You are on page 1of 32

Tectonophysics 316 (2000) 255286

www.elsevier.com/locate/tecto

Tertiary tectonic evolution of the external


East Carpathians (Romania)
L. Matenco a, *, G. Bertotti b
a Bucharest University, Faculty of Geology and Geophysics, 6 Traian Vuia str., sect. 1, 70139 Bucharest, Romania
b Department of Sedimentary Geology, Vrije Universiteit, De Boelelaan 1085, 1081 HV Amsterdam, Netherlands
Received 16 February 1999; accepted for publication 1 October 1999

Abstract
Paleostress calculation and analysis of mesoscopic structures are integrated with depth interpreted geological
profiles based on seismic studies and well correlation to derive a Tertiary tectonic model for the East Carpathians.
Following Early Miocene and older orogenic phases, the first tectonic event that affected the studied area is
characterised by a WSWENE-oriented shortening of Middle Miocene (Late Burdigalian) in age. Resulting
deformations induced ENE-ward thrusting of Tarcau and Marginal units, as well as the internal part of the
Subcarpathian nappe. A second shortening event with an EW to WSWENE contraction direction took place in the
Late Miocene (Sarmatian), characterised by further foreland thrusting of the Subcarpathian nappe and out-ofsequence deformation in the Tarcau and Marginal Folds nappes. Along strike, differences in deformation mechanisms
are controlled by the friction coefficients along the main detachment layers, by the lateral variations in the wedge
thickness and by the involvement in the northern part of the thrusting system of the thick, competent East European
platform. Tear faulting occurred in both tectonic events, the main resulting structure being the triangle zone developed
south of the Trotus valley. The Latest Miocene (Latest Sarmatian)Early Pliocene is characterised by a strikeslip
stress field with NNESSW compression and WNWESE tension axis, left-lateral faults being dominant. The last
deformation which affected the studied area is characterised by NNWSSE shortening during the Pliocene, major
deformations taking place mainly in the SW-most bending zone. 2000 Elsevier Science B.V. All rights reserved.
Keywords: East Carpathians; Tertiary; subsurface data; paleostress; tectonic model

1. Introduction
The Romanian segment of the Carpathians is a
highly arcuate orogenic belt formed in response to
subduction and continental collision between the
European and Apulian plates and related microplates during the Alpine orogeny (Sandulescu,
* Corresponding author. Tel.: +40-1-2117390;
fax: +40-1-2113120.
E-mail addresses: matl@gg.unibuc.ro (L. Matenco),
bert@geo.vu.nl (G. Bertotti)

1984; Royden, 1988; Csontos, 1995) ( Fig. 1). The


Carpathians consist of a nappe pile of crystalline
rocks with Upper Paleozoic to Mesozoic sedimentary cover and, in an external position, a
Lower Cretaceous to Tertiary thin-skinned belt.
The Alpine tectonic evolution of the Carpathians
is traditionally subdivided into Triassic to Early
Cretaceous extension followed by Middle
Cretaceous
to
Pliocene
shortening
(e.g.
Sandulescu, 1984). Three main Tertiary deformational stages are recognised (Csontos, 1995 and
references cited therein). During PaleogeneEarly

0040-1951/00/$ - see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S0 04 0 - 1 95 1 ( 9 9 ) 00 2 6 1- 9

256

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

Fig. 1. Tectonic sketch map of the Carpathians system and the location of the studied areas. 1, Central part of the East Carpathians.
2, Southern part of the East Carpathians. TF=Trotus Fault; IMF=Intramoesian Fault.

Miocene times, clockwise rotation of the Tisza


Dacia block (Csontos, 1995), part of the
PannonianCarpathians system, caused NNE
SSW to ENEWSW shortening in the internal
Moldavides nappes (Convolute flysch and
Audia/Macla nappes, Fig. 2). Middle and Late
Miocene (BadenianSarmatian) deformations led
to EW shortening, which caused further deformation of the external East Carpathians. Late
Miocene to Pliocene NWSE (to NS) shortening
in the East Carpathians led to further deformation

mainly concentrated in the external parts of the


junction zone between the East and South
Carpathians.
The evolution of the Carpathians belt is characterised by temporal changes of stress and strain
fields. This is shown by an increasing amount of
structural data (Ratschbacher et al., 1993; Fodor
et al., 1996; Huismans et al., 1997; Matenco, 1997;
Linzer et al., 1998; Zweigel et al., 1998) and
required by the arcuate shape of the belt (e.g.
Csontos, 1995). Models assuming a roughly con-

Fig. 2. Schematic structural cross-section in the central part of the East Carpathians (simplified from Stefanescu and working group,
1988). Location of the section in Fig. 1.

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

temporaneous emplacement of thrust sheets in the


various segments of the Carpathians (e.g.
Sandulescu, 1984, 1988) are unlikely because of
the absence of structures able to accommodate the
required coeval large orogen-parallel extension
(e.g. Morley, 1996). Other models, envisaging the
Carpathians as mainly due to E-ward translation
of the Intra Carpathians units (Royden, 1988;
Ellouz and Roca, 1994; Linzer, 1996) are at odds
with structural data and with the absence of largescale transcurrent movements within the South
Carpathians (e.g. Rabagia and Fulop, 1994;
Matenco et al., 1997a). Most recent models (e.g.
Royden, 1988; Csontos, 1995; Linzer, 1996) lack
a real integration between outcrop structural
analysis with seismically imaged structures buried
in the foredeep (Dicea et al., 1966; Dicea, 1995,
1996; Tari et al., 1997). As a result, important
differences exist between these models, mainly concerning the timing and especially the motion directions through time.
Several kinematic studies have been published
on the South Carpathians (e.g. Ratschbacher et al.,
1993; Matenco et al., 1997a; Schmid et al., 1998)
and on the bend zone connecting the East and
South Carpathians (e.g. Morley, 1996; Hyppolite
and Sandulescu, 1996). Much less is known in the
East Carpathians, especially in the central areas
stretching from the Slovakian and Polish
Carpathians to the Vrancea bending area (Fig. 1).
Classical studies on the external East Carpathians
deal with the external part of the chain and with
the stratigraphic evolution of the foredeep mainly
based on surface/outcrop studies (e.g. Joja et al.,
1968; Bancila, 1958; Ionesi, 1971; Sandulescu,
1984; Sandulescu et al., 1981a,b and references
cited therein), but generally lack good structural
control.
In this paper we aim to fill the gap in knowledge,
presenting new structural data from the segment
of the East Carpathians comprised between the
northern Romanian border and the Intramoesian
fault to the south. Integrating these data with
interpreted seismic profiles, we reconstruct the
kinematic evolution of the East Carpathians
during Tertiary times. We devote particular attention to less studied topics, such as the role of
strikeslip faulting, the relations between structures in the subducting plate and structures in the

257

fold-and-thrust belt, and the Tertiary kinematics


of the Trotus and Intramoesian fault systems.

2. The structure of the external East Carpathians:


depth data analysis
The East Carpathians are schematically made
up of a stack of basement nappes (internal
Dacides) with crystalline rocks and a Mesozoic
sedimentary cover, tectonically overlying the
internal (Ceahlau, Curbicortical ) and external
(Audia/Macla, Tarcau, Marginal Folds) nappe
system (Sandulescu, 1984, 1988). The latter units
are thrust over the Subcarpathian nappe, which is
itself carried over the undeformed foreland
( European, Scythian and Moesian platforms)
(Sandulescu, 1984) ( Fig. 2). Thrusting respected,
in general, a foreland propagating sequence. The
chain has a non-cylindrical shape and ages of
thrusting change along strike, possibly as a consequence of the pre-existing structural grain ( Ellouz
and Roca, 1994). The total shortening of the East
Carpathians outer units from Late Oligocene to
Present is about 180 km ( Ellouz et al., 1994).
Turbiditic and other clastic units forming the
external East Carpathians nappes ( Tarcau,
Marginal and Subcarpathian) and their undeformed foredeep have been deposited in a roughly
eastward thinning basin associated with Tertiary
thrusting of the Carpathians nappes. Sediments of
the basin fill are mainly derived from the hinterland
but, especially after the Eocene, sediment input
from the external areas became significant
(Sandulescu et al., 1981b). Sediment facies and
depositional geometries were influenced by NW
SE-trending paleo-highs inherited from Late
PermianMiddle Jurassic ( Ellouz and Roca, 1994)
and Paleogene (Sandulescu, 1992) extensional
structures which were inverted during Late
JurassicEarly Cretaceous and post-Senonian
shortening episodes, respectively.
We present the main features of the external
East Carpathians units and of their foredeep with
two regional geological maps and 14 regional
sections based on published and unpublished
seismically controlled profiles numbered from I in
the north to XIV in the south (Figs. 3 and 4).

258

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

Fig. 3. Geological structural map of the Central East Carpathians. Compiled from Sandulescu et al. (1981b); geological maps
1:200,000 and 1:50,000, published by the Geological Institute of Romania and results of this study. Thick, grey lines, SI to SV,
indicate the position of the geological sections. OHW, Oituz half-window; BHW, Bistrita half-window; GHO, Gura Humorului
outlier. Location of the map is shown in Fig. 1.

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

259

Fig. 4. Simplified geologicalstructural map of the southern part of East Carpathians (modified after Sandulescu et al., 1981b) with
the location of the depth-interpreted profiles used in the present study. OHW, Oituz half-window; VHW, Vrancea half-window; SS,
Slanic syncline; DS, Drajna syncline; BA, Breaza anticline. Location of the map is shown in Fig. 1.

