You are on page 1of 54

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/273317663

Stratigraphic modelling of platform architecture


and carbonate production: a Messinian case
study (Sorbas Basin, SE Spain)
Article in Basin Research March 2015
Impact Factor: 2.73 DOI: 10.1111/bre.12125

READS

260

5 authors, including:
Emmanuelle Vennin

Raphal Bourillot

University of Burgundy

Institut Polytechnique de Bordeaux

105 PUBLICATIONS 1,039 CITATIONS

19 PUBLICATIONS 90 CITATIONS

SEE PROFILE

SEE PROFILE

Didier Granjeon

Guy Desaubliaux

IFP Energies nouvelles

GDF SUEZ

52 PUBLICATIONS 467 CITATIONS

43 PUBLICATIONS 264 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

SEE PROFILE

Available from: Raphal Bourillot


Retrieved on: 14 June 2016

Accepted Article

Received Date : 03-Jun-2013


Revised Date : 10-Feb-2015
Accepted Date : 17-Feb-2015
Article type

: Original Article

Corresponding author mail-id : christophe.kolodka@u-bourgogne.fr

Stratigraphic modelling of platform architecture and carbonate


production: a Messinian case study (Sorbas Basin, SE Spain).
Christophe Kolodka1,2, Emmanuelle Vennin1, Raphael Bourillot3, Didier Granjeon4 & Guy
Desaubliaux2
1

Laboratoire BIOGEOSCIENCES, UMR CNRS/uB 6282, Universit de Bourgogne, Dijon,

France
2

GDF SUEZ E&P International, Paris La Dfense, France

ENSEGID, EA 4592 Goressources & Environnement Institut Polytechnique de Bordeaux,

Pessac, France
4

IFP Energies Nouvelles, Rueil-Malmaison, France

Short title: 3D modelling of Messinian carbonate platforms

Abstract
The late Messinian mixed carbonate-siliciclastic platforms of the Sorbas Basin, known as the Terminal
Carbonate Complex, record significant changes in carbonate production and geometry. Their facies
and stratigraphic architecture result from complex interactions between base-level fluctuations,
evaporite deformation/dissolution and detrital inputs. A 3D quantitative approach (with DIONISOS
software) is used to explore the basin-scale platform architecture and to quantify the carbonate
production of the Terminal Carbonate Complex. The modelling strategy consists in integrating detailed
2D field-based transects and modern carbonate system parameters (e.g. carbonate production rates,

This article has been accepted for publication and undergone full peer review but has not been
through the copyediting, typesetting, pagination and proofreading process, which may lead to
differences between this version and the Version of Record. Please cite this article as doi:
10.1111/bre.12125
This article is protected by copyright. All rights reserved.

Accepted Article

bathymetric and hydrodynamic ranges of production). This approach limits user impact and so
provides more objective output results. Tests are carried out on carbonate production rates, subsidence
and evaporite deformation/dissolution. Numerical modelling provides accurate predictions of
geometries, facies distributions and depositional sequence thicknesses, validated by field data.
Comparative statistical testing of the field transects and of the various model outputs are used to
discern the relative contribution of the parameters tested to the evolution of basin filling. The 3D
visualisation and quantification of the main carbonate producers (ooids and microbialites) are
discussed in terms of changes in base-level and detrital supply. This study demonstrates that base-level
fluctuations have the greatest impact on the carbonate budget. Evaporite deformation/dissolution
affects the type and amount of carbonate production, inducing a transition from an ooid- to
microbialite-dominated system and also has a major effect on stratigraphic architecture by inducing
the migration of depocentres. The numerical modelling results obtained using modern carbonate
system parameters could also be applied to subsurface ooid-microbialite reservoirs, and the Terminal
Carbonate Complex is a good analogue for such systems.

1. Introduction
Over the last three decades, stratigraphic forward modelling (SFM) has been used mainly as a tool for
predicting geometries and facies in petroleum geology (Warrlich et al., 2008), or as an experimental
approach for testing theoretical concepts of physical systems (Charvin et al., 2011; Burgess, 2012).
Most of those studies attempted to test the response of sedimentary architecture with respect to
empirical variables such as sediment transport or cyclic eustatic fluctuations (e.g. Burgess et al., 2008;
Smme et al., 2009; Williams et al., 2011). In carbonate sedimentary systems, SFM is a useful tool for
characterising and quantifying processes controlling platform development (Bosence & Waltham,
1990; Warrlich et al., 2002), thus enabling petroleum geoscientists to assess and predict carbonate
reservoir heterogeneities (Warrlich et al., 2008). However, the simulation outcrop data often use
interpreted field data (Burgess, 2012) and might therefore lead to circular reasoning (Warrlich et al.,
2008).

Numerous SFM programs allow 3D models to be simulated. Some programs focus on carbonate
systems (Burgess & Wright, 2003; Paterson et al., 2006; Williams et al., 2011; Barrett & Webster,
2012; Seard et al., 2013) and others on mixed carbonate-siliciclastic systems (Kendall et al., 1991;
Quiquerez et al., 2000; Warrlich et al., 2002; Gratacs et al., 2009; Saura et al., 2012). Among such
software packages, DIONISOS has proved particularly robust (in various sedimentary systems) in
deltaic systems (Burgess et al., 2008), in carbonate systems (Williams et al., 2011; Seard et al., 2013),
in complex geodynamic settings (e.g. growth-faulted margin, Alzaga-Ruiz et al., 2009; rifting/basin
inversion, Csato et al., 2012) and in foreland basins (Saura et al., 2012).

This article is protected by copyright. All rights reserved.

Accepted Article

In this study, the DIONISOS program (Granjeon, 1997; Granjeon & Joseph, 1999) is used to
investigate the contributions of the main processes controlling the stratigraphic architecture and
carbonate production of the late Messinian carbonate platforms in the Sorbas Basin. This peripheral
Mediterranean basin includes an evaporitic unit deposited during the Messinian Salinity Crisis
(Rouchy & Caruso, 2006). A short marine transgression led to the development of a mixed carbonatedetrital system, the Terminal Carbonate Complex platforms (TCC; Esteban, 1979) dominated by
ooidic, lime mud and microbialitic facies atop those evaporites. Bourillot et al. (2010a; b) postulated a
syn-sedimentary evaporite deformation/dissolution episode during TCC deposition, associated with a
turnover in the carbonate factory (i.e. change in the dominant carbonate producer), probably controlled
by evaporite brine seepages (Bourillot et al., 2010a; b).
This sedimentary system is an excellent example for studying the basin-scale relationship between the
developments of carbonate platforms and the deformation/dissolution of their evaporitic substratum.
The detailed dataset proposed by Bourillot et al. (2010a; b) is used to build a 3D model of the basin.
This study is therefore one of the first attempts at 3D process-based modelling of experimental testing
of the response of carbonate systems (dominated by ooid and microbialite production) with different
aspects of evaporite deformation.

DIONISOS is used to test the relative influence of various parameters set independently of the field
database: sequence duration, subsidence and carbonate production rates. Our modelling strategy (Fig.
1) is based on combining field data (2D cross-sections, palaeogeographic maps) from the studies of
Bourillot et al. (2010a; b) with modern sedimentary system data (carbonate production rates,
bathymetric and hydrodynamic ranges of production, etc.) so as to minimize the influence of user and
field databases. In this study, the term base-level is preferred to relative sea level because the
Sorbas Basin was disconnected from the global ocean for most of the time during the deposition of the
TCC and recorded a transition from open-marine to lacustrine conditions.

The understanding and quantification of carbonate production and stratigraphic architecture modified
by base-level fluctuations, subsidence and evaporite deformation/dissolution, applied to the TCC ooidmicrobialite system, can thus provide crucial information for the exploration of such reservoirs. It can
be applied to other Mediterranean Messinian platforms (i.e. Mallorca) and carbonate reservoir
analogues.

The present study proposes a 3D model to: (i) establish a series of modelling experiments of carbonate
production based on modern sedimentary system parameters; (ii) test the best-fit model and discuss the
relative influence of evaporite deformation and sediment supply on carbonate platform architecture
and facies distribution; (iii) decipher the contribution of various parameters to carbonate production
and basin filling; and (iv) define sequence durations.

This article is protected by copyright. All rights reserved.

Accepted Article

2. Geological setting
The Sorbas Basin in south-eastern Spain (Fig. 2a) belongs to the Internal Betic basins, which opened
during Serravallian times in conjunction with the exhumation of the Betic Metamorphic Core
Complexes (Fig. 2b; Augier et al., 2005; Jolivet et al., 2009). During the Messinian, the Sorbas Basin
recorded a ~ 130 m (Ott dEstevou & Montenat, 1990) thick evaporitic event consisting successively
of interbedded gypsum (Yesares Member) and non-evaporitic layers (e.g. marls, limestones, etc.)
resulting from a major sea-level fall induced by the restriction of the Mediterranean realm during the
Messinian Salinity Crisis (MSC; Rouchy & Saint-Martin, 1992; Riding et al., 1998). The Yesares
Member conformably overlies the marl deposits of the Abad Member in the basin centre and exhibits
onlap geometries marginwards over the fringing reef unit of the Cantera Member (Fig. 3; Bourillot et
al., 2010a). A transgressive event post-dating evaporite deposition in SE Spain allowed the
development of mixed carbonate-detrital platforms named the Terminal Carbonate Complex (TCC)
(Esteban, 1979; Riding et al., 1998; Bourillot et al., 2010a; b). This unit conformably overlies the
gypsum deposits in the basin centre and onlaps the fringing reef unit toward the margins (Bourillot et
al., 2010a; b). On the northern margin of the basin, the TCC platform records significant detrital inputs
from an alluvial fan system located at the foot of the Sierra de Los Filabres (Fig. 2b). Stratigraphically,
the TCC corresponds to a fourth order transgressive-regressive sequence (Franseen et al., 1998) that
can be divided into four smaller-scale (fifth order) depositional sequences (DS1 to DS4; Bourillot et
al., 2010a; b). These depositional sequences are separated by five sequence boundaries (SB1 to SB5)
marking base-level falls (Bourillot et al., 2010b). For each sequence, the maximum transgression is
recorded in a maximum flooding interval (MFS1 to MFS4). DS1 to DS3 were deposited during the
transgressive part of the fourth-order sequence. They consist mainly of tide-dominated ooidicoobioclastic systems and outer-platform lime muds developed in a shallow semi-open sea with normal
salinity, as attested by the presence of oligotypic stenohaline faunas (Rouchy et al., 1986; Braga et al.,
1995; Riding et al., 1998; Bourillot et al., 2010b). The upper part of DS3 records a transition from
ooidic- to microbialite-dominated systems, related to a renewed restriction of the Sorbas Basin
(Bourillot et al., 2010b). DS4 corresponds to microbialite-dominated platforms developed in a salt
lake with fluctuating brackish to hypersaline water (Bourillot et al., 2010b), followed by continental
alluvial to brackish lake Lago Mare type deposits (Zorreras Member; Fortuin & Krijgsman, 2003;
Bassetti et al., 2006). Crustal movements (subsidence and uplift) appear to have been slight for most
of the TCC deposition (Ott dEstevou, 1980) but there is evidence of syn-sedimentary
deformation/dissolution of the underlying evaporites during DS3 and DS4 (Bourillot et al., 2010a; b).
This is materialised by folds, listric faults, turtle antiforms and collapse structures affecting the TCC
strata (Bourillot et al., 2010a). The deformation seems to have been triggered by competition between:
(1) regional mechanisms resulting from the uplift of the Betic cordillera and leading to the
development of subsequent alluvial fan systems. The increase in alluvial deposits from DS1 to DS4

This article is protected by copyright. All rights reserved.

Accepted Article

(Bourillot et al., 2010a; b) could be related to more efficient freshwater discharge in the Sorbas Basin
causing the partial dissolution of evaporitic bodies; (2) local mechanisms due to heterogeneities in
sedimentary overload and antecedent topographic slopes (palaeo-reefal slopes) that together could
have

induced

spreading

and/or

gliding

processes

(Bourillot

et

al.,

2010a;

b).

The

deformation/dissolution is coetaneous with the turnover in carbonate production (from marine to


continental conditions). The subsidence related to evaporite withdrawal and the associated seepage of
saline brines seems to have played an important role in the microbialitic bloom/preservation observed
in the Sorbas Basin from the end of DS3 to DS4 (Bourillot et al., 2010a).

3. Stratigraphic Forward Modelling


DIONISOS is a 3D process-based stratigraphic forward model, which simulates the evolution of
sedimentary basins (10100s km) over millions of years (Granjeon, 1997; 2009; Granjeon & Joseph,
1999). Three main processes are taken into account: (1) accommodation variations; (2) sedimentary
supply; (3) transport. Accommodation variations are controlled by base-level changes and tectonics
(subsidence, uplift, compaction, flexure and salt diapirism). Sedimentary supply corresponds to
terrigenous sediment inflow and in situ carbonate production. Different carbonate factories can be
simulated using empirical formulae linking carbonate production to bathymetry and sediment supply,
such as in Lawrence et al. (1990), Li et al. (1993) or Granjeon (1997). Detrital supply is introduced
from a sediment source placed at the edge of the simulated area and may be either fluctuating or fixed
through time. Transport is simulated using a diffusion equation.

4. Model parameterisation
The two-parameterisation steps for the modelling strategy consisted in (a) calibrating the initial
conditions and (b) inputting the parameters controlling basin filling in DIONISOS. The initial
conditions remained invariable and formed the backbone of the model (time, space, facies equivalent
to DIONISOSs sediment classes). The dataset used here corresponded to Bourillot et al.s field
database supplemented by data from the literature (e.g. palaeogeographic maps of Ott dEstevou &
Montenat, 1990). The second step consisted in calibrating (1) basin-specific parameters based on field
data to the extent that they are local basin-specific. The field database consisted of two detailed crosssections described by Bourillot et al. (2010a; b; Fig. 4; basin margin: Cariatiz area, ~ 1 km long; basin
centre: Sorbas area, ~ 2 km long) and (2) independent parameters from bibliographic data:
palaeogeographic maps, time-stratigraphic framework and carbonate production rates and associated
parameters (bathymetric range, relative wave-energy window, salinity tolerance and response to
detrital influx). Sensitivity analyses were performed to test the most realistic combination of carbonate
production rates, subsidence and evaporite deformation/dissolution, which were kept to build a best-fit
model constituting the basis to discuss the factors controlling sedimentation.

This article is protected by copyright. All rights reserved.

Accepted Article

4.1 Initial conditions


Ages were constrained from a chronological estimation by Bourillot et al. (2010b; Fig. 5), who
proposed an age ranging between ~ 5.6 and ~ 5.45 Ma for the TCC, based on three correlations with
Mediterranean and global events (Rouchy & Caruso, 2006). The temporal framework was refined by
testing the hypothesis of Krijgsman et al. (2001) of a precessional (~ 20 kyr) or obliquity (~ 40 kyr)
control on TCC depositional sequences. Considering the four-depositional-sequence scheme of
Bourillot et al. (2010b), this implied a duration of 80 or 160 kyr for TCC deposition. Two experiments
were performed, both with TCC initiation at 5.6 Ma, and ending respectively at 5.52 Ma and 5.44 Ma.
The modelling time-step was set at 5 kyr for all the simulations, to ensure good resolution (4 to 8
steps) within each DS.
The area simulated initially was a 14.8 km 12.4 km rectangle with grid spacing of 0.1 km. The
model substratum was composed of two layers: (1) the top of the Fringing Reef Unit (Cantera
Member) forming a rim at the basin periphery and (2) the gypsum unit (Yesares Member; Fig. 6)
filling the central part of the basin and partially covering the reefal slope. The upper boundary of the
Cantera Member was an eroded karstic surface known as the Top Reef Unconformity (TRU). Distally,
it passed into a conformable surface between the Upper Abad Member marl-sapropel-diatomite
deposits and the Yesares Member (Fig. 3). This initial surface was digitized using palaeogeographic
maps and the basin cross-section of Ott dEstevou & Montenat (1990). The Yesares Member was
simulated as a single unit surrounded by two surfaces (Fig. 6) from a compilation of the
palaeogeographic maps proposed by Bourillot et al. (2010b) and back stripped cross-sections of
Montenat et al. (1987) and Ott dEstevou & Montenat (1990). We matched ante-deformation
thicknesses of evaporites (e.g. ~130 m at Los Molinos de Aguas; Ott dEstevou & Montenat, 1990;
3050 m at Cariatiz, Bourillot et al., 2010a).

