the West Troms Basement Complex.
Lower – Upper plate model used to explain among others, metamorphic differences [8].

A model involving SW‐directed convergence/accretion toward an orogenic front near


Lofoten and an oppositely NE‐dipping terrane boundary zone in the northeast (e.g. reac‐
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 311

tivated Neoarchaean migmatite zones in Ringvassøya) may explain both the contrast in
deformation styles and metamorphic grades along the geotransect. For example, an accre‐
tion/subduction‐derived shear zone situated in the Senja shear belt adjacent to eastern
Lofoten‐Vesterålen would locate them to the lower plate of two or more unified colliding
terranes (Figure 21, left). This would have formed prograde amphibolite to granulite‐
facies metamorphic assemblages, granitic melts, migmatite zones and strong ductile de‐
formation, including a detachment in the Senja shear belt that could have become the
locus of later‐stage partitioned crustal deformation. Conversely, in an upper plate posi‐
tion (Figure 21, right), localized and weakly developed but more likely retrogressed, low‐
grade metamorphosed cratonic‐marginal shear zones and supracrustal platform sequenc‐
es may appear (as in Vanna). This model would favor assembly of different crustal levels
of at least two main crustal segments (terranes) rather than spatially separated smaller
terranes [8].

4.2. Comparison between Lewisian, western Troms and Lofoten‐Vesterålen


Plate tectonic reconstructions of Precambrian units in the North Atlantic realm with
respect to Fennoscandia and Laurentia (Figure 22) have to some extent failed to demon‐
strate whether these cratons belonged to the same supercraton in the Neoarchean (2.8–
2.5 Ga) and Palaeoproterozoic (1.8‐1.6 Ga). A tectonic linkage is supported by paleomag‐
netic reconstructions [8, 119] stating that Fennoscandia was positioned close to the
Greenland/Laurentia and Superior supercraton in the late Neoarchean (Figure 22). Geo‐
logical similarities and differences between domains are important criteria for restoring
possible supercontinents, as stated by [3, 4, 119]. In this context, the studied basement
outliers west of the Scandinavian Caledonides in North Norway and the Lewisian com‐
plex in Scotland both have a pivotal central location within the marginal orogenic belts
constituting the presumed Neoarchaen supercontinent (Figure 1) [110, 120]. These units,
however, also occupy an interior position of the Caledonian orogen far from the autoch‐
thonous shield rocks and are bounded by younger faults. They are, thus, usually not
considered part of any shield areas, but instead assigned an uncertain or exotic tectonos‐
tratigraphic status [3, 121].

Based on the comparison between the Archaean‐Palaeoproterozoic basement suites in North


Norway and the Lewisian of Scotland outlined above we can discuss potential correlation of
these suites in the context of the North Atlantic realm. Such a correlation can be tested using
similarities or dissimilarities in lithology, age, supracrustal units, igneous/ petrogenetic,
structural and metamorphic features and evolutionary and tectono‐metamorphic history
(see Table 1).

4.2.1. Archaean components


Archaean crust forms the backbone of both the western Troms, Lofoten‐Vesterålen province
and the Lewisian basement complexes. These complexes reveal some broad similarities in
312 Tectonics – Recent Advances

terms of lithology and general age patterns but also some differences, which, however, are
most pronounced within each of the regions (Figure 6). Thus, we are comparing two
heteorgeneous Archaean crustal segments.

Figure 22. (a) Reconstruction of Laurentia and Fennoscandia during the Palaeoproterozoic based on the
palaeomagnetic fit of [118]. Note that the West Troms Basement Complex (including Lofoten) and the
Lewisian Complex lies within a continuous Palaeoproterozoic belt extending from the Torngat orogen
of Laurentia through the Ketilidian and Nagssugtoqidian orogens to link up with the Kola and Kareli‐
an/Belomorian provinces of Fennoscandia. The arrows show inferred movement directions of various
crustal segments relative to Laurentia. The map is modified from [4, 110], while the reconstruction of
continents at 1.83 Ga is based on [118]. Abbreviations: WTBC =West Troms Basement Complex.

In the West Troms Basement Complex we can distinguish: (i) a Mesoarchaean tonalitic do‐
main formed between 2.9 and 2.8 Ga in Ringvassøya and Vanna, overlain by (ii) the broadly
coeval, but tectonically distinct Ringvassøya greenstone belt. These two domains are sepa‐
rated by the (iii) late orogenically active Kvalsund gneiss migmatite zone from (iv) the Neo‐
archean domain of Kvaløya and Senja farther south, which formed during a short time in‐
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 313

terval between 2.72 and 2.67 Ga. The Lofoten‐Vesterålen province has broadly the same age
structure as the southern part of the West Troms Basement Complex, except for the fact that
it underwent high‐grade metamorphism at 2.63 Ga. The apparent formation of leucosome in
southern Senja at around 2.61 Ga [32] may be the expression of the same event, indicating
that Senja and Vesterålen may simply represent different crustal levels experiencing meta‐
morphism at different times, a very common pattern of crust construction and maturation
[122]. The late Archaean deformation and migmatization that seems to characterize the
Kvalsund migmatized zone separating the Dåfjord and Kvalsund gneisses [17, 32] may be
another expression of this late orogenic activity.
In the Lewisian there are also some differences in the Archaean history of the contrasting
blocks (Figure 6). Several of them appear to have formed mainly in the late Mesoarchaean,
between 2.85 and 2.80 Ga. Older ages up to 3150 Ma are recorded in the Rona and Assynt
blocks whereas Neoarchean magmatic activity seems to be restricted mainly to the Assynt
and Gruinard blocks, although there are local indications of such activity also in the Tarbert‐
West Uist and Rhiconich blocks.

The single most distint event that is fully missing in the West Troms Basement Complex and
Lofoten‐Vesterålen is the pervasive high‐grade metamorphism and subsequent rehydration
and retrogression in the Assynt block at 2.5‐2.48 Ga. However, this event is not seen in the
other Lewisian blocks except for the anorthosite body in South Harris (Figure 6). One expla‐
nation is that it was a specific terrain formed in another orogen, the alternative is that the 2.5
Ga high‐grade event reflects a process affecting lower crustal levels but not recorded higher
in the crust, in the same way as seen for example in the Superior Province [123].

A comparison of the evolution of these two crustal sectors is difficult because of the inherent
internal differences, which may reflect dependencies from crustal level, and probably also
the tectonic juxtaposition of different terranes combined with the unequal geochronological
coverage in different blocks and the technical difficulties to cleanly date the age of protolith
and orogenic tranformations of complex polymetamorphic gneisses. Hence we can conclude
that the West Troms Basement Complex, the Lofoten‐Vesterålen province and the Lewisian
have certain affinities in common (Figure 6), suggesting that they could have been linked to
some degree in the Archaean period, but the opposite conclusion is also possible.

2.40 to 1.98 Ga supracrustal rocks and mafic dyke swarms

Intrusion of the huge Ringvassøya mafic dyke swarm in the West Troms Basement Complex
(Figure 23b) occurred at c. 2.40 Ga (Figure 6) [34]. This event is part of a major mafic dyke‐
producing event that affected several Archaean cratons, and as such it does not necessarily
represent a unique stitching tool for linking these crustal domains. One argument
supporting such a role, however, is the apparent south‐westerly shift in age from ca 2.5 Ga
events in Kola, to the most widespread phase at 2.45 Ga in Kola and Karelia, and finally to
the 2.40 Ga phase in Ringvassøy [34], which is close to the age of the older Scourie dyke
generation in the Lewisian. The 2221 Ma dioritic sill intruding meta‐sedimentary rocks of
the Vanna group [29] is also the expression of a localized but very ubiquitous magmatic
phase across northern Fennoscandia [124] and also in Laurentia. No equivalents have so far
314 Tectonics – Recent Advances

been described from the Lewisian. However, there is a good temporal correlation, instead,
between mafic magmatism at 1.98 Ga in the Mjelde‐Skorelvvatn belt of the West Troms
Basement Complex [31] and emplacement of the younger generation of the Scourie dykes in
the Lewisian, both corresponding to a period of extension and rifting in Laurentia and
Fennoscandia.