260

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

While detailed structural and stratigraphic images


have been obtained for the European platform
and for the Subcarpathian nappe, steep dips and
lack of good seismic reflectors have made difficult
the unravelling of the deep structure of the external
nappes ( Tarcau, Marginal ). Subsurface images in
these units are mainly based on correlation
between the surface geology and the deep wells.
Consequently, the shallow levels of the sections
are generally well constrained, while the deeper
ones are relatively uncertain.
In this work, the time scale used for the Middle
Late Miocene to Pliocene deposits represents a
local combination of Central and Eastern
Paratethys regional stages. Correlation with the
global Tethys stages (Rogl, 1996) is schematically
drawn in Fig. 5.
2.1. Stratigraphy and sedimentology of the nappes
and of the foreland
The Tarcau (Sandulescu, 1984) or Medio-marginal (Bancila, 1955) nappe is the most internal of
the three considered units (Figs. 3 and 4). It is
composed of Lower Cretaceous to Turonian black
and variegated shales and marls followed by a
mainly turbiditic succession of Senonian to Late
Oligocene age, which becomes finer-grained from
SW to NE (Dumitrescu, 1952) ( Fig. 5). Senonian
to Middle Eocene marls and subordinate sands
are laterally replaced by limestones and sandshale
alternations. Eocene deposits change from thick,
coarse-grained sandstones ( Tarcau sandstone) to
finer-grained, often shaly turbidites in the NE.
These formations are overlain by a thin calcareous
limestone unit, the Doamna limestone, developed
especially east and north of the Bistrita valley, and
by shalymarlysandy turbidites, Globigerina

261

marls and a thin sandstone unit (Joja et al., 1968)


( Fig. 5). The Oligocene is represented by thick
sandstones,
sandymarly
turbidites
and
wildflysch/olistolite laterally continuous with menilites and bituminous marls, bituminous paper
shales interlayered with siliceous sandstones
( Kliwa) (Joja et al., 1968) ( Fig. 5). According to
Sandulescu et al. (1981a), these rocks were shed
from external zones.
The Marginal Folds nappe (Dumitrescu, 1952)
(also named External or Vrancea nappe) (Figs. 3
and 4) can be followed discontinuously along the
belt. Outcrops in the studied area are limited to
the half-windows (Bistrita, Oituz, Vrancea) or to
the rabotage outliers (sensu Sandulescu, 1984)
( Figs. 3 and 4). From N to S, the width of the
Marginal unit increases up to the Trotus valley,
and then decreases southward. The unit disappears
south of the Vrancea half-window, (Figs. 3 and 4).
Sediments of the Marginal Folds nappe are
similar to those of the eastern part of the Tarcau
nappe (Fig. 5). Thin Lower Cretaceous to
Paleocene black shales, variegated shales and other
pelagic rocks are found at the base of the nappe
units (Sandulescu et al., 1981a). Fine-grained turbidites with coarse intercalations are characteristic
for the Paleocene to Lower Eocene and are capped
by the regionally widespread Doamna limestone
( Upper Eocene). This is followed by Globigerina
beds, sandstones, bituminous rocks of Oligocene
age and by the Upper Oligocene Kliwa quartzarenites (Sandulescu et al., 1981a; Ionesi, 1971). The
youngest deposit found in the Marginal Folds
nappe is the Lower Miocene salt formation, only
locally overlain by molasse type sediments (Hirja)
(Sandulescu et al., 1981a) (Fig. 5).
The easternmost allochthonous unit is the
Pericarpathian (Mrazec and Popescu-Voitesti,

Fig. 5. General time correlation table, stratigraphic column for the Tarcau, Marginal and Subcarpathian units (modified after
Sandulescu et al., 1981a) and tectonogram of the main CretaceousTertiary tectonic events for the East Carpathians. Correlation
with Central and Eastern Paratethys for the Oligocene and Miocene ages after Rogl (1996); T, Global Tethys stages; PT, Paratethys.
Gray areas represent the ages used in this study. Note especially the differences at the Miocene/Pliocene boundary between the ages
used in the present study and the standard Tethys scale. A, B, C represent the internal, median and external sedimentary facies of
the studied units. For B and C, only variations in respect to A were drawn. 1, 2, 3 represent results of the first, second and, respectively,
third paleostress data sets in the present paper. Deformation patterns, paleostress fields and tectonic events for the CretaceousLower
Miocene and for the internal flysch and East Carpathians inner basement were taken from the results of Sandulescu (1984, 1988),
Csontos (1995), Matenco (1997), Schmid et al. (1998).

262

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

1914) or Subcarpathian (Sandulescu, 1984) nappe.


It is mainly formed by non-outcropping Eocene to
Oligocene clastics similar to those of the Marginal
Folds nappe followed by molasse type sediments
interlayered with two main evaporitic levels
( Fig. 5). The Lower Salt formation ( Early
Miocene) is found in the internal areas and is
replaced towards the east by molasse conglomerates and the Gray Schlier formation. This salt
formation acted during the Miocene deformation
as an important decollement horizon, and its lateral distribution affected the thrusting geometries
(see later). Badenian tuffs, gypsum, calcareous
sandstone
and
fine-grained
quartzarenites
(Sandulescu et al., 1981a) follow and are overlain
by Lower Sarmatian coarse deposits which are the
youngest sediments of the Pericarpathian nappe
( Fig. 5).
The Subcarpathian nappe is thrust on the East
Carpathians foreland platforms ( Fig. 6). The East
Carpathians foreland is formed by the Paleozoic
to Cretaceous coalescence of three main lithospheric blocks named the East European, Scythian and
Moesian platforms (Fig. 6). The western boundary
of the East European platform is formed by
the TornquistTeisseire zone [ TTZ, or TransEuropean Suture Zone ( TESZ)], a tectonic lineament of lithospheric importance stretching from
Sweden to the Black Sea (e.g. Zielhuis and Nolet,
1994). The TTZ separates very thick and cold East
European plate in the NE from thinner and
warmer Moesian lithosphere in the SW. In the
study area ( Fig. 6), the major lineaments related
to this zone are the CampulungBicaz fault and,
S of the Trotus fault, the PecenagaCamena fault
(Cantini et al., 1991), which juxtapose TTZ over
the Scythian platform and the North Dobrogea
orogen (Fig. 6). In the East European platform,
deep seismic reflection and refraction profiles show
thicknesses of about 10, 20 and 4045 km for the
base of the sedimentary cover, Conrad and Moho
discontinuities, respectively ( Enescu et al., 1988,

263

1992; Raileanu et al., 1994). The Paleozoic


Tertiary sedimentary cover (including the foredeep) of the East European platform, about 6
12 km (Raileanu et al., 1994), decreases towards
the east. In contrast, deep reflection seismic profiles
and seismological data show crustal thickness
values of 3440 km for the Moesian platform, with
a thicker sedimentary cover (Radulescu, 1988;
Enescu et al., 1992).
The East Carpathians foreland is flexed in front
of the fold-and-thrust belt. This flexure allowed
for the formation of an undeformed foredeep
basin. The base of the foredeep, represented by
the top Mesozoic, deepens both towards the W
and S from an average of 1500 m in the north to
5000 m in the region of Bistrita valley and to
800010,000 m south of Trotus valley (Dicea,
1995). Flexure is accompanied by NWSEtrending normal faults with offsets up to 12 km
(e.g.
VicovPaltionoasaBacau,
StrajaGura
Humorului; Dicea, 1995) ( Fig. 6 and profile 1,
Fig. 7). Reflection seismic surveys have also
revealed regional NESW- to EW-trending
transverse faults (Fig. 6) (Dicea, 1995).
Since the front of the East Carpathians is not
always parallel to the TTZ, there is a difference in
the substratum on which the external nappes and
the foredeep lie. In the northern sector, the East
Carpathians have passed the TTZ and, together
with their foredeep, overlay the thick East
European plate (e.g. Botezatu and Calota, 1983;
Guterch et al., 1986; Pinna et al., 1991). To the
south, the lithospheric transition is in a more
easterly position and the East Carpathians nappes
and foredeep are still overlying the thinner
Moesian platform (Fig. 6). These features have
important implications, which will be discussed in
a following section.
2.2. Main structures along the profiles
The deep structure of the external East
Carpathians was analysed through 14 geological

Fig. 6. Tectonic map of the East Carpathians foreland platforms (compiled and modified from Sandulescu, 1984; Sandulescu and
Visarion, 1988; Dicea, 1995; Ellouz et al., 1994). Hatched area represents the TornquistTeissere zone (Cantini et al., 1991), contour
lines represent the Bouguer anomaly (Mocanu and Radulescu, 1994). CBF, CampulungBicaz Fault; SF, StrajaGura Humorului
fault; ScF, Solca fault; SiF, Siret fault; BF, Bistrita fault; TF, Trotus fault; PCF, Peceneaga Camena fault.

264

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

profiles ( Figs. 79, and position in Figs. 3 and 4).


These profiles illustrate the studied units, from the
internal one ( Tarcau) to the external nappe
(Subcarpathian).
The Tarcau nappe is thrust by more internal
units (Audia, Macla nappes) during Aquitanian
Early Burdigalian times (Sandulescu, 1984, 1988).
Towards the east, the Tarcau nappe almost completely covers the Marginal Folds nappe which
outcrops only in tectonic windows such as the
Bistrita, Oituz and Vrancea half-windows (Figs. 3
and 4). The eastern boundary of the Tarcau nappe
is erosional. The overall structure of the Tarcau
nappe is a ramp anticline, with associated imbricate fan of thrusts (see for instance profiles VI
VIII, Fig. 8). The internal structure of the nappe
( Figs. 79) is mainly characterised by low angle
imbricate thrusts. The larger ones define digitations
(e.g. Tarcau, Tazlau) (Sandulescu, 1984). Along
the strike of the belt, backthrusts can be documented on the basis of field structures and depth
interpretations. Previous studies assume that these
faults represent overturned structures, their dipping direction changing at depth towards the hinterland ( Fig. 10A). Locally, the faults are
documented at depth and display a normal offset,
within CretaceousEocene deposits. Our interpretation assumes that these structures are in fact
sets of backthrusts, which may connect with foreland propagating thrusts along pop-ups. Partial
inversion of former extensional faults can account
for the normal offsets still preserved at depth
( Fig. 10B). Backthrusting occurred also on the
rear limb of ramp anticlines producing pop-up
structures such as in the Moldova valley region
(profile I, Fig. 7), or in the Buzau valley region
(profiles IX, X, Fig. 8).
The Marginal fold nappe is basically composed
of large-scale duplexes, hinterland-dipping to antiformal-stack types. Internal, low- to intermediatedip faults merge at depth in the major detachment
surfaces and are often associated with recumbent
folds (Figs. 79).

265

The Late Burdigalian age for the Tarcau and


Marginal nappes shortening is based on the syntectonic character of the Upper Burdigalian sediments deposited within growing synclines in the
frontal part of the Marginal Folds nappe (e.g.
Matenco, 1991; Dicea, 1995) and on the Lower
Middle Burdigalian age of the salt breccia formation and overlying complexes which are the youngest deposits below the Marginal sole thrust.
Deformation continued during Early Badenian, as
suggested by locally overthrust beds (Slanic Tuff
and Salt Formation) (Sandulescu, 1988).
The internal structure of the Subcarpathian
nappe (Figs. 79) is characterised mainly by NE
to SE-vergent thrusts and high angle reverse faults,
locally organised in imbricate fans, and by faultrelated folding. Diapirs of Lower Burdigalian salt
are common (e.g. in profiles V Fig. 7; profiles VII,
IX Fig. 8; profile XIII Fig. 9). Where the belt
strikes NNWSSE to NS (i.e. the central sectors),
deformation is accommodated by faults with relatively high angles, higher than usually observed in
fold-and-thrust belts (profiles VVIII, Figs. 7 and
8), associated with widespread tear faults and en
echelon folds ( Figs. 3 and 4). This possibly reflects
thrusting associated with dextral movements along
fault planes. The faults become flatter towards the
north, where the belt strike changes to NWSE
(profiles IIII, Fig. 7) and would therefore suggest
only thrusting. The transition from one domain to
the other occurs between the southern termination
of the Bistrita half window and the Trotus valley
( Fig. 3). The amount of internal shortening is
estimated at 57% in profile III (Matenco, 1991)
and is representative for the external nappes of
East Carpathians in the northern sector.
2.3. The contact between nappes and the
undeformed foreland
The Subcarpathian nappe is thrust towards the
E onto the East European/Scythian/Moesian platforms. It is difficult to estimate the amount of

Fig. 7. Geological profiles in the centralnorthern part of the East Carpathians derived from surface geology and interpretation of
seismic sections. The names of the nappes outcropping are indicated in the upper part of the sections. Location of the profiles in Fig. 3.