In this study, three sediment classes were defined characterising the three main carbonate compounds
of the TCC: lime muds, ooids and microbialites. Output results were displayed with facies in order to
compare field geometries and facies distribution with the simulated architecture. Thus, seven facies
(F1 to F7), corresponding to the five main facies associations proposed by Bourillot et al. (2010b)
were defined from the outer to inner platform and summarized in figure 7.
Outer platform muds (F1) were dominantly lime muds deposited in relatively low-energy and subtidal
domains (540 m deep) within brackish to normal marine waters. Ooidic and oobioclastic grain- to
rudstones (F2) accumulated within the high-energy zone, corresponding to sub-to-intertidal domains
(010 m deep) and normal marine to slightly hypersaline conditions. Lagoonal muds (F3) were
shallow and low- to moderate-energy sediments deposited in back-shoal, intertidal to shallow subtidal
domains (05 m deep), in normal to hypersaline conditions. The two microbialites (F4 and F5)
differed in terms of salinity tolerance and hydrodynamic and bathymetric ranges. The first type of
microbialite (Microbial Association 1 or MA1; see Bourillot, 2009) developed in a low-to-moderate

This article is protected by copyright. All rights reserved.

Accepted Article

energy environment with highly fluctuating salinities from brackish to hypersaline conditions (010s
m deep). The second microbialite (Microbial Association 2 or MA2) developed in a high-energy
setting with normal marine to slightly hypersaline conditions. It was associated with ooidic grainstones
(F2) and was calibrated for optimal production at depths of more than 010 m. Detrital sediments
corresponded to coarse-grained sediments (F6 and F7), indicating alluvial fan environments dominated
by gravity-driven processes.

4.2 Input parameters


4.2.1 Basin-specific parameters
Basin-specific parameters include: (A) accommodation space, (B) sea-water chemistry and (C) erosion
and transport. Except for detrital supplies, these parameters are determined by data from the literature
(see references below).
(A) Accommodation space was input into the model as a curve varying with time. This curve resulted
from the water-level curve calculated in Cariatiz by Bourillot et al. (2010a; Fig. 8a) using pinning
points (Goldstein & Franseen, 1995) and variations of subsidence (Ott dEstevou, 1980). The initial
zero level used in this model corresponded to the top of the Yesares Member at SB1. At the TCC
duration-scale, the base-level elevation was about 66 metres within amplitude of several 10s metres
from SB1 until MFS3. From MFS3 to SB5, water level decreased by 23 metres. On the depositional
sequence scale, base-level fluctuations ranged from 22 to 48 m. Basin-scale subsidence was low; Ott
dEstevou (1980) calculated a regional subsidence rate of between 0.1 and 0.2 mm y-1 during the
Messinian in the Sorbas Basin. Local subsidence was controlled mainly by evaporite
deformation/dissolution (Megias, 1985; Bourillot et al., 2010a; b), especially on the north-western
margin of the Sorbas Basin during DS3 and DS4. This local subsidence was digitized according to the
palaeogeographic sketches proposed by Bourillot et al. (2010b), showing a basinward displacement of
the evaporite mass, in the form of a subsidence map applied to the top of the evaporitic layer (top of
the Yesares Member; Figs. 6 and 9). Due to the short simulated time span, between 80 and 160 kyr,
the total subsidence deformation of the basin was less than 40 m. The compaction of TCC sediments
was considered negligible because of the shallow burial of these deposits, covered by 0160 m of
younger sediments at most (latest Messinian, Zorreras Member; Pliocene; Ott dEstevou & Montenat,
1990). Comparison with compaction curves for shallow-water carbonate sediments in southern Florida
showed that, at such depths, the porosity loss ranged between 14.5 %, representing less than 5 m of
thickness reduction (Schmoker & Halley, 1982).
(B) During TCC deposition, the salinity fluctuations were controlled both by the degree of
restriction/opening

of

AtlanticMediterranean

connections

and

by

the

syn-sedimentary

deformation/dissolution of the evaporite layers (Bourillot et al., 2010b). Bourillot et al. (2010b)
attempted to define a salinity fluctuation range from the evolution of floro-faunal associations and the
mineralogical composition of the sediment. They estimated that salinity was close to the marine

This article is protected by copyright. All rights reserved.

Accepted Article

conditions during DS1 and DS2 (3040). Salinity increased and fluctuated between ~ 30 and
120 during DS3 and between ~ 20 and 140 during DS4 (Bourillot et al., 2010b). The version of
DIONISOS used included a salinity fluctuation module. It consisted in calculating salinity for each
cell of the basin from an input salinity variation curve (Fig. 8b). It was used to display spatial
variations caused by e.g. freshwater discharge at the basin margin. Deduced from Bourillot et al.s
ranges, a salinity fluctuation curve was defined and input into the model. For DS1DS2, highfrequency salinity variations from 30 to 40 were introduced and for DS3 and DS4, they fluctuated
between 20 and 140. The salinity fluctuation curve corresponded to a salinity-tolerance window
enabling or inhibiting the production of one facies. In the model, the windows were calibrated by
using modern carbonate system values. In order to simulate the water chemistry changes due to
evaporitic dissolution, we introduced a stress function corresponding to the release of evaporitic brines
into the basin. This drove the model locally by introducing a coefficient map associated with local
evaporite deformation and mimicking the distribution of evaporitic brines. These coefficients were
high in the maximal deformation area and tended to decrease towards the undeformed zones. It thus
limited or enhanced carbonate production in each cell by multiplying the carbonate production rate by
the brine coefficient.

(C) As erosion is poorly constrained in subaqueous environment and is significant in subaerial


environments, subaerial erosion is simulated. Given the relatively short platform exposure time, its
alteration on the stratigraphy and thickness of the deposits appears not to have been significant
(Montaggioni et al., in press). The maximum erosion rate of exposed rocks as in subtropical
environments was set at 100 m Myr-1, similar to the range of values proposed by Kartson et al.
(2006).
Sediment dispersal concerns both carbonate sediment and detrital influx. It was mainly controlled here
by water-driven transport, based on the following diffusion equation (Granjeon, 1997; Granjeon &
Joseph, 1999):

The evolution of sediment flow Q (km kyr-1) along a slope S (m) varies with the water flow water
(dimensionless) and a diffusion coefficient Kwater (km2 kyr-1). K values range between 0 and 4.10 km2
kyr-1 (Granjeon & Joseph, 1999).
Carbonate production is very high favoring in situ deposition and offsetting sediment dispersal.
Consequently, carbonate particles are transported in the direction of water flow (i.e. wave-driven or
water driven), but most of them remain in the production zone, and sediment dispersal does not appear
significant considering the short simulation duration. The detrital influx source was a palaeo-river
located 9 km from the northern margin according to the field case study (Bourillot et al., 2010b; Fig.
6). Neither water discharges nor sediment supplies were constant through time and were usercontrolled. Sediment supplies were based on time-averaged sedimentation rates. Alluvial fan deposits

This article is protected by copyright. All rights reserved.

Accepted Article

correspond to event sedimentation (e.g. Blair & McPherson, 1994; flood events) and modelling input
values cannot be directly compared to analogue systems, due to the specificity of each system.
Accordingly, the calibration and quantification of detrital influx corresponded to a time-averaged
volumetric sediment supply estimated from the field study (Bourillot et al., 2010b). It indicated an
increase in detrital influx throughout TCC deposition. Values of sediment supply used in the model
fluctuated between 0 and 8 km3 Myr-1 (mean: 3.75 km3 Myr-1), which was consistent with averaged
sediment supplies of another upper Miocene (Tortonian) alluvial fan system in the Teruel Basin
(Rohais et al., 2008).

4.2.2 Independent parameters: Origin and production rates of modern carbonate facies
Carbonate production results from complex interactions between auto- and allo-cyclic processes
determining its type and amount (Schlager, 2005; Wright & Burgess, 2005). Production pathways
have a major impact on production rates, facies distribution and platform geometry (Mutti & Hallock,
2003). Different classifications have been established depending on the precipitation mode (biotic vs.
abiotic) and controlling factors of carbonate production (e.g. Wright & Burchette, 1996; James, 1997).
The shallow-water, biological production of sediments (biotically-induced/influenced and bioticallycontrolled; sensu Dupraz et al., 2009) is driven mostly by light, temperature, salinity, carbonate
saturation and nutrient concentration, whereas abiotic precipitations are largely controlled by seawater chemistry (Schlager, 2005). Schlager (2005) introduces three types of carbonate factories
independently distinguished from their precipitation pathways: (1) the tropical shallow-water factory
(T-factory) is composed mostly of light-dependent producers (e.g. corals, green and red algae) and
also includes some abiotic precipitates (marine cements, ooids and carbonate whitings); (2) the coolwater factory (C-factory) also includes light-dependent producers (red algae and symbiotic larger
foraminifers; Lees & Buller, 1972; James, 1997) but is dominated by heterotrophic producers (e.g.
bryozoans, benthic foraminifers, molluscs) and (3) the mud-mound and micrite factory (M-factory)
dominated by abiotic precipitations and micrites of microbial origin controlled by oxygen and nutrient
levels. During the Phanerozoic the M-factory developed in dysphotic or aphotic, nutrient-rich
conditions (Leinfelder et al., 1993; Neuweiler et al., 1999; Boulvain, 2001; Schlager, 2005).
Based on a controlled laboratory experiment, Bosscher & Schlager (1992) quantified the growth-rate
of coral species as a function of light intensity with depth. From these results, they established an
equation of carbonate production vs. depth based on the effect of decreasing photosynthetic activity
with light and applicable to most of the T-factory components (Schlager, 2005). In this formula, the
skeletal growth rate G varies with the maximal growth rate
saturation with depth.

(in mm yr-1) and the decrease in light

(in E m-2 s-1) corresponds to the surface light intensity;

(in E m-2 s-1) is

the saturating light intensity; k (in m-1) is the extinction coefficient and z (in m) the depth.

This article is protected by copyright. All rights reserved.

Accepted Article

On modern carbonate platforms, ooids are produced within the upper photic zone in a bathymetric
range of 012 m, in high-energy areas (Lloyd et al., 1987). Accretion and accumulation are mainly
dependent on the hydrodynamic regime (wave- or tidal-energy; e.g. Harris, 1979; Reeder & Rankey,
2008; Rankey & Reeder, 2010) and sea-water chemistry (elevated pH and total alkalinity; Rankey and
Reeder, 2010; Rankey et al., 2011). They can precipitate in normal to hypersaline conditions (Lees,
1975; Loreau, 1982). The formation of their cortices is still under discussion and may result from
abiotic and biotic processes (Davies et al., 1978; Morse & MacKenzie, 1990): some authors postulate
a chemical and physical control of ooid formation based on evidence of a close correlation between
water agitation and crystal orientation (Bathurst, 1975), while others evoke the influence of microbes
on carbonate precipitation around a nucleus (Pacton et al., 2012) or a combination of bacterial and
mechanical processes (Wilson, 1967; Folk, 1973; Morse & MacKenzie, 1990). Schlager (2005)
postulated that ooids could be included in the abiotic quantum of the T-factory. On the modern Great
Bahamas Bank, Harris (1979) estimated production rates from 0.08 to 2.74 m kyr-1 with a median of
0.93 m kyr-1 and an average of 1.12 m kyr-1 (Fig. 10).

The origins and sources of shallow-water lime muds has been a source of debate for many years
(Flgel, 1982; Morse et al., 1984; Shinn et al., 1989; Morse et al., 2007; Perry et al., 2011). Carbonate
lime muds originate from disaggregation of foraminifera and algae, bioerosion of carbonate substrate
by grazers (fish, etc.; Scoffin, 1987), biological production by fish (Perry et al., 2011) and/or direct
precipitation in the water column with or without the influence of micro-algae, also called whitings
(Yates & Robbins, 1999; Thompson, 2001; Morse et al., 2007; Perry et al., 2011). Muds are generally
deposited in low-energy settings (Morse & MacKenzie, 1990; Morse et al., 2007) and develop in a
relatively large range of salinity from normal marine to hypersaline conditions (33100 ; Mann &
Nelson, 1989). In the Great Bahamas Bank, most of the mud production occurs in shallow-water
domains (Milliman et al., 1993; < 7 m of bathymetry, Shinn et al., 1989) and about half of the
production may be exported basinward (Wilber et al., 1990). Mud deposition varies between outer and
inner platforms: (1) Open and deep subtidal muds (< 5 m of bathymetry; off-bank transport to platform
slope environments, Shinn et al., 1989) have production rates ranging from 0.26 to 0.43 m kyr-1 with a
median of 0.33 m kyr-1 and an average of 0.4 m kyr-1 around the Bahamian archipelago (Fig. 10;
Broecker & Takahashi, 1966; Robbins et al., 1997) and from 0.49 to 1.01 m kyr-1 on the Florida outer
shelf (Fig. 10; Enos, 1977); (2) muds produced and accumulated in situ within shallow restricted
lagoons (> 5 m of bathymetry) exhibit a range from 0.12 to 0.61 m kyr-1 with a median of 0.315 m kyr1

and an average of 0.33 m kyr-1 (Fig. 10) on the Bahamas (Neumann & Land, 1975), Florida

(Stockman et al., 1967; Enos, 1977) and Belize (Halley et al., 1977) shelves.

This article is protected by copyright. All rights reserved.

Accepted Article

In supratidal to shallow subtidal microbialites, carbonate precipitation is controlled both by


environmental parameters (dissolved CO2, temperature, salinity, etc.) and by bacterial metabolisms.
The combined metabolisms of autotrophic (mainly cyanobacteria) and heterotrophic (mainly sulphatereducing bacteria) lead to carbonate precipitation within microbial mats (Vasconcelos & McKenzie
1997; Reid et al., 2000; Dupraz & Visscher, 2005). Carbonate production in microbial mats is thus
both photo- and chemo-dependent: for instance, dissolved sulphate has a key role in the anaerobic
degradation of extracellular polymeric substance (EPS), which liberates calcium and increases
carbonate alkalinity in the mats, promoting carbonate precipitation (Dupraz et al., 2009). This
phenomenon is well known in open marine (Reid et al., 2000; Visscher et al., 2000) and hypersaline
lake microbialites (Dupraz et al., 2004; Vasconcelos et al., 2006). The lithifying mats are
preferentially found in shallow-water, where UV radiation might be responsible for EPS alteration by
disinhibiting carbonate precipitation (Dupraz & Visscher, 2005).

In terms of growth morphology, fabric, composition and surrounding sedimentary features, MA1
microbialites are similar to microbial mats developing in low-energy Bahamian salinas/lakes, where
salinity reaches 100150 (e.g. Storrs Lake, Big Pond; Pinckney et al., 1995; Glunk et al., 2011). In
these lakes, microbialite growth rates vary from 0.08 to 0.8 m kyr-1 (Paull et al., 1992) with a median
of 0.17 m kyr-1 and an average of 0.27 m kyr-1. MA2 microbialites are comparable to the open sea
stromatolites developing within high energy ooidic belts on the Bahamas platform (e.g. Little Darby
Island, Reid et al., 2011; Lee Stocking, Feldmann & McKenzie, 1998; Highborne Cay, Dupraz et al.,
2009) or in the hypersaline Shark Bay lagoon (Logan & Cebulski, 1970; Reid et al., 2003; Jahnert &
Collins, 2012). Bahamian columns exhibit average growth rates ranging from 0.33 m kyr-1 (Planavsky
& Ginsburg, 2009) to 1 m kyr-1 (Reid & Browne, 1991). Shark Bay microbialites present net growth
rates between 0.04 and 0.3 m kyr-1 (Jahnert & Collins, 2012). The median production rate calculated
from those values is 0.21 m kyr-1 and the average is 0.46 m kyr-1 (Fig. 10).
In our simulations, carbonate production is simulated using a water depth-dependency law (Lawrence
et al., 1990; Li et al., 1993; Granjeon, 1997; Fig. 11):

The production rate of a carbonate compound or a facies i ( ; in m kyr-1) varies as a function of its
maximal production rate (
evolves with

; in m kyr-1) and with water depth influence

the bathymetry (in m) and

(dimensionless).

a critical bathymetry (in m), which may correspond to the

fair-weather wave-base or the base of the photic zone.

This article is protected by copyright. All rights reserved.

Accepted Article

Production curves are constructed for each carbonate compound defined (Fig. 11) with median,
average, minimum and maximum production rates (Pmin, Pmed, Pave and Pmax) of modern sediments
compiled from the literature (Fig. 10). The curves were successively tested with (Pmin, Pmed, Pave and
Pmax) values in order to approach the best numerical analogue, which we call the best-fit model.