1.90‐1.85 Ga arc magmatism and convergence

The subsequent period of arc magmatism, likely connected to plate convergence and
subduction, and final collision was important in the Fennoscandian Shield [4] and also in the
Lewisian where it formed the well documented successions at Loch Maree and South Harris
in a sequence of events between 1.90 and 1.85 Ga. Granitic sheets of this age are also
important along the Laxford shear zone. In Lofoten there was a correlative event at 1.87 Ga,
which emplaced granite and local mangerite‐charnockite intrusions. Such rocks, however,
have so far not been reported from the West Troms Basement Complex, a feature that may
reflect a more distal position relative to the orogenic front near Lofoten (Figure 23c; see
below). In the Lofoten‐Vesterålen province, there is also clear evidence of meta‐supracrustal
units that post‐date the Neoarchaean gneisses [9], but Palaeoproterozoic ages have not yet
been documented by radiometric age dating.

1.80 – 1.78 Ga magmatism

The single most important and widespread magmatic event affecting the West Troms Base‐
ment Complex and Lofoten‐Vesterålen province occurred in a short burst at 1.80 ‐1.79 Ga. It
formed most of the AMCG suite in Lofoten and the major Paleoproterozoic intrusions in
Kvaløya and Senja. In Lofoten the event was pre‐ and post‐dated by high grade metamor‐
phism and ductile deformation [9, 10], whereas in western Troms there is no evidence for
much activity preceding this magmatic phase. These events can be correlated with a well‐
defined period of late orogenic magmatism in Fennoscandia [10, 125]. Interestingly, there
are no such plutons or strong metamorphic overprint of this age in the northern part of
Ringvassøya and Vanna, even though granulite facies metamorphism and partial melting
occurred at about 1.78 Ga in Sandøya, just at the edge of this block, supporting an alloch‐
thonous origin of the latter [17]. A similar situation is also characteristic of the entire Lewis‐
ian which lacks 1.80‐1.78 Ga intrusion altogether

1.80‐1.75 Ga deformation and metamorphism

Regional deformation and metamorphism are well documented in the West Troms
Basement Complex at c. 1.80‐1.75 Ga (Table 1). These processes involved high‐strain
deformation and prograde metamorphism up to granulite facies (1.78‐1.768 Ga). The
deformation was focused mainly along the boundaries to metasupracrustal belts, e.g. in the
Senja Shear Belt (Figure 10, 23c, d), and was probably also superimposed on pre‐existing
Neoarchaean structures [17, 32, 51]. The deformation started with ENE‐directed thrusting
and was followed by macroscopic upright folding and combined, late‐stage strike‐slip
shearing and SE‐directed thrusting (Figure 24) [17]. The late stages of deformation, not yet
documented by age datings (but likely younger than 1.75 Ga), were characterized by
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 315

partitioned contraction and lateral displacements. In the limbs of the mid‐stage macroscopic
folds (Figure 24b) the subsequent oblique deformation produced foliation‐parallel sinistral
strike‐slip faults and steeply‐plunging folds (Figure 24c), creating a regional lens‐shaped
structural pattern in the West Troms Basement Complex (Figure 10). The final phase of SE‐
directed thrusting (Figure 24d) was temporally linked to the strike‐slip shearing, thus
indicating partitioned transpression as the overall deformation mechanism [17].

Figure 23. Cartoon sections summarizing the Neoarchaean to Palaeoproterozoic tectonic evolution of
the West Troms Basement Complex: (a) Neoarchaean (2.92‐2.56 Ga) tonalitic gneiss forming events with
crustal accretion and thickening due to underplating. Note position of possible precursory volcanic
deposits of the Ringvassøya greenstone belt. (b) Neoarchaean and Early Palaeo proterozoic (2.4‐1.9 Ga)
crustal extension, basin formation and intrusion of the Ringvassøya mafic dyke swarm. (c) Palaeoprote‐
rozoic (1.9‐1.8 Ga) continental contraction and probable magmatic arc accretion in the southwest (in‐
cluding the Lofoten AMCG suite). (d) Section illustrating the composite result of Palaeoproterozoic
crustal contraction, accretion and continent‐continent collision with increasing transpressive defor‐
mation through time. For abbreviations, see Figure 2.

In contrast, although the deformation style is similar, there is not full evidence for
temporally equivalent deformation events in the Lewisian. The exception is titanite ages of
about 1750 Ma near the Laxford shear zone. These ages are considered as the potential
expression of a phase of regional metamorphism but the evidence in favor of such an
interpretation is dubious. In the Lewisian there are, for example, no datable dykes
interspersed with the deformation events like in the West Troms Basement Complex.
316 Tectonics – Recent Advances

Figure 24. Tectonic model for the Palaeoproterozoic deformation in the West Troms Basement Complex
(A‐D) [17] compared with the Lewisian (E) [37]. The overall framework is that of NE‐SW directed or‐
thogonal shortening and an increasing transpressive component with time. The spatial domains, named
in Figure 2, and their kinematic characters are also illustrated. (A) Early‐stage formation of NE‐directed
thrusts and a low‐angle main mylonitic foliation in the metasupracrustal belts. (B) Continued orthogonal
NE‐SW contraction produced upright macro‐folds with steep limbs. Note that the main foliation and early
thrusts were folded. (C) Late‐stage tectonism involved NE‐SW orthogonal and/or oblique to orogen‐
parallel contraction (NW‐SE) and mostly sinistral strike‐slip reactivation of steep macro‐fold limbs, e.g. in
the Senja Shear Belt. The eastern, more flat‐lying macro‐fold hinges (e.g. Ringvassøya greenstone belt)
provided the locus for potential low‐angle thrust detachments that may have accommodated partitioned
NW‐SE shortening and SE‐directed thrusting. (D) Late‐stage Palaeoproterozoic kinematic model for the
north‐eastern part of the West Troms Basement Complex, where potential low‐angle shear
zones/detachments accommodated NW‐SE directed thrust movements on flats and steep orogen‐parallel
strike‐slip/transfer‐type shear zones on ramps [17]. (E) Simplified kinematic model for the Palaeo protero‐
zoic deformation in the Lewisian Complex involving a combination of thrust and strike‐slip movements on
flats/detachments and ramps/steep transfer‐type shear zone [37].

1.70‐1.65 Ga deformation and metamorphism

The Lewisian underwent a very distinct set of events at 1.70‐1.65 Ga including deformation,
the local development of migmatites, the ubiquitous intrusion of pegmatites, and low grade
metamorphic overprints reflected in secondary titanite and rutile ages. These events largely
post‐date similar pegmatite intrusions affecting the West Troms Basment Complex and the
Lofoten‐Vesterålen province.

