Você está na página 1de 18

ELSEVIER Sedimentary Geology 119 (1998) 85–102

Meteoric water participation in the genesis of Jurassic cherts in the


Subbetic of southern Spain — a significant indicator of
penecontemporaneous emergence
M.A. Bustillo a,Ł , A. Delgado b , J. Rey c , P.A. Ruiz-Ortiz d
a
Museo Nacional de Ciencias Naturales, CSIC, C= José Gutiérrez Abascal, 2, 28006-Madrid Spain
b Estación Experimental del Zaidı́n, CSIC, Profesor Albareda, 1, 18008-Granada, Spain
c Departado de Geologı́a, Universidad de Jaén, E.U. Politécnica de Linares, 23700-Linares (Jaén), Spain
d Departado de Geologı́a, Universidad de Jaén, Facultad de Ciencias Experimentales, 23071-Jaén, Spain

Received 14 April 1997; accepted 18 February 1998

Abstract

In the eastern sector of the Betic Cordillera outcrops of the Internal Subbetic display chert lenses in the upper parts
of oolitic limestones (Camarena Formation, Mid-Jurassic). The stratigraphy indicates a sedimentary record related to
synsedimentary tectonics (tilting). The sediments were deposited in different tectonic blocks, at varying depths. The
Camarena Formation is mainly made up of shallow platform oolitic limestones and was deposited on a ‘pelagic’ swell, a
shallow isolated platform that developed far from continental areas. The chert lenses appear in those stratigraphic units
deposited in the lowermost sunken fault blocks which display biosiliceous facies overlying the Camarena Formation.
The chert mainly consists of megaquartz, formed by replacement of the oolitic host rocks without an opaline precursor.
The part of the host that has been most affected by the silicification is the carbonate cement. The petrological features
indicate a slow replacement from solutions which were poor in silica and cations. Comparing the textures and fabric of
the cherts and their host rocks, it is deduced that the silicification took place after compaction, cementation, fracturing
and local dissolution of the oolitic limestones. The stable isotopic data (δ18 O, δ13 C) obtained from the chert, oolitic host
rock and overlain biosiliceous rocks display significant differences. The host rocks (Camarena Formation) and overlying
biosiliceous rocks have isotopic signatures that are typical of marine environments or shallow-marine diagenesis. However,
the chert isotopic values are interpreted to suggest meteoric water participation in the silicification process. The timing of
the chertification is suggested to have been between the Bathonian (beginning of the deposition of the biosiliceous source
facies) and the Aptian (first appearance of this chert as clasts in breccia beds). In palaeogeographic terms the optimum
timing would have been during emergence (i.e. between the Bathonian and Tithonian). The exposed areas could have acted
as recharge areas of meteoric water into the regional aquifer system. The involvement of meteoric water in the silicification
processes gives new support to the hypothesis that pelagic swells were subjected to episodic exposure.  1998 Elsevier
Science B.V. All rights reserved.

Keywords: chert; oolitic limestones; stable isotopes; meteoric water; emergence; Bathonian–Tithonian

Ł Corresponding author. Fax: C34 (91) 564-4740; E-mail: abustillo@mncn.csic.es

0037-0738/98/$19.00  1998 Elsevier Science B.V. All rights reserved.


PII S 0 0 3 7 - 0 7 3 8 ( 9 8 ) 0 0 0 5 0 - 5
86 M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102

1. Introduction with intercalated shallow-marine facies and emer-


gence episodes associated with pelagic swells (e.g.
The Mesozoic sediments of the southern Iberian Vera et al., 1988; Garcı́a-Hernández et al., 1988,
continental margin make up part of the External 1989; Jiménez de Cisneros et al., 1993) (Fig. 1).
Zones of the Betic Cordillera. The palinspastic re- Subaerial erosion, and the reworking of sediments
constructions of these folded chains for the Middle during subsequent transgressive episodes, led to a
Jurassic show two main palaeogeographic domains: poorly preserved record of the emergence phases on
the Prebetic, marine platform to the south of the con- pelagic swells. Calcrete and bauxite deposits, as well
tinental Iberian areas; and the Subbetic, open marine as the morphologies and infills of neptunian dykes,
environments with mainly pelagic sediments, but support the swell–emergence hypothesis (Jiménez de