5. Best-fit model and validation


The best-fit model was built by conducting sensitivity analyses on the main two parameters controlling
basin sedimentation: carbonate production and subsidence (Figs. 1214). Each parameter was tested
independently. Six experiments were performed and are shown in Figure 13. Four experiments used
(1) average (Fig. 13a), (2) median (Fig. 13b), (3) minimal (Fig. 13c) and (4) maximal (Fig. 13d)
carbonate production rates and two were run with (5) no regional subsidence (Fig. 14a) and (6) no
evaporite deformation/dissolution (Fig.14b). The four experiments testing the influence of production
rates were run with regional subsidence and evaporite deformation/dissolution and the two
experiments testing regional and local subsidence were run with average production rates (Figs. 12
and 13a). This approach meant one variable could be fixed in the accommodation space/sediment
supply ratio (A/S) and the relative contribution of the different parameters tested discussed.
Figure 12 allows comparison between field case study and best-fit model, in terms of sedimentary
geometry dimensions, prograding distances, shoal thickness and distribution of patchy microbialites.
In Sorbas, from SB1 to MFS3, sedimentation is dominated by ooids (F2) developing as large
basinward prograding shoals under high-energy conditions (Fig. 12a). From MFS3 to SB5, some small
and flat ooidic shoals are observed in lagoonal mud-dominated deposits (F3). Following the overall
sedimentary succession, the best-fit model matches in terms of facies and stratigraphic architecture
with the field case study. However, in MFS3, thin MA1 and MA2 microbialites (F5) are lacking (Fig.
12d) and outer platform muds (F1) exhibit a wider landward extension in the best-fit model.
Microbialites are replaced, in our model, by extensive outer mud sedimentation, characterised by a
lower production rate (0.46 m kyr-1 for MA2 and 0.4 m kyr-1 for outer muds). It has been shown that
reduced carbonate production limits or prevents significant in situ deposition, and promote intense
sediment reworking and transport (Montaggionni et al. in press). Consequently, in the absence of
microbial trapping, grain dispersal in favoured (Jahnert & Collins, 2012) and outer mud facies (F1) are
transported inducing flat platform geometry. This low-relief platform recorded higher bathymetries,
restricting the production of ooids to the western part of the model. Ooid shoals then form flat
prograding clinoforms due to the low accommodation space during the end of DS3. In the field data,
clinoforms were steeper and their geometries could explain the absence of retrogradational outer
muds. This detailed mismatching between field data and model output could be related to very local
and heterogeneous changes in substratum relief, which cannot be displayed in the model because of

This article is protected by copyright. All rights reserved.

Accepted Article

their reduced spatial extension. This is also suggested by the local occurrence of slumps and softsediments (in DS3) observed in the field (Bourillot et al., 2010b) but not reproduced in the model,
while the overall stratigraphic architecture and facies distribution fit correctly.
In Cariatiz, from SB1 to MFS3, sedimentation is dominated by ooids (F2) in the top reef flat, forming
aggrading shoals. These ooidic belts (F2) separate shallow low-energy settings dominated by lagoonal
muds (F3) from the deep outer platform muds (F1) in the basin centre. From MFS3 to SB5, the general
trend evolves towards basin restriction. Evaporite deformation starts, increasing the available space.
Sedimentation is dominated by MA1 microbialites (F4) and lagoonal muds (F3) with increasing
detrital supplies (F6 and F7). As with Sorbas, the best-fit model matches in terms of facies and
stratigraphic architecture with the field case study. However, in DS1, MA1 microbialites (F4) did not
develop in the best-fit model (Fig. 12b and c) and DS1 is thicker in the field case study than in the
best-fit model (Fig. 12e).

5.1 Effect of carbonate production rates


Tests carried out on production rates are displayed in Figures 13 and 14. To confirm the similarity
between the field study and one of the different experiments (14), cumulated thicknesses of
sediments per depositional sequences were measured both in Cariatiz and Sorbas (Fig. 14a).
Thicknesses were tested in three virtual wells (localised in the marginal, transitional and basinal
settings of the Sorbas and Cariatiz cross-sections; Figs. 14 bd). The sensitivity tests were performed
within fixed accommodation space fluctuations (A), controlled by the field study, allowing a
discussion of the impact of sediment supply variations (S). The best-fit model (Fig. 13a), by
comparison with the field study stratigraphic architecture and facies distribution (Fig. 12e), was
obtained using average carbonate production rates (0.271.12 m kyr-1; Experiment 1, Fig. 13a). In
Cariatiz, this is confirmed by the remarkable similarity of the cumulated thickness curves for both the
field case study and best-fit model. In Sorbas, both field case study and best-fit model curves are
almost parallel, even though the Sorbas cumulated thickness curve is ~10 m lower than in the field
case study (Fig. 13a).
Among the four tests, the main difference between Experiments 1 (Fig. 13a) and 2 (Fig. 13b)
corresponding to average and median production rates respectively lay in the distribution of ooid
facies (F2). Even if production rates did not differ significantly (0.93 m kyr-1 for median rates and 1.12
m kyr-1 for average rates), the ooid facies (F2) were thinner in Experiment 2 (Fig. 13b). Experiments 3
and 4 using respectively minimal (0.040.12 m kyr-1; Fig. 13c) and maximal (0.613 m kyr-1; Fig.
13d) production rates differed significantly from the field study, considering stratigraphic architecture,
DSs cumulated thickness (Fig. 14) and facies distribution. When the model was run using minimal
production rates (Fig. 13c), ooidic belts did not develop and a thin muddy platform was produced. In
Experiment 4 (maximal production rates), ooid facies formed wide thick belts filling the basin rapidly,
except in Sorbas where muds are preserved distally during DS1. Due to the high ooid production rates,

This article is protected by copyright. All rights reserved.

Accepted Article

the Accommodation space/Sediment supply ratio became rapidly close to zero, limiting or stopping
carbonate production. MA1 microbialites (F4) were only present in Cariatiz, during DS4.

5.2 Impact of subsidence and evaporite deformation/dissolution


To determine the best-fit model, two other experiments were performed testing the impact of regional
subsidence and evaporite deformation/dissolution (Fig. 14). These tests were performed with fixed
sediment supply variations (S) based on average carbonate production rates depicted in Experiment 1
(Fig. 1213a) as this corresponds to the most realistic model by comparison with the field study of
Bourillot et al. (2010a; b). Among accommodation space variations (A), base-level fluctuations were
fixed and corresponded to those defined in the field database (Bourillot et al., 2010a; Fig. 7) enabling
to discuss the relative contribution of regional and local subsidence on available space changes. The
best-fit model (Fig. 13a), by comparison with field study stratigraphic architecture and facies
distribution (Fig. 12), was obtained when both evaporite deformation/dissolution and regional
subsidence were operative.
Experiment 5 (Fig. 15) showed that when the model was run without regional subsidence, the results
were similar to the field study and Experiment 1 (Fig. 13a). Geometries were successfully reproduced
for Cariatiz and depositional sequence thicknesses were almost identical. In Sorbas, geometries, facies
distribution and DS thicknesses are inconsistent in comparison with the field case study (Fig. 15a).
Indeed, ooidic shoals (F2) prograded farther than in Experiment 1 (Fig. 13a). Outer platform muds
(F1) were lacking and formed a thin and limited transgressive wedge during DS2. The absence of
subsidence also induced a rapid filling of accommodation space limiting the deposition of end-DS3
and DS4, turning SB4 into an extensive bypass surface and consequently sediments prograded more
than in the field study (Fig. 15a). Cumulated thickness measurements revealed a flat curve (Sorbas;
Fig. 15b) from DS3 to DS4 and confirmed this pattern.

Experiment 6 (without evaporite deformation/dissolution, Fig. 15c) indicated that in Cariatiz,


geometries and facies distribution were not consistent with field observations (Fig. 12e). Actually, ooid
shoals (F2) prograded more than in the field study (Fig. 15c). MA1 microbialites (F4) were lacking
whereas lagoonal muds (F3) were too abundant. The distribution and thickness of alluvial fan deposits
clearly differed from the field study (Fig. 15c), as detrital sediments spread out too far basinward
during DS4. Nevertheless, the comparison of simulated cumulated thicknesses with the field study, as
for the best-fit model (Experiment 1), did not present significant differences. In Sorbas, Experiment 6
(Fig. 15c) reproduced geometries and field-observed facies distributions. Consequently, the
sedimentary architecture and facies distribution were almost identical to the field study. The main
differences lay in the formation of thicker ooidic shoals (F2) during DS4 (Fig. 16c).

This article is protected by copyright. All rights reserved.

Accepted Article

These results highlight the need to invoke subsidence to explain field-observed geometries, facies
distribution and DS thicknesses in Sorbas. Indeed, water-level fluctuations alone were not sufficient to
explain the observed stratigraphic architecture, as the available space was filled too rapidly.
In Cariatiz, regional subsidence seems to have had only a minor impact as attested by the similarity
between the field study and simulated geometries (Fig. 15a). By contrast, the evaporite
deformation/dissolution appeared to drive the stratigraphic architecture and could be considered as an
essential controlling factor for geometries and facies distribution. In Experiment 6, the typical layercake pattern observed in the field case study and in Experiment 1 (Fig. 12) is not reproduced.
Base-level fluctuations and production rates were fixed for models run with (Experiment 1) and
without (Experiment 6) evaporite deformation/dissolution. A decrease in carbonate production rates
(from an ooidic-dominated: 1.12 m kyr-1 to a microbialitic-dominated system: 0.27 m kyr-1) occurred
within DS3 due to a turnover in carbonate production. This change in sedimentary dynamics induced a
decrease in sediment supply and implicitly a slight increase in A/S ratio at the basin-scale. However,
even if the S term increased at the basin-scale, it could locally (Cariatiz area) decline due to the
constant rise in detrital influxes.

Comparison of simulated cumulated thicknesses (with or without evaporite deformation/dissolution,


Experiments 1 and 6) with the field study did not present significant differences (Fig. 15b). This may
be explained by the different modes of sediment dispersal across the platform, caused by two different
palaeotopographies. Indeed, the cumulated thicknesses may locally change resulting in a
homogeneous facies distribution in Experiment 6 or a heterogeneous distribution as in Experiment 1.
This pattern is well displayed on the thickness maps (Fig. 16ad). Sediments were evenly deposited
throughout the basin when the model was run without evaporite deformation/dissolution (Experiment
6; Fig. 16d), whereas they were restricted to the northern margin in Experiment 1 (Fig. 16b).
This evidence argues for a high impact of evaporite deformation/dissolution on sedimentation. These
numerical simulations finally demonstrate that even if the evaporite deformation/dissolution was a
local phenomenon, occurring mainly on the northern margin platform (Bourillot et al., 2010b), it
seems to have impacted the sedimentary record of the entire basin.

In Sorbas, Experiment 5 highlighted the need to invoke regional subsidence to explain field-observed
geometries, facies distribution and DS thickness (Figs. 12 and 15). Indeed, base-level fluctuations
alone were not sufficient to explain the observed stratigraphic architecture, as the available space was
filled too rapidly (Fig. 15a). The impact of evaporite deformation/dissolution was weaker than in the
Cariatiz area, as observed in the field study (Bourillot et al., 2010b). Consequently, both the
stratigraphic architecture and facies distribution of Experiment 6 were almost identical to the field
study. However, even if cumulated DS thicknesses did not differ significantly (i.e. parallel curves)
from the field study, simulated DSs observed in Experiment 6 were slightly thinner than in the best-fit

This article is protected by copyright. All rights reserved.

Accepted Article

model (Experiment 1; Fig. 16b).


The experiments showed that both regional subsidence and local evaporite deformation/dissolution
were important controlling factors of the A/S ratio, even if they acted differently between the two
referenced cross-sections.

6. 3D modelling and the quantification of carbonate production


The use of stratigraphic forward modelling programs is a robust approach for predicting stratigraphic
architecture in unexplored zones. Indeed, interpolations from field-controlled cross-sections using 3D
visualisation can be used to ascertain the extent and geometry of non-outcropping or eroded
sedimentary bodies (Warrlich et al., 2002; 2008). In petroleum geoscience, this approach can be used,
for example, to trace reservoir distribution at block frontiers (Granjeon, 2010). Comparison with
outcrop analogues using a detailed database can help to assess the potential distribution of reservoir
facies (Aigner et al., 2007; Palermo et al., 2011) and predict their inherent geometrical heterogeneities
(Burchette et al., 1990; Palermo et al., 2011). In the Sorbas Basin, the main potential reservoir facies
are ooids and microbialites. 3D distribution and volume quantification of these facies are coupled to
provide a clear picture of the stratigraphic architecture of the TCC deposits.
The carbonate production and the net-to-gross ratio (N/G), i.e. the net facies volume divided by the
total sediment volume (carbonate + detrital), per time-step were calculated. Figure 17 displays the
sediment gross volume per time-step, with split facies proportions compared with base-level
fluctuations. About 82% of total carbonate production occurred during the transgressive part of the
TCC (from SB1 to MFS3; basin-scale averaged production rate: 0.06 m.kyr-1; Fig. 16). A significant
fall, ~ 73%, in the total sediment volume occurred at the base-level reversal between 5.5 and 5.495
Ma. Only ~ 18% of the carbonate sediments accumulated between the end of DS3 and the end of DS4
(basin-scale averaged production rate: 0.02 m.kyr-1). This shows that base-level fluctuations,
accounting for a great part of the accommodation space variations, exerted close control over
carbonate production and total sediment accumulation (Fig. 17).

This decrease in sediment accumulation was coeval with a change in carbonate production. The basin
was filled mostly by ooids (F2) and outer platform muds (F1) during the transgressive phase, while
microbialites (F4; mostly MA1) became dominant facies from the end of DS3. This turnover could be
explained by the renewed restriction of the Sorbas Basin at the end of TCC deposition, which could
have resulted in decreased wave energy (Bourillot et al., 2010a). This turnover in carbonate production
and a radical change in hydrodynamics together caused a decline in the volumetric accumulation
balances towards the end of DS3 (Fig. 17). During the regressive part of the TCC (MFS3-SB5), the

This article is protected by copyright. All rights reserved.

Accepted Article

accommodation space was very low, explaining the thin accumulation of sediment in the nondeformed areas such as Sorbas (Figs. 14a and 16).
Detailed quantifications of ooid (Fig. 18) and microbialite production (MA1 and MA2, Fig. 19) were
performed to assess the impact of the main controlling factors on their respective productions. Ooid
production occurred almost exclusively in the transgressive part of the fourth-order TCC sequence
(nearly 99% of the total ooid production; Fig. 18). Geometrically, ooidic facies formed wide,
homogenous and continuous belts in our simulations (Fig. 18), with the bulk production located on the
flat topography to the west of Sorbas (Fig. 18). From SB1 to SB4, they were organized as wideconnected strata, their extent and volume changing in accordance with base-level fluctuations (Fig.
18). During the first time-step, a large increase in ooidic production was observed (5.595 Ma) in
conjunction with a rapid base-level rise over the very flat morphology of the basin floor (Sorbas crosssection). The flat and gentle palaeotopography allowed the formation of a large hydrodynamic strip in
which ooidic production occurred. The 3D views show that ooid production (indicated by net-to-gross
ratio (N/G)) was more extensive during the maximum flooding of the basin (MFS1; Fig. 18). N/G
values of ooids are always higher than 0.41 during the maximum water level of each sequence (0.41,
MFS1; 0.72, MFS2; 0.53, MFS3; Fig. 18). By contrast, N/G values fall rapidly within the regressive
part of each sequence (0.197, SB3; 0.006, SB4). However, the regressive trend of DS1 reveals a
smaller fall in ooid production (N/G decreases by 0.03 from MFS1 to SB2). This pattern can be
explained by an excessive production of ooids during the first simulation time-step (Fig. 18).
Therefore, ooid production seems to have been controlled by: (1) base-level fluctuations; and (2) a
decline in hydrodynamism related to basin restriction.

Comparison of both referenced cross-sections shows that microbialites (MA1 & MA2 associations)
are less developed in the best-fit model than in the field study (Fig. 12e). High-energy microbialites
(MA2, F5) developed mainly during the transgressive part of DS3, while forming small patches in
other sequences (Fig. 19b). Their production seems to have varied with base-level fluctuations (Fig.
19). N/G does not exceed 0.22 (MFS3 = 0.212). As the maximal production of MA2 microbialites
took place in association with ooidic shoals during MFS3, its under-representation can be explained by
the disparity between MA2 microbialite and ooid production rates (0.46 and 1.12 m kyr-1). This
assumption is corroborated by a slight increase in microbialite production during DS4 (Figs. 16 and
19). During this depositional sequence, ooids were lacking due to reduced hydrodynamics in the basin,
allowing more extensive development of microbialites. Even if MA2 microbialites developed within
high-energy shoals, their production rate was half the ooid production rate (MA2: 0.46 m.kyr-1; Ooid:
1.12 m.kyr-1). Consequently, their development could have been limited by the ooids because of their
excessive productivity. In the best-fit model, MA1 microbialites (F4) developed during the last two
sequences only, whereas they are recorded in all the DSs of the field study (Figs. 12 and 19a). Figure
20 displays two-trend variations in MA1 microbialite record. Their volume exhibits a positive

This article is protected by copyright. All rights reserved.