266

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

Fig. 8. Geological profiles in the centralsouthern part of the East Carpathians derived from surface geology and interpretation of
seismic sections. Location of the profile in Fig. 4.

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

267

Fig. 9. Geological profiles in the southernmost part of the East Carpathians derived from surface geology and interpretation of
seismic sections. Location of the profiles in Fig. 4.

displacement because of insufficient information


on the platforms beneath the main thrust, but a
minimum value of 1525 km can be observed
along all profiles (Figs. 7 and 8).
The kinematics and especially the age of thrusting along the frontal thrust are controversial and

need to be described in some detail. In all profiles,


the frontal thrust cuts Lower to Middle Sarmatian
deposits and is therefore younger. The upper age
limit is less constrained because the geometry of
the frontal zone changes along strike.
In the northern profiles (from I to IV ) the

268

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

Fig. 10. Cartoon illustrating (A) previous structural interpretation of hinterland-vergent thrust faults overturned in depth to
hinterland-dipping normal faults (e.g. Stefanescu and working
group, 1988); (B) inversion mechanism of an inherited normal
fault along a frontal thrust and a backthrust, organised in a
pop-up structure.

major thrust outcrops at the surface, cutting


through Lower to Middle Sarmatian sediments
( Fig. 7). Consequently, the age for the cessation
of thrusting is poorly constrained. Some secondorder complications might be present in profiles
IIII, where the thrust is decomposed in a leading
imbricate fan. Locally, Uppermost Sarmatian
Lowermost Meotian sediments seal the frontal sole
thrust, indicating a Late Sarmatian thrusting age.
Moving southward the geometry of the frontal
zone changes and the Subcarpathian nappe is
unconformably overlain by Upper Sarmatian
deposits dipping to the E (profiles V to IX ). This
contact has been traditionally considered in subsurface interpreted profiles as stratigraphic and,
therefore, indicative of the end of thrusting.
However, the overall position, and particularly the
E-ward dip of the Upper Sarmatian reflectors,
suggest that this surface is a backthrust. In the
frontal areas, therefore, a triangle zone is formed,
with the described backthrust compensating the
displacement of the Pericarpathian fault. Younger
salt diapiric movements overprint the boundary
zone at the front of the Subcarpathian nappe
(MagirestiPerchiu line, after Sandulescu, 1984;
e.g. profile V, Fig. 7). The dip of the backthrust is
relatively shallow in the Trotus valley region, but
increases towards the south (see for instance profile
VII and VIII ). These features point to a post
Upper Sarmatian age of thrusting.

In the southernmost part of the East


Carpathians (profiles XIXIV, Fig. 9) the
Pericarpathian frontal thrust is precisely dated,
cutting Lower to Middle Sarmatian deposits but
being sealed by the Uppermost Sarmatian
sediments.
In the central bending area, Latest Miocene
Early Pliocene shortening along the frontal sole
thrust was accommodated by out-of-sequence
thrusting in the more internal parts of the orogen
and Transylvania basin (Sanders, 1998; Ciulavu,
1999).

3. Field structures
Structural data were collected in the external
nappes ( Tarcau, Marginal, Subcarpathian) of the
East Carpathians and in the adjacent foredeep
between the Moldova valley in the north, and the
Oituz valley in the south ( Fig. 3). For the regional
correlation we have used literature data (e.g.
Sandulescu, 1984, 1988; Sandulescu et al., 1981a,b;
Stefanescu and working group, 1988; Micu, 1990;
Morley, 1996), and geological maps 1:200,000 and
1:50,000 published by the Geological Institute of
Romania.
3.1. Data and methods
Brittle structures such as fault striations, folds,
tension joints, fault-related folds (fault propagation, drag folds), regional-scale faults have been
analysed in 90 stations in Upper Cretaceous to
Sarmatian sediments belonging to the thrust sheets.
Regional paleostress directions were reconstructed
using fault slip data sets collected in 67 stations
along the belt. For a similar and more complete
description of the methods used, see Matenco and
Schmid (1999). Fault planes with slickensides are
the most common structures measured. Roughly
1400 faults with direction and sense of slip were
measured, each site having between nine and 120
measurements. The slip sense was deduced from
kinematic indicators along the fault plane, such as
mineral steps, tension gashes, Riedel shears, fractures with tension planes, in-plane conjugate shear
fractures, conjugate fault planes (for a complete

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

review see Angelier, 1994), or rare shear bands


(Simpson and Schmid, 1983) in the case of more
complicated shear zones. The quality of slip sense
was classified in the field as certain (37% of fault
population), probable (54%), supposed (12%), or
unknown (7%). Relative chronological constraints
were obtained in the field in approximately 30%
of the locations, using criteria such as cross-cutting
relationships, successive striations along a fault
plane and reactivations of conjugate faults. In each
location, subsets of faultslip data consistent with
various stress directions were separated, on the
basis of the stress regime type/orientation and on
the chronological constraints. Where a sufficiently
large number of faults/subsets could be measured,
data sets were analysed using three methods:
(1) The Angelier (1984, 1989) inverse method,
that finds the best possible fit between observed
faultslip data and computed shear stresses generated on the fault planes. We have used the Delvaux
(1993) method and software, which starts from
the tensor given by the right dihedron method
(Angelier and Mechler, 1977). This tensor is first
optimised automatically ( least squares method )
and second manually, in order to obtain the best
mentioned fit (usually minimising a mean slip
deviation). The limitations of this method concerning the tensors and isotropy are discussed in
detail elsewhere (e.g. Etchecopar et al., 1981;
Angelier, 1984; Dupin et al., 1993; Pollard et al.,
1993). The method allows for the definition of
the orientation of principal stress tensors
(s s s ) and the ratio R between stress magni1
2
3
tudes [R=(s s )/(s s )] for a single deforma2
3
1
3
tion period ( Tables 1 and 2 and Figs. 11A, 13A,
14A). On the basis of the slip sense quality and
TQR indicator (tensor quality rank, Delvaux 1993;
Delvaux et al., 1997), each tensor was classified as
good (TQR1.5, 7% from the total number of
tensors), medium (0.5TQR<1.5, 63%) and poor
(0.3TQR<0.5, 21%) quality ( Table 1), while
non-reliable tensors (0.3<TQR, 9%) were rejected.
(2) The Turner (1962) PT axes method
( Figs. 11C, 13C, 14C ), that assumes for every
shear plane a contraction (P) and an extension
axis (T ) in the plane given by the slip line and the
shear plane normal. The angle between the con-

269

traction axis and the shear plane is a direct function


of the material properties.
(3) The Spang (1972) numeric dynamic analysis
method. In addition to the Turner method, the
Spang (1972) method is computing a reduced
stress tensor, the relative values of the principal
stresses being calculated from the mean values and
vectors of the bulk stress tensor.
The most reliable stress axis determinations are
those provided by the Angelier method. In cases
of ambiguous solutions we have used the other
two methods to double-check and further constrain
the stress field.
In places with a low number of measurements,
the inversion method was combined with paleostress determinations from two conjugate faults
(e.g. Angelier, 1984), minimisation of s on tension
n
joints (Delvaux, 1993) and fault-related fold axis
analysis. For the sake of simplicity, only the two
conjugate faults were plotted as faults with slip
sense (stereoplots in Figs. 11A, 13A, 14A).
For the definition of the stress field we have
used the nature of the (sub)vertical stress axis and
the value of ratio R (Delvaux et al., 1997), if
computed ( Table 1). For the regional stress field,
all the stress tensors related to a given deformational stage, as obtained by these methods, were
plotted in a single diagram ( Figs. 11B, 13B, 14B).
Statistical calculations of the cone of confidence
for each deformation stage and regional deviations
were performed, using the Wallbrecher (1986)
method and software.
Regional timing constraints were obtained
through correlating the chronological constraints
observed in the field with the deformation age of
the sediments, observed at the surface and in
interpreted geological profiles (i.e. orientation of
fault planes, similar type of kinematics and chronology). In addition, the age of the ENEWSW
compressional stress regime (see below) was correlated with the fission-track cooling ages obtained
by Sanders (1998), which according to this author
are associated with the Late Miocene shortening
in the studied area. Most of our new data were
collected in the northern part of the East
Carpathians, between the Moldova Valley and the
Oituz half-window ( Fig. 3). For regions to the
south we have used data from the literature

270

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

Table 1
Location stations and parameters of paleostress reconstruction. s s s , azimuth and dip of principal stress axes. R ratio=stress
1
2
3
ellipsoid shape factor, R=(s s )/(s s ). a represents the mean slip deviation between the measured kinematic indicator on fault
2
3
1
3
plane and the orientation of the calculated shear stress. n/N represents the number of faults generating a stress tensor versus the total
number of faults in the index. TQR=tensor quality rank (Delvaux et al., 1997), TQR=n(n/N )/a. Stress fields may vary from
extension (s vertical ), with pure extension (0.25<R<0.75) and transtension (0.75<R<1), to strikeslip stress fields (s vertical ),
1
2
with pure strikeslip (0.25<R<0.75), transtension (0.75<R<1) and transpression (0<R<0.25), to compression (s vertical ), with
3
pure compression (0.25<R<0.75) and transpression (0<R<0.25). Radial extension (s vertical, 0<R<0.25) and radial compres1
sion (s vertical, 0.75<R<1) have been rejected from the calculation, being subjectively considered non-conclusive. Sn, Senonian;
3
Pg1-lt, PaleoceneLutetian; Pg2, Eocene; Pg3, Oligocene; Bd, Burdigalian; Bn, Badenian; Sm, Sarmatian
Station

EC5-SEN
EC6-CP1
EC7-CP1
EC7-SEN
EC8-SEN
EC9-CP2
EC11a-SEN
EC11b-SEN
EC12-CP1
EC14-SEN
EC15-CP1
EC16-SEN
EC18-CP1
EC18-CP2
EC21-SEN
EC22-SEN
EC22-CP2
EC23-SEN
EC24-CP1
EC24-SEN
EC25-CP1
EC25-SEN
EC26-CP1
EC26-SEN
EC27-SEN
EC28-CP1
EC28-SEN
EC30-CP1
EC30-SEN
EC31-CP1
EC32-CP2
EC33-CP1
EC35-CP1
EC35-SEN
EC35-CP2
EC36-CP1
EC36-SEN
EC37-SEN
EC39-CP2
EC40-CP1
EC41-CP2
EC43-CP1
EC43-SEN

Latitude
(deg min s)

46 16 46
46 26 53
46 26 49
46 26 49
46 43 15
46 44 00
46 45 58
46 45 58
46 48 01
46 49 53
46 50 16
46 50 15
46 59 23
46 59 23
46 54 44
46 55 28
46 55 28
46 53 05
46 56 17
46 56 17
46 53 52
46 53 52
47 09 08
47 09 08
47 14 25
47 06 19
47 06 19
47 10 23
47 10 23
47 08 55
47 12 48
47 13 14
47 33 17
47 33 17
47 33 17
47 32 35
47 32 35
46 16 59
46 16 08
46 11 31
46 16 13
46 20 21
46 20 21

