Accepted Manuscript The Zagros fold-and-thrust belt in the Fars province (Iran): II. Thermal evolution L. Aldega, S. Bigi, E. Carminati, F. Trippetta, S. Corrado, A.M. Kavoosi PII:
S0264-8172(18)30122-3
DOI:
10.1016/j.marpetgeo.2018.03.022
Reference:
JMPG 3287
To appear in:
Marine and Petroleum Geology
Received Date: 28 September 2017 Revised Date:
14 March 2018
Accepted Date: 15 March 2018
Please cite this article as: Aldega, L., Bigi, S., Carminati, E., Trippetta, F., Corrado, S., Kavoosi, A.M., The Zagros fold-and-thrust belt in the Fars province (Iran): II. Thermal evolution, Marine and Petroleum Geology (2018), doi: 10.1016/j.marpetgeo.2018.03.022. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
RI PT
4 km 3.2 km
-4 km
SC
s.l.
AC C
EP
TE D
M AN U
Reconstruction of the eroded thickness by paleothermal indicators
ACCEPTED MANUSCRIPT 1
THE ZAGROS FOLD-AND-THRUST BELT IN THE FARS PROVINCE (IRAN): II.
2
THERMAL EVOLUTION
3
ALDEGA L.1*, BIGI S.1, CARMINATI E.1, TRIPPETTA F.1, CORRADO S.2 and KAVOOSI
5
A.M.3
RI PT
4
6 1
7
Dipartimento di Scienze della Terra, Sapienza Università di Roma, P.le Aldo Moro, 5, 00185 Roma
2
Dipartimento di Scienze, Sezione di Geologia, Università degli Studi Roma Tre, L.go S. Leonardo Murialdo 1, 00146 Roma
10 3
11
M AN U
9
SC
8
National Iranian Oil Company, Tehran, Iran
12
*corresponding author e-mail:
[email protected] tel. +39 06 49914547
13
TE D
14
Abstract
16
Temperature-dependent clay minerals and vitrinite reflectance data, surface and subsurface
17
geological constraints were used to unravel the burial evolution of the Ordovician-Quaternary
18
sedimentary successions from the inner to the outer zones of the Zagros fold-and-thrust belt in the
19
Fars province (Iran). These sedimentary successions were buried to their thermal maxima during
20
early to late diagenesis, achieving temperatures corresponding to the immature to early mature
21
stages of hydrocarbon generation. They experienced low levels of thermal maturity in the Interior
22
Fars, corresponding to vitrinite reflectance values between 0.38 and 0.66%, to mixed layers illite-
23
smectite (I-S) with an illite content between 30 and 75% and to KI values between 0.97 and 1.18
24
°∆2θ. In the Central and Coastal Fars, vitrinite reflectance ranges between 0.35 and 0.51%, the illite
25
content in I-S displays values between 20 to 87% and KI data are between 0.71 and 1.30 °∆2θ. In
26
individual anticlines, mixed layers I-S show an increase of the illite content as a function of
AC C
EP
15
1
ACCEPTED MANUSCRIPT stratigraphic age (depth), suggesting that levels of thermal maturity are controlled by sedimentary
28
burial. One dimensional thermal history models allowed us: (i) to estimate the maximum burial
29
experienced by the sedimentary successions and the amount of the sedimentary pile currently
30
removed by erosion, (ii) to determine the thickness of the ophiolite units obducted during Late
31
Cretaceous time in the High Zagros, and (iii) to define the onset of oil generation for the Albian
32
source rocks throughout the Zagros belt. Paoleothermal data were used to constrain the geometry of
33
eroded structures in a 253 km long cross-section extending from the High Zagros to the Coastal
34
Fars. Along the cross-section, lithostatic load slightly decreases towards the foreland (e.g., from
35
3.65 km to 3.2 km for the Bangestan Group) and the amount of the eroded material varies between
36
~6 km (above anticlines in the Central and Interior Fars) and ~200 m (above synclines in the
37
external part of the belt).
M AN U
SC
RI PT
27
38
KEYWORDS: mixed layers illite-smectite, vitrinite reflectance, thermal maturity, 1D thermal
40
modelling, Fars province, Zagros fold-and-thrust belt
41
TE D
39
1. Introduction
43
Contractional deformation in thrust belts can either affect the sedimentary cover detached from the
44
underlying basement in the frontal portion of the belt or may also involve basement rocks along
45
crustal-scale ramps in the inner part of the belt, thus resulting in different mechanical behavior and
46
in thin- or thick-skinned tectonics (e.g., Chapple, 1978; Coward, 1983; DeCelles and Mitra, 1995;
47
Calabrò et al., 2003; Poblet and Lisle, 2011). Both tectonic styles may develop through time and
48
superimpose to each other, implying large differences in the amounts of shortening in various
49
portions of thrust belts. Generally, either surface (e.g., attitude of beds and tectonic structures,
50
thickness of the stratigraphic pile) or subsurface (e.g., seismic lines, borehole data) geological and
51
geophysical data are used to constrain the orogen structure. In this regard, several works
52
investigating the structure of the Zagros in the Fars province (Iran) provided crustal scale cross
AC C
EP
42
2
ACCEPTED MANUSCRIPT sections using such constraints (e.g., McQuarrie, 2004; Molinaro et al., 2005; Mouthereau et al.,
54
2007; Alavi, 2007) and concluded that the belt is characterized either by thin-, thick- or by a
55
combination of thin- and thick-skinned tectonics.
56
An open issue for the Zagros belt is that surface and subsurface data do not exclude the past
57
occurrence of overthrusts now removed by erosion and provide partial information on the amount of
58
burial experienced by the sedimentary succession during mountain building. Such constraints can
59
be provided either by low temperature thermochronology (e.g., apatite fission track and U-Th/He
60
dating) or from organic and inorganic thermal indicators, including vitrinite reflectance (Ro%) and
61
illite content in mixed layer illite-smectite (I% in I-S).
62
In particular, these latter paleothermal indicators allow quantifying the maximum burial thickness
63
of the stratigraphic section and therefore provide an estimation as to the thickness of the
64
stratigraphic section removed by erosion and/or tectonics (Botti et al., 2004; Aldega et al., 2011, Di
65
Paolo et al., 2012; Meneghini et al., 2012; Carlini et al., 2013; Perri et al., 2016; Schito et al., 2017).
66
In other cases, paleothermal indicators allow estimating the original thickness and extent of thrust
67
units (Caricchi et al., 2015; Aldega et al., 2017). This is important for the frontal part of thrust belts,
68
where tectonic burial is generally limited to a few kilometers and very few techniques may be
69
applied to both sedimentary and crystalline rocks to detect such tectonic thickening (e.g., Corcoran
70
and Dorè, 2005 for a review), and for the inner part of orogens where tectonic overburden may be
71
totally removed by erosion (e.g., Ring et al., 1999; Doré et al., 2002).
72
The integration of organic and inorganic thermal indicators with surface and subsurface geological
73
data were used: (i) to constrain the burial evolution of the Ordovician-Quaternary sedimentary
74
sequences exposed in the inner and outer zones of the Zagros fold-and-thrust belt in the Fars
75
province, (ii) to define the onset of oil generation for the Albian source rocks throughout the Zagros
76
belt, (iii) to determine the thickness of the ophiolite units obducted in Late Cretaceous time in the
77
internal part of the belt. These new constraints were included in the 253 km long cross-section
AC C
EP
TE D
M AN U
SC
RI PT
53
3
ACCEPTED MANUSCRIPT 78
(described in Bigi et al., 2018) extending from the High Zagros to the Coastal Fars to provide
79
information on the original thickness of the thrust belt and its eroded part.
80
This
81
thermochronological datasets, and can help to substantially reduce the number of acceptable
82
geometric, thermal and kinematic models on the basis of the amount of maximum burial that
83
characterizes different portions of the thrust wedge.
can
be
applied
to
other
fold-and-thrust
84
belts
worldwide
integrating
RI PT
approach
2. Geological setting
86
The Zagros fold-and-thrust belt is the result of the Cenozoic convergence and Alpine-type
87
continental collision between the Central Iran domain and the Arabian plate (e.g., Mouthereau et al.,
88
2012; Navabpour and Barrier, 2012). The Zagros fold-and-thrust belt is constituted by two main
89
zones, the High Zagros Imbricated Zone to the north and the Zagros Simply Folded Belt to the
90
south, bounded by the High Zagros fault (HZF in Fig. 1A; Mouthereau et al., 2007). The Zagros
91
Simply Folded Belt is separated from the Dezful Embayment by the Mountain front fault (MFF;
92
Sepehr and Cosgrove, 2004). In addition to these tectonic subdivisions, the Zagros belt can be
93
divided into three main geological provinces bounded by major N-S trending basement faults
94
(Bahroudi and Talbot, 2003), re-sheared as strike-slip faults during the orogenic phase (e.g.,
95
Sherkati and Letouzey, 2004; Sepehr and Cosgrove, 2005; Lacombe et al., 2006; Ahmadhadi et al.,
96
2007; Carminati et al., 2014). From north-west to south-east, they are the Lorestan, the Dezful
97
Embayment and the Fars provinces (Fig.1A).
98
Crustal scale cross-sections across the Zagros (e.g., McQuarrie, 2004; Sherkati et al., 2006;
99
Mouthereau et al., 2007) and subsurface data from wells reveal an imbricate-fan geometry,
100
constituted by an uppermost Neoproterozoic to Phanerozoic sedimentary succession up to 12 km
101
thick (Fig. 2, Alavi, 2004). In detail, the Precambrian basement rocks are covered by the Hormuz
102
series constituted by evaporites (halite and anhydrite), gray trilobite-bearing dolostones with
103
interlayered mafic and felsic volcanic rocks in its upper part (Alavi, 2004). The tectono-sedimentary
AC C
EP
TE D
M AN U
SC
85
4
ACCEPTED MANUSCRIPT setting of the Ordovician-Carbonifeorus rocks is poorly constrained because of the scarcity of
105
outcrops and the presence of major regional unconformities related to several Paleozoic erosional
106
episodes (Alavi, 1994). In general, the Paleozoic succession is characterized by thickness and facies
107
changes likely associated with early Cambrian salt diapirism, basement faulting and eustatic sea
108
level changes (Berberian and King, 1981; Callot et al., 2007; Jahani et al., 2009).
109
The Permian-Triassic succession marks a significant change in sedimentation from dominantly
110
Paleozoic clastic sediments to carbonate rocks deposited in a shallow marine environment (Alavi,
111
2004). This succession contains volcanic rocks associated with the Permo-Triassic rifting and the
112
opening of the Neo-Tethys ocean. Following the rifting episode, sedimentation took place in a
113
passive continental margin and different successions were deposited in the Lorestan and Dezful
114
basins to the NW and the Fars basin to the SE. As an example, the Fars region was characterized by
115
shallow water sedimentation until Late Cretaceous time, whereas coeval deep marine sediments
116
were deposited in the Lorestan and Dezful Embayment areas (Berberian and King, 1981; Casciello
117
et al., 2009).
118
During Early Triassic time, the marine carbonate sedimentary regime persisted with the deposition
119
of the Kangan Fm. in the Fars Province (Szabo and Kheradpir, 1978). Regressive conditions
120
occurred in the Middle-Late Triassic, resulting in deposition of the evaporites of the Dashtak
121
Formation (Berberian and King, 1981) and the dolomitic limestones of the Khaneh Kat Fm. (Pyriaei
122
et al., 2010).
123
The Early Jurassic was characterized by terrigenous clastic sedimentation and by transitional to
124
open marine deposits of the Neyriz Fm. (Szabo and Kheradpir 1978), followed by limestones and
125
marls of the Khami Group (Surmeh, Fahliyan, Gadvan, Daryan; Middle Jurassic-Aptian).
126
In the Interior Fars, siltstones and iron/glauconite rich sandstones occurring in the upper parts of the
127
Fahliyan and Dariyan Fms suggest marine regression, emergence, and erosion in Neocomian and
128
late Aptian times (Setudehnia 1978; Navabpour et al., 2010). The shallow marine shales and
AC C
EP
TE D
M AN U
SC
RI PT
104
5
ACCEPTED MANUSCRIPT limestones of the Bangestan Group (Kazhdumi and Sarvak Fms; Albian-Turonian) contain late
130
Turonian conglomerates and breccias, suggesting tectonic-related uplift (Berberian and King 1981).