Accepted Article

correlation with fifth-order base-level variations (Fig. 20), increasing during the transgressive trend
and decreasing during the regressive trend water-level falls. N/G of MA1 microbialites seems to be
independent of water level, but is inversely proportional to conglomerate (F6) influx (Fig. 20).
Therefore, the amount of detrital supplies during DS3DS4 may have been responsible for a partial
limitation of microbialite development. This pattern was already evidenced on the Cariatiz crosssections, demonstrating that the development of MA1 microbialite distribution was at least partially
controlled by detrital influx, which prevented carbonate production. Bourillot et al. (2010a) assumed a
climatic control of high-frequency variations in detrital amount, possibly a reflection of alternating
wet/arid episodes. A recent study of the Guadalquivir Basin has produced evidence of a major climatic
change in pollen assemblages (oaks and thermophile taxa; Jimnez-Moreno et al., 2013). Indeed, the
latter authors indicate a transition from a cold/arid to a warm/humid climate around 5.5 Ma, which
could coincide with the increase in detrital influxes in our numerical modelling (Fig. 20). The
quantification, at basin scale, thus highlights that this disturbance in microbialite development is
recorded throughout the basin (Fig. 17). Moreover, field data (Fig. 12e) indicate that most of the MA1
microbialites were produced on the northern margin platform (Cariatiz cross-section) within the
evaporite deformation/dissolution area. As proposed by Bourillot et al. (2010b), microbialite
production may have been driven by salinity fluctuations caused by basin restriction and evaporite
recycling associated with evaporite deformation/dissolution. The increasing dissolution of evaporites
may have been responsible for brine circulation enhancing microbialite development.

7. Timing and origins of processes controlling accommodation space


variations
The platform deposits in the Terminal Carbonate Complex (TCC) all lack accurate dating due to the
absence of specific biostratigraphic markers or magnetic polarity reversals (chron C3R; Montgomery
et al., 2001; Krijgsman et al., 2001). Krijgsman et al. (2001) proposed a cyclostratigraphic age for
intra-MSC deposits in the Sorbas Basin by counting 14 gypsum-intergypsum sequences in the Yesares
Member, to which they added the three depositional sequences identified in the Sorbas Member by
Dabrio & Polo (1995). By correlating these sedimentary cycles with the insolation peaks of Laskar et
al. (1993), Krijgsman et al. (2001) determined that the Yesares Member was precessionally-controlled
and deposited between 5.96 and 5.67 Ma, while the Sorbas Member/TCC was controlled either by
precession or obliquity and therefore ended at 5.6 or 5.54 Ma, with a duration ranging between 60 and
120 kyr.
In order to discriminate among TCC durations, an experiment was carried out considering either
precessional or obliquity control of the four fifth-order depositional sequences proposed by Bourillot
et al. (2010a; b). Figure 21 presents the output results obtained by using 20 kyr (Experiment 7) and 40
kyr (Experiment 1; Figs. 12 and 13a) depositional sequences, respectively. When average production

This article is protected by copyright. All rights reserved.

Accepted Article

rates are applied to a model run with 20 kyr durations for each sequence, the basin is never filled. In
Experiment 7 (20 kyr; Fig. 21a) sediments of DS1DS2 show an aggradation. The basin is filled
mainly by outer platform muds (F1) during DS1 and DS2 and by MA1 microbialites (F4; in Sorbas
mainly) and conglomerates (F6; only in Cariatiz) during DS3 and DS4. Ooids are scarce in both
sections (Fig. 21a), except for very small shoals (< 3 m high) formed on the palaeo-reefal slope relief
in Cariatiz and in the most proximal domain of Sorbas during DS3 (Fig. 21a). As wave energy is the
same in Experiments 1 and 7, this pattern could be related to variations in accommodation space.
Using average production rates with DS durations of 20 kyr, base-level fluctuations seem too rapid and
excessive in amplitude, inducing the formation of a starved basin, and the basin remains unfilled at the
end of the simulations. As production rates are constant per time-step in DIONISOS, total sediment
production with 20 kyr DSs is half that of 40 kyr sequences. Another experiment was performed using
double the carbonate production rates with 20 kyr sequences (Experiment 8; Fig. 21b). In this
simulation, DS thicknesses are significantly similar to those measured in the best-fit model (Fig. 12),
but facies extension and distribution are both completely different. High-energy domains are
characterised by extensive development of MA2 microbialites (F5), while ooids (F2) are rare, as in
Experiment 7. Contrastingly, the best-fit model records intensive ooid production from SB1 to MFS3.
This ooid sedimentation occurs mainly during the transgressive part of the DSs (Fig.18). This could be
related to the creation of wider high-energy areas during base-level rise phases that enhance ooid
production (Figs. 12 and 13a). In recent settings, this pattern is well known in the Bahamas Islands
(e.g. Schlager et al., 1994; Kindler & Hearty, 1996; Reijmer et al., 2012). Indeed, in Pleistocene
Holocene deposits of the Bahamas Islands, Aurell et al. (1995) and Kindler & Hearty (1996)
demonstrate that oolitic production occurred exclusively during major flooding events, as attested by
the occurrence of perched beach and shoreface facies. During the regressive trend, carbonate
production is dominated by coral reef and bioclastic deposits (Kindler & Hearty, 1996). The scarcity of
ooids in Experiment 8 may be explained by very rapid water-level fluctuations (i.e., during SB1 to
MFS1, base level rises is 4.79 m.kyr-1 in Experiment 8 while it is 2.395 m.kyr-1 in Experiment 1)
preventing the formation of wide ooid belts as observed in the field study.

Investigating the origins of TCC base-level fluctuations, Bourillot et al. (2010a) emphasize that
comparable changes are recorded in the other basins of the western Mediterranean margins (e.g.
Morocco, Cunningham et al., 1995; Cabo de Gata, south-eastern Spain, Franseen et al., 1998). They
postulate an intra-Messinian reflooding of the western Mediterranean margins (Bourillot et al., 2010b).
Indeed, TCC base-level fluctuations (Fig. 7) are comparable with the eustatic fluctuations of about 70
80 m (TG12TG9) around 5.6 Ma followed by a 6070 m sea-level fall ((TG9TG6; Miller et al.,
2005). This eustatic fluctuation is coupled to high-frequency oscillations associated with the waning
and waxing of polar ice sheets, supposedly controlled by 40 kyr climatic periods (Miller et al., 2005).
Using realistic carbonate production rates, our modelling shows that the results most consistent with

This article is protected by copyright. All rights reserved.

Accepted Article

the field study are obtained when the model is run with 40 kyr DSs. This is in good agreement with a
recent study providing evidence of a predominant expression of obliquity in deposits of south-western
Spain from 6 to 5.2 Ma (Jimnez-Moreno et al., 2013).

8. Advantages and limits of DIONISOS in constraining basin-fill controlling


factors
DIONISOS successfully simulates the facies distribution, geometries and field-observed thicknesses
of the two referenced cross-sections using modern carbonate system parameters. The use of modern
observational data is common in SFM of carbonate platforms (e.g. Bosence & Waltham, 1990;
Bassant & Harris, 2008; Paterson et al., 2006; Warrlich et al., 2008; Barrett & Webster, 2012; Saura et
al., 2012; Seard et al., 2013). The sensitivity analysis performed on modern production rates appears
to be an objective strategy for simulating the evolution of carbonate platforms. Moreover, in our
simulations, production rates remain constant over time and the factors controlling carbonate
production are the only parameters that fluctuate. By coupling this approach with sensitivity analyses
of production rates, subsidence rates and sequence durations can be used to identify precisely the
factors controlling carbonate production.

The approach used in this study could be applied to other Messinian platforms of western
Mediterranean basins (Bajo-Segura Basin, SE Spain; Mallorca, Balearic Islands or Melilla, Morocco)
to confirm the factors that allegedly control sedimentation and to test the models reproducibility on a
wider scale. Our modelling strategy could also be applied to older ooid-microbialite systems to make
3D predictions about sedimentary geometries. Such systems may be major hydrocarbon reservoirs but
they generally display complex geometries (Harris et al., 2011) and such studies may further our
understanding of the parameters that determine their spatial distribution and their evolution over time.
The TCC deposits may prove a good reservoir analogue from which to assess the impact of evaporite
deformation on carbonate geometries. Indeed, some subsurface ooid-microbialite reservoirs in
northern Europe (Ca2, Zechstein, Strohmenger & Strauss, 1996; Kotarba & Wagner, 2007) or in the
Middle East (Khuff reservoirs, Alsharhan, 1993; Arab reservoirs, Ehrenberg et al., 2007) have
developed upon possibly deformed evaporite strata. Our study demonstrates that evaporite
deformation exerts a substantial control over sediment composition and dispersal.
Like most deterministic programs, DIONISOS is applied to test the impact of climate and tectonics on
sedimentary basin filling. It can successfully reproduce large-scale geometries originated by allocyclic
processes such as the response of eustatic variations on detrital-shelf morphologies (Smme et al.,
2009) or decipher the effect of base-level fluctuations and tectonics on basin filling (Csato et al.,
2012). Among examples from the carbonate realm, DIONISOS demonstrates a great capability to

This article is protected by copyright. All rights reserved.

Accepted Article

reproduce basin-scale stratigraphic architecture and facies heterogeneities. For instance, it can depict
how the folding of foreland basins controls the development of carbonate platforms by creating
extensive shallow-water areas (Saura et al., 2012) or it can support discussion of the complex facies
distribution of isolated platforms in relation to water-level fluctuations (Bassant & Harris, 2008). Our
numerical models confirm this capability to reproduce basin-scale processes that control stratigraphic
architecture and facies distribution heterogeneities. DIONISOS confirms that it is a relevant tool for
predicting reservoir distribution but it cannot display internal sedimentary body heterogeneities. While
large-scale structures and stratigraphic architecture can be successfully simulated this is difficult for
smaller features because of the cell-size limitation. For example, DIONISOS cannot display smallscale irregular structures related to the TRU (i.e. karst) or small-scale sedimentary or deformation
structures such as sand waves, collapse breccia and fault networks affecting pre-kinematic strata in
relation with evaporite deformation/dissolution (Fig. 12e). However, increased cell resolution would
entail longer computation time, possibly resulting in incoherent output results due to a limitation of the
diffusivity equation (Pierre et al., 2009). Moreover, processes such as salinity variations, brine
recycling and evaporite deformation remain user-controlled. Because this study is one of the first
attempts to integrate salinity fluctuations in a series of 3D SFM simulations, this variable was
parameterized from geological data (floro-faunal associations and mineralogy; Bourillot et al., 2010b).
To obtain more realistic simulations, it might well be necessary to induce direct salinity variations
depending on climatic, hydrogeologic and tectonic controls. Indeed, this module is fully usercontrolled as it is input manually from field observations. In future software solutions, such
deformation should be mechanical, taking into account the rheological properties of evaporites that
could induce gravity-driven deformation such as spreading processes generated by sedimentary
overload.

9. Conclusions
The contribution of various parameters (water level, subsidence and evaporite deformation) to
accommodation space variations can be deciphered by coupling averaged modern production rates,
sensitivity analyses and carbonate production quantification. At basin scale, base-level fluctuations are
the main factor controlling accommodation space. They control both stratigraphic architecture and
carbonate production. Simulation highlights their impact on ooid production, which occurs
predominantly during the transgressive part of fifth-order sequences. Numerical modelling indicates
that evaporite deformation/dissolution acts at different scales. At basin scale, it partly controls the
turnover in carbonate production in association with the basin restriction trend. It governs local
accommodation-space variations brought about by changes in facies distribution. The increase in local
accommodation space during the DS3DS4 interval induces a concentration of detrital supplies on the
northern margin of the basin. Evaporite deformation/dissolution on the northern margin platform

This article is protected by copyright. All rights reserved.

Accepted Article

greatly impacts the stratigraphic architecture, forming aggrading-prograding geometries, whereas


sedimentary bodies prograde in the rest of the basin. The recycling of evaporitic brines substantially
affects the proliferation of microbialites, as our modelling results demonstrate that they develop
preferentially in the areas of maximum deformation. The negative correlation between detrital influx
and microbialites combined with high-frequency variations in detrital supply indicate a climatic
control over sedimentation, which could reflect alternating wet/arid episodes. Our approach also
makes it possible to specify sequence durations and to provide evidence of an obliquity control over
TCC medium-scale depositional sequences (fifth order), possibly through glacio-eustatism.
The stratigraphic architecture and facies distribution of a Messinian carbonate platform system have
been successfully reproduced using averaged modern carbonate production rates with depth-dependent
changes. This basin modelling is based on two detailed cross-sections, reflecting both the central and
marginal parts of the Sorbas Basin. Field and numerical model comparisons are focused on these two
areas and extrapolated at the basin-scale.

Nevertheless, DSs thickness and local discrepancy in clinoform geometries between model output and
field data may be related to fluctuating local parameters (e.g. evaporite dissolution, soft-sediment
deformation, carbonate sediment dispersal, and specific accretion rates). Thus, initial basement
topography and production rates are the main controls on platform physiography.
This reconstruction demonstrates the potential reproducibility of such approaches for basin-scale
carbonate platform modelling. The 3D approach remains essential for appraising the relative
contributions from the major controlling factors. Coupled with modern sedimentary system
parameterisation, this could prove a very useful approach for exploration/production strategies,
through the prediction of stratigraphic architectures, even from incomplete datasets. This study
demonstrates how DIONISOS can help to assess carbonate production. Finally, TCC platforms are a
good analogue for subsurface hydrocarbon reservoirs in the Middle East and northern Europe,
developed above evaporite strata potentially subjected to deformation/dissolution phenomena.

Acknowledgements
This work is a contribution by the Systmes, Environnements et Dynamique Sdimentaire team of
the Biogosciences Laboratory (UMR CNRS/uB 6282) and the Gosciences, Hydrosciences,
Matriaux, Construction of the ENSEGID. Dionisos software has been developed at IFP Energies
Nouvelles. The authors thank Sarah Jane Mairet (Universit de Bourgogne) for her contribution to the
preliminary 2D Dionisos simulations. Christopher Sutcliffe and Carmela Chateau-Smith are thanked
for the English checking. Georg Warrlich, Isabel Montanez and an anonymous reviewer are greatly
thanked for their constructive comments which helped to improve the manuscript.

This article is protected by copyright. All rights reserved.

Accepted Article

References
AIGNER, T., BRAUN, S., PALERMO, D. & BLENDINGER, W. (2007) 3D Geological modelling of
a carbonate shoal complex: reservoir analogue study using outcrop data. First Break, 25, 6572.
ALSHARHAN, A.S. (1993) Facies and sedimentary environment of the Permian carbonates (Khuff
Formation) in the United Arab Emirates. Sedimentary Geology, 84, 89-99.
ALZAGA-RUIZ, H., GRANJEON, D., LOPEZ, M., SERANNE, M. & ROURE, F. (2009)
Gravitational collapse and Neogene sediment transfer across the western margin of the Gulf of
Mexico: insights from numerical models. Tectonophysics, 470, 21-41.
AUGIER, R., JOLIVET, L. & ROBIN, C. (2005) Late orogenic doming in the Eastern Betic
Cordilleras: final exhumation of the Nevado-Filabride Complex and its relation to basin
Genesis. Tectonics, 24, TC4003.
AURELL, M., MCNEILL, D.F., GUYOMARD, T. & KINDLER, P. (1995) Pleistocene shallowingupward sequences in New Providence, Bahamas: signature of high-frequency sea-level
fluctuations in shallow carbonate platforms. Journal of Sedimentary Research, 65, 170-182.
BARRETT, S.J. & WEBSTER, J.M. (2012) Holocene evolution of the Great Barrier Reef: Insights
from 3D numerical modeling. Sedimentary Geology, 265-266, 56-71.
BASSANT, P. & HARRIS, P.M. (2008) Analyzing sequence architecture and reservoir quality of
isolated carbonate platforms with forward stratigraphic modeling. In: Controls on Carbonate
Platform and Reef Development (Ed. by J. Lukasik and J.A. Toni Simo) SEPM Special
Publication, 89, pp. 343-359. SEPM, Tulsa, Oklahoma.
BASSETTI, M.A., MICULAN, P. & SIERRO, F.J. (2006) Evolution of depositional environments
after the end of Messinian Salinity Crisis in Njar Basin (Se Betic Cordillera). Sedimentary
Geology, 188, 279-295.
BATHURST, R.G.C. (1975) Carbonate sediments and their diagenesis. Developments in
Sedimentology 12. Elsevier, Amsterdam.
BENSON, R.H. RAKIC-EL BIED, K., BONADUCE, G. (1991) An important current reversal (influx)
in the Rifian corridor (Morocco) at the Tortonian-Messinian boundary: the end of Tethys
Ocean. Paleoceanography, 6, 164-192.
BLAIR, T.C. & MCPHERSON, J.G. (1994) Alluvial fans and their natural distinction from rivers
based on morphology, hydraulic processes, sedimentary processes, and facies assemblages.
Journal of Sedimentary Research, 64, 450-189.
BLOM, W.M. & ALSOP, D.B. (1988) Carbonate mud sedimentation on a temperate shelf: Bass
Basin, Southeastern Australia. Sedimentary Geology, 60, 269-280.
BOSENCE, D.W.J. & WALTHAM, D.A. (1990) Computer modelling of the internal architecture of
carbonate platforms. Geology, 18, 26-30.

This article is protected by copyright. All rights reserved.