The late stages of magmatism and deformation in the Lewisian at 1.70‐1.65 Ga involved
localized steep ductile reverse (Figure 17) and dextral oblique shear zones developed by
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 317

strain partitioning likely due to reactivation of steep pre‐existing shear zones [77]. This
partitioned deformation was interpreted to have been related to major flat‐lying detachment
zones (Figure 24e) that generated dip‐slip thrust movement on flat portions, i.e. the Laxford
shear zone, and strike‐slip shear zones on steep oblique ramps, i.e. the Canisp shear zone
[37]. The 1.70‐1.65 Ga stage corresponds to a retrogression of the amphibolite/granulite
facies conditions to greenschist facies, indicating exhumation of the rocks from mid to upper
crustal levels [79]. This event was associated with the formation of steep‐plunging
asymmetric folds, retrogressed cleavage, widespread fluid‐flow and quartz vein
precipitation, and multiple strike‐slip shear zones [37].

A similar crustal model is invoked for the late‐stage partitioned deformation in the West
Troms Basement Complex [17], which, despite lack of critical age dating may correspond to
the 1.70‐1.67 Ga event in the Lewisian (Figure 24d). In this model, potential flat‐lying thrust
detachments (shear zones) are present in the northern part of the region.

In summary, the Palaeoproterozoic deformation in the Lewisian Complex and that of the
West Troms Basement Complex and Lofoten‐Vesterålen, display obvious similarities in
terms of tectonic style and partitioning of the deformation, even though they may not be
fully temporally correlative (Figure 23). They show: (i) Long‐term protracted deformation
character, (ii) presence of crustal scale ductile shear zones (iii) partitioned transpressive
crustal deformation, e.g. thrusts and orogen‐parallel strike‐slip shear zones, and (iv) spatial
changes in metamorphic grades, i.e. major tectono‐metamorphic breaks such as the regional
metamorphic boundary in Lofoten and the Laxford shear zone in the Lewisian. These simi‐
larities are all consistent with comparable tectonic assembly processes caused by accretion
followed by crustal convergence and orogen‐oblique/parallel displacement of crustal seg‐
ments, and a terminal phase of crustal/differential uplift and reactivation [8, 17].

4.3. Linking Palaeoproterozoic terranes and events in the North Atlantic realm
A major problem when trying to restore Precambrian plate tectonics is the nature and
processes of assembly of lower crustal blocks or terranes [3, 4, 7, 8]. In terms of the North
Atlantic realm Williams et al. (1991)[126] proposed that at the end of the Neoarchean there
was a supercontinent, Kenorland, whose breakup led to the formation of several micro‐
continents which were reassembled together with juvenile terranes in the Palaeoproterozoic
(Figure 22). Others, however, have argued for the existence of several micro‐continents at
the end of the Neoarchaean instead of one single supercontinent [33].

Most workers agree that the Karelia and Superior cratons of Fennoscandia and Laurentia
were in close vicinity to each other or connected in the Neoarchaean [24]. The outline of
these cratons (Figure 22) is a result of cycles of collision, granitoid intrusion, extension,
rifting and basin formation, and if several of these events can be correlated between various
cratons, then it is possible to reconstruct former crustal assemblages or supercratons [33]. In
particular, timing of large igneous provinces and associated episodes of continental breakup
and supracrustal deposits can be used for such analyses, whereby the most detailed record
known is that of the Laurentian cratonic fragments [127]. Similar Palaeoproterozoic
318 Tectonics – Recent Advances

configurations have also been discussed in the literature, and various models presented [8,
110, 128, 129, 130]. Following breakup of a potential Neoarchaean supercraton, oceanic arcs
started to converge from c. 2.0 Ga, with eventually accretion of the cratons along sutures
that follow the grain of the Palaeoproterozoic orogens (Figure 22). The model proposed by
[110] suggests a rather familiar configuration of the various Fennoscandia/Baltica and
Laurentia cratons at the beginning of the Palaeoproterozoic (Figure 22), and despite being a
speculative model it addresses the need for more detailed research within these cratons and
especially along their margins.

Recent paleomagnetic reconstruction of the Palaeoproterozoic [8, 128] suggests the presence
of several large colliding plates, including the North Atlantic and west Greenland plates, the
Central Greenland Craton and the Fennoscandian (Baltic‐Kola) plate, with the Lewisian
somewhere in between (Figure 25). Most workers link the Lewisian to the Palaeoproterozoic
Nagssugtoqidian belt in Greenland [5, 129, 131], and consider that this belt may have coun‐
terparts both in North America and/or the Fennoscandian Shield [3, 4, 110]. A link between
the Lewisian of NW Scotland and the Lappland‐Kola and Karelia craton of northern Fen‐
noscandia would then place the West Troms Basement Complex and Lofoten‐Vesterålen
province exactly along the line of intersection between these major Palaeoproterozoic oro‐
genic belts (Figure 22). A similar reconstruction [129] supports a correlation of Palaeoprote‐
rozoic orogens in Greenland and Fennoscandia at the c. 1.8 Ga supercontinent stage.

The scenario proposed by Park (2005) [8] gives a valid plate setting for the end of the
Neoarchaean (Figure 25a) and explains the subsequent Palaeoproterozoic tectono‐
metamorphic events in the Lewisian Complex and tentatively, also the deformation events
in the West Troms Basement Complex and Lofoten‐Vesterålen province. At ca. 1.9‐1.87 Ga,
volcanic arcs were created between North American craton and Central Greenland
craton/Kola craton due to the subduction/ accretion of the oceanic crust located between
them (Figure 25b). The calk‐alkaline plutonic intrusions within the Loch Maree Group, the
South Harris Igneous complex, and potentially, the earliest phases of magma intrusions in
the Lofoten igneous province and West Troms Basement Complex, and accompanied
convergent deformation and granulite facies metamorphism manifest this regional
accretionary event [8]. At ca. 1.87 Ga, the Central Greenland craton and Kola‐Karelian
craton collided and was under‐thrusted beneath the the North American craton in a NW‐
SE direction within the Lapland‐Kola belt, and resulting in the main phase of deformation
(Figure 25c). Granulite facies metamorphism occurred in the down‐going slab due to
under‐thrusting (lower plate). The line of collision between juvenile terranes was likely
oriented in the same direction as the Palaeoproterozoic Nagssugtoqidian belt and the
orientation of the collision could be given by the orientation of the main NW‐SE trend of
the Laxford shear zone [8]. At ca. 1.8 Ga, subduction of oceanic crust to the SW of this new
continent may have created a volcanic arc trending NW‐SE (Figure 25d), and this arc may
have been involved with renewed collision at ca. 1.75 Ga (Figure 25d, e), corresponding to
the main stages of deformation in the West Troms Basement Complex. There, the intrusion
of the calk‐alkaline Hamn norite (1.8 Ga) and the Ersfjord granite (1.79 Ga) may have been
related to this phase. Similar calk‐alkaline intrusive rocks exist further south of the
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 319

Lewisian complex [132, 133], known as the ‘Malin block’ and this block was thought to be
part of a belt comprising the Labradorian of NE Canada, the Ketilidian of South Greenland
and the Gothian of Scandinavia (see Figure 22). This belt became tectono‐magmatic active
at c. 1.8 Ga and an event at c. 1.7 Ga could be the result of igneous activity and deformation
related both to the latest Palaeoproterozoic events in the Lewisian and in the West Troms
Basement Complex.