Fig. 1. (A) Map showing the distribution of units of the Betic Cordillera (modified from Garcı́a-Hernández et al., 1989). (B)
Palaeogeographic reconstruction of the southern continental margin of the Iberian plate for the Middle–Late Jurassic showing the
differentiation into swells and troughs (from Garcı́a-Hernández et al., 1989). The study area is indicated.
M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102 87

Cisneros et al., 1993; Molina et al., 1995; Martı́n- present in the upper part of the limestones of the
Algarra and Vera, 1995). However, this hypothesis of Middle Jurassic, Camarena Formation. Although sili-
the genesis of neptunian dykes has been questioned cification of limestones in the Betic External Zones
by Winterer and Sarti (1994). has been previously studied and attributed to re-
The deposits that accumulated on the innermost placement in a marine environment (Bustillo and
pelagic swells of the Jurassic Subbetic basin and in Ruiz-Ortiz, 1981, 1987; Ruiz-Ortiz et al., 1989), sili-
surrounding pelagic areas are named Internal Sub- cification in the Vélez Blanco area is different in
betic (Fig. 1). In their easternmost outcrops, near appearance. This, along with the existence of sig-
Vélez Blanco (Fig. 2), large lenses of chert are nificant stratigraphic discontinuities related to large

Fig. 2. Geological sketch of study area showing different outcrops in which the oolitic limestones of the Camarena Formation are
silicified.
88 M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102

emergence episodes of pelagic swells (Jiménez de using a vacuum line similar to that described by
Cisneros et al., 1993; Rey, 1993; Molina et al., 1995; Clayton and Mayeda (1963). Isotopic ratios were
Martı́n-Algarra and Vera, 1995), could point towards measured in a Finnigan MAT 251 mass spectrome-
an origin related to subaerial exposure. ter. The experimental error for carbonates (δ13 C and
Using stratigraphic, petrological, and geochemi- δ18 O) was < š0.05‰, using Carrara and EEZ-1
cal data, this study aims to analyse the chertified as internal standards that were previously compared
oolitic limestones of the Camarena Formation and with NBS-18 and NBS-19. For silicates, commercial
to differentiate them from other types of chert in CO2 , that was previously calibrated vs. V-SMOW,
the area. A second goal is to draw conclusions on SLAP and GIPS water was used as internal standard.
the diagenetic environments of these rocks and place It yielded δ18 O D C5:15 š 0:2‰ (V-SMOW) for
them in a palaeogeographic model for the Subbetic NBS-30 (biotite) and a value of δ18 O D C9:7 š 1 for
basin. In this context, the possible relationship of NBS-28 (quartz).
the silicifications with meteoric phreatic waters is
discussed based on isotopic data. 3. Stratigraphy and sedimentological
interpretation
2. Sampling and methodology
Three stratigraphic columns (Fig. 3) show the
Rock samples for the study of cherts from the main features of the Jurassic and Lower Cretaceous
Camarena Formation were collected from five local- stratigraphic units in different parts of the studied
ities (a, a0 , b, c and d, in Fig. 2). Additional sections area. The lower Lias is represented by platform
were studied to investigate the presence of rede- limestones and dolostones of the Gavilán Formation
posited chert clasts from the Camarena Formation in (Lias; pre-Domerian p.p.), followed by the Camarena
Cretaceous or Tertiary rocks (e, f, g, m, in Fig. 2). Formation (Bajocian–Bathonian p.p.). The first im-
Radiolarites, marls and marly limestones bearing ra- portant discontinuity (R.1, Fig. 3) appears at the
diolarians that could have been a local source of contact of these two formations. The associated hia-
silica were sampled at localities n, i and j in Fig. 2. tus spans at least the upper Lias, represented by
Chert associated with Jurassic host rocks, but differ- pelagic deposits in the nearby sectors (Rey, 1993).
ent from the Camarena Formation, were sampled at The overlying Camarena Formation consists
localities k, l and n in Fig. 2 for comparison with mainly of about 300 m of shallow platform oolitic
the mineralogy and textural features of the Camarena limestones, deposited on a ‘pelagic swell’ (Garcı́a-
Formation. Hernández et al., 1988; Rey, 1993). Texturally, the
The mineralogy of the samples was determined rocks are oolitic=peloidal grainstones=packstones.
by X-ray diffraction using a Philips PV1710 diffrac- The stratigraphic differences among the sections start
tometer. The petrographic observations were made at the top of the Camarena Formation. The sedimen-
by transmitted light microscopy. Major element ox- tary record is continuous in the southwestern sectors,
ides (Table 1) were determined using atomic absorp- such as the Gabar, Rambla Seca, Piedras Berme-
tion spectrometry and X-ray fluorescence, depending jas, Cerro Gordo and Torre Charcón (Fig. 2) (see
on the elements concerned. Rambla Seca–Piedras Bermejas column, in Fig. 3).
Isotope measurements (Table 2) were carried out Peloid-rich layers containing organisms of pelagic
in the Stable Isotope Laboratory at the Estación affinity (abundant ‘filaments’, thin-shelled bivalves,
Experimental del Zaidı́n (CSIC, Granada). Carbon and belemnites) are intercalated with oolitic lime-
dioxide was evolved from the carbonates using 100% stones in the upper part of the Camarena Formation.
phosphoric acid for 12 h in a thermostatic bath at These facies have been recently assigned to the
25ºC (McCrea, 1950). All samples were treated with early-middle Bathonian (Aguado and Rey, 1996). It
cold 1 : 5 HCl to remove carbonates. Oxygen was is in the oolitic grainstone layers from the upper part
extracted from the cherts by reaction with ClF3 , us- of the Camarena Formation that the cherts studied
ing the method described by Borthwick and Harmon in this paper occur. The Camarena Formation in the
(1982). It was then purified and converted to CO2 southwestern sector is topped by the next sedimen-
M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102 89