Accepted Article

BOSSCHER, H. & SCHLAGER, W. (1992) Computer-simulation of reef growth. Sedimentology, 39,


503-512.
BOULVAIN, F. (2001) Facies architecture and diagenesis of Belgian Late Frasnian carbonate
mounds. Sedimentary Geology. 145, 269-294.
BOURILLOT, R. (2009) Evolution des plates-formes carbonates pendant la Crise De Salinit
Messinienne: de la dformation des vaporites aux communauts microbialithiques (Sud-est
de l'Espagne). PhD Thesis, Universit de Bourgogne, Dijon.
BOURILLOT, R., VENNIN, E., ROUCHY, J.-M., DURLET, C., ROMMEVAUX, V., KOLODKA,
C. & KNAP, F. (2010a) Structure and evolution of a Messinian mixed carbonate-siliciclastic
platform: the role of evaporites (Sorbas Basin, South-east Spain). Sedimentology, 57, 477-512.
BOURILLOT, R., VENNIN, E., ROUCHY, J.-M., BLANC-VALLERON, M.M., CARUSO, A. &
DURLET, C. (2010b) The end of the Messinian Salinity Crisis in the Western Mediterranean:
insights from the carbonate platforms of South-eastern Spain. Sedimentary Geology, 229, 224253.
BRAGA, J.C., MARTIN, J.M. & RIDING, R. (1995) Controls on microbial dome fabric development
along a carbonate-siliciclastic shelf-basin transect, Miocene, SE Spain. Palaios, 10, 347-361.
BROECKER, W.S. & TAKAHASHI.T (1966) Calcium carbonate precipitation on Bahama Banks.
Journal of Geophysical Research, 71, 1575-1602.
BURCHETTE, T.P., WRIGHT, V.P. & FAULKNER, T.J. (1990) Oolitic sand body depositional
models and geometries, Mississippian of Southwest Britain: implications for petroleum
exploration in carbonate ramp settings. Sedimentary Geology, 68, 87-115.
BURGESS, P.M. (2012) A brief review of developments in stratigraphic forward modelling 20002009. In: Regional Geology and Tectonics: Principles of Geologic Analysis, 1st Edition. (Ed.
by D.G. Roberts and A.W. Bally), pp. 379-406. Elsevier Science, Amsterdam.
BURGESS, P.M. & WRIGHT, V.P. (2003) Numerical forward modeling of carbonate platform
dynamics: an evaluation of complexity and completeness in carbonate strata. Journal of
Sedimentary Research, 73, 637-652.
BURGESS, P.M., STEEL, R.J. & GRANJEON, D. (2008) Stratigraphic forward modeling of basinmargin clinoform systems: implications for controls on topset and shelf width and timing of
formation of shelf-edge deltas. In: Recent Advances in Models of Siliciclastic Shallow-Marine
Stratigraphy (Ed. by G.J. Hampson, R.J. Steel, P.M. Burgess and R.W. Dalrymple), SEPM
Special Publication, 90, pp. 35-45. SEPM, Tulsa, Oklahoma.
CHARVIN, K., HAMPSON, G.J., GALLAGHER, K.L., STORMS, J.E.A. & LABOURDETTE, R.
(2011) Characterization of controls on high-resolution stratigraphic architecture in wavedominated shoreface-shelf parasequences using inverse numerical modeling. Journal of
Sedimentary Research, 81, 562-578.

This article is protected by copyright. All rights reserved.

Accepted Article

CHIVAS, A.R., TORGERSEN, T. & POLACH, H.A. (1990) Growth-rates and Holocene development
of stromatolites from Shark Bay, Western-Australia. Australian Journal of Earth Sciences, 37,
113-121.
CUEVAS CASTELL, J.M., BETZLER, C., RSSLER, J., HSSNER, H. & PEINL, M. (2007)
Integrating outcrop data and forward computer modelling to unravel the development of a
Messinian carbonate platform in SE Spain (Sorbas Basin). Sedimentology, 54, 423-441.
CUNNINGHAM, K.J., BENSON, R.H., RAKIC-EL BIED, K. & MCKENNA, L.W. (1997) Eustatic
implications of Late Miocene depositional sequences in the Melilla Basin, Northeastern
Morocco. Sedimentary Geology, 107, 147-165.
CSATO, I., GRANJEON, D., CATUNEANU, O. & BAUM, G. R. (2012) A Three-dimensional
stratigraphic model for the Messinian Crisis in the Pannonian Basin, Eastern Hungary. Basin
Research, 25, 121-148.
DABRIO, C.J. & POLO, M.D. (1995) Oscilaciones eustaticas de alta frecuencia en el Neogeno
Superior de Sorbas (Sorbas, Sureste de Espana). Geogaceta, 18, 75-78.
DAVIES, J.A., JAMES, R.O. & LECKIE, J.O. (1978) Surface ionization and complexation at the
oxide/water interface. Journal of Colloid Interface Science, 63, 480-499.
DILL, R.F., SHINN, E.A., JONES, A.T., KELLY, K. & STEINEN, R.P. (1986) Giant subtidal
stromatolites forming in normal salinity water. Nature, 324, 55-58.
DILL, R.F., KENDALL, C.G. & SHINN, E.A. (1989) Giant subtidal stromatolites and related
sedimentary features, Lee Stocking Island, Exumas, Bahamas. 28th International Geological
Congress Field Trip Guidebook. Washington, D.C., American Geophysical Union.
DRONKERT, H. (1985) Evaporite models and sedimentology of Messinian and recent evaporites.
GUA Papers of Geology, Series 1, 24.
DUPRAZ, C. & VISSCHER, P.T. (2005) Microbial lithification in marine stromatolites and
hypersaline mats. TRENDS in Microbiology, 13, 429-438.
DUPRAZ, C., VISSCHER, P.T., BAUMGARTNER, L.K. & REID, R.P. (2004) Microbe-mineral
interactions: early carbonate precipitation in a hypersaline lake (Eleuthera Island, Bahamas).
Sedimentology, 51, 745-765.
DUPRAZ, C., REID, R.P., BRAISSANT, O., DECHO, A.W., NORMAN, R.S. & VISSCHER, P.T.
(2009) Processes of carbonate precipitation in modern microbial mats. Earth-Science Reviews,
96, 141-162.
ENOS, P. (1977) Holocene sediment accumulations of the South Florida shelf margin. In: Quaternary
Sedimentation in South Florida (Ed. by P. Enos and R.D. Perkins), Geological Society of
America, Memoir, 147, pp. 1-130. Geological Society of America, Boulder, Colorado.
EHRENBERG, A.N., NADEAU, P.H. & AQRAWI, A.A.M. (2007) A comparison of Khuff and Arab
reservoir potential throughout the Middle East. AAPG Bulletin, 91, 275-286.

This article is protected by copyright. All rights reserved.

Accepted Article

ESTEBAN, M. (1979) Significance of the Upper Miocene coral reefs of the Western Mediterranean.
Palaeogeography Palaeoclimatology Palaeoecology, 29, 169-188.
FELDMANN, M. & MCKENZIE, J.A. (1998) Stromatolite-thrombolite associations in a modern
environment, Lee Stocking Island, Bahamas. Palaios, 13, 201-212.
FLGEL, E. (1982) Microfacies Analysis of Limestones. Springer-Verlag, Berlin.
FOLK, R.L. (1973) Carbonate petrography in the post-Sorbian age. In: Evolving Concepts in
Sedimentology (Ed. By R.N. Ginsburg), pp. 118-159. Johns Hopkins Press, Baltimore,
Maryland.
FORTUIN, A.R. & KRIJGSMAN, W. (2003) The Messinian of the Njar Basin (SE Spain):
Sedimentation, depositional environments and paleogeographic evolution. Sedimentary
Geology, 160, 213-242.
FORTUIN, A.R., KRIJGSMAN, W., HILGEN, F.J. & SIERRO, F.J. (2000) Late Miocene
Mediterranean desiccation: topography and significance of the Salinity Crisis erosion surface
on-land in Southeast Spain: Comment. Sedimentary Geology, 133, 167-174.
FRANSEEN, E.K., GOLDSTEIN, R.H. & FARR, M.R. (1998) Quantitative controls on location and
architecture of carbonate depositional sequences: Upper Miocene, Cabo De Gata Region, SE
Spain. Journal of Sedimentary Research, 68, 283-298.
GAUTIER, F., CLAUZON, G., SUC, J.-P., CRAVETTE, J. & VIOLANT, D. (1994) Age et dure de
la Crise de Salinit Messinienne. Compte Rendu de lAcadmie des Sciences de Paris, Srie 2,
318, 495-510.
GLUNK, C., DUPRAZ, C., BRAISSANT, O., GALLAGHER, K.L., VERRECCHIA, E.P. &
VISSCHER, P.T. (2011) Microbially mediated carbonate precipitation in a hypersaline lake,
Big Pond (Eleuthera, Bahamas). Sedimentology, 58, 720-738.
GOLDSTEIN, R.H. & FRANSEEN, E.K. (1995) Pinning points - a method providing quantitative
constraints on relative sea-level history. Sedimentary Geology, 95, 1-10.
GRANJEON, D. (1997) Modlisation stratigraphique dterministe: conception et applications d'un
modle diffusif 3D multilithologique. PhD Thesis, Universit de Rennes 1.
GRANJEON, D. (2009) 3-D stratigraphic modeling of sedimentary basins. AAPG Annual Convention
and Exhibition, Denver, Colorado.
GRANJEON, D. (2010) Dionisos 3D stratigraphic modelling of sedimentary basins. AAPG R&D
Studies, 5, 4-5.
GRANJEON, D. & JOSEPH, P. (1999) Concepts and application of a 3D multiple lithology, diffusive
model in stratigraphic modeling. In: Numerical Experiments in Stratigraphy: Recent Advances
in Stratigraphic and Sedimentologic Computer Simulations (Ed. by J.W. Harbaugh, W.L.
Watney, E.C. Rankey, R. Slingerland, R.H. Goldstein and E.K. Franseen), SEPM Special
Publication, 62, pp. 197-210. SEPM, Tulsa, Oklahoma.

This article is protected by copyright. All rights reserved.

Accepted Article

GRATACS, O., BITZER, K., CASAMOR, J.L., CABRERA, L., CALAFAT, A., CANALS, M. &
ROCA, E. (2009) Simulating transport and deposition of clastic sediments in an elongate
basin using the SIMSAFADIM-CLASTIC program: the Camarasa artificial lake case study
(NE Spain). Sedimentary Geology, 222, 16-26.
HALLEY, R.B., SHINN, E.A., HUDSON, J.H. & LIDZ, B. (1977) Recent and relict topography of
Boo Bee patch reef, Belize. 3rd International Coral Reef Symposium, 2, Miami, 29-35.
HARRIS, P.M. (1979) Facies anatomy and diagenesis of a Bahamian ooid shoal. Sedimenta VII. The
University of Miami, The Comparative Sedimentology Laboratory.
HARRIS, P.M., PURKIS, S.J. & ELLIS, J. (2011) Analyzing spatial patterns in modern carbonate
sand bodies from Great Bahama Bank. Journal of Sedimentary Research, 81, 185-206.
JAHNERT, R.J. & COLLINS, L.B. (2012) Characteristics, distribution and morphogenesis of subtidal
microbial systems in Shark Bay, Australia. Marine Geology, 303, 115-136.
JAMES, N.P. (1997) The cool-water carbonate depositional realm. In: Cool-Water Carbonates (Ed. by
N.P. James and J.A.D. Clarke), SEPM Special Publication, 56, pp. 1-20. SEPM, Tulsa,
Oklahoma.
JIMNEZ-MORENO, G., PREZ-ASENSIO, J.N., LARRASOAA, J.C., AGUIRRE, J., CIVIS, J.,
RIVAS-CARBALLO, M.R., VALLE-HERNNDEZ, M.F., & GONZLES-DELGADO,
J.A. (2013) Vegetation, sea-level and climate changes during the Messinian salinity crisis.
GSA Bulletin, 125, 432-444.
JOLIVET, L., FACCENNA, C. & PIROMALLO, C. (2009) From mantle to crust: stretching the
Mediterranean. Earth and Planetary Science Letters, 285, 198-209.
KARTSON, D., NMETH, K., SZKELY, B., RUSZKICZAY-RDIGER, Zs. & PCSKAY, Z.
(2006) Incision of a river curvature due to exhumed Miocene volcanic landforms: Danube
Bend, Hungary. Int. J. Earth Sci., 95, 929-944.
KENDALL, C., STROBEL, J., CANNON, R., BEZDEK, J. & BISWAS, G. (1991) The simulation of
the sedimentary fill of basins. Journal of Geophysical Research, 96, 6911-6929.
KINDLER, P. & HEARTY, P.J. (1996) Carbonate petrography as an indicator of climate and sea-level
changes: new data from Bahamian Quaternary units. Sedimentology, 43, 381-399.
KOTARBA, M. & WAGNER, R. (2007) Generation potential of the Zechstein Main Dolomite (Ca2)
carbonates in the Gorzw Wielkopolski-Midzychd-Lubiatw area: geological and
geochemical approach to microbial-algal source rock. Przegld Geologiczny, 55, 1025-1036.
KRIJGSMAN, W., HILGEN, F.J., RAFFI, I., SIERRO, F.J. & WILSON, D.S. (1999) Chronology,
causes and progression of the Messinian Salinity Crisis. Nature, 400, 655-662.
KRIJGSMAN, W., FORTUIN, A., HILGEN, F. & SIERRO, F. (2001) Astrochronology for the
Messinian Sorbas Basin (SE Spain) and orbital (precessional) forcing evaporite cyclicity.
Sedimentary Geology, 140, 43-60.

This article is protected by copyright. All rights reserved.

Accepted Article

LASKAR, J., JOUTEL, F. & BOUDIN, F. (1993) Orbital, precessional, and insolation quantities for
the Earth from -20 Myr to +10 Myr. Astronomy and Astrophysics, 270, 522-533.
LAWRENCE, D.T., DOYLE, M. & AIGNER, T. (1990) Stratigraphic simulation of sedimentary
basins - concepts and calibration. AAPG Bulletin, 74, 273-295.
LEES, A. (1975) Possible influence of salinity and temperature on modern shelf carbonate
sedimentation. Marine Geology, 19, 159-198.
LEES, A. & BULLER, A.T. (1972) Modern temperate-water and warm-water shelf carbonate
sediments contrasted. Marine Geology, 13, 67-73.
LEINFELDER, R.R., NOSE, M., SCHMID, D.U. & WERNER, W. (1993) Microbial crusts of the
Late Jurassic: composition, palaeoecological significance and importance in reef construction.
Facies, 29, 195-229.
LI, Y.Y., LERCHE, I. & PERLMUTTER, M.A. (1993) Global cyclostratigraphy: a model of
carbonate growth patterns. Marine and Petroleum Geology, 10, 620-631.
LISITZIN, A.P. & RODOLFO, K.S. (1972) Sedimentation in the world. SEPM Special Publication,
17, SEPM, Tulsa, Oklahoma.
LLOYD, R.M., PERKINS, R.D. & KERR, S.D. (1987) Beach and shoreface ooid deposition on
shallow interior banks, Turks and Caicos Islands, British West-Indies. Journal of Sedimentary
Petrology, 57, 976-982.
LOGAN, B.W. & CEBULSKI, D.E. (1970) Sedimentary environments of Shark Bay, Western
Australia. In: Carbonate Sedimentation and Environments, Shark Bay, Western Australia (Ed.
by B.W. Logan, G.R. Davies, J.F. Read and D.E. Cebulski), AAPG Memoir, 13, pp. 1-37,
American Association of Petroleum Geology, Tulsa, Oklahoma.
LOGAN, B.W., HOFFMAN, P. & GEBELEIN, C.D. (1974) Algal mats, cryptalgal fabrics and
structures, Hamelin Pool, Western Australia. AAPG Bulletin, 22, 140-194.
LOREAU, J.-P. (1982) Les sdiments aragonitiques et leur gense. Mmoire du Musum dHistoire
Naturelle de Paris, srie C, 47.
MANN, C.J. & NELSON, W.M. (1989) Microbialitic structures in Storrs Lake, San Salvador Island,
Bahama Islands. Palaios, 4, 287-293.
MARTIN, J.M. & BRAGA, J.C. (1994) Messinian events in the Sorbas Basin in Southeastern Spain
and their implications in the recent history of the Mediterranean. Sedimentary Geology, 90,
257-268.
MEGIAS, A.G. (1985) Relaciones tectonosedimentarias entre arrecifes y evaporitas del Mio-Plioceno
de las cuencas de Almeria y Sorbas. Trabajos des geologia, 15, 153-157.
MILLER, K.G., KOMINZ, M.A., BROWNING, J.V., WRIGHT, J.D., MOUNTAIN, G.S., KATZ,
M.E., SUGARMAN, P.J., CRAMER, B.S., CHRISTIE-BLICK, N. & PEKAR, S.F. (2005)
The Phanerozoic record of global sea-level change. Science, 310, 1293-1298.