Figure 25. Plate tectonic setting of the Lewisian complex during the Palaeoproterozoic, based on the
North Atlantic reconstruction [8, 128]. Abbreviations: CGC=Central Greenland craton, Goth = Gothian
belt, Kar = Karelia craton, Ket = Ketilidian belt, Kola= Kola craton, Lap‐Kol = Lapland‐Kola belt, Lew =
Lewisian, NAC = North Atlantic craton, Nag = Nagssugtoqidian belt, NI = north Ireland, NS = north
Scotland. (a) Distribution of cratons and orogenic belts during the Mesoproterozoic. (b) 2.0 Ga: Subduc‐
tion and creation of a volcanic arc in oceanic crust between two continental plates (NAC and CGC/Kol)
followed by accretion of oceanic/arc elements along the leading edge of the NAC. (c)1.87 Ga: Collision
of the two continents followed by underthrusting of the CGC/Kola craton beneath the NAC, causing the
early Palaeoproterozoic deformation and metamorphism. At the same time, collision occurs in the
Lapland/Kola belt to the SE caused by collision with the Karelia craton. Note the NW‐SE movement
direction. (d) 1.80 Ga: Development of a volcanic arc in oceanic crust SW of the amalgamated continent
created in b. (e) 1.75 Ga: Collision between the “Malin block” and the continent, causing late‐
Palaeoproterozoic deformation, metamorphism and granitic melt formation in the Lewisian complex.
320 Tectonics – Recent Advances

5. Conclusions
(1) The West Troms Basement Complex is underlain by Archaean gneisses (2.92‐2.56 Ga),
metasupracrustal rocks (2.4‐1.9 Ga) and mafic dyke swarms (2.4 and 2.2 Ga) that were
variably reworked, metamorphosed and intruded by felsic and mafic plutons at c.1.8 Ga.
Along strike to the southwest, Neoarchaean high‐grade gneisses in the Lofoten‐Vesterålen
province display magmatic protolith ages of between 2.85 and 2.7 Ga and record a high‐grade
metamorphic event at c. 2.64 Ga. The Neoarchaean basement rocks have been intruded by a
huge 1.8 Ga magmatic suite composed of anorthosites, mangerites, charnockites, gabbros and
granites, which corresponds in age with the 1.81–1.77 Ga, NW‐trending granitoids in the older
part of the Transscandinavian igneous belt of southern Sweden. A similar present structural
position of the basement high in the Lofoten‐Vesterålen province and the West Troms
Basement Complex invokes they are along‐strike correlatives.

(2) The Lewisian rocks of NW Scotland comprise a series of Neoarchaean blocks thought to
have been amalgamated during a multistage and complex set of Palaeoproterozoic collision
events between 1.97 and 1.67 Ga, producing a variety of block‐bounding accretional and
intrablock shear zones. This province also records Neoarchaean crustal deformation and
metamorphism at intervals of 2.7‐2.6 Ga and 2.49‐2.40 Ga followed by episodes of crustal
rifting and mafic dyke intrusion (2.4 and 2.0 Ga), deposition of continental margin‐like
metasedimentary sequences between 2.0 and 1.9 Ga ago upon the substratum of
Neoarchaean gneisses and later on subjected to major orogenic deformation and
metamorphism (c. 1.85 and 1.70 Ga).

(3) The West Troms Basement Complex, the Lofoten‐Vesterålen province and the Lewisian
rocks of Scotland are thus very similar crustal regions in terms of lithology, age, igneous,
structural and metamorphic features and tend to share a similar tectono‐magmatic and
evolutionary history, but there are also sharp differences such as the lack of 1.80 magmatism
in the Lewisian and the c.100 m.y. difference in the timing of the latest Palaeoproterozoic
deformation overprints in the two regions.

(4) Reconstructing Palaeoproterozoic plate scenarios is a difficult task. Nevertheless, paleo‐


magnetic restorations suggest the presence of several large colliding plates, including the
North Atlantic and western Greenland plates, the Central Greenland craton and the Fen‐
noscandian (Baltic‐Kola) Shield, with the Lewisian somewhere in between. In this context,
the Lewisian has been temporally linked to the Palaeoproterozoic Nagssugtoqidian belt in
Greenland and may have its counterpart in North America and/or the Fennoscandian
Shield. A link between the Lewisian of NW Scotland and the Lappland‐Kola and Karelia
craton of northern Fennoscandia would locate the West Troms Basement Complex and
Lofoten‐Vesterålen province directly along the line of intersection between these major
Palaeoproterozoic orogenic belts at the c. 1.8 Ga supercontinent stage.

(5) Tentative similar Palaeoproterozoic terrane models (1.80‐1.67 Ga) can be invoked for the
basement outliers in northern Norway and the Lewisian Complex. The continental assembly
may have involved either multiple small terranes or crustal rejuvenation of one or two large
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 321

terranes. The latter model is based on component similarities and metamorphic variations
and can be explained by the presence of at least two different crustal blocks and/or depth
portions assembled along crustal scale ductile shear zones. The juxtaposition included arc‐
magmatism and accretion of Neoarchaean continental terranes in the vicinity of the Fen‐
noscandia‐Laurentia border, followed by uplift and reworking.

Author details
Steffen G. Bergh1, Per Inge Myhre, Kåre Kullerud and Erling K. Ravna
Dept. of Geology, University of Tromsø, Tromsø, Norway

Fernando Corfu
Dept. of Geosciences, University of Oslo, Blindern, Oslo, Norway

Paul E.B. Armitage


Consulting Ltd, Rochester, England, UK

Klaas B. Zwaan
Geological Survey of Norway, Trondheim, Norway

Robert E. Holdsworth
Dept. of Earth Sciences, University of Durham, UK

Anupam Chattopadhya
Dept. of Geology, University of Delhi, Delhi, India

Acknowledgement
This paper is based on extensive research in the West Troms Basement Complex over the
last decade. The still ongoing work is an interdisciplinary study at the University of Tromsø,
the University of Oslo and the Geological Survey of Norway, aimed at resolving Precambri‐
an regional tectonic questions. The work in the Lewisian was undertaken during the first
author’s sabbatical leave at University of Durham in 2006‐07, and the Department of Earth
Science is thanked for hosting his visit and providing necessary infrastructural facilities for
the research. SB also wishes to thank Dr. R.W. Wilson and Prof. Ken McCaffrey for construc‐
tive collaboration during the project.

6. References
[1] Gaal G, Gorbatschev R (1987) An outline of the Precambrian evolution of the Baltic
Shield. Prec Res 35: 15‐52.
[2] Gorbatschev R, Bogdanova S (1993) Frontiers in the Baltic Shield. Prec Res 64: 3‐21.
[3] Hölttä P, Balagansky V, Garde A, Mertanen S, Peltonen P, Slabunov A, Sorjonen‐Ward P,
Whitehouse M (2008) Archaean of Greenland and Fennoscandia. Episodes 31: 13‐19.