131
The Late Cretaceous sedimentation in the Interior Fars was affected by the southwestward
132
obduction of ophiolites in a subduction setting (e.g., Breton et al. 2004), which caused significant
133
variations in sedimentary facies, sedimentation patterns and accommodation space, and determined
134
the shift of depocentres (Pyriaei et al., 2010). During this time, sedimentation was characterized by
135
neritic carbonates (Ilam Fm.) followed by deeper water conditions with the deposition of marls and
136
shales of the Gurpi Fm. in most of the Zagros. Since Maastrichtian time, marly sedimentation of the
137
Gurpi Fm. was coeval with rudist-dominated platform carbonate sedimentation (Tarbur Fm.;
138
Setudehnia 1972; Vaziri-Moghaddam et al., 2005) in the Interior Fars. This lateral facies change has
139
been related to the uplift associated with ophiolites emplacement (Pyriaei et al., 2010).
140
During the Cenozoic, sedimentation changed from an open marine to a continental environment
141
with increasing clastic input. The Tarbur limestones were overlain by the uppermost Maastrichtian
142
to Paleogene evaporitic deposits of the Sachun Fm., consisting of gypsiferous limestones,
143
dolostones and red marls deposited in a sabkha environment (Alavi, 2004). Heteropic to the Sachun
144
evaporitic limestones are the neritic shales and marls of the Pabdeh Fm., deposited on top of the
145
Gurpi basinal sediments in the Coastal Fars (Fig. 2).
146
During the Eocene-Oligocene, shallow water limestones of the Jahrum and Asmari Formations
147
were deposited in the Fars Province. Supratidal, sabkha-like conditions, developed locally in the
148
Early Miocene, as the Razak Formation of the Interior Fars.
149
The Middle Miocene clastic sedimentation marks the transition of the Fars province to a foreland
150
basin setting (Khadivi et al., 2010) with the deposition of marls, shales and sandstones of the
151
Mishan Fm., conglomerates, calcarenites and cross-bedded sandstones of the Agha Jari Fm. and
152
molasse-type conglomerates (Bakhtiyari Fm., Fig. 2).
153
Strontium isotope stratigraphy shows that marine foreland deposits of the Mishan Fm. are strongly
154
diachronous with ages ranging between 17 Ma and 1.1 Ma becoming progressively younger from
AC C
EP
TE D
M AN U
SC
RI PT
129
6
ACCEPTED MANUSCRIPT the Dezful Embayment to the Fars province (Pirouz et al., 2015). Magnetostratigraphic and
156
chronostratigraphic analysis pointed out that the onset of continental clastic sedimentation was
157
largely diachronous as well throughout the Zagros foreland basin. The deposition of the Bakhtiyari
158
conglomerates began in the early Miocene in the High Zagros (Fakhari et al., 2008), after 14.8 Ma
159
close to the High Zagros fault in the NE Fars (Khadivi et al., 2010), after 3.6-3.2 Ma in the Central
160
Fars (Ruh et al., 2014) and an age of 3.0 Ma was proposed for the coastal areas (Homke et al.,
161
2004).
RI PT
155
SC
162
2.1 Stratigraphy: thickness variations
164
Thickness variations of the main lithostratigraphic units can be inferred from borehole stratigraphic
165
logs shown in figure 3. Along NE-SW and N-S-trending directions, main thickness changes occur
166
in the Triassic Dashtak Fm, the Khami Group and the Asmari-Jahrum Fms.
167
The pre-folding thickness of the Dashtak Fm. varies from 550 m in the Coastal Fars to 850 m in the
168
Interior Fars indicating a northward thickening of the Triassic sediments (Figs. 3B and C). In the
169
Sefid Zakhur-1 and Cham-e-noori-1 wells, located at the crest of anticlines, about 1,400 to 2,000 m-
170
thick successions of shallow-marine limestones and dolomitic limestones interbedded with
171
evaporites and local intercalations of shales belonging to the Dashtak Fm. have been intercepted.
172
These large thickness variations (∆ = 600 m in Fig. 3B, ∆ = 1,300 m in Fig. 3C) are also reported in
173
the isopach and facies distribution maps by Koop and Stoneley (1982) and Pyriaei et al. (2010) and
174
are interpreted as the result of thrust faulting in the crestal domain of anticlines (Najafi et al., 2014).
175
The Khami Group shows an increasing thickness toward the Interior Fars similar to that observed
176
for the Dashtak Fm. In the Varavi-1 well, the Jurassic to Lower Cretaceous succession is 1050 m
177
thick whereas it reaches 1324 m and 1446 m of thickness in the Aghar-2 and Sefid Zakhur-1 well
178
respectively (Fig. 3).
179
The Asmari-Jahrum Fm. displays a more complex thickness variability. Maximum values were
180
measured in the Kuh-e-Sim anticline where a 700m thick succession is exposed (Carminati et al.,
AC C
EP
TE D
M AN U
163
7
ACCEPTED MANUSCRIPT 2013) and in the Aghar-2 well where 652 m of Nummulites-bearing limestones and dolomitic
182
limestones were intercepted. Twenty-five kilometers to the south, the thickness of the Asmari-
183
Jahrum Fm. reduces to 274 m in the Sefid-Zakhur-1 well and to about 150 m in the Safid Baghun-1
184
well. In the Coastal Fars, thicknesses ranging from 250 to 350 m were calculated from Hangam
185
(sheet 20861E), Dehram (20861W) and Kangan (20867W) geological maps (scale 1:100.000).
186
No major thickness changes occur in the coastal areas along a WNW-WSE trending direction
187
parallel to the strike of anticlines for the Triassic-Upper Cretaceous succession (Fig. 3D)
RI PT
181
SC
188
3. Methods
190
3.1 X-ray diffraction of clay minerals
191
Clay minerals contained in sedimentary rocks are heterogeneous assemblages of detrital material
192
coming from various source rocks, and, at paleotemperatures >70°C (Środoń, 1999), of
193
superimposed diagenetic modifications of this sediment. Clay minerals undergo diagenetic and very
194
low-grade metamorphic reactions when sedimentary basins subside in response to burial and/or
195
tectonic loading. Reactions in clay minerals are irreversible under normal diagenetic and anchizonal
196
conditions, so that uplifted and exhumed sequences generally retain indices and fabrics indicative of
197
thermal maturity and maximum burial (Árkai, 2002). Clay minerals are mainly sensitive to
198
temperature, and the use of mixed layers illite-smectite (I-S) and the transformation sequence
199
dismectite-random-ordered mixed layers I-S (R0)-ordered mixed layers I-S (R1 and R3)-illite-di-
200
octahedral K-mica (muscovite) as indicator of maximum paleotemperature condition is generally
201
accepted (Burst, 1959; Hower et al., 1976; Pollastro 1990; Aldega et al., 2007a, b, Corrado et al.,
202
2010 a, b; Izquierdo-Llavall et al., 2013, Schito et al., 2016). In fact, changes in the composition of
203
mixed layering, layer expandability, and I-S ordering are empirically related to temperature changes
204
due to burial (Hoffman and Hower, 1979; Pollastro and Barker, 1986; Botti et al., 2004).
205
Another parameter successfully applied worldwide for determining the grade of diagenesis and
206
very-low metamorphism of clay-rich and clastic sedimentary rocks is the Kübler index (Kübler,
AC C
EP
TE D
M AN U
189
8
ACCEPTED MANUSCRIPT 1967; KI). It has been commonly used as an empirical measure of the changes in sharpness of the
208
X-ray 10 Å basal reflection of illite-dioctahedral K-white mica. The 10Å peak width at half-peak-
209
height is commonly considered to be primarily a function of the average illite crystallite thickness
210
normal to (001) and several authors have shown how KI values decrease as metamorphic grade
211
increases (e.g., Jaboyedoff et al., 2001; Warr and Cox, 2016; Potel et al., 2016).
212
X-ray diffraction (XRD) analysis of the <2 µm grain size fraction has been carried out on 38 surface
213
samples with a Scintag X1 X-ray system (CuKα radiation) at 40 kV and 45 mA. Oriented air-dried
214
and ethylene-glycol solvated samples were scanned from 1 to 48 °2θ and from 1 to 30 °2θ
215
respectively with a step size of 0.05 °2θ and a count time of 4 s per step. The illite content in mixed
216
layers I-S was determined according to Moore and Reynolds (1997) using the delta two-theta
217
method after decomposing the composite peaks between 9-10 °2θ and 16-17 °2θ. The I–S ordering
218
type (Reichweite parameter, R; Jagodzinski 1949) was determined by the position of the I001-S001
219
reflection between 5 and 8.5 °2θ (Moore and Reynolds 1997).
220
The KI determinations, measured on oriented air-dried (AD) and ethylene-glycol solvated (EG) <2
221
µm mounts were made by first subtracting the background from the raw data, followed by peak
222
fitting using Pearson 7 functions. The FWHM (full-width-half-maximum) of the deconvoluted
223
~10Å peak was used as expression of illite “crystallinity”. Half-peak widths were converted to the
224
Crystallinity Index Standard (CIS) scale (Warr and Rice, 1994).
225
Peaks in relative close position were selected for clay mineral quantitative analysis in order to
226
minimize the angle-dependent intensity effect. Composite peaks were decomposed using Pearson
227
VII functions and the WINXRD Scintag associated program. Integrated peak areas were
228
transformed into mineral concentration by using mineral intensity factors as a calibration constant
229
(for a review, see Moore and Reynolds 1997).
AC C
EP
TE D
M AN U
SC
RI PT
207
230 231
3.2 Organic matter optical analysis
9
ACCEPTED MANUSCRIPT Vitrinite is derived from the thermal degradation of organic macerals of continental origin that are
233
present in post-Silurian sediments (Stach et al., 1982). Its reflectance strictly depends on the thermal
234
evolution of the host sediments and is correlated with the stages of hydrocarbon generation and
235
other thermal parameters in sedimentary environments (Durand, 1980). Thus, it is one of the most
236
widely used parameter to calibrate basin modeling and provides consistent and reliable constraints
237
on maximum burial depths (Dow, 1977; Mukhopadhyay, 1994; Corrado et al., 2009; 2010a).
238
Specimens for vitrinite reflectance were prepared according to standardized procedures described in
239
Bustin et al. (1990). Picked kerogen particles were cold set into epoxy resin blocks and polished
240
using carborundum papers and isopropanol as lubricant. After washing the sample in order to
241
remove debris, three alumina powders of decreasing grain size (1, 0.3, 0.01 µm) were used to polish
242
the samples. Random reflectance was measured under oil immersion (ne 1.518, at 23ºC), with a
243
Zeiss Axioskop 40 A pol microscope-photometer system and calibrated against standards of
244
certified reflectance. On each sample, measurements were performed on vitrinite or bitumen
245
unaltered fragments. Mean vitrinite (Ro%) and bitumen (Rb%) reflectance values were calculated
246
from the arithmetic mean of these measurements. Rb values have been converted into vitrinite
247
reflectance equilvalent values (Roeq%) according to Jacob and Hiltmann (1985).
248
EP
TE D
M AN U
SC
RI PT
232
4. Results
250
4.1 Clay mineralogy
251
High Zagros
252
X-ray diffraction analyses refer to the Lower Cretaceous Kazdhumi and Gadvan Fms (Tab.1). The
253
Kazdhumi Fm. (KAZ2 and KAZ3) in the Arsenjan area shows a mineralogical assemblage
254
composed of kaolinite (72% mean value), illite (15%), mixed layers I-S (7%) and chlorite (6%).
255
The underlying Gadvan Fm. (GAD1) consists of an illite- and kaolinite-rich assemblage (48% and
256
38% respectively) with subordinate amounts of mixed layers I-S (14%). Traces of calcite and
257
gypsum have been observed in the XRD pattern.
AC C
249
10
ACCEPTED MANUSCRIPT 258
Both the Kazdhumi and the Gadvan Fms display short range-ordered (R1) mixed layer I-S with an
259
illite content of about 75% and Kübler index data ranging from 0.92 and 1.01 °∆2θ which
260
corresponds to the first stages of the late diagenetic zone (Fig. 4A and Tab. 1; Merriman and Frey,
261
1999).
RI PT
262
Zagros Simply Folded belt: Interior Fars
264
A suite of 10 samples, belonging to the Upper Cretaceous-Miocene portion of the sedimentary
265
succession, has been collected in the Sarvestan area (Tab.1). Seven samples are from the
266
hangingwall of the Sarvestan fault (Fig. 1B) whereas three samples (MIS9, RAZ2, RAZ1) have
267
been collected some 40 km SW of the fault.