Longitude
(deg min s)

26 36 06
26 20 32
26 24 37
26 24 37
26 27 16
26 27 28
26 21 56
26 21 56
26 17 55
26 24 49
26 23 37
26 22 15
26 18 02
26 18 02
26 22 29
26 23 11
26 23 11
26 24 02
26 13 45
26 13 45
26 08 36
26 08 36
26 00 19
26 00 19
26 04 22
26 17 41
26 17 41
26 15 37
26 15 37
26 13 12
26 20 47
26 11 43
25 57 11
25 57 11
25 57 11
25 51 15
25 51 15
26 37 11
26 35 19
26 25 34
26 27 35
26 24 02
26 24 02

Direct inversion ( TENSOR)


s
1

340/03
104/24
221/23
257/14
25/29
165/14
348/21
63/09
260/03
87/39
262/16
200/06
73/17
161/08
35/03
196/22
170/04
61/07
55/01
24/12
262/10
10/20
255/19
61/13
29/11
275/12
222/10
75/19
31/07
250/03
296/09
251/03
217/07
214/05
159/28
234/14
181/23
170/21
140/00
210/21
175/20
237/35
210/22

233/80
3/23
324/27
118/72
242/55
258/10
192/67
309/69
170/03
253/50
353/05
60/82
166/11
258/43
287/80
68/57
264/40
187/78
325/08
279/50
170/13
226/66
163/04
167/49
284/53
183/09
111/64
338/21
246/81
159/10
203/16
160/03
310/23
320/73
269/33
326/08
314/58
324/67
230/11
119/03
83/06
146/01
66/63

71/10
236/56
97/53
349/12
125/17
20/72
81/09
156/19
37/86
352/07
100/73
291/05
288/69
63/46
126/09
296/23
76/49
330/09
152/82
123/37
29/74
105/13
61/70
321/38
127/34
57/75
316/23
204/61
121/05
357/80
55/72
22/86
111/66
123/16
38/44
85/74
82/21
76/10
50/79
22/69
338/69
55/55
306/14

Ratio

0.00
0.36
0.20
0.49
0.54
0.45
0.75
0.52
0.34
0.12
0.26
0.50
0.31
0.58
0.13
0.59
0.13
0.37
0.07
0.50
0.17
0.50
0.08
0.26
0.40
0.33
0.00
0.38
0.44
0.12
0.21
0.74
0.48
0.40
0.52
0.00
0.50
0.51
0.13
0.26
0.52
0.18
0.83

17.55
13.72
19.63
10.74
15.50
13.70
9.71
16.56
17.85
12.49
7.49
18.42
10.63
11.41
8.4
13.30
12.82
14.33
5.04
11.76
9.73
7.19
15.81
7.56
12.43
15.41
12.35
16.26
13.91
12.65
10.44
17.46
13.52
18.37
8.94
16.18
11.93
13.29
13.00
5.95
18.47
4.17
16.06

n/N

TQR

Rock age

6/06
16/22
9/12
7/09
6/06
19/21
11/13
7/08
6/06
14/20
20/27
7/08
8/10
10/12
25/30
11/13
13/17
7/10
5/05
10/11
16/25
5/05
9/15
6/08
11/18
11/17
10/10
7/08
10/12
12/18
16/19
13/20
14/17
14/18
14/15
11/15
8/11
16/20
13/19
16/22
12/16
12/13
16/17

C-0.34
B-0.84
C-0.34
B-0.50
C-0.39
B-1.25
B-0.95
C-0.36
C-0.33
B-0.78
A-1.97
C-0.33
B-0.60
B-0.73
A-2.48
B-0.69
B-0.77
C-0.34
B-0.99
B-0.77
B-1.05
B-0.69
C-0.34
B-0.59
B-0.54
C-0.46
B-0.80
C-0.37
B-0.59
B-0.63
B-1.29
C-0.48
B-0.85
B-0.59
B-1.46
B-0.51
C-0.48
B-0.96
B-0.68
A-1.95
C-0.48
A-2.65
B-0.93

Pg3
Pg3
Pg1-lt
Pg1-lt
Pg3
Pg3
Pg3
Pg3
Pg3
Pg3
Pg2
Pg3
Pg3
Pg3
Pg3, Bd
Pg3
Bd
Pg3
Pg3
Pg3
Pg1-lt
Pg1-lt
Sn
Sn
Sn
Pg2
Pg2
Pg2
Pg2
Pg3
Bd
Pg3
Pg3
Pg3
Pg3
Pg3
Pg3
Bd
Pg3
Pg3
Pg3
lt, Sm
lt, Sm

271

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286


Table 1 (continued )
Station

EC44a-CP2
EC44b-SEN
EC44b-CP2
EC46-CP1
EC46-SEN
EC47-CP2
EC48-SEN
EC48-CP2
EC49-CP2
EC49-CP1
EC50-CP1
EC50-SEN
EC50-CP2
EC51-SEN
EC52-CP1
EC53-CP1
EC53-SEN
EC54-SEN
EC55-CP2
EC56-SEN
EC57-SEN
EC58-SEN
EC59-SEN
EC60-SEN
EC60-CP2
EC61-CP1
EC61-CP2
EC62-CP1
EC63-CP1
EC64-CP1
EC64-SEN
EC65-SEN
EC66-SEN
EC67-CP1

Latitude
(deg min s)

46 20 10
46 19 54
46 19 54
46 29 54
46 29 54
46 33 39
46 34 37
46 34 37
46 26 21
46 26 21
46 27 08
46 27 08
46 27 08
46 31 37
46 33 26
46 35 23
46 35 23
46 37 35
46 40 28
46 48 55
46 42 54
46 46 59
46 53 34
46 56 21
46 56 21
46 56 14
46 56 14
47 03 46
47 35 00
47 32 32
47 32 32
47 27 45
47 32 15
47 35 10

Longitude
(deg min s)

26 14 08
26 21 28
26 21 28
26 29 04
26 29 04
26 27 39
26 27 02
26 27 02
26 22 12
26 22 12
26 11 46
26 11 46
26 11 46
26 30 30
26 31 06
26 29 32
26 29 32
26 26 24
26 26 34
26 08 36
26 12 11
26 06 24
26 03 36
26 05 56
26 05 56
26 07 21
26 07 21
26 03 17
25 40 50
25 48 08
25 48 08
25 48 26
25 52 49
25 51 24

Direct inversion ( TENSOR)


s
1

177/48
37/38
125/22
250/25
220/40
130/06
30/19
330/11
310/15
236/23
250/22
47/30
152/01
215/06
288/34
80/04
45/04
41/23
145/25
33/20
29/09
235/26
18/29
355/23
160/09
203/21
323/04
205/16
212/19
255/05
43/12
248/13
205/11
232/23

83/03
172/42
226/26
355/28
20/48
35/40
209/71
105/75
69/61
331/12
155/12
227/60
62/08
313/53
197/02
172/26
299/75
225/67
240/11
219/70
264/74
81/62
240/53
192/66
255/30
97/35
65/72
115/01
121/02
164/08
228/78
1/60
358/78
331/20

(Morley, 1996; Hyppolite and Sandulescu, 1996;


Zweigel et al., 1998; Ciulavu, 1999) integrated with
our own observations.
3.2. Central East Carpathians
We present evidence for three major Tertiary
tectonic events in the studied units, illustrated in
an old-to-young succession. For each stage we will
discuss first the general stress parameters and the
fault data, and second the significant associations
of structures observed in outcrops. Large, mapscale structures that can be correlated with these

350/42
286/24
0/55
126/51
121/10
227/50
300/00
238/11
213/24
87/64
39/65
317/00
251/82
120/37
104/56
342/64
136/14
131/02
351/63
124/02
121/13
330/11
120/21
88/06
55/58
318/47
232/18
22/74
25/71
15/80
133/01
151/27
114/05
98/59

Ratio

0.47
0.69
0.25
0.33
0.34
0.06
0.30
0.29
0.50
0.30
0.12
0.75
0.35
0.31
0.40
0.56
0.14
0.53
0.50
0.59
0.53
0.50
0.63
0.43
0.43
0.03
0.00
0.20
0.29
0.30
0.18
0.65
0.68
0.40

9.29
17.17
13.55
21.10
15.58
16.37
14.45
12.37
12.75
6.99
9.69
11.82
13.70
16.39
14.06
18.73
13.84
15.75
8.17
15.28
16.72
10.81
14.99
12.11
13.72
12.25
16.41
13.65
8.51
12.82
11.81
13.31
18.08
11.55

n/N

TQR

Rock age

15/18
8/08
8/14
21/24
8/11
13/19
7/07
14/15
11/14
8/09
7/07
12/12
12/15
17/21
8/10
16/19
13/18
25/33
10/17
21/27
8/11
10/13
15/20
9/14
12/16
7/07
9/10
18/20
9/13
8/10
21/27
8/09
7/07
16/17

B-1.34
B-0.52
C-0.33
A-0.87
C-0.37
B-0.54
C-0.48
B-1.05
B-0.67
A-1.01
B-0.72
B-1.01
B-0.70
B-0.83
C-0.45
B-0.71
B-0.67
B-1.20
B-0.71
B-1.06
C-0.34
B-0.71
B-0.75
C-0.47
B-0.65
B-0.57
B-0.51
B-1.18
B-0.73
B-0.51
B-1.38
B-0.53
C-0.38
B-1.30

lt
Sn
Sn
Pg3
Pg3
Pg3
Pg3
Pg3
lt
lt
lt
lt
lt
Pg3, Bd
Pg3
Pg3
Pg3
Pg3
Pg3
lt
lt
Pg1
lt
Pg1
Pg1
Sn
Sn
Sn
Pg3
Pg3
Pg3
Pg3
Bn
Pg1-lt

deformations are presented in a later section of


the paper.
3.2.1. WSWENE shortening
The first data set (stereoplots in Fig. 11) is
characterised by a compressional stress regime
with WSWENE-oriented s (Fig. 11). A fairly
1
good concentration of the obtained tensors is
observed (Fig. 11B). The slip deviation factor ( b)
which compares the observed faults with the
shear plane determined by the tensor is low, suggesting that deformation was accommodated along
well-grouped fault directions ( Fig. 11C and D).