1 Corresponding Author
322 Tectonics – Recent Advances

[4] Lahtinen R, Garde A, Melezhik VA (2008) Palaeoproterozoic evolution of Fennoscandia


and Greenland. Episodes 31: 20‐28.
[5] Van Gool JAM, Connelly J N, Marker M, Mengel FC (2002) The Nagssugtoqidian orogen
of West Greenland: tectonic evolution and regional correlation from a West Greenland
perspective. Can J Earth Sci 39: 665–686.
[6] Sidgren A.‐S, Page L, Garde AA (2006) New hornblende and muscovite 40Ar/39Ar
cooling ages in the central Rinkian fold belt, West Greenland. Geol Surv Den Greenl
Bull 11: 115–123.
[7] Friend PD, Kinny PD (2001) A reappraisal of the Lewisian Gneiss Complex:
geochronological evidence for its tectonic assembly from disparate terranes in the
Proterozoic. Cont Min Pet 142:198–218.
[8] Park R G (2005) The Lewisian terrane model: a review. Scott J Geol 41:105–118.
[9] Griffin WL, Taylor PN, Hakkinen JW, Heier KS, Iden IK, Krogh EJ, Malm O, Olsen KI,
Ormaasen DE, Tveten E (1978) Archaean and Proterozoic crustal evolution in Lofoten‐
Vesterålen, N. Norway. J Geol Soc Lond 135: 629‐647.
[10] Corfu F (2004) U‐Pb age, setting and tectonic significance of the anorthosite‐mangerite‐
charnockite‐granite suite, Lofoten‐Vesterålen, Norway. J Pet 45:1799‐1819.
[11] Olesen O, Torsvik T, Tveten E, Zwaan KB, Løseth H, Henningsen T (1997) Basement
structure of the continental margin in the Lofoten‐Lopphavet area, northern Norway:
constraints from potential field data, on‐land structural mapping and paleomagnetic
data. Norw J Geol 77:15‐30.
[12] Blystad P, Brekke H, Færseth RB, Larsen BT, Skogseid J, Tørudbakken B (1995)
Structural elements of the Norwegian continental shelf. Part II: The Norwegian Sea
Region. Norw Petr Dir Bull 8: 1‐45.
[13] Henkel H (1991) Magnetic crustal structures in Northern Fennoscandia. Tectonophysics
192: 57‐79.
[14] Gorbatschev R (2004) The Transscandinavian Igneous Belt – introduction and
background. In Högdahl K, Andersson UB, Eklund O, editors. The Transscandinavian
Igneous Belt (TIB) in Sweden: a review of its character and evolution. Geol Surv Finland
Specl Paper 37: 9–15.
[15] Lahtinen R, Korja A, Nironen M (2005) Palaeoproterozoic tectonic evolution of the
Fennoscandian Shield. In: Lehtinen M, Nurmi P, Rämö T, editors. The Precambrian
Bedrock of Finland‐ Key to the evolution of the Fennoscandian Shield. Elsevier Science
BV: 418–532.
[16] Corfu F, Armitage PEB, Kullerud K, Bergh SG (2003) Preliminary U‐Pb geochronology
in the West Troms Basement Complex, North Norway: Archaean and Palaeoproterozoic
events and younger overprints. Geol Surv Norway Bull 441: 61‐72.
[17] Bergh SG, Kullerud K, Armitage PEB, Zwaan KB, Corfu F, Ravna EJK, Myhre PI (2010)
Neoarchaean to Svecofennian tectono‐magmatic evolution of the West Troms Basement
Complex, North Norway. Norw J Geol 90: 21‐48.
[18] Bamford D, Nunn K, Prodehl C, Jacob B (1978) Crustal structure of Northern Britain.
Geophys J Int 54: 43‐60
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 323

[19] Park RG, Cliff RA, Fettes DJ, Stewart AD (1994) Precambrian rocks in northwest
Scotland west of the Moine Thrust: the Lewisian Complex and the Torridonian. Geol
Soc Lond Spec Rep 22:6‐22
[20] Woodcock N, Strachan RA (2000) Geological History of Britain and Ireland. Blackwell
Science Ltd.
[21] Park RG (1995) Palaeozoic Laurentia‐Baltica relationships: a view from the Lewisian. In
Coward MP, Ries AC, editors, Early Precambrian processes. Geol Soc Lond Spec Publ
105:299–310.
[22] Bergh SG, Kullerud K, Holdsworth RE, Corfu F, Armitage PEB, McCaffrey K, Ravna E,
Wilson RW, Chattopadhyay A (2007) The West Troms Basement Complex, North
Norway: A segment of the Lewisian crust assembled by multiple Archaean through
Palaeoproterozoic collision events. Abstr, Geol Soc Am, Continental Tectonics and
Mountain Building, Peach & Horne Centennial Meeting, NW Scotland.
[23] Bleeker W, Ernst R (2006) Short‐lived mantle generated magmatic events and their dyke
swarms: The key unlocking Earthʹs paleogeographic record back to 2.6 Ga. In Hanski E,
Mertanen S, Rämö T, Vuollo J, editors. Dyke Swarms ‐ Time Markers of Crustal
Evolution. Balkema Publishers, Rotterdam, 3‐26.
[24] Mertanen S, Korhonen F (2011) Paleomagnetic constraints on an Archean‐
Paleoproterozoic Superior‐Karelia connection: New evidence from Archean Karelia.
Prec Res 186:193‐204.
[25] Bridgwater D, Austrheim H, Hansen BT, Mengel F, Pedersen S, Winter J (1990) The
Proterozoic Nagssugtoqidian mobile belt of southeast Greenland: a link between the
Eastern Canadian and Baltic Shields. Geosci Can 17:305–310.
[26] Connelly JN, Van Gool JAM, Mengel FC (2000) Temporal evolution of a deeply eroded
orogen: the Nagssugtoqidian Orogen, West Greenland. Can J Earth Sci 37: 1121‐1142.
[27] Armitage PEB, Bergh SG (2005) Structural development of the Mjelde‐Skorelvvatn Zone
on Kvaløya, Troms: a metasupracrustal shear belt in the Precambrian West Troms
Basement Complex, North Norway. Norw J Geol 85:117‐132.
[28] Kullerud K, Corfu F, Bergh SG, Davidsen B, Ravna E K (2006) U‐Pb constraints on the
Archaean and Early Proterozoic evolution of the West Troms Basement Complex, North
Norway. Abstr Bull Geol Soc Finl Spec Issue I, pp. 79.
[29] Bergh SG, Kullerud K, Corfu F, Armitage PEB, Davidsen B, Johansen HW, Pettersen T,
Knudsen S (2007) Low‐grade sedimentary rocks on Vanna, North Norway: a new
occurrence of a Palaeoproterozoic (2.4‐2.2 Ga) cover succession in northern
Fennoscandia. Norw J Geol 87: 301‐318.
[30] Myhre PI, Heaman LM, Bergh SG (2010) Svecofennian (c. 1780 Ma) metamorphic zircon
ages from the West Troms Basement Complex, northern Norway. In NGF Abstr Proc
Geol Soc Norway. pp. 128‐129.
[31] Myhre PI, Corfu F, Bergh S (2011) Palaeoproterozoic (2.0–1.95Ga) pre‐orogenic
supracrustal sequences in the West Troms Basement Complex, North Norway. Precamb
Res 186: 89‐100.
324 Tectonics – Recent Advances