This article is protected by copyright. All rights reserved.

Accepted Article

MILLIMAN, J.D., FREILE, D., STEINEN, R.P. & WILBER, R.J. (1993) Great Bahama Bank
aragonitic muds - mostly inorganically precipitated, mostly exported. Journal of Sedimentary
Petrology, 63, 589-595.
MONTAGGIONI, L.F., BORGOMANO, J., FOURNIER, F. & GRANJEON, D. (Accepted Article)
Quaternary atoll development: New insights from the two-dimensional stratigraphic forward
modelling of Mururoa Island (Central Pacific Ocean). Sedimentology, doi: 10.1111/sed.12175.
MONTENAT, C. (1990) Les bassins Nognes du domaine btique oriental (Espagne). Tectonique et
sdimentation dans un couloir de dcrochement. Premire partie : tude Rgionale.
Documents et Travaux IGAL, Paris.
MONTENAT, C., OTT DESTEVOU, P., LAROUZIERE, F.D. DE. & BEDU, P. (1987) Originalit
godynamique des bassins Nognes du domaine btique oriental (Espagne). Notes et
Mmoires Total-CFP., 21, 11-50.
MONTGOMERY, P., FARR, M.R., FRANSEEN, E.K. & GOLDSTEIN, R.H. (2001) Constraining
controls on carbonate sequences with high-resolution chronostratigraphy: Upper Miocene,
Cabo De Gata region, SE Spain. Palaeogeography Palaeoclimatology Palaeoecology, 176,
11-45.
MORSE, J.W. & MACKENZIE, F.T. (1990) Geochemistry of sedimentary carbonates. Developments
in Sedimentology 48. Elsevier, Amsterdam.
MORSE, J.W., THURMOND, W., BROWN, E. & OSTLUND, H.G. (1984) The carbonate chemistry
of Great Bahamas Bank waters: after 18 years another look. Journal of Geophysical Research,
89, 3604-3614.
MORSE, J.W., ARVIDSON, R.D. & LTTGE, A. (2007) Calcium carbonate formation and
dissolution. Chemical Reviews, 107, 342-381.
MLLER, D.W. & HS, K.J. (1987) Event stratigraphy and paleoceanography in the Fortuna Basin
(Southeast Spain): a scenario for the Messinian Salinity Crisis. Paleoceanography, 2, 679696.
MUTTI, M. & HALLOCK, P. (2003) Carbonate systems along nutrient and temperature gradients:
some sedimentological and geochemical constraints. International Journal of Earth Sciences,
92, 465-475.
NEUMANN, A.C. & LAND, L.S. (1975) Lime mud deposition and calcareous algae in Bight of
Abaco, Bahamas: a budget. Journal of Sedimentary Petrology, 45, 763-786.
NEUWEILER, F., GAUTRET, P., THIEL, V., LANGE, R., MICHAELIS, W. & REITNER, J. (1999)
Petrology of Lower Cretaceous carbonate mud mounds (Albian, N Spain): insights into
organomineralic deposits of the geological record. Sedimentology, 46, 837-859.
OTT DESTEVOU, P. (1980) Evolution dynamique du bassin Nogne de Sorbas (cordillires
btiques orientales-Espagne). PhD Thesis, Universit de Paris VII, Paris.

This article is protected by copyright. All rights reserved.

Accepted Article

OTT DESTEVOU, P. & MONTENAT, C. (1990) Le bassin de Sorbas-Tabernas. In: Les Bassins
Nognes du domaine btique oriental (Espagne). Tectonique et Sdimentation dans un
Couloir de Dcrochement. Premire Partie : tude Rgionale (Ed. by C. Montenat), 1213,
pp. 101128. Documents et Travaux IGAL, Paris.
PACTON, M., ARIZTEGUI, D., WACEY, D., KILBURN, M.R., ROLLION-BARD, C., FARAH, R.
& VASCONCELOS, C. (2012) Going Nano: a new step toward understanding the processes
governing freshwater ooid formation. Geology, 40, 547-550.
PALERMO, D., AIGNER, T., NARDON, S. & BLENDINGER, W. (2011) Three-dimensional facies
modeling of carbonate sand bodies: outcrop analog study in an epicontinental basin (Triassic,
Southwest Germany). AAPG Bulletin, 94, 475-512.
PATERSON, R.J., WHITAKER, F.F., JONES, G.D., SMART, P.L., WALTHAM, D. & FELCE, G.
(2006) Accommodation and sedimentary architecture of isolated icehouse carbonate
platforms: insights from forward modeling with CARB3D(+). Journal of Sedimentary
Research, 76, 1162-1182.
PAULL, C.K., NEUMANN, A.C., BEBOUT, B., ZABIELSKI, V. & SHOWERS, W. (1992) Growthrate and stable isotopic character of modern stromatolites from San-Salvador, Bahamas.
Palaeogeography Palaeoclimatology Palaeoecology, 95, 335-344.
PERRY, C.T., SALTER, M.A., HARBORNE, A.R., CROWLEY, S.F., JELKS, H.L. & WILSON,
R.W. (2011) Fish as major carbonate mud producers and missing components of the tropical
carbonate factory. Proceedings of the National Academy of Sciences of the United States of
America, 108, 3865-3869.
PIERRE, A., JONES, G., HARRIS, P.M. & DURLET, C. (2009) Simulating the stratigraphic
evolution of a Jurassic carbonate ramp outcrop analogue using the forward stratigraphic model
DIONISOS. AAPG Annual Convention. Denver, Colorado.
PINCKNEY, J., PAERL, H.W. & BEBOUT, B.M. (1995) Salinity control of benthic microbial mat
community production in a Bahamian hypersaline lagoon. Journal of Experimental Marine
Biology and Ecology, 187, 223-237.
PLANAVSKY, N. & GINSBURG, R.N. (2009) Taphonomy of modern marine Bahamian
microbialites. Palaios, 24, 5-17.
QUIQUEREZ, A., ALLEMAND, P. & DROMART, G. (2000) DIBAFILL: a 3-D two-lithology
diffusive model for basin infilling. Computers & Geosciences, 26, 1029-1042.
RANKEY, E.C. & REEDER, S.L. (2010) Controls on platform-scale patterns of surface sediments,
shallow Holocene platforms, Bahamas. Sedimentology, 57, 1545-1565.
RANKEY, E.C., REEDER, S.L. & GARZA-PEREZ, J.R. (2011) Controls on links between
geomorphical and surface sedimentological variability: Aitutaki and Maupiti atolls, South
Pacific Ocean. Journal of Sedimentary Research, 81, 885-900.

This article is protected by copyright. All rights reserved.

Accepted Article

REEDER, S.L. & RANKEY, E.C. (2008) Interactions between tidal flows and ooid shoals, Northern
Bahamas. Journal of Sedimentary Research, 78, 175-186.
REID, R.P. & BROWNE, K.M. (1991) Intertidal stromatolites in a fringing Holocene reef complex,
Bahamas. Geology, 19, 15-18.
REID, R.P., VISSCHER, P.T., DECHO, A.W., STOLZ, J.F., BEBOUT, B.M., DUPRAZ, C.,
MACINTYRE, L.G., PAERL, H.W., PINCKNEY, J.L., PRUFERT-BEBOUT, L., STEPPE,
T.F. & DESMARAIS, D.J. (2000) The role of microbes in accretion, lamination and early
lithification of modern marine stromatolites. Nature, 406, 989-992.
REID, R.P., JAMES, N.P., MACINTHYRE, I.G., DUPRAZ, C. & BURNE, R.V. (2003) Shark Bay
stromatolites: microfabrics and reinterpretation of origins. Facies, 49, 299-324.
REID, R.P., FOSTER, J.S., RADTKE, G. & GOLUBIC, S. (2011) Modern marine stromatolites of
Little Darby Island, Exuma Archipelago, Bahamas: environmental setting, accretion
mechanisms and role of euendoliths. In: Advances in Stromatolite Geobiology (Ed. by J.
Reitner, N.-V. Query and G. Arp), pp. 77-89. Springer-Verlag, Berlin.
REIJMER, J.J.G., PALMIERI, P. & GROEN, R. (2012) Compositional variations in calciturbidites
and calcidebrites in response to sea-level fluctuations (Exuma Sound, Bahamas). Facies, 58,
493-507.
RIDING, R., BRAGA, J.C., MARTIN, J.M. & SANCHEZ-ALMAZO, I.M. (1998) Mediterranean
Messinian Salinity Crisis: constraints from a coeval marginal basin, Sorbas, Southeastern
Spain. Marine Geology, 146, 1-20.
ROBBINS, L.L., TAO, Y. & EVANS, C.A. (1997) Temporal and spatial distribution of whitings on
Great Bahama Bank and a new lime mud budget. Geology, 25, 947-950.
ROHAIS, S., VENTRA, D. & DE BOER, P.L. (2008) Quantifying climatic and tectonic forcing
alluvial-fan stratigraphy by 3D numerical modeling and comparison with outcrop examples.
AAPG Annual Convention. San Antonio, Texas.
ROUCHY, J.-M. & SAINT-MARTIN, J.-P. (1992) Late Miocene events in the Mediterranean as
recorded by carbonate-evaporite relations. Geology, 20, 629-632.
ROUCHY, J.-M. & CARUSO, A. (2006) The Messinian Salinity Crisis in the Mediterranean Basin: a
reassessment of the data and an integrated scenario. Sedimentary Geology, 188, 35-67.
ROUCHY, J.-M., SAINT-MARTIN, J.-P., MAURIN, A. & BERNET-ROLLANDE, M.-C. (1986)
Evolution et antagonisme des communauts bioconstructrices animals et vgtales la fin du
Miocne en Mditerrane occidentale : biologie et sdimentologie. Bulletin du Centre de
Recherche Exploration et Production Elf Aquitaine, 10, 333-348.
SAURA, E., EMBRY, J.-C., VERGS, J., HUNT, D.W., CASCIELLO, E. & HOMKE, S. (2012)
Growth fold controls on carbonate distribution in mixed foreland basins: insights from the
Amiran foreland Basin (NW Zagros, Iran) and stratigraphic numerical modelling. Basin
Research, 25, 149-171.

This article is protected by copyright. All rights reserved.

Accepted Article

SCHLAGER, W. (2005) Carbonate sedimentology and sequence stratigraphy. SEPM Concepts in


Sedimentology, 8, SEPM, Tulsa, Oklahoma.
SCHLAGER, W., REIJMER, J.J.G. & DROXLER, A.W. (1994) Highstand shedding of carbonate
platforms. Journal of Sedimentary Research, 64, 274-281.
SCHMOKER, J.W. & HALLEY, R.B. (1982) Carbonate porosity versus depth a predictable relation
for South Florida. AAPG Bulletin, 66, 2561-2570.
SCOFFIN, T.P. (1987) An introduction to carbonate sediments and rocks. Chapman & Hall, NewYork.
SEARD, C., BORGOMANO, J., GRANJEON, D. & CAMOIN, G. (2013) Impact of environmental
parameters on coral reef development and drowning: Forward modeling of the last deglacial
reefs from Tahiti (French Polynesia; IODP Expedition #310). Sedimentology, doi:
10.1111/sed.12030.
SHINN, E.A., STEINEN, R.P., LIDZ, B.H. & SWART, P.K. (1989) Whitings, a sedimentologic
dilemma. Journal of Sedimentary Research, 59, 147-161.
SMME, T., HELLAND-HANSEN, W. & GRANJEON, D. (2009) Impact of eustatic amplitude
variations on shelf morphology, sediment dispersal, and sequence stratigraphic interpretation:
icehouse versus greenhouse systems. Geology, 37, 587-590.
STOCKMAN, K.W., GINSBURG, R.N. & SHINN, E.A. (1967) Production of lime mud by algae in
South Florida. Journal of Sedimentary Petrology, 37, 633-648.
STROHMENGER, C. & STRAUSS, C. (1996) Sedimentology and playnofacies of the Zechstein 2
carbonate (Upper Permian, Northwest 2063 Germany). In: Approaches to Sequence
Stratigraphy. (Ed. by R. Gaupp and A.A. Van de Weerd). Sedimentary Geology, 102, 55-77.
THOMPSON, J.B. (2001) Microbial whitings. In: Microbial Sediments (Ed. by R. Riding and S.M.
Awramik), pp. 250-269. Springer-Verlag, Berlin.
VAN DE POEL, H.M. (1994) Messinian marginal-marine and continental facies and their
stratigraphy in the Eastern Almeria province (SE Spain) Strata. Actes du Laboratoire de
Gologie Sdimentaire de lUniversit Paul Sabatier. Toulouse.
VASCONCELOS, C. & MCKENZIE, J.A. (1997) Microbial mediation of modern dolomite
precipitation and diagenesis under anoxic conditions (Lagoa Vermelha, Rio de Janeiro,
Brazil). Journal of Sedimentary Research, 67, 378-390.
VASCONCELOS, C., VISSCHER, P.T., WARTHMANN, R. & MCKENZIE, J.A. (2006) Formation
of lamination in modern stromatolites from Lagoa Vermelha, Brazil: an example for
Precambrian relics? Geochimica et Cosmochimica Acta, 70, A669- A669.
VISSCHER, P.T., REID, R.P. & BEBOUT, B.M. (2000) Microscale observations of sulfate reduction:
correlation of microbial activity with lithified micritic laminae in modern marine
stromatolites. Geology, 28, 919-922.

This article is protected by copyright. All rights reserved.

Accepted Article

WARRLICH, G., WALTHAM, D. & BOSENCE, D. (2002) Quantifying the sequence stratigraphy
and drowning mechanisms of atolls using a new 3-D forward stratigraphic modelling program
(CARBONATE 3D). Basin Research, 14, 379-400.
WARRLICH, G., BOSENCE, D., WALTHAM, D., WOOD, C., BOYLAN, A & BADENAS, B.
(2008) 3D stratigraphic forward modelling for analysis and prediction of carbonate platform
stratigraphies in exploration and production. Marine and Petroleum Geology, 25, 35-58.
WILBER, R.J., MILLIMAN, J.D. & HALLEY, R.B. (1990) Accumulation of bank-top sediment on
the Western slope of Great Bahama Bank - rapid progradation of a carbonate megabank.
Geology, 18, 970-974.
WILLIAMS, H., BURGESS, P., WRIGHT, V., DELLA PORTA, G. & GRANJEON, D. (2011)
Investigating carbonate platform types: multiple controls and a continuum of geometries.
Journal of Sedimentary Research, 81, 18-37.
WILSON, R.C.L. (1967) Particle nomenclature in carbonate sediments. Neues Jahrb. Geol. Palontol.
Monatsh., 68, 498-510.
WRIGHT, V.P. & BURCHETTE, T.P. (1996) Shallow-water carbonate environments. In:
Sedimentary Environments, Processes, Facies and Stratigraphy (Ed. by H.G. Reading), pp.
325394. Blackwell Scientific Publications, Oxford.
WRIGHT, V.P. & BURGESS, P.M. (2005) The carbonate factory continuum, facies mosaics and
microfacies: an appraisal of some of the key concepts underpinning carbonate sedimentology.
Facies, 51, 17-23.
YATES, K.K. & ROBBINS, L.L. (1999) Radioisotope tracer studies of inorganic carbon and Ca in
microbially derived Caco3. Geochimica Et Cosmochimica Acta, 63, 129-136.

Figure captions
Fig. 1. Model workflow used to assess parameters controlling carbonate production. Two different
data types (blue rectangles) are input into the model to parameterize it (green rectangle). They
correspond to field database and literature data, which are either based on modern carbonate systems
or published data focusing on the Sorbas Basin (palaeogeographic maps and cross-sections).
Sensitivity analyses (Experiments 1 to 8; orange rectangle) are performed to obtain different output
models that are then compared to the field study enabling a best-fit model (pink rectangle) to be
defined. Finally the best-fit model helps to characterise the processes controlling sedimentation (red
rectangle) and to realize feedback (green arrow) to the field study in order to assess their timing and
relative contribution of factors controlling basin filling.
Fig. 2. (a) Study areas (yellow stars) displayed on a simplified geological map of south-eastern Spain
(modified from Cuevas Castell et al., 2007; Bourillot et al., 2010a). (b) Detailed geological map of the

This article is protected by copyright. All rights reserved.