272

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

273

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

Table 2
Statistical calculations of the general parameters of the stress fields for each determined stage. Stress axes, R ratio, and slip deviation
are statistically averaged from the results obtained from the direct inversion method (Direct inversion), the numeric dynamic analysis
method (Numeric), and the PT axes method (PT axes). Statistical parameters obtained are preferred orientation (PO), concentration
parameter (CO), cone of confidence (CC ) and spherical aperture (SA) ( Wallbrecher, 1986)
Direct Inversion

Numeric

PT Axes

Statistic parameters
PO (%)

CP

CC ()

SA ()

Stage CP1
s
1
s
2
s
3
Ratio R
Slip deviation

241/15
338/04
64/80

12.81

238/11
148/03
44/78
0.56

245/08
155/02
52/82

70.70
72.49
84.91

6.56
6.99
12.74

15.14
14.59
10.43

32.77
31.63
22.86

Stage CP2
s
1
s
2
s
3
Ratio R
Slip deviation

149/09
245/14
26/70

12.81

148/16
245/24
28.61
0.38

148/15
244/20
23.65

83.73
75.06
79.24

11.52
7.40
8.89

14.49
21.04
18.94

23.79
29.96
27.10

Stage SEN
s
1
s
2
s
3
Ratio R
Slip deviation

30/06
252/82
117/07

13.71

38/08
252/81
129/05
0.64

36/08
250/81
132/4

65.63
79.75
61.66

5.65
9.60
5.07

14.09
10.37
15.07

35.89
26.74
38.26

The ratio R indicates pure compression for the


majority of the stations. A strikeslip component
associated with the inverse-dextral faults can be
defined only for four stations ( EC28, 30, 31, 33,
Fig. 11A). Generally, the s orientation tends to
1
be perpendicular to the dominant strike of the
belt, i.e. from EW to ENEWSW in the south to
NESW oriented s in the north (Fig. 11A). The
1
age of the youngest sediments in which paleostress
measurements were performed is Lower
Burdigalian.
Faults and folds that can be associated with

this WSWENE compression are common and


form the dominant grain of the belt. They trend
predominantly NNESSW to NS in the southern
and central sectors and NWSE in the northern
ones (Fig. 11A). Forethrusts dominate and are
commonly associated with backthrusts. Similar
features are observed in representative outcrops.
In station EC1, located near the base of the
Marginal Folds nappe (Fig. 12A) two sets of contractional structures are recorded. The older ones
are foreland-vergent folds often associated with
thrusts in their cores. These folds are truncated by

Fig. 11. (A) Geologicalstructural map of the studied area with paleostress tensors associated with the Late BurdigalianSarmatian
compressional event. Mean regional stress values of s =241/1515, 15, s =338/0415, s =64/8010, 15 and 10 being
1
2
3
the aperture of the cone for 95% confidence, were computed using the Wallbrecher (1986) method. Structures (faults, folds, nappe
contacts) active at the time are evidenced with thicker black lines. Stereoplots represent the paleostress results for the WSWENE
(CP1) shortening event. Only faults with a certain sense of movement have been plotted (70% of fault population). (B) Principal
stress axes derived from the direct inversion method. (C ) Projection of the measured compressional (small circles) and tensional
(small squares) axes for each fault in the measured set, and projection of mean compression, tension and medium directions, computed
using the Turner method. (D) Hanging-wall movements for all faults, and principal stress axes computed for the whole faults set
using the numeric dynamic analysis (NDA) method.

274

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

Fig. 12. (A) Sketch and stereoplot of the outcrop structures in station EC1. Note the development of hanging-wall anticlines and
recumbent folding and subsequent out-of-sequence thrusting. Stratigraphy: Lower menilites and bituminous marls formation
(Oligocene). Unit: Tarcau nappe. (B) Sketch and stereographic projections of the outcrop structures in station EC15. Stratigraphy:
Doamna limestone ( Eocene). Unit: Marginal Folds nappe. Description in the text. (C ) Sketch and stereographic projection of the
outcrop structures in station EC68. Stratigraphy: Upper gypsum formation. Unit: Subcarpathian nappe. Description in the text.

thrusts faults with correlative ramp folds. Vergence


of the younger thrusts is towards the ENE. The
shortening direction is ENEWSW for both sets
of structures, thereby suggesting that they belong
to the same deformation phase.
In station EC15 ( Fig. 12B), the Eocene
Doamna limestone forms an asymmetric anticline
(approximately 800 m wavelength) with E-ward
vergence and slight S-ward axial plunge. Secondorder structures such as parasitic folds, low-angle
thrust faults on the normal limb, high-angle inverse
faults on the steep limb and thrust-ramping along
the hinge are common. Faults formed in the initial
stages of deformation accommodating layer parallel shortening, and were subsequently passively
rotated during folding. Shortening directions are
compatible with those derived for the large-scale
fold.

In station EC68 ( Trotus valley) (Fig. 12C ),


outcrop-scale folding is associated with E-ward
thrusting. All the structures are formed in the
hinge of an approximately 200 m wavelength
anticline.
3.2.2. Strikeslip regime
The second data set (stereoplots of Fig. 13) is
characterised by a strikeslip regime with NNE
SSW-oriented s and WNWESE-oriented s
1
3
( Fig. 13). The concentration of the obtained tensors is good (Fig. 13B). The slip deviation from
the theoretical general tensor is low, reflecting
deformation along fault planes with fairly constant
strike along the belt (Fig. 13C and D). Most R
ratios demonstrate a pure strikeslip character.
Tensional/strikeslip R values are found mainly
in or near the Comanesti depression. A

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

275

Fig. 13. Results of the paleostress analysis for the Late Sarmatian strikeslip (SEN ) event. Mean regional stress values are
s =30/0614, s =252/8210, and s =117/0715. Conventions as in Fig. 11.
1
2
3

276

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

compressive/strikeslip character is observed in a


few stations located close to the contact with
Subcarpathian nappe, or in the nappe itself (e.g.
EC21, EC28, EC53, Fig. 13A). The youngest sediments in which these measurements were performed are Upper Burdigalian.
Outcrop-scale structures developed during this
stage are common and are mainly strikeslip faults
( Fig. 13A). Sinistral transcurrent faults dominate.
They strike ENEWSW in the central sectors and
NESW in the northern ones. Dextral, NNE
SSW conjugate faults are subordinate (Fig. 13A).
Large-scale structures associated with this stage
( Fig. 13) can be observed over the entire studied
area. Sinistral faults clearly dominate and are E
W to WNWESE directed in the area of Trotus
valley, and EW to WSWENE north of this
valley. Conjugate, NS oriented dextral faults were
formed locally as reactivation of the older roughly
NS trending thrust contacts.
3.2.3. NNWSSE shortening
The third data set (stereoplots of Fig. 14) is
characterised by a compressional regime with
roughly NNWSSE-oriented s (Fig. 14). The ori1
entations of the calculated tensors have a certain
degree of dispersion, reflected by high slip deviation from the general theoretical shear planes
( Fig. 14C and D). R values show pure compressional to strikeslip compressional character.
Three stations ( EC48, 29, 61) have a pure strike
slip character (Fig. 14A). The age of the youngest
sediments in which such tensors were obtained is
Lower Miocene.
Outcrop-scale faults ( Fig. 14A) are thrusts with
SE vergence and secondary thrusts with NW
vergence. Large-scale structures associated with
this deformation ( Fig. 14A) are rare. However,
SSE-vergent thrusts can be documented in the
Comanesti area, in the Tarcau nappe and around
Piatra Neamt.
3.3. Southern East Carpathians
Based on the work carried out in the area by
several authors (e.g. Hyppolite and Sandulescu,
1996; Morley, 1996; Zweigel et al., 1998) three
deformation stages can be identified.

In the Early MioceneSarmatian, NWSE compression led to the major thrusting and folding of
the Moldavides nappes ( Fig. 15). North of the
Buzau Valley, the maximum stress axes trend
WNWESE, and are thus parallel with the transport direction. These conclusions are compatible
with results obtained by Morley (1996) according
to whom regional Miocene shortening is characterised by a stress field with roughly EW contraction
direction north of the Buzau Valley, changing to
NWSE south of it (Fig. 15). These data are also
compatible with our observations from the northern segment of the East Carpathians.
No strikeslip stress field has been explicitly
reported for the external East Carpathians bending
area. However, a large number of published paleostress stations have a clear strikeslip character
and spatially coincide with large-scale transpressive
structures. For instance, a large part of the tensors
obtained by Hyppolite and Sandulescu (1996)
( Fig. 15) displays a strikeslip character with NE
SW compression axes. The age of this stress field
is considered by the authors as Miocene without
further specification. In addition, according to
Zweigel et al. (1998), brittle deformation of basement and Mesozoic sedimentary cover in the
internal part of the East Carpathians bend zone
occurred exclusively in strikeslip with NWSE
contraction direction and extensional modes. The
same authors recognised, in the external part of
the East Carpathians bend zone, two groups of
brittle structures. The first and larger group has
contraction axes oriented WNWESE to NNW
SSE, interpreted to be coeval with the Paleogene
Miocene shortening. The second group is younger,
has NS- to NNESSW-oriented contractional
axis, and is interpreted to be coeval with the Late
MioceneQuaternary shortening. Roughly 40% of
these paleostress stations have a strikeslip character with NWSE to NNESSW, both groups being
compatible with our observations and with the
strikeslip stress field with NWSE to NS direction observed in the external South Carpathians
(see also Ratschbacher et al., 1993; Matenco et al.,
1997a; Linzer et al., 1998). Since there is no real
evidence in the subsurface for any kind of
Paleogene tectonics (e.g. Stefanescu and working
group, 1988; Ionescu, 1994; Dicea, 1995, 1996; and

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

277

Fig. 14. Results of the paleostress analysis for the Pliocene shortening (CP2) event. Mean regional stress values are
s =149/0914, s =245/1421, s =26/7019. Conventions as in Fig. 11.
1
2
3

278

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

Fig. 15. Compilation of paleostress measurements performed in the southern East Carpathians bending area (after Morley, 1996;
Hyppolite and Sandulescu, 1996).

review of Sandulescu, 1988), we are correlating


the first group with the Middle to Late Miocene
shortening depicted by our study in the external
East Carpathians.

The Pliocene ( Wallachian) deformation stage is


characterised in the southern part of the East
Carpathians by roughly NS to NNESSW compressional axes (Hyppolite and Sandulescu, 1996)

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

as demonstrated by paleostress tensors in several


localities of the external nappes, especially between
the Dambovita and the Buzau Valleys (Fig. 15).