Fig. 3. Stratigraphy columns showing regional facies changes.

tary break (R.2, Fig. 3), followed by 15 to 30 m of and by radiolarian-rich marls and marly limestones
rhythmically alternating layers of marls and marly (Charco Formation). The R.3 can probably be cor-
limestones rich in thin-shelled bivalves (‘filaments’) related with the break at the end of the Bathonian:
and radiolarians. Aguado and Rey (1996) have at- one of the most important sedimentary breaks in the
tributed these rocks to the early-middle Bathonian. Betic External Zones (e.g. Vera et al., 1988; Garcı́a-
Intercalated with these rhythmites are calcarenite Hernández et al., 1989). The radiolarian facies of the
layers with oolites in their lower part and oriented Charco Formation were most likely deposited during
‘filaments’ in their upper part. They have been inter- the early Callovian (Aguado and Rey, 1996), and are
preted as turbidites (Rey, 1993). The rhythmites are overlain by marly nodular limestones of Tithonian
followed by another sedimentary break (R.3, Fig. 3) age (Fig. 3).
90 M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102

The chert appears only in this type of sections, in emergences in the Almirez sector, and simultaneous
the highest part or at the top of the oolitic limestones, deepening in the Rambla Seca–Piedras Bermejas.
just below the R.2 break, partially replacing the The Gigante, and other intermediate sectors, were
carbonate and always occurring in oolitic grainstone likely places of shallow-marine sedimentation, pri-
layers. It forms brownish and wine-red lens-shaped marily shaped by storms.
bodies that tend to follow the bedding, preserving At the beginning of the Cretaceous, mainly in the
the original structure of the rock (bedding, parallel late Berriasian, the entire region was characterized
and cross-lamination, etc.). The chert bodies may by deposition of alternating pelagic marls and marly
reach up to 3 m in thickness and are generally less limestones (Carretero Formation). These lithologies
than 10 m long. Their surface is frequently vertically are highly folded and badly exposed, making an ex-
grooved, giving the lenses a columnar appearance. act calculation of thickness difficult, but they never
In the Gigante sector (Fig. 3), the oolitic lime- seem to be more than 100 m thick. From the end
stones of the Camarena Formation are capped by of the Barremian until the early Cenomanian, the
an erosive surface (R.2) and overlain unconformably dark-greenish marls of the Fardes Formation were
by packstones with hummocky bedding rich in ‘fila- deposited. Their thickness is variable, but never more
ments’ and peloids. These facies are coeval with the than 300 m. Associated with these facies, and along a
aforementioned rhythmites (with turbidite intercala- fault escarpment (Aguado et al., 1991), are turbidite
tions) and are interpreted as possible storm deposits. intercalations, consisting of breccia and calcarenite
In fact, the microfacies of the beds interpreted as beds made up of Camarena Formation debris. Com-
tempestites in this sector are similar to the turbidite mencing with the top of the Aptian, these deposits
microfacies in the Rambla Seca–Piedras Bermejas contain chert clasts of silicified limestones studied
sector. These lithologies are capped again by break in this work (Fig. 3). The abundance of silicification
at the end of the Bathonian (R.3, Fig. 3), overlain by in the Cretaceous turbidites has led to development
a breccia with nodular limestone clasts of probable of criteria that enable us to recognize breccia clasts
Tithonian age. that originated from erosion of silicified Camarena
In the Almirez sector (Figs. 2 and 3), the up- oolitic limestones. These criteria are: (1) the sim-
per part of the Camarena Formation consists of ilarity in appearance between breccia chert clasts
cross-bedded crinoidal limestones, which overlie and chert lenses in the Camarena Formation; (2) the
oolitic limestones. The top of the formation shows morphology of the clasts that have sharp boundaries
an irregular morphology of karstic appearance, fos- and are not rounded (Fig. 4); (3) the orientation of