268
In the hangingwall units, we observe an increase of the illite content in mixed layer I-S and a
269
decrease of Kübler index values as a function of stratigraphic age (depth; Fig. 4A). Random ordered
270
I-S (R0) with high expandability (30-45% of illitic layers), that characterize the Miocene deposits,
271
converts into short range ordered structures (R1) with an illite content of 60-70% in the Upper
272
Cretaceous Gurpi Fm. Kübler index measurements display values ranging from 0.97 to 1.18 °∆2θ
273
consistent with early to late diagenetic conditions (Tab. 1). Forty kilometers SW of the Sarvestan
274
fault, the Razak Fm. shows random ordered I-S with low illite content (30%) indicating early
275
diagenetic conditions (Fig. 5A).
M AN U
TE D
EP
AC C
276
SC
263
277
Zagros Simply Folded belt: Central Fars
278
Clay mineralogy data for the Central Fars are from the Kuh-e-Sim (Aldega et al., 2014), Kuh-e-
279
Meymand and Kuh-e-Surmeh anticlines (Tab.1; Fig. 3A for anticlines location). In the surroundings
280
of the Kuh-e-Sim anticline, Oligocene-Miocene lithologies of the Mishan, Guri and Champeh Fms
281
and the Mol member of the Gachsaran Fm. display random-ordered (R0) mixed layers I-S with an
282
illite content between 21% and 52% (Fig. 4A) indicating early diagenetic conditions according to
283
Merriman and Frey (1999). 11
ACCEPTED MANUSCRIPT The Mishan Fm. shows an illite- and chlorite-rich composition with subordinate amounts of
285
palygorskite and mixed layer I-S. Red shales from the Mol Member are mainly characterized by
286
phyllosilicates (illite, chlorite and mixed layer chlorite-smectite), carbonate group minerals (calcite
287
and ankerite), quartz and albite. In the Guri Fm., illite and chlorite are the most abundant clay
288
minerals (85%) and prevail on mixed layer I-S (15%). Marls from the Champeh Fm. show an illite-
289
rich composition (40-51%) followed by chlorite (24-34%) and low contents of random ordered
290
mixed layer I-S (3-5%). Palygorskite, kaolinite and mixed layer chlorite-smectite (C-S) were
291
occasionally observed.
292
Four samples from the Pabdeh, Champeh and Mishan Fms and the Mol member (Tab. 1) were
293
collected in the surroundings of the Kuh-e-Meymand anticline. The Mishan Fm. is composed of
294
illite (35%), chlorite (28%) and mixed layers I-S (11%) and C-S (26%). The underlying Mol
295
member (MOL2) is characterized by an illite-rich composition (63%) and subordinate amounts of
296
chlorite (18%) and mixed layers C-S (19%) with a chlorite content of 55%. The mineralogical
297
assemblages of the Champeh Member is mainly constituted by abundant illite and mixed layer I-S
298
and subordinate amounts of chlorite, palygorskite and kaolinite (Tab. 1). The greenish silt-rich
299
marls of the upper part of the Pabdeh Fm. are composed of illite (50%), short range-ordered I-S
300
(26%), chlorite (17%) and kaolinite (7%).
301
Mixed layers I-S in the Paleocene to Miocene succession of the Kuh-e-Meymand anticline are
302
either random ordered (R0) or short range ordered (R1) structures with an illite content between
303
50% and 70% indicating early to late diagenetic conditions (Tab. 1). Kübler index measurements
304
show values between 0.95 and 1.11 °∆2θ (Tab. 1), which are consistent with levels of thermal
305
maturity observed by mixed layers I-S.
306
In the Kuh-e-Surmeh anticline, we collected five samples from the Mishan Fm. down to the Silurian
307
shales of the Ghakum Fm. Mixed layered minerals indicate an increase of the illite content in mixed
308
layer I-S as function of the stratigraphic age (depth). High expandable random ordered I-S (R0)
309
typical of the Mishan Fm. progressively evolves to structures with higher amount of illite layers (up
AC C
EP
TE D
M AN U
SC
RI PT
284
12
ACCEPTED MANUSCRIPT to 50-55%) in the Pabdeh and Gurpi Fms (PAB6, GUR7; Fig. 5B), to short-range ordered R1 I-S in
311
the Gadvan Fm. (GAD2) and to long-range ordered R3 I-S in the Ghakum Fm. (GAH1). The illite
312
content in mixed layer I-S at the bottom of the succession is about 85% indicating late diagenetic
313
conditions (Fig. 4A and Tab. 1; Merriman and Frey, 1999). KI data decrease as level of thermal
314
maturity increases with values ranging from 0.71 to 1.08 °∆2θ (Tab. 1).
RI PT
310
315
Zagros Simply Folded belt: Coastal Fars
317
Most samples were collected in the surroundings of the Kuh-e-Asalujeh anticline (Fig. 3A for
318
anticline location) spanning from the Kazdhumi to the Mishan Fms (Tab.1).
319
The Kazdhumi Fm. is composed of a kaolinite-rich assemblage that constitutes at least the 50% of
320
the overall composition, and subordinate amounts of illite, mixed layers I-S and chlorite (Fig. 5C).
321
Mineralogy of the <2µm grain-size fraction of the overlying Gurpi Fm. shows illite (38%) and
322
mixed layer I-S (62%). Moving up in the stratigraphic column, younger deposits display the
323
occurrence of palygorskite, which represents the main component of the clay fraction (82% mean
324
value) in the Pabdeh Fm. The remaining 18% is made up of illite and mixed layers I-S.
325
The Mishan Fm. is constituted by palygorskite (25-54%), chlorite (14-22%), illite (11-19%) and
326
mixed layers I-S (5%). Smectite occurs in the sediments cropping out in the north-western part of
327
the Kuh-e-Asalujeh anticline with amounts of 50%.
328
Random ordered mixed layers I-S or discrete smectite characterize the Miocene deposits indicating
329
low levels of thermal alteration which correspond to the early diagenetic zone. KI values between
330
0.52 and 0.60 °∆2θ point to higher levels of thermal maturity. We interpreted these values as the
331
signature of detrital K-micas produced from uplift and erosion of the Zagros belt. Thus, KI data for
332
Miocene deposits cannot be used for the reconstruction of the burial history of the Kuh-e-Asalujeh
333
anticline but provide information on provenance and thermal condition of the source rock.
AC C
EP
TE D
M AN U
SC
316
13
ACCEPTED MANUSCRIPT 334
The Cretaceous samples of the Kazdhumi Fm. at the core of the anticline display short-range
335
ordered mixed layers I-S with an illite content ranging from 70 to 75% and KI values of 0.95-1.01
336
°∆2θ which indicate the beginning of the late diagenetic zone (Fig. 4A).
337
4.2 Organic matter optical analysis
339
A suite of 22 samples was collected from the Cretaceous to Miocene deposits throughout the Fars
340
Province. Optical analysis provided results for 10 samples as the organic matter content was very
341
low as indicated by the scarce presence of bitumen and vitrinite-like fragments (Tab.2). Samples
342
from the Arsenjan area in the High Zagros are devoid of vitrinite-like fragments as are those
343
collected in the coastal areas close to the Kuh-e-Asalujeh anticline. In the Interior Fars, rare
344
fragments of reworked vitrinite-huminite and inertinite macerals and stains of bitumen are observed
345
in the Gurpi (GUR3, GUR4, GUR5; Tab.2) and in the Mishan Fms (MIS9). Vitrinite macerals
346
display mean vitrinite reflectance values between 0.38-0.49% indicating immature to early mature
347
stages of hydrocarbon generation consistently with Roeq% values between 0.55 and 0.66% of
348
bitumen.
349
In the Central Fars, the Champeh Fm. and the Mol member are very poor in organic matter content
350
and vitrinite-huminite macerals are scarce. Very few collinite fragments of small size display a
351
mean reflectance value of 0.39% (Tab. 2) indicating the immature stage of hydrocarbon generation.
352
The Mishan Fm. (MIS6, MIS7, MIS10) contains both inertinite and vitrinite-huminite macerals
353
associated with finely dispersed pyrite. Mean vitrinite reflectance values (between 0.32% and
354
0.50%) indicate the immature/early mature stage of hydrocarbon generation.
AC C
EP
TE D
M AN U
SC
RI PT
338
355 356
4.3 Synthesis of paleothermal data
357
Organic and inorganic paleothermal indicators indicate that the Late Cretaceous to Miocene
358
sedimentary succession experienced similar levels of thermal maturity in early-late diagenetic
359
conditions independently from their position in the fold-and-thrust belt (Fig. 4 A and C). Focusing 14
ACCEPTED MANUSCRIPT on Upper Cretaceous-Eocene source rocks, a slight decrease of levels of thermal maturity toward
361
the coastal areas is observed (Fig. 4B). Miocene deposits generally show immature stages of
362
hydrocarbon generation suggesting shallow burial depths (Fig. 4C). In each anticline structure,
363
mixed layers I-S show a general increase of the illite content as function of the stratigraphic age
364
(Fig. 4A).
RI PT
360
365
5. Thermal modeling
367
Simplified reconstructions of the burial and thermal history of the Lower Cretaceuos succession in
368
the Arsenjan area and of the Ordovician to Quaternary sedimentary successions in the Kuh-e-
369
Surmeh and Kuh-e-Asalujeh anticlines were constrained by inorganic thermal indicators and
370
performed using the software package Basin Mod® 1-D (Figs. 6, 7 and 8). The main assumptions
371
for modeling are that: (1) rock decompaction factors apply only to clastic deposits, according to
372
Sclater and Christie’s method (1980); (2) seawater depth variations in time are assumed as not
373
relevant, because thermal evolution is mainly affected by sediment thickness rather than by water
374
depth (Butler, 1992); (3) thermal modeling is performed using LLNL Easy %Ro method based on
375
Burnham and Sweeney (1989) and Sweeney and Burnham (1990); and (4) geothermal gradients
376
between 15 and 24°C/km were tested on the basis of thermochronometric studies in the High
377
Zagros (Gavillot et al., 2010), results from tectonic modelling (Mouthereau et al., 2006),
378
microthermometry of fluid inclusions (Ceriani et al., 2011) and measurements from deep wells for
379
the Fars area (Bordenave, 2008).
380
Thicknesses have been calculated from geological maps and/or from subsurface stratigraphy. Age
381
constraints for Miocene to Quaternary clastic deposits are from Pirouz et al. (2015), Khadivi et al.
382
(2010), Ruh et al. (2014), and Homke et al. (2004). Mixed layers I-S from outcrop samples were
383
projected to a pseudo/paleo-depth based on stratigraphic position and measured thicknesses of the
384
units. Then mixed layers I-S, representative of individual samples, were converted into vitrinite
385
reflectance-equivalent values by the correlation of vitrinite reflectance and expandability (reverse of
AC C
EP
TE D
M AN U
SC
366
15
ACCEPTED MANUSCRIPT 386
illite content in I-S) data based on the kinetic model of vitrinite maturation of Burnham and
387
Sweeney (1989) and the kinetics of the I-S reaction determined by Hillier et al. (1995).
388
5.1 High Zagros
390
Remnants of thrust sheets of ophiolitic mantle and crustal rocks and of their sedimentary cover are
391
preserved in the Neyriz area (Fig. 1B). In the Arsenjan area, ophiolite units occur only as
392
radiolarite-rich covers and clasts of pillow to massive lavas are observed in the Gurpi Fm. (Haines
393
and Reynolds, 1980). From these geological evidences, in order to determine the thickness of
394
ophiolite units in the inner sector of the Zagros fold and thrust belt, we propose a burial and thermal
395
model for the Arsenjan area that includes overthusting of ophiolite bodies during the Late
396
Cretaceous and their partial erosion (Figs. 6A and B).
397
The reconstructed evolution depicted in figure 6A begins in the Jurassic with sedimentation of 900
398
m thick continental-shelf deposits of the Khami Group followed by 560 m of shallow marine shales
399
and limestones of the Bangestan Group during the Cretaceous. The Late Cretaceous sedimentation
400
was followed by the emplacement of 1,100 m thick ophiolite units that overthrust atop the
401
Bangestan Group during the latest Turonian. This thrusting event exposed the ophiolite units to
402
erosion, removing about 200 m of rocks during the Maastrichtian and leading to the accumulation
403
of ophiolite clasts in the Upper Cretaceous-Paleocene conglomerates and in the Gurpi Fm. Erosion
404
has been set in the Maastrichtian as rudist-dominated platform carbonate sedimentation of the
405
Tarbur Fm caps the ophiolite body in the Neyriz area (James and Wind, 1965; Moghadam and
406
Stern, 2015).
407
During the Paleocene, ~300 m-thick red shale and evaporites of the Sachun Fm. were deposited in
408
the Arsenjan area. They graded up with dolostones of the Jahrum Fm. in the Eocene and evolved to
409
shallow marine Nummulites-bearing limestones of the Asmari Fm documenting persistent sabkha-
410
like conditions until Oligocene time.