4. An integrated tectonic model


In this section we integrate the data derived
from surface structural work and the available
subsurface information into a kinematic model of
the Central East Carpathians during Miocene and
Pliocene times. Since Neogene to Quaternary structures are partly controlled by Paleogene deformation (e.g. Ellouz and Roca, 1994), a brief summary
is given on the Paleogene evolution.
4.1. Paleogene tectonics
The Paleogene history of the East Carpathians
is characterised by extensional deformation mainly
of Eocene age, stretching being accommodated by
NS-trending, mainly eastward-dipping normal
faults (Sandulescu, 1992). Some of these faults
have been imaged in seismic lines across the belt
( Figs. 7 and 8) where they appear to have been
subsequently inverted. Thick, coarse-grained sandstones in the western nappes ( Tarcau sandstones)
were deposited in the proposed extensional basin.
This extensional episode is correlative in geometry
and time with the Paleogene extension documented
in the South Carpathians, at the Getic/Danubian
nappes contact (Schmid et al., 1998; Matenco and
Schmid, 1999) and within the Getic Depression
(Matenco, 1997; Rabagia and Matenco, 1999).
4.2. Late BurdigalianSarmatian thrusting
The first tectonic event that affected the studied
area was ENEWSW- to ESEWNE-directed
shortening (Fig. 16A) responsible for the ENE- to
ESE-vergent thrusts present in all studied nappes.
Deformation began in Middle Miocene (Late
Burdigalian) times and continued until the
Sarmatian.
In the initial, Middle Miocene (Late
BurdigalianBadenian)
stages,
deformation
affected the more internal nappes of the system,
the Tarcau and Marginal Folds nappes. Foreland-

279

vergent imbricate thrusts, backthrusts, duplexes


and associated ramp folds with several hundred
metres to some kilometres offset developed during
this stage ( Figs. 79). The distribution and thickness of the competent Tarcau sandstone exerted a
fundamental control on the geometry of shortening. Large values are observed in the internal parts
of the Tarcau nappe where this horizon is largely
developed. North of the Bistrita valley or towards
the foreland, where this horizon is thinner, thrusts
are more closely spaced and the amount of internal
shortening higher ( Figs. 79). In the Marginal
Folds nappe (Figs. 79) deformation is accommodated by more tightly imbricated thrust faults,
large duplex systems and recumbent folding with
overturned foreland limbs. Detachment folds are
common in and above the Lower Cretaceous black
shales, which is the lowest detachment horizon in
the Tarcau and Marginal units (e.g. profile 7,
Fig. 8).
Beginning with the Sarmatian, the Tarcau nappe
was transported onto the Marginal Folds nappe
and thrusting continued its forward propagation,
affecting the Subcarpathian nappe with NNE
SSW- to NNWSSE-oriented thrusts and folds
( Figs. 3, 4 and 16A). Out-of-sequence thrusts and
backthrusts also formed in the more internal
Tarcau and Marginal units. Sedimentary basins
developed in a piggy-back fashion during the Early
to Middle Sarmatian, the most important being
the Comanesti basin ( Fig. 11A).
During the Sarmatian, the Subcarpathian nappe
was eventually thrust over the undeformed foreland along the Pericarpathian fault. The termination of movements along this thrust fault is
diachronous. It is Late Sarmatian north of the
Trotus fault and in the southernmost part of the
East Carpathians, but it becomes younger, Earliest
Pliocene, in the central bending area (Figs. 3 and
4). Differences in thrusting age were taken up by
various transfer zones, as for instance the Trotus
fault, a deep basement fault forming the boundary
between the East European and the Scythian platforms (Sandulescu and Visarion, 1988) ( Fig. 6),
which acted as a sinistral transfer zone during this
time interval.
A significant coeval structure is the northern
limit of the Bistrita half-window, where the

280

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

Fig. 16. Sketch of the Tertiary tectonic evolution of the Romanian East Carpathians. Description in the text.

Tarcau/Marginal contact is right-laterally transferred (Fig. 11A). This dextral movement affects the
trend of faults and fold axes, which swing from
NNWSSE to NESW and then back to NNW
SSE. Tear faults continue both towards more
internal areas and towards the foreland (Fig. 11A),
and are kinematically linked with a major WE
trending fault segmenting the foredeep basement
( Fig. 6), its northern block being roughly 2000 m
uplifted (Dicea, 1995). A major change in the
thrusting geometry occurs across this fault. The
large-scale antiformal stack developed in the
Marginal nappe and correlative ramp faults in the
Tarcau nappe (profile II, Fig. 7) present north of
the fault are replaced to the south by hinterlanddipping duplexes developed in the Marginal Folds
nappe associated with an imbricated thrust system
in the Tarcau nappe (profile III, Fig. 7).

4.3. Latest SarmatianEarly Meotian strikeslip


The second deformation stage developed in a
strikeslip regime with NNESSW to NS compressional axes (Fig. 16B). Structures are not
homogeneous across the entire Romanian East
Carpathians and two different zones can be
recognised.
In the northern zone, extending from the
Moldova Valley in the north to the Oituz halfwindow in the south ( Fig. 13A), strikeslip deformation was predominantly accommodated by
ENEWSW-trending sinistral faults. Sinistral
faults of regional importance develop at the southern termination of the Bistrita half-window and
along the Trotus valley ( Fig. 13A). The northern
sinistral fault has an offset of ca. 10 km and is
kinematically linked with the reactivation of a

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

deeper structure affecting the thrusted foreland


platform (Bistrita fault, Fig. 6; see also Dicea,
1995). The second major structure is the EW
Trotus fault, separating the Scythian and the
Moesian platforms. Its crustal importance is suggested by the 30 km sinistral displacement shown
by the Bouguer anomaly contours across the fault
zone (Fig. 6). The sinistral reactivation of the
Trotus fault caused widespread deformation in the
overlying flysch units (Fig. 13A). Faults belonging
to this system cut older thrusts both in outcrop
and in map view (e.g. the contact between Audia
and Tarcau nappes in the west and the frontal
thrust of the Subcarpathian nappe in the east,
Fig. 13A).
In the S-most part of the East Carpathians,
Latest SarmatianEarly Meotian deformation was
mainly accommodated by WNWESE- to NW
SE-trending dextral strikeslip faults (Fig. 4).
They are often transpressional, forming positive
flower structures (e.g. the Moreni, Cimpina and
Valenii de Munte lineaments, Fig. 9), the associated EW-trending anticlines being arranged en
echelon. In the frontal part of the Subcarpathian
nappe south of Buzau, EW-trending thrusts are
common (Fig. 4), their activation being related to
the reactivation of older pre-existing normal faults
during the Latest Sarmatian times (see also
Rabagia and Matenco, 1999).
Strain was partitioned between the sinistral
faults developing in the northern part of the
Romanian East Carpathians and the dextral faults
in the southern one. This resulted in a movement
towards the SSE of the central, intermediate sectors, and an increase of curvature of the southern
part of the fold belt (Fig. 4). As a result, contractional structures developed during this time span in
the bending area, compatible with the different
age, Latest SarmatianEarly Meotian, of the frontal sole thrust. In the internal zones, the SSE-ward
movement of the bending area is accompanied by
extension and strikeslip movements (e.g.
Girbacea and Frisch, 1998; Ciulavu, 1999).
The total displacement accommodated by
Latest SarmatianEarly Meotian strikeslip is of
the order of some tens of kilometres. Deformation
ended in the Early Pliocene, as documented by

281

Meotian sediments sealing numerous transcurrent


faults.
4.4. PliocenePleistocene thrusting
The last tectonic event affecting the area is the
roughly NNWSSE-oriented shortening (Fig. 16C ).
Deformation in the northern part of the studied area
is diffuse and mainly represented by small-scale
thrusting with SSE vergence (Fig. 14A). Thrusts
related to this stage cut Late Sarmatian strikeslip
fault and are therefore Pliocene or younger.
Large-scale structures associated with this stage
are common in the southern sectors, i.e. from the
Buzau Valley southward ( Fig. 4). Here, deformation in the Subcarpathian Nappe is mainly characterised by out-of-sequence, EW-trending thrusts,
which affect deposits as young as the Upper
PliocenePleistocene ( Fig. 9). Offsets up to 2000 m
are locally observed along these thrust faults (e.g.
the Apostolache, Urlati, Valenii de Munte, Valea
Lunga, Moreni structures, Fig. 9). The total
amount of deformation accommodated during this
interval is of the order of 1520 km (see also
Matenco et al., 1997).

5. Structural evolution and the shape of the East


Carpathians
On a regional map view, the Romanian East
Carpathians show two quite abrupt changes in the
dominant trends. The overall structural grain of
the belt is NWSE in northern Romania and in
adjacent Ukraine, changes to a NS direction in
correspondence with the Moldova valley, and
assumes an EW direction at the junction with the
South Carpathians ( Fig. 1). In both cases an oroclinal bending of the already structured fold-andthrust belt is unlikely, because of the absence/
insufficient amount of accommodating extensional
features (Marshak, 1988) in the external parts of
the fold-and-thrust belt and in the adjacent foreland (for a complete description see Morley, 1996;
Zweigel et al., 1998). Therefore, the observed
changes should relate to pre-contractional features
such as irregular plate boundaries or lateral
changes in the characteristics of the sedimentary

282

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

package entering the subduction zone and involved


in thrusting. Additionally, they could be at least
partly related to the polyphase tectonic history.
We propose that the change in direction taking
place across the Moldova valley (Fig. 3) is associated with lateral changes of the sediments previously lying on top of the subducting plate and
involved in thrusting. According to this view,
changes in directions (oroclines) are passive features associated with differential wedge geometries.
Two parameters are known to exert a primary
influence on thrusting geometry and thus on the
width of the accretionary wedge, the thickness of
the sediments involved in thrusting and the friction
at the main detachment surface (Fig. 17A). Thick
sedimentary packages create wide wedges and
widely spaced thrusts (e.g. Marshak and
Wilkerson, 1992). On the other hand, low friction
surfaces cause low internal shortening (ramping

Fig. 17. (A) Lateral variations in thrusting geometry due to


variations in wedge thickness and decollement friction (after
Nieuwland and Walters, 1993). (B) Flexural modelling results
(after Matenco et al., 1997b), which indicate possible mechanisms of slab break-off and slab tearing for the East
Carpathians foreland platforms and their distal components.

associated with a high number of backthrusts) and


high foreland advancement (Nieuwland and
Walters, 1993). Both processes seem to apply in
the East Carpathians. Indeed, in the northern part,
where the Tarcau nappe is very thick, especially at
the PaleoceneEocene level (e.g. thick Tarcau
sandstone) (Fig. 17A), thrust sheets are widely
spaced. In the same region, the basal detachment
layer is either the Lower Cretaceous black shales
in the Tarcau and Marginal nappes, or Lower
Burdigalian and Badenian evaporitic (mainly salt)
in the Subcarpathian nappe. In the northern part
of the East Carpathians, on the contrary, the
wedge is narrower and has a high degree of internal
shortening (duplexes, antiformal stacking). This is
compatible with the lower thicknesses of the
Tarcau sandstone in the region and of the shales
and evaporites acting as detachment layer. Along
strike, changes in the thickness of the Tarcau
sandstone have been attributed to Paleogene
stretching events (Sandulescu, 1992; Matenco,
1997).
Other changes in deformation geometries are
related to lateral variations of the lithospheric
characteristics of the thrusted platforms entering
the subduction zone. In the late Miocene, the most
advanced East Carpathians nappes (central sectors) reached the East European block north of
the Trotus valley. The introduction into the system
of lithospheric block with up to 50 km thick crust
and very thick lithosphere (Radulescu et al., 1976;
Zielhuis and Nolet, 1994; Guterch et al., 1994)
imposed changes on thrusting geometries
( Fig. 16A). The most important was the onset of
substantial uplift in the rear part of the wedge,
associated with the activation of regional
backthrusts in the internal part of the orogen
(Sanders, 1998), and further in the Transylvania
basin (e.g. Ciulavu, 1999). Similar structures are
also found further to the south, but in a much
more external position (i.e. Tarcau and
Subcarpathian nappes, Matenco, 1997). Such
changes are not observed in the southern segments
of the chain where the thinner Moesian plate is
involved in subduction.
One of the most interesting features of the East
Carpathians is the SW bend area, where the struc-

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

tural grain changes from NS north of the Buzau


Valley, to NESW and further EW southward
( Fig. 4). Almost all published studies (e.g.
Sandulescu, 1984; Morley, 1996; Hyppolite and
Sandulescu, 1996; Zweigel et al., 1998) assume
that individual structures in the external East
Carpathians nappes are continuous and follow the
belt bend across the SE corner. The low values
deduced in these studies for the required orogenparallel extension (1520% after Morley, 1996;
Zweigel et al., 1998) through surface kinematics
or mapping are moreover not supported by the
subsurface images, i.e. no such types of structure
were observed by extensive exploration studies in
the external East Carpathians nappes (e.g.
Stefanescu and working group, 1988; profiles A9
A22, Ionescu, 1994; Dicea, 1995, 1996, or unpublished data of Petrom RA). This observation and
our data suggest that the main structures that
developed at the surface of the oroclinal bending
formed during different episodes of deformation.
During the Middle to late Miocene shortening,
thrusts and backthrusts with NNESSW strike
were formed in the northern part of the bend area
( Fig. 4). Southward, oblique EW thrusting
occurred, possibly in association with dextral
translation along the thrust planes. Further WSWward prolongation of the Pericarpathian line
formed during the following, latest Miocene episode, in a strikeslip regime with NS shortening
direction (Fig. 16B). Later, Pliocene south-vergent
out-of-sequence thrusting (Fig. 16C ) enlarged the
S-ward offset of the mentioned thrusts and
accounts for the apparent continuity of the structures across the East Carpathians SE corner.