[32] Myhre PI (2011) U‐Pb geochronology along a Meso‐Neoarchean geotransect in the West
Troms Basement Complex, North Norway. Unpubl PhD thesis, University of Tromsø,
Norway.
[33] Bleeker W (2003) The late Archaean record: a puzzle in ca. 35 pieces. Lithos 71: 99‐134.
[34] Kullerud K, Skjerlie KP, Corfu F, DeLaRosa J (2006) The 2.40 Ga Ringvassøy mafic
dykes, West Troms Basement Complex, Norway: The concluding act of Early
Palaeoproterozoic continental breakup. Prec Res 150:183‐200.
[35] Tveten E (1978) Geological map of Norway, bed rock map Svolvær 1:250 000, Geological
Survey of Norway
[36] Andresen A, Tull JF (1983) The age of the Lødingen granite and its possible regional
significance. Norw J Geol 63:269‐276.
[37] Coward MP, Park RG (1987) The role of mid‐crustal shear zones in the Early Proterozoic
evolution of the Lewisian. In Park RG, Tarney J, Editors. Evolution of the Lewisian and
Comparable Precambrian High Grade Terrains. Geol Soc Lond Spec Publi 27: 127‐138
[38] Zwaan KB (1995) Geology of the Precambrian West Troms Basement Complex, northern
Norway, with special emphasis onthe Senja Shear Belt: a preliminary account. Geol
Surv Norw Bull 427: 33‐36.
[39] Kullerud K, Skjerlie KP, Corfu F, DeLaRosa J (2006) The 2.40 Ga Ringvassøy mafic
dykes, West Troms Basement Complex, Norway: The concluding act of Early
Palaeoproterozoic continental breakup. Prec Res 150:183‐200.
[40] Zwaan KB, Fareth E, Grogan PW (1998) Geologic map of Norway, bed rock map
Tromsø, M 1:250.000. Geol Surv Norw.
[41] Zwaan KB, Tucker RD (1996) Absolute and relative age relationships in the
Precambrian West Troms Basement Complex, northern Norway (Abstract). 22nd Nord
Geol Wint Meet, Finland, pp 237.
[42] Zozulya D, Kullerud K, Ravna EK, Corfu F, Savchenko Y (2009) Geology, age and
geochemical constraints on the origin of the Late Archaean Mikkelvik alkaline massif,
West Troms Basement Complex in Northern Norway. Norw J Geol 89:327‐340
[43] Zwaan KB (1989) Berggrunnsgeologisk kartlegging av det prekambriske
grønnsteinsbeltet på Ringvassøy, Troms. Geol Surv Norw Report 89:101
[44] Motuza G, Motuza V, Beliatsky B, Savva E (2001) Volcanic rocks of the Ringvassøya
Greenstone Belt (North Norway):Implication for the Stratigraphy and Tectonic Setting. J
Conf (Abstract) EUG XI, 6: 578.
[45] Binns RE, Chroston PN, Matthews DW (1980) Low‐grade sediments on Precambrian
gneiss on Vanna, Troms, Northern Norway. Geol Surv Norw Bull 359: 61‐70.
[46]Andresen A (1979) The age of the Precambrian basement in western Troms, Norway.
Geol För Stockh Förh 101: 291‐298.
[47] Krill AG, Fareth E (1984) Rb‐Sr whole‐rock ages from Senja, North Norway. Norw J
Geol 64:171‐172.
[48] Condie KC (2005) TTGs and adakites: are they both slab melts? Lithos 80: 33-44
[49] Henderson I, Kendrick M (2003) Structural controls on graphite mineralisation, Senja,
Troms. Geol Surv Norw Report 2003.011: 111pp.
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 325

[50] Doré AG, Lundin ER, Fichler C, Olesen O (1997) Patterns of basement structure and
reactivation along the NE Atlantic margin. J Geol Soc Lond 154: 85‐92.
[51] Gjerløw E (2008) Petrologi og alder av høymetamorfe mafiske bergarter i det vestlige
gneiskomplekset i Troms. Unpubl MS thesis, Univ Tromsø, 90 pp.
[52] Herr W, Wolfe R, Kopp E, Eberhardt P (1967) Development and recent applications of
the Re/Os dating method. In Radioactive dating and methods of low‐level counting. Int
Atomic Energy Agency, Vienna, pp 499‐508.
[53] Heier KS, Compston W (1969) Interpretation of Rb‐Sr age patterns in high grade
metamorphic rocks, north Norway. NorGeol Tidsskr 49:257‐283.
[54] Heier KS (1960) Petrology and geochemistry of high‐grade metamorphic and igneous
rocks on Langøy, Northern Norway. Nor Geol Tidsskr 207:1‐246.
[55] Mjelde R, Sellevoll MA, Shimamura H, Iwasaki T, Kanazawa T (1993) Crustal structure
beneath Lofoten, N. Norway from vertical incidence and wide‐angle seismic data.
Geoph J Int 114:116‐126.
[56] Taylor PN (1975) An Early Precambrian Age for Migmatitic Gneisses from Vikan i Bø,
Vesterålen, North Norway. Earth Plan Sci Lett 27:35‐42.
[57] Jacobsen SB, Wasserburg GJ (1978) Interpretation of Nd, Sr and Pb isotope data from
Archaean migmatites in Lofoten‐Vesterålen, Norway. Earth Plan Sci Lett 41:245‐253.
[58] Wade SJR (1985) Radiogenic Isotope Studies of Crust‐forming Processes in the Lofoten‐
Vesterålen Province of North Norway. Unpubl PhD Thesis, Univ Oxford, vol. I:1‐285,
vol. II:1‐292.
[59] Taylor PN (1974) Isotope Geology and Related Geochemical Studies of Ancient High‐
Grade Metamorphic Basement Complexes: Lofoten and Vesterålen, North Norway.
Unp PhD thesis, Oxford Univ:1‐187.
[60] Markl G (1998) The Eidsfjord anorthosite, Vesterålen, Norway: field observations and
geochemical data. Nor Geol Tidsskr434: 51‐73.
[61] Olsen KI (1978) Metamorphic petrology and fluid‐inclusion studies of granulites and
amphibolite‐facies gneisses on Langøy and W Hinnøy, Vesterålen, N Norway. Unpubl
Cand Real. thesis, Univ Oslo, 214 pp.
[62] Malm O, Ormaasen DE (1978) Mangerite‐charnockite Intrusives in the Lofoten‐
Vesterålen Area, North Norway: Petrography, Chemistry and Petrology. Nor Geol
Tidsskr 338: 83‐114.
[63] Ormaasen, D.E. 1977: Petrology of the Hopen mangerite‐charnockite intrusion, Lofoten,
north Norway. Lithos 10, 291‐310.
[64] Markl G, Frost BR, Bucher K (1998) The origin of anorthosites and related rocks from
the Lofoten Islands, Northern Norway: I. Field relations and estimation of intrinsic
variables. J Petrol 39:1425‐1452.
[65] Markl G, Frost BR (1999) The origin of anorthosites and related rocks from the Lofoten
Islands, Northern Norway: II. Modelling of parental melts for anorthosites. J Petrol 40:
61‐77.
[66] Brueckner, H.K. 1971: The age of the Torset granite, Langöy, Northern Norway. Norsk
Geologisk Tidsskrift 51, 85‐87.
326 Tectonics – Recent Advances

[67] Romer RL, Kjøsnes B, Korneliussen A, Lindahl I, Skyseth T, Stendal M, Sundvoll B