AC C
EP
TE D
M AN U
SC
RI PT
389
16
ACCEPTED MANUSCRIPT From the Early Miocene onwards, a large amount of marls and siliciclastic rocks (Razak, Agha Jari,
412
and Bakthiyari Fms; Khadivi et al., 2010) buried the sedimentary succession at depths of ~5.1 km.
413
Maximum burial took place in the Late Miocene when both the Bangenstan and the Khami Groups
414
(in particular the Kazdhumi and the Gadvan Fms) experienced early mature stages of hydrocarbon
415
generation (Fig. 6B). The onset of hydrocarbon generation occurred in the Early Eocene for the
416
Gadvan Fm. and during the middle Miocene for the Kazdhumi Fm.
417
Since the Late Miocene, a 3650 m-thick overburden composed of Late Cretaceous to Miocene
418
sedimentary cover and ophiolite units began to be eroded and the outcropping succession exhumed.
419
The burial history has been reconstructed by a constant geothermal gradient of 20°C/km which
420
provide the best fit matching paleothermal data (Fig. 6C) and is consistent with surface and
421
subsurface thickness of the stratigraphic section and (U‐Th)/He thermochronometric data by
422
Gavillot et al. (2010).
423
A sensitivity analysis was also performed assuming different geothermal gradients (between 15 and
424
24°C/km) derived from thermochronological data from the High Zagros (Gavillot et al., 2010),
425
results from tectonic modelling (Mouthereau et al., 2006), microthermometry of fluid inclusions
426
(Ceriani et al., 2011) and measurements from deep wells for the Fars area (Bordenave, 2008). The
427
overburden calculated atop the Bangestan Group varies from 3,000 m to 5,250 m as a function of
428
decreasing geothermal gradient (Tab. 3). A 5250 m-thick overburden (case A in Tab.3) is not
429
geologically valid as it implies the deposition of more than 3 km-thick siliciclastic deposits eroded
430
since the Early Pliocene. An overburden of 3000 m (case C) is stratigraphically acceptable but the
431
thermal maturity curve calculated by a geothermal gradient of 24°C/km overestimates levels of
432
thermal maturity of the Bangestan Group. Simulating ophiolite thrust sheets thicker than 1,100 m
433
during the Late Cretaceous overestimates I-S derived vitrinite reflectance equivalent values, leading
434
to a present-day thermal maturity curve that does not match paleothermal constraints.
SC
M AN U
TE D
EP
AC C
435 436
RI PT
411
5.2 Central Fars 17
ACCEPTED MANUSCRIPT The burial history of the Kuh-e-Surmeh anticline in the Central Fars begins in the Late Ordovician
438
with the deposition of the ca. 400 m-thick Gahkum Fm. (Fig. 7A). Hiatuses and erosional episodes
439
influenced Silurian to Carboniferous sedimentation in several areas of the Zagros and Central Fars
440
as a consequence of sea level changes (Alavi, 1994). During Permian and Triassic times,
441
sedimentary facies in the Central Fars changed from clastic to carbonate platform sediments with
442
the deposition of about 1,300 m of oolitic shallow-water carbonates interbedded with evaporites
443
belonging to the Dehram Group and 300 m of dolomitic limestones and evaporites of the Dashtak
444
Fm. In the Early Jurassic, about 200 m of thin-bedded dolostones and shales of the Neyriz Fm. were
445
followed by a 500 m-thick succession of Jurassic-Lower Cretaceous continental shelf deposits
446
belonging to the Khami Group. Anoxic conditions prevailed until the Early Cretaceous leading to
447
the deposition of dark bituminous shales (Kazhdumi Fm.).
448
Late Cretaceous sedimentation was characterized by neritic conditions with the deposition of the
449
youngest formations of the Bangenstan Group (Ilam-Sarvak Fms) followed by deeper water
450
conditions with marls and shales of the Gurpi and Pabdeh Fms (about 900 m).
451
During Oligocene time, the marls and shales of the Pabdeh Fm. graded up and interfingered the
452
shallow marine limestones of the Asmari Fm. From Early Miocene onwards, sedimentation rates
453
increased and a large amount of marls and siliciclastic rocks (Gachsaran, Mishan and Agha Jari
454
Fms) buried the sedimentary succession of the Kuh-e-Surmeh anticline at depths of ~6.3 km in the
455
early Pleistocene. At that time, maximum levels of thermal maturity were experienced and recorded
456
by mixed layer I-S (Fig. 7C). Upper Cretaceous-Miocene rocks were thermally immature whereas
457
the underlying deposits of the Khami Group experienced early mature stages of hydrocarbon
458
generation at depths of 3.5 km (Fig. 7B). The Gahkum Fm. experienced maximum temperatures of
459
145 °C in mid-mature stage of hydrocarbon generation at time of maximum burial while it entered
460
the early mature stage of hydrocarbon generation during the Late Cretaceous (at about 80 Ma) prior
461
to Zagros folding.
AC C
EP
TE D
M AN U
SC
RI PT
437
18
ACCEPTED MANUSCRIPT This burial reconstruction, modelled assuming a geothermal gradient of 20°C/km (case B in tab. 3),
463
accounts for the erosion of 700 m-thick Agha Jari Fm. since the early Pleistocene and indicates that
464
sedimentary load was the main factor affecting levels of thermal maturity. Furthermore, thermal
465
modelling of figure 7A is consistent with magnetostratigraphic data for growth strata for the Kuh-e-
466
Ghol Ghol anticline (about 40 km to the SE of the Kuh-e-Surmeh anticline) where initiation of
467
deformation took place in the Late Pliocene and continued until the early Pleistocene (Ruh et al.,
468
2014).
469
Other attempts to simulate the burial evolution of the Kuh-e-Surmeh anticline lead to a less accurate
470
thermal maturity curve implying: i) the accumulation of large amounts of siliciclastic deposits (>2
471
km thick) from Miocene to early Pleistocene (case A in tab. 3), or ii) the absence of erosion during
472
the Quaternary (case C in tab. 3). Both hypothesis do not satisfy geological data such as the present-
473
day thickness of the Agha Jari Fm. in adjacent areas (~1000 m thick) and the exhumation of the
474
Kuh-e-Surmeh anticline that is very unlikely to have occurred without erosion.
475
The most accurate reconstruction indicates a maximum overburden of 3,200 m atop the Bangestan
476
Group.
477
TE D
M AN U
SC
RI PT
462
5.3 Coastal Fars
479
Similar burial and thermal conditions can be reconstructed for the Cretaceous to Quaternary
480
sedimentary succession cropping out in the Kuh-e-Asalujeh anticline in Coastal Fars.
481
Figures 8A and 8B display that the base of the Bangestan Group and the top of the Khami Group
482
experienced burial depths in the order of 3.2 km in early mature stage of hydrocarbon generation
483
during the early Pleistocene when large supply of clastic material lead to the deposition of about
484
1,000 m of the Bakthiyari Fm. Upper Cretaceous to lower Pleistocene deposits overlying the
485
Bangestan Group were thermally immature as indicated by the present-day maturity curve
486
constrained by inorganic thermal indicators (Fig. 8C). Erosion in Quaternary times removed about
487
200 m of the sedimentary succession. The type of evolution outlined here allows the best calibration
AC C
EP
478
19
ACCEPTED MANUSCRIPT 488
against measured data, suggesting that levels of thermal maturity are controlled by sedimentary
489
burial. Also for this area, burial and thermal models built by assuming geothermal gradients of 15
490
and 24°C/km do not satisfy geological and paleothermal data (Tab. 3; Fig. 8C).
491
6. Discussion
493
6.1 Maximun ophiolite thickness in the Arsenjan area
494
Short range ordered I-S for the Kazdhumi and Gadvan Fms in the Arsenjan area and thermal model
495
of figure 6A allow us to conclude that a 3,650 m thick overburden atop the Kazdhumi Fm. was the
496
main factor responsible for the observed thermal maturity values.
497
In the Arsenjan area, ophiolite thrust sheets mainly consist of radiolarites and lack of oceanic
498
mantle and crustal rocks, whereas the present-day Neyriz ophiolite crustal sequence is represented
499
by harzburgites and by a ~700–900 m thick, highly fragmented sheeted dike complex with pillow to
500
massive lavas associated with radiolarites and Upper Cretaceous pelagic limestones. The occurrence
501
of different portions of ophiolite units (sedimentary vs. magmatic rocks) in the two areas suggests
502
that the Arsenjan ophiolites were close to the obduction front (to the E-SE of Arsenjan no more
503
ophiolites crop out), whereas the Neyriz ophiolites were in a more internal position.
504
Short-range ordered I-S measured in the Arsenjan area allow us to constrain the original thickness
505
of the ophiolite sheet close to its front. In fact, inorganic paleothermal indicators display levels of
506
thermal maturity consistent with a 1,100 m thick ophiolite thrust sheet that was partially eroded
507
during the Maastrichtian (Fig. 6A). Erosion has been estimated in 200 m, explaining the
508
accumulation of ophiolite clasts in the Upper Cretaceous-Paleocene conglomerates and within the
509
Gurpi Fm. Simulating the emplacement of thicker ophiolite bodies during the Late Cretaceous or
510
higher amounts of erosion, leads to an overestimation of the I-S derived vitrinite reflectance
511
equivalent values and related levels of thermal maturity. This result allows us to exclude the
512
obduction of large and thick slices of ophiolite units in the Interior Fars and to suggest that the
AC C
EP
TE D
M AN U
SC
RI PT
492
20
ACCEPTED MANUSCRIPT 513
extent of the Maastrichtian erosional phase, was negligible in comparison to the removal of 3,500
514
m-thick ophiolite units in Oman (Aldega et al., 2017).
515
6.2 Maximum burial in the Fars province
517
Inorganic and organic thermal indicators show that the sedimentary cover experienced similar levels
518
of thermal maturity from the internal to the external zones, consistent with early-late diagenetic
519
conditions. Only a slight reduction of the overburden thickness can be observed from hinterland to
520
foreland. In the most internal area, where the Bangestan Group (Kazdhumi Fm.) crops out, a
521
currently eroded overburden of 3.65 km has been calculated (case B in tab. 3).
522
In the Sarvestan area, Carminati et al. (2016) pointed out similar burial depths for the Bangestan
523
Group through the interpretation of seismic lines and 1D thermal modelling. In particular, the
524
Sarvak Fm. was buried to depths of 3.6 km by a large amount of marls and siliciclastic rocks in
525
Early Pliocene times.
526
In the Central and Coastal Fars, organic and inorganic indicators from the Paleocene to Miocene
527
deposits display low levels of thermal maturity in early diagenetic conditions and slightly lower
528
burial depths. In fact, in the Kuh-e-Surmeh and Kuh-e-Asalujeh anticlines, 1D thermal models
529
evidenced a lithostatic load of about 3.2 km for the top of the Bangestan Group occurring in early
530
Pleistocene times. The decrease of sedimentary load and of the associated levels of thermal maturity
531
towards the foreland is consistent with the findings of recent papers regarding source rocks
532
evaluation in the Persian Gulf. Mashhadi et al. (2015) pointed out that the Kazdhumi Fm. and the
533
overlying Gurpi and Pabdeh Fms are immature in the Central Persian Gulf with vitrinite reflectance
534
values <0.5% suggesting lower burial depths than those recorded in the Coastal Fars.
535
Sfidari et al. (2016), on the basis of vitrinite reflectance measurements and Rock-Eval pyrolysis
536
data, reconstructed the thickness of the sedimentary pile from the Kazdhumi Fm. to the Mishan Fm.
537
prior to the onset of folding in the Zagros Simply Folded Belt indicating a decreasing thickness
538
from the Interior Fars towards the Persian Gulf. These thicknesses (3000-3500 for the Interior Fars,
AC C
EP
TE D
M AN U
SC
RI PT
516
21
ACCEPTED MANUSCRIPT 539
1500-2000 for the Coastal Fars and <1000 m for the Central Persian Gulf) are lower than those
540
calculated by our thermal models because the thickness of syn-orogenic deposits (Agha Jari and
541
Bakhtiyari Fms) eroded during the uplift of the anticlines is not considered.
542
6.3 Timing of thermal maturation
544
Even if the Bangestan Group (in particular the Kazdhumi Fm.) experienced similar burial depths
545
(3.65 km in the Arsenjan area, 3.2 km in the Central and Coastal Fars) and levels of thermal
546
maturity (%I in I-S between 68 and 75%, Fig. 4A and Tab.1) in different sectors of the Zagros fold-
547
and-thrust belt, 1D thermal models point out that the onset of hydrocarbon generation was largely
548
diachronous throughout the Zagros foreland basin depending on timing of deformation and the age
549
of synorogenic deposits.