silized by a laterally discontinuous bed of pro- chert clasts parallel to cross-bedding (Fig. 4); (4)
toglobigerinid-rich pelagic limestones, Callovian– the divergence in bedding between silicified lime-
Oxfordian in age. Therefore, a stratigraphic disconti-
nuity is placed at the top of the Camarena Formation,
and the hiatus includes the Bathonian and Callovian
pro parte (R.2 C R.3). Overlaying the bed of pelagic
limestones, or directly on the Camarena Formation,
nodular limestones of Ammonı́tico Rosso facies (Up-
per Ammonı́tico Rosso Formation), lower Tithonian
in age, occur. The presence of faults, N-100-E and
N-10-E in strike, in the Camarena Formation which
do not affect the Ammonı́tico Rosso Formation, in-
dicates faulting episodes during the late Bathonian–
earliest Tithonian.
The differences in the stratigraphic record in the
studied sectors have been interpreted by Rey (1993,
Fig. 4. Chert clast oriented parallel to the cross-bedding which is
1995) as a result of synsedimentary tilting. This marked by the holes (moulds of probable marly chips) and other
would have given rise to shallowing, and possible breccia limestone clasts.
M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102 91

stone boulders and the breccia; and (5) petrographic crystals in the mosaic are variable in shape and size
differences. (up to 800 µm), and in some cases have marked cleav-
age. The contacts between the allochems are point,
4. Petrology tangential and sometimes concavo-convex, depend-
ing on the samples. Some ooids display spalled cor-
Although the petrological study is focussed on tices that are healed with mosaic cements. Therefore,
the cherts formed in the oolitic limestone of the mosaic cements are formed after the physical com-
Camarena Formation, other types of chert found paction (Choquette and James, 1987).
in the area are briefly described. The purpose of The most common form of quartz in the chert
these descriptions is to show that the cherts of the is megaquartz (crystal size from 20 to 220 µm)
Camarena Formation are different from other cherts (Fig. 5b,c) that occurs as mosaics or forms pal-
of the area, and therefore easily distinguishable from isades around the allochems (palisadic quartz). In
other chert clasts in Cretaceous turbidites. some places the megaquartz is zoned, with succes-
sive idiomorphic growth stages (Fig. 5d). Locally,
4.1. Description and in very small amounts, there is fibrous quartz
(length-fast chalcedony) and microcrystalline quartz
X-ray diffraction patterns of the cherts from the (<20 µm in size). The samples with fibrous and
Camarena Formation indicate that they are com- microcrystalline quartz are those that have the low-
posed of quartz and variable amounts of low Mg est crystallinity index values. There are iron oxides
calcite corresponding to relicts of the host rock. The on the surface of all the allochems and when the
crystallinity index of the quartz (Murata and Nor- amounts of these oxides are high (>1%), they are
man, 1976) falls in the range of 7 to 8.5. The bulk also found between the quartz crystals.
chemical composition of the cherts (Table 1) cor- The part of the host rock which is most affected
roborates the mineralogical data. Clays are absent, by the silicification corresponds to what was for-
and despite the reddish-brown colour of the chert the merly the cement (Fig. 5b). Where fibrous quartz
total Fe2 O3 is not high in most samples (Table 1). is found, it comprises the first layers around the
The host rock is an oolitic grainstone (Fig. 5a). The allochems, followed by the palisadic megaquartz
allochems are mainly ooids but there are also echi- and finally by the mosaic megaquartz which oc-
noidal fragments, oncoids, pelloids, coated grains, curs farthest away from the allochems. The ooids
coral fragments and some benthonic foraminifera and and other allochems, when affected by silicification,
bivalves. The different types of calcite cement are: show signs of replacement in their outer layers and
acicular drusy cement, syntaxial overgrowth on echi- sometimes towards the interior (called ‘islands of
noidal fragments and mosaic cements. The calcite advance’, Grigor’ev, 1965) (Fig. 5c).