(1992) The Archaean‐Proterozoic boundary beneath the Caledonides of northern
Norway and Sweden: U‐Pb, Rb‐Sr and ε Nd isotope data from the Rombak‐Tysfjord
area. Geol Surv Norw Report 91.225, 67 pp.
[68] Markl G, Höhndorf A (2003) Isotopic constraints on the origin of AMCG‐suite rocks on
the Lofoten Islands, N Norway. Min Petrol 78:149‐171.
[69] Markl G (2001) REE constraints on fractionation processes of massive‐type anorthosites
on the Lofoten Islands, Norway. Min Petrol 72:325‐351.
[70] Sutton J, Watson J (1951) The pre‐Torridonian metamorphic history of the Loch
Torridon and Scourie areas in the North‐west Highlands, and its bearing on the
chronological classification of the Lewisian. Q J Geol Soc Lond 106:241‐307
[71] Sutton J, Watson J (1962) Further observations on the margins of the Laxfordian
Complex of the Lewisian near Loch Laxford, Sutherland. Trans Roy Soc Edin 65: 89‐106
[72] Kinny PD, Friend CRL, Love GJ (2005) Proposal for a terrane‐based nomenclature for
the Lewisian Gneiss Complex of NW Scotland. J Geol Soc Lond 162:175‐186.
[73] Evans CR,Tarney J (1964) Isotopic ages of Assynth dykes. Nature Lond 204: 638‐641
[74] Evans CR (1965) Geochronology of the Lewisian basement near Lochinver, Sutherland.
Nature 207: 54–56.
[75] Park RG (1970) Observations on Lewisian chronology. Scott J Geol 6: 379‐399.
[76] Heaman LM, Tarney J (1989) U‐Pb baddeleyite ages for the Scourie dyke swarm,
Scotland: evidence for two distinct intrusion events. Nature 340: 705‐708
[77] Attfield P (1987) The structural history of the Canisp Shear Zone. In: Park RG, Tarney J,
editors. Evolution of the Lewisian and Comparable Precambrian High‐Grade Terrains.
Geol Soc Lond Spec Publ 27: 165–173.
[78] Coward MP (1984) Major shear zones in the Precambrian crust; examples from NW
Scotland and southern Africa and their significance. In Kroner A, Greiling SR, editors.
Prec Tect Illustr, Stuttgart, pp 207‐235.
[79] Park RG, Crane A, Niamatullah M (1987) Early Proterozoic structure and kinematic
evolution of the southern mainland Lewisian. In Park RG, Tarney J, editors, Evolution
of the Lewisian and Comparable Precambrian High Grade Terrains. Geol Soc Lond
Spec Publ 27: 139‐151.
[80] Goodenough KM, Park RG, Krabbendam M, Myers J, Wheeler J, Loughlin SC,Crowley
QG, Friend CRL, Beach A, Kinny P, Graham R (2010) The Laxford Shear Zone: an end‐
Archaean terrane boundary? Spec Publ Geol Soc 335:103‐120.
[81] Davies FB (1976) Early Scourian structures in the Scourie‐Laxford region and their
bearing on the evolution of the Laxford Front. J Geol Soc Lond 132: 543–554.
[82] Chattopadhya A (2007) Digital mapping and analysis of continental basement shear
zones. Unpubl report, Royal Soc Lond 56pp
[83] Giletti BJ, Moorbath S, Lambert RStJ (1961) A geochronological study of the
metamorphic complexes of the Scottish Highlands. Q J Geol Soc Lond 117: 233‐272
[84] Moorbath S, Welke H, Gale NH (1969) The significance of lead isotope studies in ancient
high‐grade metamorphic basement complexes, as exemplified by the Lewisian rocks of
northwest Scotland. Earth Planet Sci Lett 6: 245‐256
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 327

[85] Pidgeon RT, Bowes DR (1972) Zircon U‐Pb ages of granulites from the Central Region of
the Lewisian, northwesternScotland. Geol Mag 109: 247‐258
[86] Chapman HJ, Moorbath S (1977) Lead isotope measurements from the oldest recognised
Lewisian gneisses of north‐west Scotland. Nature 268: 41‐42
[87] Whitehouse MJ, Moorbath S (1986) Pb‐Pb systematics of Lewisian gneisses ‐
implications for crustal differentiation. Nature 319: 488‐489
[88] Cohen AS, OʹNions RK, OʹHara MJ (1991) Chronology and mechanism of depletion in
Lewisian granulites. Contrib Mineral Petrol 106: 142‐153
[89] Friend CRL, Kinny PD (1995) New evidence for protolith ages of Lewisian granulites,
northwest Scotland. Geology 23: 1027‐1030
[90] Kinny PD, Friend CRL (1997) U‐Pb isotopic evidence for the accretion of different
crustal blocks to form the Lewisian Complex of northwest Scotland. Contrib Mineral
Petrol 129:326‐340
[91] Love GP, Kinny PD, Friend, CRL (2004) Timing of magmatism and metamorphism in
the Gruinard Bay area of the Lewisian Gneiss Complex: comparisons with the Assynt
Terrane and implications for terrane accretion. Contr Mineral Petrol 146:620‐636
[92] Whitehouse MJ, Kemp AIS (2010) On the difficulty of assigning crustal residence,
magmatic protolith and metamorphic ages to Lewisian granulites: constraints from
combined in situ U‐Pb and Lu‐Hf. In Law RD, Butler RWH, Holdsworth RE,
Krabbendam M, Strachan RA, editors. Continental Tectonics and Mountain Building:
The Legacy of Peach and Horne. Geol Soc Lond Spec Publ 335: 81‐101.
[93] Corfu F, Heaman LM, Rogers G (1994) Polymetamorphic evolution of the Lewisian
complex, NW Scotland, as recorded by U‐Pb isotopic compositions of zircon, titanite
and rutile. Contrib Mineral Petrol 117:215‐228
[94] Corfu F (2007) Comment to paper: Timing of magmatism and metamorphism in the
Gruinard Bay area of the Lewisian gneiss complex: comparison with the Assynt Terrane
and implications for terrane accretion by D.J. Love, P.D. Kinny and C.R.L. Friend (Contr
Mineral Petrol (2004) 146:620‐636). Contr Min Petr 153: 483–488 doi: 10.1007/s00410‐006‐
0157‐5
[95] Zhu XK, OʹNions RK, Belshaw NS, Gibb AJ (1997) Lewisian crustal history from in situ
SIMS mineral chronometry and related metamorphich textures. Chem Geol 136: 205‐
218.
[96] Whitehouse MJ, Claesson S, Sunde T, Vestin J (1997) Ionmicroprobe U‐Pb zircon
geochronology and correlation of Achaean gneisses from the Lewisian complex of
Gruinard Bay, northwestern Scotland. Geoch Cos Acta 61: 4429‐4438
[97] Corfu F, Crane A, Moser D, Rogers G (1998) U‐Pb zircon systematics at Gruinard Bay,
northwest Scotland: implications for the early orogenic evolution of the Lewisian
Complex. Contrib Mineral Petrol 133: 329–345
[98] Whitehouse MJ, Bridgwater D (2001) Geochronological constraints on
Palaeoproterozoic crustal evolution and regional correlations of the northern Outer
Hebridean Lewisian Complex, Scotland. Precamb Res 105: 227–245.
[99] Whitehouse MJ (2003) Rare earth elements in zircon: a review of applications and case
studies from the Outer Heebrides Lewisian Complex, NW Scotland. In Vance D, Muller
328 Tectonics – Recent Advances