550
In particular, the onset of hydrocarbon generation for the Kazdhumi Fm. occurred in middle
551
Miocene times at about 13 Ma in the Arsenjan area (Fig. 6B) and at 11 Ma in the Interior Fars
552
(Carminati et al., 2016). In the Kuh-e-Sim anticline, hydrocarbon generation occurred after the early
553
Pliocene (Aldega et al., 2014) rejuvenating towards the external part of the belt. In the Kuh-e-
554
Surmeh anticline, the Kazdhumi Fm. enter the early mature stages of hydrocarbon generation in the
555
early Pleistocene (Fig. 7B) and in the late Pleistocene in the Coastal Fars (Fig. 8B).
556
This is a consequence of foreland propagating thrusting inducing progressive flexural subsidence
557
from the High Zagros to the Cosatal Fars with sedimentation of the Bakthiyari conglomerates (14.8
558
Ma close to the High Zagros fault in NE Fars; Khadivi et al., 2010; 3.6-3.2 Ma in the Central Fars,
559
Ruh et al., 2014; 3.0 Ma for the Coastal Fars, Homke et al., 2004).
AC C
EP
TE D
M AN U
SC
RI PT
543
560 561
6.4 Geological and paleothermal constraints for reconstructing the geometry of eroded portion of
562
the Zagros fold-and-thrust belt
563
22
ACCEPTED MANUSCRIPT Surface (geological maps and stratigraphic sections) and subsurface (borehole) stratigraphic data
565
were used to draw a cross section across the Zagros fold-and-thrust belt using 3DMove 2014
566
software by Midland Valley (http://www.mve.com/software/move) that is extensively described in
567
Bigi et al. (2018).
568
The surface geological data were extracted from a series of 1:100.000 sheets of the Geological Map
569
of Iran: Kangan (no. 20867W), Hangam (no. 20861E), Firoozabad (no. 6547), Kushk (no. 6647),
570
Sarvestan (no. 6648), Arsenjan (no. 6649) and Abadeh-e-Tashk (no. 6649) and derive from 2011
571
and 2012 field campaigns. Measured stratigraphic sections and boreholes stratigraphy from NIOC
572
(National Iranian Oil Company) internal reports and thickness and facies information from the
573
paleogeographic maps of Koop and Stoneley (1982) and Pyriaei et al. (2010) integrate the dataset.
574
Borehole stratigraphic data come from several wells drilled in the area such as: Assaluyeh West-1
575
well (located 11 km ESE of the section), Kangan-1A well (40 km NW), Nar-1 well (6 km NW),
576
Varavi-1 well, Safid Baghun-1 well (9 km ESE), Cham-e-Noori-1 well (45 km ESE), Sefid Zakhur-
577
1 well (6 km ESE), Aghar-2 well (25 km ESE), Sim-1 well (1 km ESE), Toudej-1 well (71 km
578
ESE), Sarvestan-1 well (11 km ESE) (Fig. 3).
579
The cross-section of figure 9 highlights a geometry of the Zagros fold-and-thrust belt controlled by
580
a complex tectonic evolution that includes thin-skinned shortening of a thick sedimentary pile
581
characterized by facies and thickness variations associated with pre-subduction tectonics.
582
Major thickness changes occur in the Dashtak Fm., Khami and Bangestan Groups, beneath the Kuh-
583
e-Asalujeh and Pazan anticlines in the Coastal Fars, beneath the Daryau anticline in the Central
584
Fars, and in the Sarvestan plain in the Interior Fars. These thickness variations are thought to be
585
associated with E-W and NNW-SSE normal fault systems (Navabpour et al., 2010) generated by
586
Paleozoic-Mesozoic rifting-related extension. In the Interior Fars (from the Sarvestan plain to the
587
north), lateral facies and thickness changes were due to differential flexure of the Arabian plate
588
controlled by the Cretaceous obduction of ophiolites, cropping out in the Neyriz and Arsenjan area
589
(Pyriaei et al., 2010; Carminati et al., 2016). In agreement with Bigi et al. (2018), we propose that
AC C
EP
TE D
M AN U
SC
RI PT
564
23
ACCEPTED MANUSCRIPT thin-skinned thrust faults within the Gachsaran and Dashtak Fms either inverted pre-existing
591
extensional structures or rotated them as in the case of the Kuh-e-Surmeh, Kuh-e-Gareh and Kuh-e-
592
Amhadi anticlines (Fig. 9). In these areas, back-thrusts developed in response of thickness
593
variations of the Bangestan Group and Dashtak Fm. which are controlled by Cretaceous and
594
Triassic normal faults as extensively described by Bigi et al. (2018).
595
Paleothermal indicators and 1D thermal modelling were included in the cross-section providing
596
maximum burial depths experienced by the sedimentary successions from the inner to the outer
597
zones of the Zagros fold-and-thrust belt. Eroded portion of structural units has been reconstructed
598
by projecting the calculated overburden onto the geological cross-section. This allowed us to infer
599
the geometry of eroded structures in different areas of the chain providing information on the
600
original thickness of the thrust belt (Fig. 9).
601
Paleothermal data exclude the occurrence of significant overthrusting during growth of the orogenic
602
wedge, coherently with the low shortening value (7% on average) that characterizes the cross-
603
section (Bigi et al., 2018) and consistent with shortening evaluations from previous works (e.g.,
604
Alavi, 2007; Mouthereau et al., 2007). Burial and thermal model of figure 8A clearly indicates 3.2
605
km of overburden for the Bangestan Group exposed at surface in the Kuh-e-Asalujeh anticline, thus
606
constraining exhumation and erosion to more than 3 km. This anticline is surrounded by synclines
607
characterized by much lower exhumation (down to 200 m, Fig. 9) and outcrops of Miocene-
608
Pleistocene rocks. In the Kuh-e-Surmeh anticline, thermal modelling (Fig. 7A) indicates similar
609
overburden (3.2 km) for the Bangestan Group and documents more than 6 km of exhumation north
610
of the cross section trace (Fig. 1B) where Silurian strata crop out. This amount of exhumation is
611
consistent with the vertical throw of the Surmeh fault involving the basement, in agreement with
612
previous studies (Mouthereau et al., 2006; 2007; Bigi et al., 2018) and with active seismicity
613
(Berberian, 1995; Ansari and Zamani, 2014; Mouthereau et al., 2012).
614
From the Kuh-e-Surmeh anticline to the Sarvestan plain, the eroded portion of the thrust belt
615
increases progressively from ca. 2 km to ca. 4.5 km. The Sarvestan plain represents a structural low
AC C
EP
TE D
M AN U
SC
RI PT
590
24
ACCEPTED MANUSCRIPT 616
where eroded overburden varies between ca. 400 m and 2.3 km. North-east of the Sarvestan plain,
617
structural elevation is constant, with erosion of 3.5 km, until the Kuh-e-Siah anticline characterized
618
by 5 km of erosion and higher structural elevation.
620
7. Conclusions
621
Main findings can be summarized as follows:
622
•
RI PT
619
Vitrinite reflectance measurements and temperature-dependent clay minerals show that the sedimentary successions of the Zagros fold-and-thrust belt, in the Fars province (Iran),
624
experienced similar levels of thermal maturity from the internal (Interior Fars) to the
625
external zones (Coastal Fars) in early-late diagenetic conditions. •
M AN U
626
SC
623
1-D thermal models display a slightly decreasing lithostatic load towards the foreland from 3.65 km to 3.2 km for the Bangestan Group and indicate that sedimentary burial is the main
628
factor affecting thermal maturity. Only, in the Arsenjan area of the High Zagros, the
629
overburden includes thin tectonic slices of ophiolite units, maximum 1,100 m-thick. •
deposits more than on the amount of maximum burial.
631 632
The onset of oil generation depends on timing of deformation and the age of synorogenic
•
Organic and inorganic thermal indicators were used to reconstruct the eroded portion of the
EP
630
TE D
627
Zagros thrust belt that varies between 200 m in synclines and 6 km in anticlines, confirming
634
minor shortening amounts (average shortening is 7%).
635
AC C
633
636
Acknowledgements
637
Financial support from the Darius Programme and Progetti di Ateneo 2017 to Eugenio Carminati,
638
from Progetti di Ateneo 2016 to Luca Aldega, and from PRIN2015 (Project 2015EC9PJ5_001) to
639
C. Doglioni are acknowledged. The National Iranian Oil Company is thanked for assistance in
640
fieldwork organization and logistics and for allowing the publication of this work and of the
641
associated data. Ali Shaban and Hossain Narimani are thanked for sharing the fieldwork campaign. 25
ACCEPTED MANUSCRIPT 642
Midland Valley is acknowledged for providing educational license for Move software. Frederic
643
Mouthereau, Dale R. Issler and Jeremy Powell provided constructive criticisms to an early version
644
of the manuscript. We are grateful to A. Ceriani and three anonymous reviewers for detailed
645
revisions and helpful suggestions.
RI PT
646
References
648
Ahmadhadi, F., Lacombe, O., Daniel, J.M., 2007. Early reactivation of basement faults in Central
649
Zagros (SW Iran): evidence from pre-folding fracture populations in the Asmari Formation and
650
Lower Tertiary paleogeography. In: Lacombe, O., Lavé, J., Roure, F., Vergés, J., (Eds.), Thrust
651
belts and foreland basins: from fold kinematics to hydrocarbon systems, Springer-Verlag, p. 205-
652
228.
655 656
M AN U
654
Alavi, M., 1994. Tectonics of the Zagros orogenic belt of Iran: new data and interpretations. Tectonophysics 229, 211-238.
Alavi, M., 2004. Regional stratigraphy of the Zagros fold-thrust belt of Iran and its proforeland
TE D
653
SC
647
evolution. Am. J. Sci. 304, 1-20.
Alavi, M., 2007. Structures of the Zagros fold-thrust belt in Iran. Am. J. Sci. 307, 1064-1095.
658
Aldega, L., Corrado, S., Grasso, M., Maniscalco, R., 2007a. Correlation of diagenetic data from
659
organic and inorganic studies in the Apenninic-Maghrebian fold-and-thrust belt: a case study
660
from Eastern Sicily. J. Geol. 115 (3), 335-353.
662
AC C
661
EP
657
Aldega, L., Botti, F., Corrado, S., 2007b. Clay mineral assemblages and vitrinite reflectance in the Laga Basin (Central Apennines, Italy): What do they record? Clays Clay Min. 55, 504-518.
663
Aldega, L., Corrado, S., Di Paolo, L., Somma, R., Maniscalco, R., Balestrieri, M.L., 2011. Shallow
664
burial and exhumation of the Peloritani Mts. (NE Sicily, Italy): Insight from paleo-thermal and
665
structural indicators. Geol. Soc. Am. Bull. 123, 132-149.
26
ACCEPTED MANUSCRIPT 666
Aldega, L., Corrado, S., Carminati C., Shaban, A., Sherkati, S., 2014. Thermal evolution of the
667
Kuh-e-Asmari and Sim anticlines in the Zagros fold-and-thrust belt: implications for
668
hydrocarbon generation. Mar. Petrol. Geol. 57, 1-13. Aldega, L., Carminati, E., Scharf, A., Mattern, F., Al-Wardi, M., 2017. Estimating original
670
thickness and extent of the Semail Ophiolite in the eastern Oman Mountains by paleothermal
671
indicators. Mar. Petrol. Geol. 84, 18-33.
RI PT
669
Árkai, P., 2002. Phyllosilicates in very low-grade metamorphism: transformation to micas. In:
673
Mottana, A., Sassi, F.P., Thompson, J.B., Guggenheim, S., (Eds.), Micas: Crystal Chemistry and
674
Metamorphic Petrology: Reviews in Mineralogy and Geochemistry, The Mineralogical Society
675
of America, Washington, D.C., v. 46, pp. 463-478.
679 680 681 682 683
M AN U
678
S0210.
Bahroudi, A., Talbot, C.J., 2003. The configuration of the basement beneath the Zagros basin. J. Petrol. Geol. 26, 257–282.
TE D
677
Ansari, S., Zamani, A., 2014. Short-term seismic crustal deformation of Iran. Ann. Geophys. 57 (2),
Berberian, M., King, G.C.P., 1981. Towards the paleogeography and tectonic evolution of Iran. Can. J. Earth Sci. 18, 210–265.