Accepted Article

Sorbas Basin (modified from Ott dEstevou & Montenat, 1990; Bourillot et al., 2010a). Black stars
correspond to the location of Cariatiz and Sorbas cross-sections used in this study.
Fig. 3. Stratigraphy of the Sorbas Basin (modified from Martin & Braga, 1994; Tor.: Tortonian; Zan.:
Zanclean; Mb.: Member). Ages correspond to the relative datings of the top member boundaries, (1:
Krijgsman et al., 2001; 2: Rouchy & Saint-Martin, 1992; Rouchy & Caruso, 2006; 3: Ott dEstevou,
1980; Fortuin et al., 2000).
Fig. 4. Intra-basin correlations of the Sorbas and Cariatiz cross-sections. It highlights the impact of
evaporite deformation on stratigraphic architecture in Cariatiz (modified from Bourillot et al., 2010b).
Fig. 5. Chronostratigraphical review of the Tortonian and Messinian series of the Sorbas Basin and
their equivalent in the Central Mediterranean Basin as proposed by several authors (1: Krijgsman et
al., 1999; 2: Krijgsman et al., 2001; 3: Riding et al., 1998; 4: Gautier et al., 1994; 5: Benson et al.,
1991; 6: Mller & Hs, 1986; 7: Dronkert, 1985; 8: Rouchy & Saint-Martin, 1992; 9: Bourillot et al.,
2010b). It demonstrates the large temporal uncertainties in deposition and duration of the Sorbas
Member (in yellow; Terminal Carbonate Complex; modified from Krijgsman et al., 2001).
Fig. 6. (a) 3D view of the model basement used for the initial stage of the simulation (SB1; 5.6 Ma).
The Yesares Member was simulated as a single unit surrounded by two surfaces (Top Reef
Unconformity and Top Yesares Member surface). It shows the location of the referenced crosssections of Cariatiz and Sorbas. This basinal view also depicts the location of the detrital influx source.
Colour chart depicts initial bathymetry of the model. Positive values reflect relief and negatives values
correspond to depressions in the basin floor. (b) Detail of reconstitution of the Top Reef Unconformity
surface (TRU) based on palaeogeographic maps of Ott dEstevou & Montenat (1990). (c) Detail of
reconstitution the Top Yesares Member surface based on palaeogeographic maps of Ott dEstevou &
Montenat (1990). (d) Palaeogeographical map of Bourillot et al. (2010b) indicating local uncertainties
(black square corresponds to the simulated area; G: Gochar; C: Cariatiz; S: Sorbas; H: Hueli).
Fig. 7. Comparison between the five facies association of Bourillot et al. (2010b) and the six facies
used in this work. Environmental parameters are detailed for each facies.
Fig. 8. (a) Base-level fluctuations associated with depositional sequences and sequence boundaries
(SBn; colour used referred to field study); this curve is based on the Relative water-level variations
proposed by Bourillot et al. (2010a). (b) Salinity fluctuation curve input in the model and deduced
from floro-faunal associations described by Bourillot et al. (2010b) and mineralogical contents.
Fig. 9. Numerical thickness maps of the evaporitic layer (Yesares Mb.) depicting the deformation
steps through time (digitized after Bourillot et al., 2010b): (a) before (from 5.6 to 5.5 Ma). (b and c)
during (5.48 Ma and 5.46 Ma, respectively). (d) after (5.44 Ma).

This article is protected by copyright. All rights reserved.

Accepted Article

Fig. 10. Production rates recorded in modern carbonate systems for each carbonate compound of the
TCC (ooids, muds and microbialites). Values displayed in parentheses respectively correspond to
minimal, median, averages and maximal production rates (Pmin, Pmed, Pave and Pmax). Colours
correspond to the main carbonate compounds in accordance with colour chart used in facies modelling
(see Fig. 7 for details).
Fig. 11. Carbonate production profiles defined for the main carbonate compounds recognized in the
TCC (ooids, muds and microbialites) according to the formula of Granjeon (1997) depicted in (a).
Those profiles link carbonate production rates (solid line) with water depth. Each curve is defined
from a maximal production rate (Pmax) decreasing with depth until zero from a critical boundary,
which can be either the fair weather wave base (a; b and c) or the photic zone base (c). (a) Ooidic
production in yellow. (b) Mud production (pink: lagoonal muds; green: outer platform muds). (c)
Microbial production (light purple: MA1; dark blue: MA2). Values used to define maximal production
rates correspond to average production rates and are detailed in the Figure 10.
Fig. 12. Global and detailed views of the Best-fit model (a-c) compared with the field study (d-e) of
Bourillot et al. (2010a; b). The two referenced cross-sections are depicted and corresponded to Sorbas
on the left and Cariatiz on the right (output time step; black lines: 5 kyr). (a) Global view of the Sorbas
cross-section; From SB1 to MFS3, sedimentation is dominated by ooids (F2) developing in highenergy zones, which form large shoals prograding basinward. From MFS3 to SB5, some small and flat
ooidic shoals are observed in lagoonal mud (F3). (b) Global view of Cariatiz cross-section. From SB1
to MFS3, sedimentation is dominated by ooids (F2) in the top reef flat forming aggrading shoals.
These ooidic belts (F2) separate shallow low-energy settings dominated by lagoonal muds (F3) from
the deep centre basin where outer platform muds (F1) are accumulated. From MFS3 to SB5, general
trend evolves towards basin restriction. Evaporite deformation starts, increasing the available space.
Sedimentation is dominated by MA1 microbialites (F4) and muds (F3) with increasing detrital
supplies. (c) Detailed views Cariatiz cross-section showing a good matching with field data (e).
Displayed colour chart for facies and sequence boundaries (according to the field study Bourillot et
al., 2010a). The brown arrow depicts the location of the source detrital influxes and the white arrows
display their dispersal (Output time step: 5kyr; V.E.: Vertical Exaggeration).
Fig. 13. Sensitivity analyses performed on production rates for the Cariatiz and Sorbas cross-sections.
As the field database parameters control the accommodation space variations, the unknown parameter
corresponds to the sediment supply. For simulations (a) to (d), model matching is depicted by a
comparison between numerical sediment supply (Sm) and field study sediment supply (Sfd). (a)
Experiment 1: average production rates ranged between 0.27 and 1.12 m kyr-1. This test seems to be
the closest to the field study as indicated by the model-matching ratio. (b) Experiment 2: median
production rates ranged between 0.17 and 0.93 m kyr-1. The matching ratio shows that numerical

This article is protected by copyright. All rights reserved.

Accepted Article

sediment supply is lower than expected. (c) Experiment 3: minimal production rates ranged from 0.04
to 0.12 m kyr-1. The sediment supply is the lowest and appears to be less than observed in the field
study. (d) Experiment 4: maximal production rates ranged from 0.61 to 3 m kyr-1. Numerical sediment
supply is greater than in the field study. Displayed colour chart of the different facies is available for
each view. Arrows depict maximal extension of DS2 ooidic shoals (yellow, Best-fit model,
Experiment 1; blue, tested experiment) and detrital facies (white, Best-fit, Experiment 1; black, tested
experiment; Output time step, black lines: 10 kyr).
Fig. 14. Comparison of cumulated thicknesses of depositional sequences for the Sorbas and Cariatiz
cross-sections. (a) Curves are depicted to test the similarity between sensitivity analyses performed on
production rates and the field study (Sorbas, left; Cariatiz, right). For the Cariatiz cross-section the
closest curve to the field study (diamond shapes) is given by average production rates (circles). For the
Sorbas cross-section the field case study curve is ranged by an underestimated average production rate
and an overestimated maximal production rate. (b) Location of three virtual wells respectively in
marginal, transitional and basinal settings on both referenced cross-sections. (c) Measurement methods
of cumulated thicknesses in Cariatiz. (d) Measurement methods of cumulated thicknesses in Sorbas.
Fig. 15. Cross-section view of simulated geometries when the model is run without regional
subsidence (Experiment 5; a) and without evaporite deformation/dissolution (Experiment 6; c). (b)
Cumulated thicknesses of depositional sequences of both experiments compared with the field case
study. In Sorbas, even if the cumulated thicknesses of the other two experiments (a, without
subsidence; b, without evaporite deformation) do not show significant differences. They differ in
terms of thicknesses, geometries and facies distribution from the field case study. In Cariatiz, both
experiments are close to the field case study but geometries are inconsistent for the Experiment 6.
Arrows depict maximal extension of DS2 ooidic shoals (yellow, Best-fit model, Experiment 1; blue,
tested experiment) and detrital facies (white, Best-fit, Experiment 1; black, tested experiment; Output
time step, black lines: 5 kyr).
Fig. 16. Thickness maps displaying depocentre evolution in Experiment 1 (with evaporite
deformation/dissolution; ab) and Experiment 6 (without evaporite deformation/dissolution; cd) from
SB1 to MFS3 (a and c) and from MFS3 to SB5 (b and d). It depicts a northern restriction of
depocentres (brown arrows) when model is run with evaporite deformation/dissolution (b) while they
are diffused when the model is run without evaporite deformation/dissolution (d). Colour chart reveals
a decrease in total sediment accumulation during the MFS3-SB5 interval, as maximal thickness is 44
m rather than 66 m from SB1 to MFS3.
Fig. 17. Total volume of sediments accumulated in the TCC. Gross volume is calculated for each time
step (5 kyrs) with respect to the proportion of facies. Base-level fluctuations (modified from Bourillot
et al., 2010a) are also displayed revealing that base level is one of the main controlling factors of

This article is protected by copyright. All rights reserved.

Accepted Article

accommodation space variations with respect to DSs and their boundaries (SB: coloured dots; MFS:
blue circles).
Fig. 18. Quantification of ooidic production (F2) and 3D distribution of ooidic bodies. (a) Total
volume of ooids accumulated in the TCC (yellow bars) with respect to DSs and their boundaries (SB:
coloured dots; MFS: blue circles). Net-to-gross ratio (light grey and squares) is also depicted and can
reach 90% (5.595 Ma) of the total accumulated sediments. Both accumulated volume and net-to-gross
ratio attest to a strong control of water-level fluctuations (dotted black line) on ooidic production. (b
d) 3D views of ooidic facies distribution. These model outputs indicate that the maximal production
occurs during transgressive trend of each DSs.
Fig. 19. Quantification and distribution of MA1 and MA2 microbialites (F4 and F5). (a) Total volume
of MA1 microbialites (light purple bars) with respect to DSs and their boundaries (SB: coloured dots;
MFS: blue circles). These microbialites only developed during the regressive trend of the fourth order
TCC sequence (DS3-DS4). Their total volume seems to be mainly water-level controlled, as attested
by the positive correlation between water-level fluctuations (dotted black line) and the volume of MA1
microbialites (F4). (b) Total volume of MA2 microbialites (dark purple bars) with respect to DSs and
their boundaries (SB: coloured dots; MFS: blue circles). Both volume of MA2 microbialites and netto-gross ratio (light grey line and squares) are controlled mainly by water-level fluctuations (dotted
black line). Net-to-gross values do not exceed 22% of total accumulated sediments.
Fig. 20. MA1 microbialites (F4) versus detrital facies (conglomerates, F6; and sands, F7) net-to-gross
ratio comparisons with respect to water-level fluctuations (dotted black line), DSs and their boundaries
(SB: coloured dots; MFS: blue circles). This comparison indicates a negative correlation, on the basin
scale, between the amount of detrital input (conglomerates, dark-brown line and diamonds; sand, lightbrown line and squares) and the production of MA1 microbialites (orange line and squares). This
pattern is indicative of carbonate production being inhibited by detrital inputs.
Fig. 21. Tests performed on the duration of depositional sequences. (a) Experiment 7 corresponds to
simulation runs combining average production rates (from 0.27 to 1.12 m kyr-1) and ~20 kyr DSs. (b)
Experiment 8 depicts results of ~20 kyr DSs runs with production rates twice as high (between 0.54
and 2.24 m kyr-1). Both experiments are inconsistent with the field study. Arrows depict maximal
extension of DS2 ooidic shoals (yellow, Best-fit model, Experiment 1; blue, tested experiment) and
detrital facies (white, Best-fit, Experiment 1; black, tested experiment; Output time step; black lines: 5
kyr).

This article is protected by copyright. All rights reserved.

- initial lithologies
- base-level curve
- local subsidence
maps (evap. def.)
- sea-water chemistry

- accumulation rates
- regional subsidence
maps
- initial surface

Siliciclastic input:
Carbonate production:
- supply
- accumulation rates
- diffusion coefficient
- bathymetry
- wave-energy
- salinity

PARAMETER UNCERTAINTIES

matching field geometries

MODEL PARAMETERIZATION

ivity analy
nsit
se
e
s
s

MODEL TESTING
Carbonate accumulation rates:
- minimum
- median
- average
- maximum

Impact of subsidence:
- with/without
regional and local (evap.
deformation) subsidence

BEST-FIT MODEL
Stratigraphic architecture (geometries, facies distribution,
thickness of depositional sequences)

QUANTIFICATION OF CONTROLLING FACTORS


=> respective contribution of each controlling factors on
carbonate production

This article is protected by copyright. All rights reserved.

performing statistical analyses on accumulation rates

LITERATURE DATA

ivity analy
nsit
se
se
s

Accepted Article

FIELD DATA

ult
ares fa

sin

Vera

Ve
ra

Cariatiz

asin

sb
rba

Sorbas

So

Tabernas

Palom

ba

Sierra de Los Filabres

Si

a
err

Ca

bre

ra
3700'000'''W

milla

Si
er

ra

E
S
10 km

ed

o
de

N
W

ite
Se rra
a nea

ta

lt

00200'000'''W

fau

lC
ab

Almeria

ja

ta

Carboneras

sin

de

N
a
err

a
rb

Ga

Njar

a
a Alh

Sier

(a)

Neogene volcanic
rocks
Miocene
reef complexes

Betic basement
Neogene and
Quaternary sediments

Evaporites
Study
areas

N
W

E
S

*
*
* Cariatiz

Sierra de Los Filabres


*
Gchar

*
*

v v

Sorbas

v v

v
v
v

v
v
v
v

v vv
v
v v
v v
v
v
v
v
v
v
v

v
v
v

Sierra de
Cabrera

v
v
v

Tabernas basin

Sierra Alhamilla

5 km

Zorreras Member (Messinian)


Red continental deposits

Sorbas Member (Messinian):


Mixed carbonates/siliciclastics,
ooids and microbialites

Postevap.
Evap.

Pliocene marine
shelly sandstones

Preevap.

Pliocene to Recent
continental deposits

v v

Yesares Member (Messinian):


massive gypsum and
interbedded carbonates-clastics
Cantera Member (Messinian):
Biohermal Unit
Fringing-Reef Unit

This article is protected by copyright. All rights reserved.

* Hueli
* *
**
*
*
*
* * *
Pre-evaporitic

Post-evaporitic

Accepted Article

SPAIN

v
v

v
v

v v

v v
v

(b)
v

Abad Member (Messinian):


Limestones-marls (L. Abad)
Diatomites-sapropels-marls (U. Abad)
Azagador Member (Tort.-Messinian):
Shallow marine calcarenites

Tortonian conglomerates
and turbiditic sandstones
Betic basement

Age
Zan.

Zorreras Mb.

?
Sorbas Mb.
?

Cantera
Mb.

5.33 Ma

(3)

~ 5.45 Ma (2)
~ 5.6 Ma

(2)

Yesares Mb.

Az

ag

ad

5.96 Ma

(1)

6.70 Ma

(1)

7.24 Ma

(1)

Upper

or

Mb

.
Lower

Messinian

~100 m

Accepted Article

Abad Mbr.

Tort.

Erosional contact
Normal contact

Mixed carbonate siliciclastic deposits,


microbialites and laminites

Unknown type of contact

Gypsum

Limestone-marl alternations
(Lower Abad Mb.)
Diatomite-sapropel-marl
alternations (Upper Abad Mb.)

Coral fringing reefs

Shallow marine calcarenites

Coral and algal bioherms,


calcarenites

Conglomerates and
turbiditic sandstones

Marine shelly sandstones


Continental deposits: conglomerates,
sandstones and lacustrine limestones

SORBAS

CARIATIZ

SE

NW

SB5
DS4

SB4

DS3

SB3

DS2

SB2?

?
?

DS1
V

V
V
V

V
V

SB1

V
V

Yesares Mbr.

LEGEND
V

Alluvial fan
Evaporites /
inter-evaporites
Sabkha

V
V

SL

V
V

20 m
500 m

U
TR a) Upper Abad Mbr.
M
(6

Lagoon / mud flat


High energy shoal
Outer platform,
mud dominated
Outer platform,
microbialite dominated

SL

Reef Unit

Solution-collapse
breccias
Reef blocks

Sequence boundaries (SB)

Microbialites

Spiculitic levels

Marker beds

Siliceous sponges

This article is protected by copyright. All rights reserved.

Top Reef Unconformity (TRU)

20 m
250 m

SORBAS BASIN

Upper

Upper

Upper

Lower

Lower

Lower

5.0
Zanclean
U. Evap.

5.5

L. Evap.

Age (Ma)

Accepted Article

CENTRAL
MEDITERRANEAN
BASIN

6.0

6.5

Messinian

THIS STUDY

marls
sapropels
diatomites

7.0
Tortonian

7.5
Azagador
Member

Abad
Member

Yesares
Member

Sorbas
Member

This article is protected by copyright. All rights reserved.

Zorreras
Member

Pliocene

hiatus

14.8

Cantera Mb.

Sorbas

-60
-50
-40
-30
-20
-5
-10

Yesares M

yN
x

0
2

Detr
ital
influ
km
xs

ourc

Cariatiz

b.