Table 1
Maxima and minima values (‰) for major elements in cherts from the Camarena Formation

Section a, 5 samples Section a0 , 2 samples Section b, 3 samples Section c, 4 samples Section d, 7 samples
max. min. max. min. max. min. max. min. max. min.
SiO2 49.23 36.57 38.71 34.74 88.90 53.20 61.92 38.07 70.00 32.91
Al2 O3 0.14 0.00 0.06 0.00 0.62 0.00 0.16 0.00 0.16 0.01
Fe2 O3 0.11 0.00 1.06 0.07 2.43 0.10 0.49 0.06 0.29 0.02
(total)
MnO 0.01 0.00 0.03 0.00 0.05 0.01 0.01 0.01 0.01 0.01
MgO 0.22 0.19 0.17 0.01 0.23 0.00 0.24 0.14 0.38 0.11
CaO 36.50 27.89 36.50 34.00 24.80 4.07 35.80 22.87 45.50 17.44
Na2 O 1.21 0.09 0.08 0.06 0.54 0.08 0.42 0.02 0.17 0.02
K2 O 0.01 0.00 0.07 0.01 0.06 0.01 0.01 0.00 0.01 0.01
P2 O5 0.02 0.00 0.01 0.00 0.03 0.00 0.03 0.01 0.03 0.01
92 M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102
M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102 93