Berberian, M., 1995. Master “blind” thrust faults hidden under the Zagros folds: active basement
EP
676
SC
672
tectonics and surface tectonics surface morphotectonics. Tectonophysics 241, 193–224. Bigi S., Carminati E., Aldega L., Trippetta F., Kavoosi M.A., 2018. Zagros fold and thrust belt in
685
the Fars Province (Iran): I. Control of thickness/rheology of sediments and pre-thrusting
686
tectonics
687
j.marpetgeo.2018.01.005
AC C
684
on
structural
style
and
shortening.
Mar.
Petrol.
Geol.
doi:
10.1016/
688
Bordenave, M.L., 2008. The origin of the Permo-Triassic gas accumulations in the Iranian Zagros
689
foldbelt and contiguous offshore areas: a review of the Paleozoic petroleum system. J. Petrol.
690
Geol. 31, 3-42.
27
ACCEPTED MANUSCRIPT 691
Botti, F., Aldega, L., Corrado, S., 2004. Sedimentary and tectonic burial evolution of the Northern
692
Apennines in the Modena-Bologna area: constraints from combined stratigraphic, structural,
693
organic matter and clay mineral data of Neogene thrust-top basins. Geodin. Acta 17 (3), 185-
694
203. Breton, J.P., Béchennec,
F., Le Métour, J., Moen-Maurel, L., Razin, P., 2004. Eoalpine
RI PT
695 696
(Cretaceous) evolution of the Oman Tethyan continental margin: insights from a structural field
697
study in Jabal Akhdar (Oman Mountains). GeoArabia. 9, 41–58.
700 701
SC
699
Burnham, A.K., Sweeney, J.J., 1989. A chemical kinetic model of vitrinite maturation and reflectance. Geochimica et Cosmochimica Acta, 53, 2649–2657.
Burst, J.F., Jr. 1959. Post diagenetic clay mineral-environmental relationships in the Gulf Coast
M AN U
698
Eocene. Clays Clay Minerals 6, 327-341.
Bustin, R.M., Barnes, M.A., Barnes, W.C., 1990. Determining levels of organic diagenesis in
703
sediments and fossil fuels. In: McIleareath, I.A., Morrow, D.W., (Eds.), Diagenesis: Geoscience
704
Canada Reprint, 4th Series, p. 205–226.
TE D
702
Butler, R.W.H., 1992. Hydrocarbon maturation, migration and tectonic loading in the western Alps.
706
In: England, W.A., Fleet, A.J., (Eds.), Petroleum Migration: Geological Society of London,
707
Special Publications, v. 59, p. 227–244.
EP
705
Calabrò, R.A., Corrado, S., Di Bucci, D., Robustini, P., Tornaghi, M., 2003. Thin-skinned vs.thick-
709
skinned tectonics in the Matese Massif, Central-Southern Apennines (Italy). Tectonophysics 377,
710
269-297.
AC C
708
711
Callot, J.P., Jahani, S., Letouzey, J., 2007. The role of pre-existing diapirs in fold and thrust belt
712
development. In: Lacombe, O., Roure, F., Lavé, J., Verges, J., (Eds.), Thrust Belt and Foreland
713
Basin, Springer, Berlin, p. 307-323.
714
Caricchi, C., Aldega, L., Corrado S., 2015. Reconstruction of maximum burial along the Northern
715
Apennines thrust wedge (Italy) by indicators of thermal exposure and modeling. Geol. Soc. Am.
716
Bull. 127 (3-4), 428-442. 28
ACCEPTED MANUSCRIPT 717
Carlini, M., Artoni, A., Aldega, L., Balestrieri, M.L., Corrado, S., Vescovi, P., Bernini, M., Torelli,
718
L., 2013. Exhumation and reshaping of far-travelled/allochthonous tectonic units in mountain
719
belts. New insights for the relationships between shortening and coeval extension in the western
720
Northern Apennines (Italy). Tectonophysics 608, 267–287. Carminati, E., Aldega, L., Bigi, S., Corrado, S., D’Ambrogi, C., Mohammadi, P., Shaban, A.,
722
Sherkati, S., 2013. Control of Cambrian evaporites on fracturing in fault-related anticlines in the
723
Zagros fold-and-thrust belt. Int. J. Earth Sci. 102, 1237-1255.
RI PT
721
Carminati, E., Aldega, L., Trippetta, F., Shaban, A., Narimani, H., Sherkati, S., 2014. Control of
725
folding and faulting on fracturing in the Zagros (Iran): the Kuh-e-Sarbalesh anticline case. J.
726
Asian Earth Sci., 79, 400-414.
M AN U
SC
724
727
Carminati, E., Aldega, L., Bigi, S., Minelli, G., Shaban, A., 2016. Not so simple "simply-folded
728
Zagros": the role of pre-collisional extensional faulting, salt tectonics and multi-stage thrusting
729
in the Sarvestan transfer zone (Fars, Iran). Tectonophysics 671, 235-248. Ceriani, A., Calabrò R., Di Giulio A., Buonaguro R., 2011. Diagenetic and thermal history of the
731
Jurassic-Tertiary succession of the Zagros Mountains in the Dezful Embayment (SW Iran):
732
constraints from fluid inclusions. Int. J. Earth Sci. 100, 1265-1281.
735 736
(8), 1189-1198.
EP
734
Chapple, W.M., 1978. Mechanics of thin-skinned fold-and-thrust belts. Geol. Soc. Am. Bull. 89
Corcoran, D.V., Doré, A.G., 2005. A review of techniques for the estimation of magnitude and
AC C
733
TE D
730
timing of exhumation in offshore basins. Earth-Sci. Rev. 72, 129–168.
737
Corrado, S., Aldega, L., Balestrieri, M.L., Maniscalco, R., Grasso, M., 2009. Structural evolution of
738
the sedimetary accretionary wedge of the alpine system in Eastern Sicily: thermal and
739
thermochronological constraints. Geol. Soc. Am. Bull. 121, 1475-1490.
740 741
Corrado, S., Aldega, L., Zattin, M., 2010a. Sedimentary vs. tectonic burial and exhumation along the Apennines (Italy). J. Virtual Explor. 36. doi:10.3809/jvirtex.2009.00232.
29
ACCEPTED MANUSCRIPT 742
Corrado, S., Invernizzi, C., Aldega, L., D'Errico, M., Di Leo, P., Mazzoli, S., Zattin, M., 2010b.
743
Testing the validity of organic and inorganic thermal indicators in different tectonic settings from
744
continental subduction to collision: the case history of the Calabria-Lucania border (southern
745
Apennines, Italy). J. Geol. Soc. 167, 985-999.
748 749
to deep in the crust. J. Struct. Geol. 5, 113–123.
RI PT
747
Coward, M.P., 1983. Thrust tectonics, thin skinned or thick skinned, and the continuation of thrusts
DeCelles, P.G., Mitra, G., 1995. History of the Sevier orogenic wedge in terms of critical taper models, northeast Utah and southwest Wyoming. Geol. Soc. Am. Bull. 107 (4), 454-462.
SC
746
Di Paolo, L., Aldega, L., Corrado, S., Mastalerz, M., 2012. Maximum burial and unroofing of Mt.
751
Judica recess area in Sicily: implication for the Apenninic-Maghrebian wedge dynamics.
752
Tectonophysics 530–531, 193–207.
M AN U
750
Doré, A.G., Cartwright, J.A., Stoker, M.S., Turner, J.P., White, N.J., 2002. Exhumation of the
754
North Atlantic margin: Introduction and background. In: Doré, A.G., Cartwright, J.A., Stoker,
755
M.S., Turner, J.P., White, N., (Eds.), Exhumation of the North Atlantic Margin: Timing,
756
Mechanisms and Implications for Petroleum Exploration. Geological Society of London Special
757
Publication, v. 196, p. 1–12.
TE D
753
Dow, W.G., 1977. Kerogen studies and geological interpretation. J. Geochem. Explor. 7, 79–99.
759
Durand, B., 1980. Sedimentary organic matter and kerogen. Definition and quantitative importance
760
of kerogen. In: Durand, B., (Ed.), Kerogen: Insoluble Organic Matter from Sedimentary Rock.
761
Editions Technip, p. 13–34.
AC C
EP
758
762
Fakhari, M.D., Axen, G.J., Horton, B.K., Hassanzadeh, J., Amini, A., 2008. Revised age of
763
proximal deposits in the Zagros foreland basin and implications for Cenozoic evolution of the
764
High Zagros. Tectonophysics 451, 170–185.
765
Gavillot, Y., Axen, G.J., Stockli, D.F., Horton, B.K., Fakhari, M.D., 2010. Timing of thrust activity
766
in the High Zagros fold–thrust belt, Iran, from (U–Th)/He thermochronometry. Tectonics 29,
767
TC4025. 30
ACCEPTED MANUSCRIPT 768 769
Haines, S.J., Reynolds, P.H., 1980. Early development of Tethys and Jurassic ophiolite displacement. Nature 283, 561-563. Hillier, S., Màtyàs, J., Matter, A., Vasseur, G., 1995. Illite/smectite diagenesis and its variable
771
correlation with vitrinite reflectance in the Pannonian Basin. Clays Clay Minerals 43, 174-183.
772
Hoffman, J., Hower, J., 1979. Clay mineral assemblages as low-grade metamorphic
773
geothermometers: application to the thrust faulted disturbed belt of Montana, USA. In: Scholle,
774
P. A., Schluger, P.S., (Eds.), Aspects of diagenesis. SEPM Special Publication, v. 26, p. 55– 79.
RI PT
770
Homke, S., Vergés, J., Garcés, M., Emami, H., Karpuz R., 2004. Magnetostratigraphy of Miocene–
776
Pliocene Zagros foreland deposits in the front of the Push-e Kush Arc (Lurestan Province, Iran).
777
Earth Planet. Sc. Lett. 225 (3–4), 397-410.
M AN U
SC
775
Hower, J., Eslinger, E.V., Hower, M.E., Perry, E.A., 1976. Mechanism of burial metamorphism of
779
argillaceous sediment, mineralogical and chemical evidence. Geol. Soc. Am. Bull. 87, 725-737.
780
Izquierdo-Llavall, E., Aldega, L., Cantarelli, V., Corrado, S., Gil-Peña, I., Invernizzi, C., Casas-
781
Sainz, A., 2013. On the origin of cleavage in the Central Pyrenees: structural and paleo-thermal
782
study. Tectonophysics 608, 303-318.
784
Jaboyedoff, M., Bussy, F., Kübler, B., Thelin Ph., 2001. Illite “crystallinity” revisited. Clays Clay Min. 49, 156-167.
EP
783
TE D
778
Jacob, H., Hiltmann, W., 1985. Disperse bitumen solids as an indicator for migration and maturity
786
within the scope of prospecting for petroleum and natural gas - A model for NW Germany.
787
DGMK, Forschungsbericht 267, 1-54.
788 789
AC C
785
Jagodzinski, H., 1949. Eindimensionale Fehlordnung in Kristallen und ihr Einfluss auf die Röntgen Interferenzen. Acta Cryst. 2, 201-207.
790
Jahani, S., Callot, J.P., Letouzey, J., Frizon de Lamotte, D., 2009. The eastern termination of the
791
Zagros Fold-and-Thrust Belt, Iran: Structures, evolution, and relationships between salt plugs,
792
folding, and faulting. Tectonics 28, TC6004.
31
ACCEPTED MANUSCRIPT 793 794
James, G.A., Wynd, J.G., 1965. Stratigraphic nomenclature of Iranian oil consortium agreement area. Am. Assoc. Petroleum Geol. Bull. 49, 2182-2245. Khadivi, S., Mouthereau, F., Larrasoaña, J.C., Vergés, J., Lacombe, O., Khademi E., Beamud, E.,
796
Melinte-Dobrinescu, M., Suc, J.P., 2010. Magnetochronology of synorogenic Miocene foreland
797
sediments in the Fars arc of the Zagros Folded Belt (SW Iran). Basin Res. 22, 918-932.
799 800 801
Koop, W.J., Stoneley, R., 1982. Subsidence history of the Middle East Zagros basin, Permian to recent. Philos. Trans. R. Soc. Lond. A. 305, 149-168.
Kübler, B., 1967. La cristallinité de l`illite et les zones tout a fait superieures du métamorphisme.
SC
798
RI PT
795
In: étages tectoniques, Colloque de Neuchâtel 1966, a La Baconniere, Neuchâtel 105–121. Lacombe, O., Mouthereau, F., Kargar, S., Meyer, B., 2006. Late Cenozoic and modern stress fields
803
in the western Fars (Iran): implications for the tectonic and kinematic evolution of Central
804
Zagros. Tectonics 25, TC1003.