W, Villa JM, editors. Geochronology: Linking the Isotopic Record with Petrology and
Textures. Geol Soc Lond Special Publ 220: 49‐64.
[100] Mason AJ, Brewer TS (2005) A re‐evaluation of a Laxfordian terrane boundary in the
Lewisian Complex of South Harris, NW Scotland. J Geol Soc Lond 162: 401–408.
[101] Kelly NM, Hinton RW, Harley SL, Appleby SK (2008) New SIMS U‐Pb zircon ages
from the Langavat Belt, South Harris, NW Scotland: implications for the Lewisian
terrane model. J Geol Soc Lond 165: 967‐981.
[102] Love GJ, Friend CRL, Kinny PD (2010) Palaeoproterozoic terrane assembly in the
Lewisian Gneiss Complex on the Scottish mainland, south of Gruinard Bay: SHRIMP U‐
Pb zircon evidence. Precamb Res 183: 89‐111.
[103] Humphries FJ, Cliff RA (1982) Sm‐Nd dating and cooling history of Scourian
granulites, Sutherland. Nature 295: 515‐517
[104] Mason AJ, Parrish RR, Brewer TS (2004) U‐Pb geochronology of Lewisian orthogneiss
in the Outer Hebrides, Scotland: implications for the tectonic setting and correlation of
the South Harris Complex. J Geol Soc Lond 161: 45–54.
[105] Tarney J, Weaver BL (1987) Geochemistry of the Scourian complex: petrogenesis and
tectonic models. In: Park RG, Tarney J, editors. Evolution of the Lewisian and
comparable Precambrian high‐grade terrains. Geol Soc Spec Publ 27, pp 45‐56
[106] Weaver BL, Tarney J (1981) The Scourie dyke suite: petrogenesis and geochemical
nature of the Proterozoic sub‐continental mantle. Contrib Min Petrol 78: 175‐188.
[107] Park RG, Tarney J (1987) The Lewisian complex: a typical Precambrian high‐grade
terrain? Geol Soc Lond Spec Publ 27: 13‐25 DOI: 10.1144/GSL.SP.1987.027.01.03
[108] Chapman HJ (1979) 2390 Myr Rb‐Sr whole‐rock for the Scourie dykes of north‐west
Scotland. Nature 277: 642‐643
[109] Waters FG, Cohen AS, OʹNions RK, OʹHara MJ (1990) Development of Archaean
lithosphere deduced from chronology and isotope chemistry of Scourie dykes. Earth
Planet Sci Lett 97: 241‐255
[110] Park RG, Tarney J, Connelly JN (2001) The Loch Maree Group: Palaeoproterozoic
subduction‐accretion complex in the Lewisian of NW Scotland. Precambr Research
105, 205‐226.
[111] Park RG (2002) The Lewisian geology of Gairloch. Geol Soc Lond Mem 26: 76pp.
[112] Whitehouse MJ, Bridgwater D, Park RG (1997) Detrital zircons from the Loch Maree
Group, Lewisian complex, NW Scotland: confirmation of a Palaeoproterozoic
Laurentia‐Fennoscandian connection. Terra Nova, 9: 260‐263.
[113] Cliff RA, Gray CM, Huhma H (1983) A Sm–Nd isotopic study of the South Harris
Igneous complex, the Outer Hebrides. Contrib. Miner. Petrol. 82, 91–98.
[114] Barooah BC, Bowes DR (2009) Multi‐episodic modification of high‐grade terrane near
Scourie and its significance in elucidating the history of the Lewisian Complex. Scott J
Geol 45: 19–41.
[115] Van Breemen O, Aftalion M, Pidgeon RT (1971) The age of the granite injection
complex of Harris, Outer Hebrides. Scott. J. Geol. 7: 139–152.
Was the Precambrian Basement of Western Troms and Lofoten-Vesterålen in Northern Norway Linked to the
Lewisian of Scotland? A Comparison of Crustal Components, Tectonic Evolution and Amalgamation History 329

[116] Pidgeon RT, Aftalion M (1972) The geochronological significance of discordant U‐Pb
ages of oval‐shaped zircons from a Lewisian gneiss from Harris, Outer Hebrides. Earth
Planet Sci Lett 17: 269‐274.
[117] Buchan KL, Mortensen JK, Card KD, Percival JA (1998) Paleomagnetism and U‐Pb
geochronology of diabase dyke swarms of Minto block, Superior Province, Quebec,
Canada. Can J Earth Sci 35: 1054‐1069.
[118] Pesonen LJ, Elming S‐u, Mertanen S, Pisarevsky S, DʹAgrella‐Filho MS, Meert JG,
Schmidt PW, Abrahamsen,N, Bylund G (2003) Palaeomagnetic configuration of
continents during the Proterozoic: Tectonophysics 375: 289–324.
[119] Mertanen S, Vuollo JI, Huhma H, Arestova NA, Kovalenko A (2006) Early
Paleoproterozoic‐Archean dykes and gneisses in Russian Karelia of the Fennoscandian
Shield–New paleomagnetic, isotope age and geochemical investigations: Precambr Res
144: 239–260.
[120] Karlstrom KE, Williams M, McLelland J, Geissman JW, Åhall K‐I (1999) Refining
Rodinia: Geological evidence for the Australia‐Western U.S. connection in the
Proterozoic. GSA Today 9: 2‐7.
[121] Koistinen T, Stephens M B, Bogatchev V, Nordgulen Ø, Wennerström M, Korhonen J
(compilers) (2001) Geological map of the Fennoscandian Shield, scale 1:2 000 000. Espoo,
Geol Surv Finl; Geol Surv Norw, Geol Surv Sweden, Ministry of Natural Resources of
Russia, Moscow.
[122] Krogh TE (1993) High precision U‐Pb ages for granulite metamorphism and
deformation in the Archean Kapuskasing structural zone, Ontario: implications for
structure and development of the lower crust. Earth Planet Sci Letters 119: 1‐18.
[123] Moser DE, Heaman LM (1997) Proterozoic zircon growth in Archean lower‐crustal
xenoliths, southern Superior craton – a consequence of Matachewan ocean opening,
Contrib Min Petrol 128: 164‐175.
[124] Hanski E, Huhma H, Vaasjoki M (2001) Geochronology of northern Finland: a
summary and discussion. In Vaasjoki M, editor. Radiometric age determinations from
Finnish Lappland and their bearing on the timing of Precambrian volcano‐sedimentary
sequences. Geol Surv Finl Spec Paper 33: 255‐279.
[125] Nironen M (1997) The Svecofennian Orogen: a tectonic model. Precambr Res 86: 21‐44.
[126] Williams H, Hoffman PF, Lewry JF, Monger JWH, Rivers T (1991) Anatomy of North
America: thematic geological portrayals of the continent: Tectonophysics 187: 117–134.
[127] Ernst R, Bleeker W (2010) Large igneous provinces (LIPs), giant dyke swarms, and
mantle plumes: significance for breakup events within Canada and adjacent regions
from 2.5 Ga to the Present. Can J Earth Sci 47: 695‐739.
[128] Buchan KL, Mertanen S, Park RG, Pesonen LJ, Elming S‐A, Abrahamsen N, Bylund G
(2000) The drift of Laurentia and Baltica in the Proterozoic: a comparison based on key
palaeomagnetic poles. Tectonophysics 319: 167‐198.
[129] Connelly JN, Van Gool JAM, Mengel FC (2000) Temporal evolution of a deeply eroded
orogen: the Nagssugtoqidian Orogen, West Greenland. Can J Earth Sci 37: 1121‐1142.
330 Tectonics – Recent Advances

[130] Pesonen LJ, Elming S, Mertanen S, Pisarevsky S, DʹAgrella‐Filho MS, Meert JG,
Schmidt PW, Abrahamsen N, Bylund G (2003) Palaeomagnetic configuration of
continents during the Proterozoic. Tectonophysics 375: 289‐324.
[131] Connelly JN, Mengel FC (2000) Evolution of Archean components in the
Paleoproterozoic Nagssugtoqidian orogen, West Greenland. Geol Soc Amer Bull 112:
747‐763.
[132] Muir RJ, Fitches WR, Maltman AJ, Bentley MR (1994) Precambrian rocks of the
southern, Inner Hebrides ‐ Malin Sea region: Colonsay, West Islay, Inishtrahull and
Iona. In Gibbons W, Harris AL editors. A Revised Correlation of Precambrian Rocks in
the British Isles., Geol Soc Lond Spec Rep 22:54‐58
[133] Daly JS, Muir RJ, Cliff RA (1991) A precise U‐Pb zircon age for the Inishtrahull syenite
gneiss, County Donegal, Ireland. J Geol Soc Lond 148: 639‐642.

You might also like