In some samples from sites b (Almoyas), c (Torre radiolarites easy to distinguish from the cherts in-
Charcón) and d (NE Cerro Gordo) (Fig. 2), there is cluded in the oolitic limestones of the Camarena
a different type of chert that is the result of silicifi- Formation. Finally, the chert clasts from the Ca-
cation of a different host rock (Fig. 5e,f). This host marena Formation are also distinguished from the in
rock is a detrital calcarenite that fills centimetre-sized situ silicification of the Cretaceous oolitic turbidites
cracks and pockets within the oolitic limestone. The because: (1) the ooids do not continue across the
clasts are fragments of the oolitic limestone itself, chert–limestone boundary (Fig. 6b); and (2) there are
ooids and other allochems from the oolitic lime- significant differences in the framework and contacts
stones, and pieces of large calcite crystals (up to 350 between the ooids in the chert clast and those in the
µm in size) (Fig. 5e). The matrix comprises micrite oolitic turbidite. The contacts are point or tangential
and smaller fragments of the aforementioned clasts. in the chert clast, whereas they are interpenetrating
The detrital calcarenite was formed by the disag- in the turbidite (Fig. 6b).
gregation of the oolitic limestone. The occurrence
of clasts of large calcite crystals indicates that this 4.2. Interpretation
disaggregation took place after the later cementation
of the oolitic limestone. The most frequent fabric between the allochems
In this new type of chert, the silicified oolitic is palisadic megaquartz followed by mosaic quartz.
limestone clasts have the same quartz textures as the This fabric is common when silica fills voids, but
original silicified oolitic limestone, and these tex- may also occur in replacement of limestones by
tures are conditioned by the allochems and cements quartz (Wilson, 1966). In the chert studied, the fol-
contained in the fragment. The calcite crystal clasts lowing points indicate that the textures were pro-
are silicified as individual quartz crystals, main- duced as a result of replacement.
taining their outlines, and therefore the megaquartz (1) Direct observation of the replacement of cal-
acquires a detrital appearance varying greatly in cite megacrystals by megaquartz in partially replaced
size. Micrite tends to be unaffected, remaining as a areas of the limestone (Fig. 5b).
residue. (2) The existence of small calcite crystal ghosts
Most cherts in the Bathonian to Aptian marine within some of the megaquartz crystals (Fig. 6c).
limestones consist of microcrystalline mosaic quartz, (3) In a cementation process, crystal size com-
sometimes with minor amounts of opal-CT iden- monly increases towards the centre of the void. In
tified by XRD. Radiolarians and, in some cases, the samples studied there is no such increase in the
sponge spicules are observed. In the Aptian oolitic size of the megaquartz crystals away from the al-
turbidites (Rambla Seca–Piedras Bermejas section; lochems (Fig. 6d), although without other evidence
Fig. 3), the in situ silicification is different and this feature alone would not prove a replacement
length-slow chalcedony appears (Fig. 6a). On the origin.
other hand, the radiolarites of the Charco Formation Therefore, although in certain areas there is ev-
are comprised of microcrystalline quartz with many idence of local cementation processes, replacement
radiolarian ghosts, and in many places opal-CT ap- is the general mechanism invoked for the genesis
pears as the major mineral. These petrological and of these cherts which enclose many relicts of the
mineralogical characteristics make these cherts and host rock. Most of the relicts are allochems, but in

Fig. 5. (a) General view of an oolitic grainstone of the Camarena Formation. In the central part, a ooid with spalled cortex healed
with mosaic cement is observed (arrow). Crossed nicols. (b) Chert from Camarena Formation. The megaquartz replaces the mosaic
cement of the host rock (S D sparite). Crossed nicols. (c) The megaquartz occurs as mosaic or forming palisades around the allochems.
The allochems are replaced in their outer layers and in places (islands of advance) in the interior. Crossed nicols. (d) Close-up view
of megaquartz displaying successive idiomorphic growth stages. Plane-polarized light. (e) Contact between the oolitic limestone of the
Camarena Formation (lower part) and the detrital calcarenite (upper part) formed by disaggregation of the oolitic limestones. Some
clasts are fragments of the large calcite crystals derived from cements. Micrite occurs between the clasts. Plane-polarized light. (f) Chert
formed by the silicification of a piece of rock similar to that described in Fig. 6e. Plane-polarized light.
94 M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102
M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102 95