6. Overview and conclusions


The tectonic evolution of the foreland of the
East Carpathians can be described in terms of
Late Burdigalian to Sarmatian (Late Miocene)
shortening, Latest Sarmatian (Latest Miocene)
Early Meotian ( Earliest Pliocene) strikeslip and
Pliocene shortening.
The ENEWSW-directed shortening acting in
Late Burdigalian to Sarmatian times was responsible for the major nappe emplacement in the East

283

Carpathians. Deformation generally propagated


from the more internal to the more external units
and in the Sarmatian the molasse deposits of the
Subcarpathian nappe were thrusted on top of the
foreland platforms. These frontal movements were
accommodated with foreland-vergent thrusts in
the north and with a backthrust associated with a
triangle zone in the south. The oroclinal shape of
the thrust belt must have been initiated in the later
deformation moments due to the irregular plate
boundaries and the lateral variations in thickness
of the sedimentary wedge involved in shortening.
During Late BurdigalianSarmatian times, the
thrust front in the northern part of the belt reached
and moved over the TTZ and the East European
Platform ( Ellouz and Roca, 1994). As a consequence of the interactions with the East European
plate, backthrusts were activated in the internal
parts of the orogen and important exhumation
took place. Fission track analysis (Sanders, 1998)
show that exhumation and cooling became important in the internal East Carpathians at 1113 Ma
(Late Miocene). The Late BurdigalianSarmatian
tectonic phase is mainly responsible for the present
day geometry of the East Carpathians and to the
up to 4 km accelerated exhumation of the rocks
(e.g. Sanders, 1998).
From the Late Sarmatian, the stress field
changed to a strikeslip configuration with NNE
SSW-directed compression in the north to NSdirected compression in the south. Widespread
left-lateral shearing occurred along numerous E
W-oriented faults in the north, and NWSE-trending dextral transpressive structures in the south,
accommodating an ESE-ward movement of the
intervening central sectors. These findings are compatible with the flexural modelling results, which
account for large-scale tearing and slab break-off
mechanisms in the southern part of the East
Carpathians (Matenco et al., 1997b, Fig. 17B).
However, in accordance with other authors (e.g.
Sandulescu, 1984; Csontos, 1995) and differently
from Linzer (1996), we could find no evidence of
regional major strikeslip faults cross-cutting the
external parts of the fold-and-thrust belt. Stress
and deformation stages similar to those presented
in this study have been documented in the South
Carpathians, particularly concerning the Late

284

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

Burdigalian to Sarmatian and the Late Sarmatian


phases (Matenco et al., 1997a). During the
Pliocene, the entire area was subject to an overall
NNWSSE compressional stress regime, the effects
of which are observed mainly to the south, in the
bending area.
The regional importance of Pliocene deformation is further demonstrated by recent paleostress
work carried out in the bending area (Hyppolite
and Sandulescu, 1996; Morley, 1996), in the southern part of the Moesian Platform (Bergerat and
Pironkov, 1996) and with the present-day stress
field of the Carpatho-Pannonian region (Gerner,
1992).
Most of the Middle to Late Miocene evolution
of the East Carpathians and surrounding areas
was related to the westward subduction of the
distal parts of the East European, Scythian and
Moesian platforms and associated roughly
E-vergent thrusting (Sandulescu, 1984; Royden,
1988; Csontos, 1995). Processes and phenomena
taking place in more recent times are less agreed
upon. In particular, the role of the retreating
subduction zone during Pliocene times is still being
debated (e.g. Linzer, 1996 and references cited
therein). Our studies have demonstrated that the
Latest Sarmatian strikeslip shearing was distributed over large parts of the East Carpathians,
favouring a 4050 km ESE-ward escape of the
central sectors, and that only limited deformation
took place during the Pliocene.

Acknowledgements
Part of this work was made possible by the
financial support of Shell Romania Exploration
B.V. and Free University, Amsterdam. Discussions
with D. Ciulavu, R. Huismans and C. Sanders
have been inspiring. S. Cloetingh, C. Dinu and O.
Dicea are thanked for useful ideas and permanent
support. B. Sperner, D. Delvaux and E.
Wallbrecher are thanked for providing the software
used for paleostress data processing. J.P. Brun, G.
Tari and L. Ratschbacher are thanked for constructive reviews, which have substantially
improved the quality of this work. This is publica-

tion no. 991101 of the Netherlands School of


Sedimentary Geology.

References
Angelier, J., 1984. Tectonic analysis of fault slip data sets.
J. Geophys. Res. 89, 58355848.
Angelier, J., 1989. From orientation to magnitudes in paleostress determination using fault slip data. J. Struct. Geol.
11, 3750.
Angelier, J., 1994. Fault slip analysis and paleostress reconstruction. In: Hancock, P.L. ( Ed.), Continental Deformation.
Pergamon Press, pp. 53100.
Angelier, J., Mechler, P., 1977. Sur une methode graphique de
recherche des contraintes principales egalement utilisable en
tectonique et en seismologie: la methode des die`dres droits.
Bull. Soc. Geol. Fr. ser. 7e XIX (6), 13091318.
Bancila, I., 1955. Paleogenul zonei mediane a flisului. Acad.
RPR Bull. St., Agron., Biol., Geol., Geogr. VII, 12011234,
(in Romanian).
Bancila, I., 1958. Geologia Carpatilor Orientali. Ed. stiintifica,
Bucharest. 367 pp. (in Romanian).
Bergerat, F., Pironkov, V., 1996. The Moesian Platform as a
key for understanding the geodynamical evolution of the
Carpatho-Balkan Alpine system. In: C.S.S. and B.E. ( Eds.),
Stratigraphy and Evolution of Peri-Tethyan Platforms, Peritethys Memoir 3, 129150.
Botezatu, R., Calota, C., 1983. Asupra anomaliei cmpului geomagnetic situata la nord-est de Suceava. Stud. Cerc. Geol.
Geofiz. Geogr., Ser. Geofiz. 9, 322 (in Romanian).
Cantini, P., Faggioni, O., Pinna, E., Soare, A., Stanicaa, D.,
Stanicaa, M., 1991. Structure of the transition from the Hercynian lithosphere to the East European platform in Romania ( TornquistTeisseyre zone). Atti del 10 Convegno
Annuale del Gruppo Nazionale di Geofisica della Terra
Solida, Roma, 68 Novembre, 601613.
Ciulavu, D., 1999. Tertiary tectonics of the Transylvanian basin.
Ph.D. Thesis, Vrije Universiteit, Faculty of Earth Sciences,
Amsterdam, 154 pp.
Csontos, L., 1995. Tertiary tectonic evolution of the Intra-Carpathian area: a review. Acta Vulcanol. 7, 113.
Delvaux, D., 1993. The TENSOR program for paleostress
reconstruction: examples from the east and the Baikal
region, Central Asia. Terra Nova 5, abstract suppl, 216.
Delvaux, D., Moeys, R., Stapel, G., Petit, C., Levi, K.., Miroshnichenko, A., Ruzhich, V., Sankov, V., 1997. Paleostress
reconstruction and geodynamics of the Baikal region. Tectonophysics 282, 14, 138.
Dicea, O., 1995. The structure and hydrocarbon geology of the
Romanian East Carpathians border from seismic data.
Petrol. Geosci. 1, 135143.
Dicea, O., 1996. Tectonic setting and hydrocarbon habitat of
the Romanian external Carpathians. In: Ziegler, P.A., Horvath, F. ( Eds.), Peritethys Memoir 2: Structure and Pros-

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286


pects of Alpine Basins and Forelands. In: Memoires du
Museum national dHistoire naturelle, Paris, pp. 403425.
Dicea, O., Ali Mehmed, E., A, D., 1966. Contribution a` letude
de la tectonique de profondeur de la zone du flysch externe
et de la zone neoge`ne du nord des Carpathes Orientales.
Rev. Roum. Geol., Geophys. Geogr., Ser. Geophys. 10,
177182.
Dumitrescu, I., 1952. Etude geologique de la region comprise
entre lOituz et la Coza. Ann. Com. Geol. XXIV, 195270.
Dupin, J.M., Sassi, W., Angelier, J., 1993. Homogenous stress
hypothesis and actual fault slip: a distinct element analysis.
J. Struct. Geol. 15 (8), 10331043.
Ellouz, N., Roca, E., 1994. Palinspastic reconstructions of the
Carpathians and adjacent areas since the Cretaceous: a
quantitative approach. In: Roure, F. (Ed.), Peri-Tethyan
Platforms. Editions Technip, Paris, pp. 5178.
Ellouz, N., Roure, F., Sandulescu, M., Badescu, D., 1994. Balanced cross sections in the eastern Carpathians (Romania):
a tool to quantify Neogene dynamics. In: Roure, F., Ellouz,
N., Shein, V.S., Skvortsov, I. ( Eds.), Geodynamic Evolution
of Sedimentary Basins. Int. Symp. Proc., Moskow, 1992.
Technip, Paris, pp. 305325.
Enescu, D., Pompilian, A., Bala, A., 1988. Distribution of the
seismic wave velocities in the lithosphere of some regions in
Romania. Rev. Roum. Geol. Geophys. Geogr., Ser. Geophys. 32, 311.
Enescu, D., Danchiv, D., Bala, A., 1992. Lithosphere structure
in Romania II. Thickness of Earth crust. Depth-dependent
propagation velocity curves for P and S waves. Stud. Cerc.
Geol. Geofiz. Geogr., Ser. Geofiz. 30, 319.
Etchecopar, A., Vasseur, G., Daignieres, M., 1981. An inverse
problem in microtectonics for the determination of stress
tensors from fault striation analysis. J. Struct. Geol. 3 (1),
5165.
Fodor, L., Csontos, L., Bada, G., Gyorfi, I., Benkovics, L.,
1996. Tertiary tectonic evolution of the Pannonian basin
system and neighboring orogens: a new synthesis of paleostress data. Mitteilungen der Gesellschaft der Geologie und
Bergbaustudenten in Osterreich, pp. 105106.
Gerner, P., 1992. Recent stress field of Transdanubia, west Hungary, ALCAPA: Geological Evolution of the Internal Alps,
Carpathians and of the Pannonian Basin. Terra Abstr. 25.
Girbacea, R., Frisch, W., 1998. Slab in the wrong place: lower
lithospheric mantle delamination in the last stage of the
Eastern Carpathians subduction retreat. Geology 26 (7),
611614.
Guterch, A., Grad, M., Materzok, R., Perchuk, E., 1986. Deep
structure of the earths crust in the contact zone of the Paleozoic and Precambrian platforms in Poland ( TornquistTeissere zone). Tectonophysics 128, 251279.
Guterch, A., Grad, M., Janik, T., Materzok, R., Luosto, U.,
Yliniemi, J., Luck, E., Schulze, A., Forste, K., 1994. Crustal
structure of the transition zone between Precambrian and
Variscan Europe from new seismic data along LT-7 profile
(NW Poland and eastern Germany). C.R. Acad. Sci. Paris
319, 14891496.
Huismans, R.S., Bertotti, G., Ciulavu, D., Sanders, C., Cloe-