M AN U
802
Mashhadi, Z.S., Rabbani, A.R., Kamali, M.R., 2015. Geochemical characteristics and hydrocarbon
806
generation modeling of the Kazhdumi (Early Cretaceous), Gurpi (Late Cretaceous) and Pabdeh
807
(Paleogene) formations, Iranian sector of the Persian Gulf. Mar. Petrol. Geol. 66, 978-997.
809
McQuarrie, N., 2004. Crustal scale geometry of the Zagros fold–thrust belt, Iran. J. Struct. Geol. 26, 519-535.
EP
808
TE D
805
Meneghini, F., Botti, F., Aldega, L., Boschi, C., Corrado, S., Marroni, M., Pandolfi, L., 2012. Hot
811
fluid pumping along shallow-level collisional thrusts: the Monte Rentella Shear Zone, Umbria
812
Apennine, Italy. J. Struct. Geol. 37, 36-52.
AC C
810
813
Merriman, R.J., Frey, M., 1999. Patterns of very low-grade metamorphism in metapelitic rocks. In:
814
Frey, M., Robinson, D., (Eds.), Low grade Metamorphism, Blackwell, Oxford, UK, p. 61-107.
815
Moghadam, H.S., Stern, R.J., 2015. Ophiolites of Iran: keys to understanding the tectonic evolution
816
of SW Asia: (II) Mesozoic ophiolites. J. Asian Earth Sci. 100, 31–59.
32
ACCEPTED MANUSCRIPT 817
Molinaro, M., Leturmy, P., Guezou, J.-C., Frizon de Lamotte, D., Eshraghi, S.A., 2005. The
818
structure and kinematics of the southeastern Zagros fold-thrust belt, Iran: From thin-skinned to
819
thick-skinned tectonics. Tectonics 24, 1–19.
821 822 823
Moore, D.M., Reynolds, R.C. Jr., 1997. X-Ray Diffraction and the identification and analysis of clay minerals: Oxford, UK, Oxford University Press, 378 pp.
RI PT
820
Mouthereau, F., Lacombe, O., Meyer, B., 2006. The Zagros Folded Belt (Fars, Iran): Constraints from topography and critical wedge modelling. Geophys. J. Int. 165, 336-356.
Mouthereau, F., Tensi, J., Bellahsen, N., Lacombe, O., De Boisgrollier, T., Kargar S., 2007.
825
Tertiary sequence of deformation in a thin-skinned/thick-skinned collision belt: The Zagros
826
Folded Belt (Fars, Iran). Tectonics 26, TC5006.
M AN U
SC
824
827
Mouthereau, F., Lacombe, O., Vergés, J., 2012. Building the Zagros collisional orogen: Timing,
828
strain distribution and the dynamics of Arabia/Eurasia plate convergence. Tectonophysics 532–
829
535, 27-60.
Mukhopadhyay, P.K., 1994. Vitrinite reflectance as maturity parameter: Petrographic and molecular
831
characterization and its applications to basin modeling. In: Mukhopadhyay, P.K., Dow, W.G.,
832
(Eds.). Vitrinite Reflectance as a Maturity Parameter: Applications and Limitations: American
833
Chemical Society Symposium Series, v. 570, p. 1–24.
EP
TE D
830
Najafi, M., Yassaghi, A, Bahroudi, A., Vergés, J., Sherkati, S., 2014. Impact of the Late Triassic
835
Dashtak intermediate detachment horizon on anticline geometry in the Central Frontal Fars, SE
836
Zagros fold belt, Iran. Mar. Petrol. Geol. 54, 23-36.
AC C
834
837
Navabpour, P., Angelier, J., Barrier, E., 2010. Mesozoic extensional brittle tectonics of the Arabian
838
passive margin, inverted in the Zagros collision (Iran, interior Fars). Geol. Soc. London Spec.
839
Publ. 330, 65–96.
840 841
Navabpour, P., Barrier, E., 2012. Stress states in the Zagros fold-and-thrust belt from passive margin to collisional tectonic setting. Tectonophysics 581, 76-83.
33
ACCEPTED MANUSCRIPT 842
Perri, F., Caracciolo, L., Cavalcante, F., Corrado, S., Critelli, S., Muto, F., Dominici, R., 2016.
843
Sedimentary and thermal evolution of the Eocene-Oligocene mudrocks from the southwestern
844
Thrace Basin (NE Greece). Basin Research 28, 319-339. Pirouz, M., Simpson, G., Bahroudi, A., Azhdari, A., 2011. Neogene sediments and modern
846
depositional environments of the Zagros foreland basin system. Geol. Mag. 148 (5-6), 838-853.
847
Pirouz, M., Simpson, G., Chiaradia, M., 2015. Constraint on foreland basin migration in the Zagros
848
RI PT
845
mountain belt using Sr isotope stratigraphy. Basin Res. 27, 714-728.
Piryaei, A., Reijmer, J.J.G., Van Buchem, F.S.P., Yazdi-Moghadam, M., Sadouni, J., Danelian, T.,
850
2010. The influence of Late Cretaceous tectonic processes on sedimentation patterns along the
851
northeastern Arabian plate margin (Fars Province, SW Iran). In: Leturmy, P., Robin, C., (Eds.),
852
Tectonic and Stratigraphic Evolution of Zagros and Makran during the Mesozoic–Cenozoic.
853
Geol. Soc. London Spec. Publ. 330, p. 211–251.
M AN U
855
Poblet, J., Lisle, R.J., 2011. Kinematic evolution and structural styles of fold-and-thrust belts. Geol. Soc. London Spec. Publ. 349, 1–24.
TE D
854
SC
849
Pollastro, R.M., 1990. The illite-smectite geothermometer. Concepts, methodology, and
857
applications to basin history and hydrocarbon generation. In: Nuccio, V.F., Barker, C.E., (Eds.),
858
Applications of thermal maturity studies to energy exploration: SEPM Rocky Mountain section,
859
p. 1-18.
EP
856
Pollastro, R.M., Barker, C.E., 1986. Application of clay-mineral, vitrinite reflectance, and fluid
861
inclusion studies to the thermal and burial history of the Pindale anticline, Green River Basin,
862
Wyoming. In: Gautier, D.L., (Ed.), Roles of organic matter in sediment diagenesis: SEPM
863
Special Publication, v. 38, p. 73-83.
AC C
860
864
Potel, S., Maison, T., Maillet, M., Sarr, A.C., Doublier, M.P., Trullenque, G., Ferreiro Mählmann
865
R., 2016. Reliability of very low-grade metamorphic methods to decipher basin evolution: Case
866
study from the Markstein basin (Southern Vosges, NE France. App. Clay Sci. 134, 175-185.
34
ACCEPTED MANUSCRIPT 867
Ring, U., Brandon, M.T., Willett, S.D., Lister, G.S., 1999. Exhumation processes. In: Ring, U.,
868
Brandon, M.T., Lister, G.S., Willett, S.D., (Eds.), Exhumation Processes: Normal Faulting,
869
Ductile Flow and Erosion: Geol. Soc. London Spec. Publ. v. 154, p. 1–27. Ruh, J.B., Hirt, A.M., Burg, J.-P., Mohammadi, A., 2014. Forward propagation of the Zagros
871
Simply Folded Belt constrained from magnetostratigraphy of growth strata. Tectonics 33, 1534–
872
1551.
874
Sclater, J.G., Christie, P.A.F., 1980. Continental stretching: An explanation of post–Mid Cretaceous subsidence on the central North Sea Basin. J. Geophys. Res. 85, 3711–3739.
SC
873
RI PT
870
Schito A., Corrado, S., Aldega, L., Grigo, D., 2016. Overcoming pitfalls of vitrinite reflectance
876
measurements in the assessment of thermal maturity: The case history of the lower Congo basin.
877
Mar. Petrol. Geol. 74, 59-70.
M AN U
875
Schito, A., Corrado, S., Trolese, M., Aldega, L., Caricchi, C., Cirilli, S., Grigo, D., Guedes, A.,
879
Romano, C., Spina, A., Valentim, B., 2017. Assessment of thermal evolution of Paleozoic
880
successions of the Holy Cross Mountains (Poland). Mar. Petrol. Geol. 80, 112-132.
883 884 885 886 887 888
Petrol. Geol. 21, 829–843.
Sepehr, M., Cosgrove, J.W., 2005. Role of the Kazerun Fault Zone in the formation and
EP
882
Sepehr, M., Cosgrove, J.W., 2004. Structural framework of the Zagros Fold–Thrust Belt, Iran. Mar.
deformation of the Zagros Fold‐Thrust Belt, Iran. Tectonics 24 (5), TC5005. Setudehnia, A., 1972. Iran du Sud-Ouest: Lexiqu Strat. Internat., Centre Nat.Rech. Scientifique,
AC C
881
TE D
878
Paris, III, Asie, Fasc.9b, p. 289-376. Setudehnia, A. 1978. The Mesozoic sequence in south-west Iran and adjacent areas. J. Petrol. Geol. 1, 3–42.
889
Sfidari, E., Zamanzadeh, S.M., Dashti, A., Opera, A., Tavakkol, M.H., 2016. Comprehensive source 890
rock evaluation of the Kazhdumi Formation, in the Iranian Zagros Foldbelt and adjacent 891
offshore. Mar. Petrol. Geol. 71, 26-40. 892
Sherkati, S., Letouzey, J., 2004. Variation of structural style and basin evolution in the central 35
ACCEPTED MANUSCRIPT
896 897 898 899 900 901 902 903
insights from seismic data, field observation, and sandbox modeling. Tectonics 25, TC4007. Środoń, J., 1999. Nature of mixed-layer clays and mechanisms of their formation and alteration. Annu. Rev. Earth Planet. Sci. 27, 19-53.
RI PT
895
Sherkati, S., Letouzey, J., Frizon de Lamotte, D., 2006. Central Zagros fold-thrust belt (Iran): New
Stach, E., Mackowsky, M.Th., Teichmüller, M., Taylor, G.H., Chandra, D., Teichmüller, R., 1982. Stach’s Textbook of Coal Petrology. Berlin-Stuttgart, Gerbrüder Borntraeger, 535 p.
Szabo, F., Kheradpir, A., 1978. Permian and Triassic stratigraphy, Zagros basin, southwest Iran. J.
SC
894
Zagros (Izeh zone and Dezful embayment), Iran. Mar. Petrol. Geol 21, 535–554.
Petrol. Geol. 1, 57–82.
Sweeney, J.J., Burnham, A.K., 1990. Evaluation of a simple model of vitrinite reflectance based on
M AN U
893
chemical kinetics. Am. Assoc. Petroleum Geol. Bull. 74, 1559–1570. Vaziri-Moghaddam, H., Safari, I.A., Taheri, A., 2005. Microfacies, paleoenvironments and
905
sequence stratigraphy of the Tarbur formation in kherameh area, SW Iran. Carbonates and
906
Evaporites 20, 131-137.
907 908
TE D
904
Warr, L.N., Rice A.H.N., 1994. Inter-laboratory standardization and calibration of clay mineral crystallinity and crystallite size data. J. Metamorph. Geol. 12, 141–152. Warr, L.N., Cox, S.C., 2016. Correlating illite (Kübler) and chlorite (Árkai) “crystallinity” indices
910
with metamorphic mineral zones of the South Island, New Zealand. Appl. Clay Sci. 134, 164-
911
174.
913
AC C
912
EP
909
914
Figure and table captions
915
Fig.1 - A) Structural setting of the Zagros fold-and-thrust belt showing the major fault zones, the
916
geological provinces and the fieldwork area (modified and redrawn after Pirouz et al., 2011). HZF:
917
High Zagros fault, MFF: Mountain front fault, ZF: Zagros front. B) Geological maps of the Fars
918
province with sampling location (modified and redrawn after Mouthereau et al., 2007). 36
ACCEPTED MANUSCRIPT 919 920
Fig.2 - Stratigraphic correlation chart of the Fars province showing lateral lithology and facies
921
changes (modified after Sepehr and Cosgrove, 2004 for the Mesozoic-Cenozoic part).
922
Fig. 3 – Lithotratigraphic correlation charts of borehole data across the Fars province. (A) boreholes
924
location, (B-C) thickness variations along NE-SW and N-S trending directions, (D) thickness
925
changes in the Coastal Fars along WNW-ESE trending direction. For the Sim-1 well, data are
926
available from depths deeper than 2,000m. Values rely to measure depths and the ground level is
927
not considered (courtesy of NIOC).