some places large calcite crystals (sparite) (Fig. 5b) line tests until the dissolved silica rises above 6 mg=l.
and acicular drusy calcite (Fig. 6d) are also found. At that point, quartz can nucleate and grow directly.
In the process of silicification, quartz could have In places where microcrystalline quartz and length-
been formed directly without an opaline precursor. fast chalcedony where found, the silica concentration
This would be indicated by the lack of lepisphere of the solution may have been higher.
ghosts and the coarse quartz crystal size (Maliva and Finally, silicification is not an early process. Com-
Siever, 1988). This has led us to denote the process paring the textures and structures of the cherts and
as chertification. their host rocks, and all the features of the host rock
The general fabric of the chert resulted from affected by chertification, it can be deduced that the
the allochems of the host rock being much more silicification process was subsequent to:
difficult to replace than the large calcite crystals (1) The physical compaction of the host rock.
of the cement (sparite). A probable explanation of There are no important differences between the
this is that the silica solutions moved preferentially shape and type of contact of the ooids included
along the contacts between these calcite crystals or in the chert and in the host rocks. Concavo-convex
through their structural defects. At the beginning contacts and spalled ooid cortices (Fig. 6e) are also
of the silicification process, ooids and allochems found in the chert.
appear to resist the penetration by silica solutions (2) Complete cementation of the oolitic limestone.
forming a framework from which the silicification This is deduced from the presence of acicular drusy
continues. The quartz textures were influenced in cement around the allochems, syntaxial overgrowth
part by the textures of the calcite cements that have on echinoid fragments, and large sparite crystals in
been replaced. As a result, the acicular drusy calcite the chert. In addition, sparite found in the tectonic
cement around the allochems could have favoured fractures has also been replaced indicating that the
the formation of palisadic quartz, and less commonly silicification was later (Fig. 6f).
length-fast chalcedony, while the sparite favoured the (3) Disaggregation processes of the oolitic lime-
formation of mosaics of megaquartz. When the sili- stones by fracturing and=or dissolution, that pro-
cification advanced, the ooids were also marginally duces fill-in of detrital calcarenites in cracks and
replaced. At first, the outer lamellae of the ooid pockets. The detrital calcarenites were also replaced
cortex were replaced by quartz. Later, as the silici- by quartz (Fig. 5e,f).
fication progressed inwards, ‘islands’ of advancing
megaquartz were formed. As silicification continued, 5. Isotopic composition
coalescing ‘islands of advance’ gave rise to a frame
of megaquartz crystals in the oolite. According to Quartz in the oolitic limestones of the Camarena
Hesse (1987), the inward replacement prevails when Formation has a wide range of isotopic values, from
the ooids have nuclei other than siliciclastic grains C29.1‰ to C35.6‰ (V-SMOW). The δ18 O cal-
(quartz and feldspar). cite values range between -1.59‰ and C0.99‰
The fact that most of the chert consists of (PDB), and the δ13 C varies between C0.99‰ and
megaquartz and that it occasionally has stages of C1.19‰. The chert clasts of the Camarena For-
idiomorphic growth indicates slow replacement by mation redeposited in the Cretaceous show δ18 O
solutions low in silica and cation concentration (Mil- values for quartz ranging from C31.3‰ to C34.0‰
lot, 1970). According to Knauth (1992), groundwater (V-SMOW), values similar to those of the Camarena
moving through a carbonate unit, may dissolve opa- Formation. The δ18 O and δ13 C values of calcite are

Fig. 6. (a) Aptian turbidite. Large clast in the lower part at right is a radiolarian-rich clast and the oolitic matrix is silicified in situ mainly
by chalcedony. Crossed nicols. (b) Right part is a chert clast from Camarena Formation. Left part is matrix of Aptian turbidite. The ooids
do not cross the chert–limestone boundary. The type of contact among the ooids differs in the clast and the matrix. Plane-polarized light.
(c) Crystal ghosts of calcite (arrow) in megaquartz crystals. Plane-polarized light. (d) Acicular drusy calcite around ooids included in
the chert. Crossed nicols. (e) Ooids with spalled cortices occur in the chert. Crossed nicols. (f) Tectonic fracture filled with sparite and
replaced by megaquartz at the same time as the cement of the host rock. Crossed nicols.
96 M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102

Table 2
Oxygen and carbon isotopic compositions of silica and coexisting carbonates

Samples Code Section Mineralogy Calcite Silica


δ18 O (V-SMOW)‰
δ13 C (PDB)‰ δ18 O (PDB)‰
1 VCA-1 a Ca 55, Q 45 C1.71 0.52 C31.7
2 VCA-2 a Ca 55, Q 45 C1.19 0.55 C29.1
3 VCA-3 a Ca 65, Q 35 C1.15 1.36 C32.3
4 VCA-5 a Ca 60, Q 40 C1.74 0.67 C31.1
5 VCA-8 a Ca 65, Q 35 C1.58 0.19 C31.9
6 VCA-9 a0 Ca 60, Q 30 C1.38 0.05 C32.6
7 VCA-10 a0 Ca 95, Q 5 C1.85 0.03 C33.8
8 VCA-11 a0 Ca 100 C1.29 C0.11 –
9 VAL-6 b Ca 35, Q 65 C2.37 1.30 C33.9
10 VAL-7 b Ca 50, Q 50 C1.46 1.45 C35.2
11 VAL-8 b Ca 15, Q 85 C2.44 0.74 C35.6
12 VTC-1a c Ca 100 C2.10 0.77 –
13 VTC-1b c Ca 40, Q 60 C2.32 0.74 C33.0
14 VTC-2a c Ca 100 C1.83 1.16 –
15 VTC-2b c Ca 100 C2.08 0.71 –
16 VTC-2c c Ca 100 C1.98 0.82 –
17 VTC-3 c Ca 50, Q 50 C1.97 0.45 C30.1
18 VG-1 d Ca 55, Q 45 C1.41 1.49 C33.1
19 VG-2 d Ca 60, Q 40 C1.41 1.18 C29.9
20 VG-3 d Ca 70, Q 30 C2.34 C0.38 C30.8
21 VG-5 d Ca 100, Q Tr C1.98 0.30 –
22 VG-6 d Ca 30, Q 70 C1.19 1.59 C30.2
23 VG-7 d Ca 45, Q 55 C2.53 C0.99 C34.1
24 VB-3 j Q 100, Ca Tr C1.46 0.34 C33.3
25 VB-5 j Q 100, Ca Tr C1.35 0.26 C32.9