~ 120

Top Upper Abad Mb.

Top Reef Unconformity

(a)

Vertical exaggeration x10

10
20
30
40
50
>50
Bathymetry (m)

N
W

CARIATIZ

12.4 km

Accepted Article

12.4 km

CARIATIZ

SORBAS

SORBAS

(c)

(b)
14.8 km
N
E

W
S

G
?

V
V
V

V
V

V
V

Shore line

5 km

Evaporites
Well known
areas (outcrops)

Detrital influx source

Locations
Studied
cross sections

simulated area

This article is protected by copyright. All rights reserved.

(d)

ENVIRONMENTAL PARAMETERS
BATHYMETRY

ENERGY

SALINITY

Low-energy

Normal marine

Low-to-moderate energy

Brackish to hypersaline

High-energy

Normal marine to hypersaline

Outer platform 1
(muddy)

Outer platform
muds (F1)

Subtidal domain (5-40 m)

Outer platform 2
(microbialitic)

MA1 microbialites
(F4)

Supratidal-to-subtidal
domain (0-10s m)

Ooids (F2)

Ooidic/skeletal/
sand shoal

MA2 microbialites
(F5)
Lagoonal muds
(F3)

Lagoon/mud flat

Inter-to-subtidal domain
(0-10 m)
Inter-to-subtidal domain
(0-5 m)

Conglomerates
(F6)

Alluvial fan

Continental domain

Sand (F7)

Top Gypsum

Top TCC

BASE-LEVEL FLUCTUATIONS (m)

simulation window

80
70
60
50
40
30
20
10
0
SB1
SB2
SB3
SB4
SB5
DS1
DS2
DS3
DS4

ne
ari
nm
e
p
O

Basin restriction
+
evaporitic deformation
Re
lag stric
sal oon ted
t la /
ke

(a)

120
100

Hypersaline

80
60

Normal
marine

40

Brackish

20

SALINITY ()

Accepted Article

MODEL FACIES
FIELD CASE STUDY
MODEL FACIES
ENVIRONMENTS
(This study)
(Bourillot et al., 2010a)

SB1
SB2
SB3
SB4
SB5
DS1
DS2
DS3
DS4
Top Gypsum
Top TCC

(b)

This article is protected by copyright. All rights reserved.

Low-to-moderate energy Normal marine to hypersaline

Accepted Article

T = 5.6-5.5 Ma (ante-deformation)

0
2
8
14
20
26
32
38
44
50
56
62
68
74
80
>80

CARIATIZ

T = 5.48 Ma (syn-deformation)

SORBAS

SORBAS

1 km

(a)

CARIATIZ

T = 5.44 Ma (post-deformation)

SORBAS

Evaporite
thickness (m)

(c)

1 km

(b)

T = 5.46 Ma (syn-deformation)

CARIATIZ

CARIATIZ

SORBAS

1 km

(d)

This article is protected by copyright. All rights reserved.

1 km

Carbonate compound (Pmin; Pmed; Pave; Pmax)

Microbial associations 1 (0.08; 0.17; 0.27; 0.88)

Microbial association 2 (0.04; 0.21; 0.46; 3)

Lagoonal muds (0.12; 0.315; 0.33; 0.61)

Outer platform muds (0.12; 0.33; 0.4; 1.01)

Harris, 1979 (Bahamas)


Median
Average
Paull et al. (1992; Storrs Lake, Bahamas)
Median
Average
Planavsky & Ginsburg (2009; Lee Stocking, Bahamas)
Logan et al. (1974; Hamelin Pool, Shark Bay, Australia);
Dill et al. (1986; 1989; Bahamas)
Reid & Browne (1991; Exuma Cays, Bahamas)
Jahnert & Collins (2012; Hamelin Pool, Shark Bay, Australia)
Chivas et al. (1990; Shark Bay, Australia)
Median
Average

Outer platform muds Lagoonal muds

Production rate (m kyr-1)

Ooids

0,1

MA1

0,01

MA2

Accepted Article

Ooids (0.08; 0.93; 1.12; 2.74)

10

Enos (1977; Florida inner shelf)


Neumann & Land (1975; Bight of Abaco, Bahamas)
Stockman et al. (1967; Florida mud bank)
Halley et al. (1977; Belize lagoon)
Median
Average
Enos (1977; Florida outer shelf)
Lisitzin & Rodolfo (1972; Cuban slope)
Blom & Alsop (1988; Temperate shelf; Bass Basin, Australia)
Robbins et al. (1997; Bahamas)
Milliman et al. (1993; Bahamas)
Broecker & Takahashi (1966; Bahamas)
Median
Average

This article is protected by copyright. All rights reserved.

0.2 0.4 0.6 0.8 1

Production Rate (m kyr-1)

1.2 1.4

fair weather
wave base

1.2 1.4

20

10

Pi = Pi,max

. e -b / bi,c

Water depth (m)

30
40
50
60
70

60
70
80
90
100

(a)

60
70

90
100

MA1
microbialites
MA2
microbialites

(c)

(b)

(a) MODEL OUTPUT: SORBAS CROSS-SECTION

50

80
Outer platform
Muds
Lagoonal Muds

photic zone
base

40

Pmax(MA2)

100

50

30

Pmax(MA1)

Ooids

40

Pmax(Om)

90

20

30

Pmax(Lm)

80

1.2 1.4

fair weather
wave base

10

20

Pi = Pi,max

0.2 0.4 0.6 0.8 1

0
fair weather
wave base
Water depth (m)

10

Water depth (m)

0.2 0.4 0.6 0.8 1

Production Rate (m kyr-1)

Pmax(Oo)

Accepted Article

Production Rate (m kyr-1)

(b) MODEL OUTPUT: CARIATIZ CROSS-SECTION

SE

20 m

20 m

100 m

100 m

V.E. x 4

V.E. x 4

(d) FIELD DATA: SORBAS CROSS-SECTION

NW

(c) DETAILED CARIATIZ CROSS-SECTION

?
20 m

20 m

100 m

100 m

V.E. x 4

V.E. x 3

TCC ENVIRONMENTS (FIELD DATA)


Outer platform 1 (muddy)
Ooidic/skeletal/sand shoal
Lagoon/mud flat
Outer platform 2 (microbialitic)
Alluvial fan

FACIES (MODEL OUTPUT)


F1: Outer platform muds
F2: Ooids
F3: Lagoonal muds
F4: MA1 microbialites
F5: MA2 microbialites
F6: Conglomerates
F7: Sands

(e) FIELD DATA: CARIATIZ CROSS-SECTION


SE

NW

SEQUENCE BOUNDARIES
SB1

SB2

SB3

SB4

SB5

V.E. x 3

This article is protected by copyright. All rights reserved.

20 m
100 m

Accepted Article

SORBAS CROSS-SECTION
W

20 m
100 m
W

20 m
100 m
W

20 m
100 m
W

20 m
100 m

CARIATIZ CROSS-SECTION

(a) EXPERIMENT 1: AVERAGE ACCUMULATION RATES


E

F1: 0.4 m kyr-1


F2: 1.12 m kyr-1
F3: 0.33 m kyr-1

F4: 0.27 m kyr-1


F5: 0.46 m kyr-1

SE

Sm= Sfd

F6: 5.5 km3 Myr-1


F7: 1.5 km3 Myr-1

(b) EXPERIMENT 2: MEDIAN ACCUMULATION RATES


E

F1: 0.33 m kyr-1


F2: 0.93 m kyr-1
F3: 0.315 m kyr-1

F4: 0.17 m kyr-1


F5: 0.21 m kyr-1

SE

Sm< Sfd

F6: 5.5 km3 Myr-1


F7: 1.5 km3 Myr-1

(c) EXPERIMENT 3: MINIMAL ACCUMULATION RATES


E

F1: 0.12 m kyr-1


F2: 0.08 m kyr-1
F3: 0.12 m kyr-1

F4: 0.08 m kyr-1


F5: 0.04 m kyr-1

SE

Sm<< Sfd

F6: 5.5 km3 Myr-1


F7: 1.5 km3 Myr-1

(d) EXPERIMENT 4: MAXIMAL ACCUMULATION RATES


E

F1: 1.01 m kyr-1


F2: 2.74 m kyr-1
F3: 0.61 m kyr-1

F4: 0.88 m kyr-1


F5: 3 m kyr-1

LEGEND:
Maximal extension of detrital facies
(tested experiment)

SE

Sm>> Sfd

F6: 5.5 km3 Myr-1


F7: 1.5 km3 Myr-1

NW

20 m
100 m
NW

20 m
100 m
NW

20 m
100 m
NW

20 m
100 m

Maximal extension of DS2 ooidic shoals (Best-fit model; Experiment 1)


Maximal extension of DS2 ooidic shoals
(tested experiment)

This article is protected by copyright. All rights reserved.

Maximal extension of detrital facies


(Best-fit model; Experiment 1)

(c)

transition

basin

(b)

basin

transition

margin
SB5

SB5

basin transition margin

(c)

SB4

100
90
80
70
60
50
40
30
20
10
0

20 m

500 m

500 m

SORBAS CROSS-SECTION

350 m350 m 100 m

CUMULATED THICKNESS (m)

100 m

CUMULATED THICKNESS (m)

Accepted Article

(d)
margin

100
90
80
70
60
50
40
30
20
10
0

20 m

SB4

SB3

SB3

SB2

SB2

SB1
10 m

CARIATIZ CROSS-SECTION (a)

time
DS1
DS2
DS3
DS4
DS1
DS2
minimal rates
median rates
field case study
average rates

margin

SB1

transition

basin

SB5

SB5

SB4

SB4

(d)

SB3
SB3
time
DS3
DS4
maximal rates

This article is protected by copyright. All rights reserved.

SB2

SB2

SB1

SB1

10 m

CARIATIZ CROSS-SECTION

(a) EXPERIMENT 5: WITHOUT REGIONAL SUBSIDENCE

F1: 0.4 m kyr-1


F2: 1.12 m kyr-1
F3: 0.33 m kyr-1

20 m
100 m

SE

F4: 0.27 m kyr-1


F5: 0.46 m kyr-1

NW

F6: 5.5 km3 Myr-1


F7: 1.5 km3 Myr-1

20 m
100 m

100
90
80
70
60
50
40
30
20
10
0

field case study


without subsidence
without deformation

DS1

20 m
100 m

DS2

time
DS4

DS3

CUMULATED THICKNESS (m)

(b) CUMULATED THICKNESS


CUMULATED THICKNESS (m)

Accepted Article

SORBAS CROSS-SECTION

100
90
80
70
60
50
40
30
20
10
0

DS1

DS2

DS3

time
DS4

WITHOUT REGIONAL SUBSIDENCE


(c) EXPERIMENT 6:a)WITHOUT
EVAPORITIC DEFORMATION

F1: 0.4 m kyr-1


F2: 1.12 m kyr-1
F3: 0.33 m kyr-1

SE

F4: 0.27 m kyr-1


F5: 0.46 m kyr-1

LEGEND:
Maximal extension of detrital facies
(tested experiment)

F6: 5.5 km3 Myr-1


F7: 1.5 km3 Myr-1

NW

20 m
100 m

Maximal extension of DS2 ooidic shoals (Best-fit model; Experiment 1)


Maximal extension of DS2 ooidic shoals
(tested experiment)

This article is protected by copyright. All rights reserved.

Maximal extension of detrital facies


(Best-fit model; Experiment 1)

Accepted Article

EXPERIMENT 1: WITH EVAPORITIC DEFORMATION

EXPERIMENT 6: WITHOUT EVAPORITIC DEFORMATION

SB1-MFS3: 5.6 - 5.5

SB1-MFS3: 5.6 - 5.5


0
6
12
18
24
30
36
42
48
54
60
>60

Cariatiz

Sorbas

Cariatiz

Sorbas

thickness (m)

(a)

(c)

1 km

MFS3-SB5: 5.5 - 5.44

MFS3-SB5: 5.5 - 5.44

Cariatiz

Cariatiz

0
4
8
12
16
20
24
28
32
36
40
>40
thickness (m)

Sorbas

(b)

1 km

1 km

Sorbas

(d)

This article is protected by copyright. All rights reserved.

1 km

DS3

DS4

Gross volume: 7.411e9 m

700e6
60
600e6
50
500e6
40

400e6

30

300e6

20

200e6

10

100e6

F7
VOLUME EVOLUTION (m)

BASE-LEVEL FLUCTUATIIONS (m)

70

DS2

F6
F5
F4
F3
F2
F1

5.45
5.44

5.46

5.5

5.49
5.48
5.47

5.53
5.52
5.51

5.54

5.57
5.56
5.55

5.58

5.6
5.59

0
TIME (Ma)

DS2

DS3

DS4
SB4

70
60
1
0.9
40 0.8
0.7
30 0.6
0.5
20 0.4
0.3
10 0.2
0.1
0 0
50

(a)

(b)

(c)

(d)

5.6
5.59
5.58
5.57
5.56
5.55
5.54
5.53
5.52
5.51
5.5
5.49
5.48
5.47
5.46
5.45
5.44

BASE-LEVEL FLUCTUATIONS (m)


NET-TO-GROSS VARIATION (%)

DS1

450e6
400e6
350e6
300e6
250e6
200e6
150e6
100e6
50e6
0

VOLUME OF OOIDS (m)

Sorbas

Cariatiz
N/G = 0.006

MFS3
N/G = 0.530

SB3
N/G = 0.197

MFS2
N/G = 0.723

TIME (Ma)
Sorbas

Sorbas

MFS1
S1
Cariatiz
N/G = 0.414

SB1
1

z
x y
500m

(b)

SB1
1

z
x y
500m

N/G = 0.377

MFS1
N/G = 0.414

N/G = 0.414

SB1

MFS1

SB2
Cariatiz
N/G = 0.377

(c)

This article is protected by copyright. All rights reserved.

z
x y
500m

(d)

SB2

Accepted Article

DS1

Vertical exaggeration x7

DS2

DS3

DS4

50

70e6

40

1
0.9
300.8
0.7
0.6
200.5
0.4
0.3
10
0.2
0.1
0 0

60e6
50e6
40e6
30e6
20e6
10e6
5.45

5.44

5.46

5.47

5.48

5.5

5.49

5.51

5.53

5.52

5.54

5.55

5.57

5.56

5.58

5.6

0
5.59

NET-TO-GROSS VARIATION (%)

BASE-LEVEL FLUCTUATIONS (m)

60

VOLUME OF MA1 MICROBIALITES (m)

TIME (Ma)

(a)

DS2

DS3

DS4

70

160e6

1
0.9
50 0.8
0.7
40
0.6
30 0.5
0.4
20 0.3
0.2
10
0.1
0 0
60

140e6
120e6
100e6
80e6
60e6
40e6

TIME (Ma)

5.44

5.45

5.46

5.47

5.49

5.48

5.5

5.51

5.53

5.52

5.54

5.55

5.56

5.58

5.57

5.6

20e6
5.59

BASE-LEVEL FLUCTUATIONS (m)


NET-TO-GROSS VARIATION (%)

DS1

VOLUME OF MA2 MICROBIALITES (m)

Accepted Article

DS1
70

(b)

This article is protected by copyright. All rights reserved.

DS3

DS4

1
60

0.9
0.8
0.7
0.6

50

0.5

30

0.4
0.3
0.2

20

40

10

0.1

5.45

5.44

5.46

5.47

5.48

5.5

5.49

5.51

5.52

5.53

5.54

5.55

5.57

5.56

5.58

5.6

5.59

NET-TO-GROSS VARIATION (%)

DS2

WATER LEVEL FLUCTUATIONS (m)

Accepted Article

DS1

TIME (Ma)
Sand
Conglomerate
MA1 microbialites

SORBAS CROSS-SECTION

CARIATIZ CROSS-SECTION

(a) EXPERIMENT 7: DURATION OF DEPOSITIONAL SEQUENCES ~20kyr


W

20 m
100 m

F1: 0.4 m kyr-1


F2: 1.12 m kyr-1
F3: 0.33 m kyr-1

SE

F4: 0.27 m kyr-1


F5: 0.46 m kyr-1

NW

F6: 5.5 km3 Myr-1


F7: 1.5 km3 Myr-1

20 m
100 m

(b) EXPERIMENT 8: DURATION OF DEPOSITIONAL SEQUENCES ~20kyr


W

20 m
100 m

F1: 0.8 m kyr-1


F2: 2.24 m kyr-1
F3: 0.66 m kyr-1

SE

F4: 0.54 m kyr-1


F5: 0.92 m kyr-1

LEGEND:
Maximal extension of detrital facies
(tested experiment)

NW

F6: 5.5 km3 Myr-1


F7: 1.5 km3 Myr-1

20 m
100 m

Maximal extension of DS2 ooidic shoals (Best-fit model; Experiment 1)


Maximal extension of DS2 ooidic shoals
(tested experiment)

This article is protected by copyright. All rights reserved.

Maximal extension of detrital facies


(Best-fit model; Experiment 1)

You might also like