M AN U
928
SC
RI PT
923
Fig. 4 – Distribution of organic and inorganic thermal indicators across the Zagros fold-and-thrust
930
belt. (A) illite content in mixed layers illite-smectite in the Fars province; (B) illite content in mixed
931
layers illite-smectite for Albian to Eocene rocks; (C) vitrinite reflectance for Miocene rocks. Data
932
for the Kuh-e-Sim anticline are from Aldega et al., 2014.
TE D
929
933
Fig. 5 - Representative ethylene-glycol-solvated (grey line) and air-dried (black line) diffraction
935
patterns of the <2 µm grain-size fraction: A) Razak Fm.; B) Gurpi Fm.; C) Kazdhumi Fm.
936
Acronyms: Chl-chlorite; I-illite; I-S-mixed-layer illite-smectite; K-kaolinite; Pal-palygorskite; Qtz-
937
quartz; Cal-calcite; Gy-gypsum; Ab-albite. R0 and R1 refer to mixed layers illite-smectite stacking
938
order.
AC C
939
EP
934
940
Fig. 6 - One-dimensional burial and thermal models of the outcropping sedimentary succession of
941
the Arsenjan area in the High Zagros. A-B) Burial evolution in the last 200 Ma and 20 Ma
942
estimating the thickness of the ophiolite thrust sheet and the amount of Maastrichtian erosion C)
943
Present-day thermal maturity data plotted against maturity curves calculated by a series of
944
geothermal gradients. 37
ACCEPTED MANUSCRIPT 945
Fig.7 - One-dimensional burial and thermal models of the outcropping sedimentary succession of
947
the Kuh-e-Surmeh anticline and adjacent areas: A) in the last 450 Ma; B) in the last 20 Ma; C)
948
Present-day thermal maturity data plotted against maturity curves calculated by a series of
949
geothermal gradients.
RI PT
946
950
Fig. 8 - One-dimensional burial and thermal models of the outcropping sedimentary succession of
952
the Kuh-e-Asalujeh anticline and adjacent areas: A) in the last 180 Ma; B) in the last 20 Ma; C)
953
Present-day thermal maturity data plotted against maturity curves calculated by a series of
954
geothermal gradients.
M AN U
SC
951
955
Fig. 9 – Cross section along the Zagros fold-and-thrust belt (refer to Bigi et al., 2018 for a full
957
discussion on geological constraints). The currently eroded portion of sedimentary successions
958
(dashed line when inferred) and maximum burials (bars) were constrained by inorganic and organic
959
thermal indicators from this study.
960
TE D
956
Tab. 1- X-ray quantitative analysis of the <2µm grain-size fraction. Sm=smectite, Pal=palygorskite;
962
I= illite; I-S= mixed layer illite-smectite; C-S = mixed layer chlorite-smectite; K= kaolinite; Chl=
963
chlorite; Qtz= quartz; Cal= calcite; Dol= dolomite; Ab= albite; Ank= ankerite, Hem= hematite;
964
Kfs= k-feldspar; Gy= gypsum; Go=goethite; Hem=hematite; R= stacking order (Jagodzinski,
965
1949); %I in I-S= illite content in mixed layer illite-smectite; %C in C-S= chlorite content in mixed
966
layer chlorite-smectite; KI =Kübler index data for air dried (AD) and ethylene-glycol-solvated (EG)
967
mounts; N.D.=not determined. Data from the Kuh-e-Sim anticline and partly from the Sarvestan
968
area are from Aldega et al. (2014) and Carminati et al. (2016) respectively.
AC C
EP
961
969
38
ACCEPTED MANUSCRIPT 970
Tab. 2. Organic matter maturity and petrographic analysis. Ro% values with an asterisk indicate
971
mean reflectance of reworked fragments or too scarce data, both not suitable for 1D thermal
972
modeling; Roeq% values in italics. – absent; s.d. = standard deviation; nr = number of
973
measurements.
RI PT
974
Tab. 3. Synthesis of maximum overburden atop the Bangestan Group and thickness of the eroded
976
units derived from 1D thermal modelling and calculated by a series of geothermal gradients
977
available from the literature.
AC C
EP
TE D
M AN U
SC
975
39
ACCEPTED MANUSCRIPT
R 1 1 1 0 0 0 0 0 0 1 1 1 0 0 0 0 0 0 0 1 0 1 0 0 0 1 3 0 1 0 0 0 1 1
RI PT
Arsenjan Arsenjan Arsenjan Sarvestan Sarvestan Sarvestan Sarvestan Sarvestan Sarvestan Sarvestan Sarvestan Sarvestan Sarvestan Kuh-e-Sim Kuh-e-Sim Kuh-e-Sim Kuh-e-Sim Kuh-e-Sim Kuh-e-Sim Kuh-e-Sim Kuh-e-Sim Kuh-e-Meymand Kuh-e-Meymand Kuh-e-Meymand Kuh-e-Meymand Kuh-e-Surmeh Kuh-e-Surmeh Kuh-e-Surmeh Kuh-e-Surmeh Kuh-e-Surmeh Kuh-e-Asalujeh Kuh-e-Asalujeh Kuh-e-Asalujeh Kuh-e-Asalujeh Kuh-e-Asalujeh Kuh-e-Asalujeh Kuh-e-Asalujeh Kuh-e-Asalujeh
SC
Kazdhumi Kazdhumi Gadvan Agha Jari Mishan Razak Razak Razak Sachun Pabdeh Gurpi Gurpi Gurpi Mishan Mishan Mishan Mishan Mol Guri Champeh Champeh Mishan Mol Champeh Pabdeh Mishan Pabdeh Gurpi Gadvan Gahkum Mishan Mishan Gachsaran Pabdeh Pabdeh Gurpi Kazdhumi Kazdhumi
M AN U
KAZ2 KAZ3 GAD1 AJ6 MIS9 RAZ2 RAZ1 RAZ3 SAC1 PAB2 GUR4 GUR3 GUR5 MIS4 MIS5 MIS6 MIS7 MOL1 GURI1 CH1 CH2 MIS10 MOL2 CH3 PAB3 MIS15 PAB6 GUR7 GAD2 GAH1 MIS14 MIS12 GAS8 PAB5 PAB4 GUR6 KAZ4 KAZ6
X-ray quantitative analysis of the <2µm grain-size fraction (%wt.) Sm Pal I I-S C-S K Chl Other 13 8 72 7 Qtz, Cal 16 7 72 5 Qtz, Cal 48 14 38 Cal, Gy 5 51 9 35 Qtz, Cal 29 27 14 30 Cal, Qtz, Ab 5 45 15 7 28 Cal, Qtz, Ab 44 21 11 7 17 Cal, Qtz, Ab 5 42 6 12 35 Qtz, Cal 91 6 1 2 Dol, Ab, Gy 34 53 6 7 Cal, Qtz 27 37 36 Cal, Qtz 20 24 43 7 6 Qtz, Cal 49 11 40 Qtz, Cal 45 18 37 Qtz, Cal, Dol, Ab 17 40 16 27 Qtz, Cal, Dol, Ab 40 24 3 33 Qtz, Cal, Dol, Ank Qtz, Cal, Ank, Ab 17 37 4 42 53 21 26 Qtz, Cal, Ank, Hem 47 15 38 Qtz, Cal, Dol, Hem 13 40 5 18 24 Qtz, Cal 51 3 12 34 Qtz, Cal, Ank 35 11 26 28 Qtz, Cal 63 19 18 Qtz, Ank 16 30 34 6 14 Qtz, Cal 50 26 7 17 Qtz, Cal 43 22 5 9 21 Cal, Qtz, Ab, Gy 46 24 19 5 6 Qtz, Cal 22 42 36 Cal, Qtz, Gy 20 10 64 6 Cal 32 7 61 Qtz, Kfs, Go 50 25 11 14 Cal, Qtz, Ab 54 19 5 22 Cal, Qtz, Ab, Dol 61 11 18 10 Cal, Qtz, Ab, Dol 85 5 10 Qtz, Cal 80 6 14 Qtz, Cal 38 62 Cal, Qtz, Gy 10 14 69 7 Cal, Qtz, Ank 33 12 50 5 Cal
TE D
Area
EP
Formation
AC C
Sample
%I in I-S 75 75 75 30 30 30 40 45 50 60 60 70 38 21 40 46 52 43 32 60 50 70 20 50 55 80 85 20 60 35 40 40 70 75
%C in C-S 60 55 70 52 55 70 55 -
KI (°∆2θ) AD EG 1.01 0.98 1.01 0.97 0.97 0.92 N.D. N.D. 1.18 1.15 N.D. N.D N.D N.D. N.D. N.D. N.D. N.D. 1.08 1.06 1.06 1.00 1.09 1.04 0.99 0.97 N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. 1.04 1.00 1.30 1.25 1.11 1.06 0.98 0.95 0.92 0.88 1.05 1.02 1.08 1.03 0.86 0.83 0.73 0.71 0.60 0.54 0.56 0.52 1.06 1.00 1.12 1.10 1.10 1.07 1.12 1.10 1.01 0.99 0.99 0.95
ACCEPTED MANUSCRIPT
Latitude
Longitude
Formation
Microscopic description of organic content
KAZ2 KAZ3 AJ5 RAZ3 GUR3
29.74359 29.84496 29.20972 29.20478 29.16244
53.17986 53.29696 53.34074 53.32447 53.31382
Kazdhumi Kazdhumi Agha Jari Razak Gurpi
GUR5 GUR4
29.44980 29.17279
53.28020 52.95759
Gurpi Gurpi
MIS9 MIS4
29.98442 28.85008
52.93960 52.86995
Mishan Mishan
MIS6
28.73050
52.94020
Mishan
MIS7 MOL1 CH1 CH2
28.62716 28.70915 28.79090 28.71220
53.12663 53.03830 52.85669 53.03550
Mishan Mol Champeh Champeh
MIS10 PAB3 PAB6 PAB5
29.67107 28.69535 28.52659 27.94757
52.75314 52.74102 52.68062 52.15743
Mishan Pabdeh Pabdeh Pabdeh
GUR6 KAZ6 KAZ4 KAZ5
27.95062 27.92971 27.61396 27.61329
52.16220 52.35970 52.53680 52.53554
Gurpi Kazdhumi Kazdhumi Kazdhumi
Scattered impregnated areas Scattered impregnated areas Barren Barren Bitumen and scarse solid fragments of reworked vitrinite and inertinite macerals Scattered fragments of vitrinite-huminite macerals Small and squared huminite-vitrinite fragments, and inertinite macerals with reflectance >0.9% Rare migrated bitumen Scarce dispersed organic matter of continental origin with both inertinite (Ro% >1%) and huminite-vitrinite group fragments (collinite) Scarce organic matter of continental origin with inertinite (Ro% >2%) and huminite-vitrinite group fragments Rare fragments of fusinite and collinite Few fragments of small sized collinite Barren Rare inertinite group macerals in small fragments, one fragment of collinite Collinite Barren Barren Scarce migrated material with abundant framboids of pyrite Scarce fragments of inertinite macerals R>0.7% Barren Barren Stains of migrated bitumen, inhomogeneous and small-sized
SC
M AN U
TE D EP AC C
Ro% and Roeq% +s.d. 0.66±0.05
nr
0.38±0.14 0.49±0.11
3 9
0.55±0.10 0.34±0.07
3 5
0.50±0.03
5
0.32±0.06 0.39±0.06 0.51*
4 3 1
0.35* -
1 -
-
-
-
-
RI PT
Sample
12
ACCEPTED MANUSCRIPT
High Zagros
Central Fars
Coastal Fars
Geothermal gradient (°C/km)
Thickness of maximum overburden atop the Bangestan Group (m)
15 20 24 15 20 24 15 20 24
~ 5250 ~ 3650 ~ 3000 ~ 4700 ~ 3200 ~ 2800 ~ 5500 ~ 3200 ~ 3000
Timing of maximum burial and subsequent erosion
Late Miocene
early Pleistocene
early Pleistocene
A
Thickness of eroded units (m) ~ 5250A ~ 3650B ~ 3000C ~ 2200A ~ 700B ~ 0C ~ 1500A ~ 200B ~ 0C
RI PT
Area
the model is not geologically acceptable as implies 1500 to 3000 m thick eroded siliciclastic deposits best solution, constrained by paleothermal data and consistent with surface and subsurface thickness of the stratigraphic section C the solution is stratigraphically acceptable but overestimates levels of thermal maturity for Cretaceous and older rocks
AC C
EP
TE D
M AN U
SC
B
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT HIGHLIGHTS
EP
TE D
M AN U
SC
RI PT
Thermal evolution of the Zagros fold-and-thrust belt Decrease of lithostatic load towards the foreland A quantitative approach to determine the amount and extent of eroded structures
AC C
• • •