285

tingh, S., C, D., 1997. Structural evolution of the Transylvania basin (Romania): a sedimentary basin in the bend zone
of the Carpathians. Tectonophysics 272, 249268.
Hyppolite, J.C., Sandulescu, M., 1996. Paleostress characterization of the Wallachian phase in its type area, southeastern
Carpathians, Romania. Tectonophysics 263, 235249.
Ionescu, N., 1994. Exploration history and hydrocarbon prospects in Romania. In: Popescu, B. ( Ed.), Hydrocarbon of
Eastern Central Europe. Habitat, Exploration and Production History. Springer-Verlag, pp. 220250.
Ionesi, L., 1971. Flisul paleogen din bazinul vaii Moldovei. Ed.
Acad. RSR, 200 pp. (in Romanian).
Joja, T., Mutihac, V., Muresan, M., 1968. CrystallineMesozoic
and Flysch Complexes of the East Carpathians (Northern
Sector). In: Guide to Excursion 46 AC, Romania, Int. Geol.
Congr., XXIII Session, Prague, 563.
Linzer, H.G., 1996. Kinematics of retreating subduction along
the Carpathian arc, Romania. Geology 24, 167170.
Linzer, H.-G., Frisch, W., Zweigel, P., Girbacea, R., Hann,
H.-P., Moser, F., 1998. Kinematic evolution of the Romanian Carpathains. Tectonophysics 297, 14, 133156.
Marshak, S., 1988. Kinematics of orocline and arc formation
in thin-skinned orogens. Tectonics 7, 7386.
Marshak, S., Wilkerson, M.S., 1992. Effect of overburden thickness on thrust belt geometry and development. Tectonics
11, 560566.
Matenco, L., 1991. Aplicatii ale metodelor sectiunilor balansate
in pinzele externe ale Carpatilor Orientali. Master Thesis,
University of Bucharest, Bucharest, 91 pp. (in Romanian).
Matenco, L., 1997. Tectonic evolution of the Outer Romanian
Carpathians: Constraints from kinematic analysis and flexural modelling. Ph.D. Thesis, Vrije Universiteit, Faculty of
Earth Sciences, Amsterdam, 160 pp.
Matenco, L., Schmid, S., 1999. Exhumation and uplift of the
Danubian nappes system (South Carpathians) during the
Early Tertiary: inferences from kinematic and paleostress
analysis at the Getic/Danubian nappes contact. Tectonophysics, submitted.
Matenco, L., Bertotti, G., Dinu, C., Cloetingh, S., 1997a. Tertiary tectonic evolution of the external South Carpathians
and the adjacent Moesian platform (Romania). Tectonics
16, 896911.
Matenco, L., Zoetemeijer, R., Cloetingh, S., Dinu, C., 1997b.
Lateral variations in mechanical properties of the Romanian
external Carpathians: inferences of flexure and gravity modelling. Tectonophysics 282, 147166.
Micu, M.C., 1990. Neogene geodynamic history of the Eastern
Carpathians. Geol. Carpathica 41 (1), 5964.
Mocanu, V.I., Radulescu, F., 1994. Geophysical features of
the Romanian territory. In: Berza, T. (Ed.), Geological
Evolution of the AlpineCarpathianPannonian System,
ALCAPA II, Field Guidebook, Rom. J. Tect. Reg. Geol.,
1736.
Morley, C.K., 1996. Models for relative motion of crustal
blocks within the Carpathian region, based on restorations
of the outer Carpathian thrust sheets. Tectonics 15, 885904.
Mrazec, L., Popescu-Voitesti, I., 1914. Contributions a la con-

286

L. Matenco, G. Bertotti / Tectonophysics 316 (2000) 255286

naissance des nappes du Flysch Carpathique en Roumanie.


Ann. Inst. Geol. Roum. V, 495527.
Nieuwland, D.A., Walters, J.V., 1993. Geomechanics of the
South Furious field. An integrated approach towards solving complex structural geological problems, including analogue and finite-element modelling. Tectonophysics 226,
143166.
Pinna, E., Soare, A., Stanicaa, D., 1991. Transition zone of
East Carpathians thrust belt, foredeep basin and East European Platform System. XVI EGS Gen. Ass. Ann. Geophys.
(Suppl. to Vol. 9), 45.
Pollard, D.D., Saltzer, S.D., Rubin, A.M., 1993. Stress inversion methods: are they based on faulty assumptions?
J. Struct. Geol. 15 (8), 10451054.
Rabagia, T., Fulop, A., 1994. Syntectonic sedimentation history
in the Southern Carpathians foredeep. In: Berza, T. ( Ed.),
Geological Evolution of the AlpineCarpathianPannonian
System, ALCAPA II, Abstr. Vol., Rom. J. Tect. Reg.
Geol., 48.
Rabagia, T., Matenco, L., 1999. Tertiary tectonic and sedimentological evolution of the South Carpathians foredeep: tectonic versus eustatic control. Marine Petrol. Geol.
Radulescu, D.P., Cornea, I., Sandulescu, M., Constantinescu,
P., Radulescu, F., Pompilian, A., 1976. Structure de la
croute terrestre en Roumanie. Essai dinterpretation des
etudes seismiques profondes. Ann. Inst. Geol. Geofiz. 50,
536.
Radulescu, F., 1988. Seismic models of the crustal structure in
Romania. Rev. Roum. Geol. Geophys. Geogr., Ser. Geophys. 32, 1317.
Raileanu, V., Diaconescu, C., Radulescu, F., 1994. Characteristics of Romanian lithosphere from deep seismic reflection
profiling. Tectonophysics 239, 165185.
Ratschbacher, L., Linzer, H.G., Moser, F., Strusievicz, R.O.,
Bedelean, H., Har, N., Mogos, P.A., 1993. Cretaceous to
Miocene thrusting and wrenching along the central South
Carpathians due to a corner effect during collision and orocline formation. Tectonics 12, 855873.
Rogl, F., 1996. Stratigraphic correlation of Paratethys Oligocene and Miocene. Mitteilungen der Gesellschaft der Geologie und Bergbaustudenten in Osterreich, Vol. 41, pp. 6573.
Royden, L.H., 1988. Late Cenozoic Tectonics of the Pannonian
Basin System. In: Royden, L.H., Horvath, F. (Eds.), The
Pannonian Basin, a Study in Basin Evolution, AAPG
Memoir 45, 2748.
Sanders, C., 1998. Tectonics and erosion, competitive forces in
a compressive orogen: a fission track study of the Romanian

Carpathians. Ph.D. Thesis, Vrije Universiteit, Amsterdam,


204 pp.
Sandulescu, M., 1984. Geotectonica Romaniei. Ed. Technica,
Bucharest. 450 pp. (in Romanian).
Sandulescu, M., 1988. Cenozoic Tectonic History of the Carpathians. In: Royden, L.H., Horvath, F. ( Eds.), The Pannonian Basin, a Study in Basin Evolution, AAPG Memoir
45, 1725.
Sandulescu, M., 1992. Reunion extraordinaire de la Societe
Geologique de France en Roumanie. Guide des Excursions.
Inst. Geol. Geofiz., Bucharest.
Sandulescu, M., Visarion, M., 1988. La structure des plateformes situees dans lavant-pays et au-dessous des nappes
du flysch des Carpathes orientales. St. tehn. econ., Geofiz.
15, 6267.
Sandulescu, M., Krautner, H.G., Balintoni, I., Russo-Sandulescu, D., Micu, M., 1981a. The structure of the East
Carpathians (MoldaviaMaramures area). Carp.-Balc.
Assoc., XII Congr.
Sandulescu, M., Stefanescu, M., Butac, A., Patrut, I., Zaharescu, P., 1981b. Genetical and structural relations between
flysch and molasse ( The East Carpathians). Carp.-Balc.
Assoc., XII Congr.
Schmid, S.M., Berza, T., Diaconescu, V., Froitzheim, N.,
Fuegenschuh, B., 1998. Orogen-parallel extension in the
South Carpathians during the Paleogene. Tectonophysics
297, 209228.
Simpson, C., Schmid, S.M.., 1983. An evaluation of criteria to
deduce the sense of movement in sheared rocks. Geol. Soc.
Am. Bull. 94, 12811288.
Spang, J.H., 1972. Numerical method for dynamic analysis of
calcite twin lamellae. Geol. Soc. Am. Bull. 83, 467472.
Stefanescu, M., working group 1988. Geological cross sections
at scale 1:200,000 A9-14. Inst. Geol. Geofiz., Bucharest.
Tari, G., Dicea, O., Faulkerson, J., Georgiev, G., Popov, S.,
Stefanescu, M., Weir, G., 1997. Cimmerian and Alpine stratigraphy and structural evolution of the Moesian platform
(Romania/Bulgaria). In: Robinson, A.G. ( Ed.), Regional
and petroleum geology of the Black Sea and surrounding
regions, AAPG Memoir 68, 6390.
Turner, F.J., 1962. Compression and tension axes deduced
from {0112} twinning in calcite. J. Geophys. Res. 67,
16601670.
Wallbrecher, E., 1986. Tektonische und gefugeanalytische
Arbeitsweisen, Stuttgart, 224 pp.
Zielhuis, A., Nolet, G., 1994. Deep seismic expression of an
ancient plate boundary in Europe. Science 265, 7981.
Zweigel, P., Ratschbacher, L., Frisch, W., 1998. Kinematics of
an arcuate fold-thrust belt: the southern East Carpathians
(Romania). Tectonophysics 297, 177207.

You might also like