26 R-21 a Ca 20, Op CT 40, Q 40 C1.74 1.39 C37.2


27 R-10 i 40 C0.36 4.07 C36.4
28 R-11b i Q 100 C0.35 3.75 C37.0
29 R-25 j Ca 5, Q 95 C1.83 1.10 C36.3
30 R-26 j Op CT 65, Q 15, Ca 20 C2.14 0.68 C35.7
Op CT 80, Q 15, Ca 5
31 VVA-1 f Ca 25, Q 75 C1.79 0.09 C32.9
32 R-18 m Ca 30, Q 70 C2.19 0.20 C32.2
33 R-14 m Ca 25, Q 75 C0.71 1.55 C31.3
34 R-15 m Q 100 – – C34.0

Samples: 1 to 25, cherts and host rocks of the Camarena Formation; 26 to 30, radiolarian bearing limestones and radiolarites of the
Charco Formation; 31 to 34, chert clasts from the Camarena Formation in the Cretaceous turbidites. Ca D calcite, Q D quartz, Op D
opal CT, Tr D traces.

the same as those of the Camarena Formation (Ta- the silica–water system with a temperature range of
ble 2; Figs. 7 and 8). 200–500ºC. Its extrapolation to lower temperatures
In the Charco Formation, the siliceous phases seems to be justified, as its yields similar values
have high δ18 O values (C35.7‰ to C37.2‰; to those of Kita et al. (1985) for amorphous silica
V-SMOW); the δ18 O of the calcite ranging from and Labeyrie (1974) for biogenic silica for the 34–
-0.68‰ to -4.07‰ (PDB) and the δ13 C values rang- 93ºC and 0–30ºC, respectively. For the calcite–water
ing between C0.36‰ and C2.14‰. system we have used the equation proposed by An-
The interpretation of the data has been carried derson and Arthur (1983) on the basis of known
out using the Clayton et al. (1972) equation for isotopic fractionation factors (Epstein et al., 1953;
M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102 97

Fig. 8. Carbon and oxygen isotopic composition of analysed


calcite. The encircled points represent calcite precipitated in a
shallow-marine diagenetic environment.

Fig. 7. Silica δ18 O values vs. the δ18 O of calcite in samples et al., 1974; Price et al., 1995). For this study we
of silicified limestones (cherts) of the Camarena Formation, have considered meteoric water with a δ18 O value of
radiolarian-rich facies of the Charco Formation and clasts of sili- -4‰ to -8‰ (SMOW), which would be typical of
cified limestones (cherts) of the Camarena Formation, including intermediate latitudes (Rozanski et al., 1993).
breccias that were redeposited in the Cretaceous. The shaded
The quartz δ18 O values of the Camarena Forma-
rectangle indicates the theoretical isotopic compositions of sam-
ples containing calcite and quartz stabilized in equilibrium with tion are lower than those typical of marine silica that
normal marine water at surface temperature. The horizontal and formed under surface temperatures, which would be
vertical arrows represent the theoretical isotopic composition of close to C40‰ (Labeyrie, 1974; Kolodny and Ep-
quartz and calcite, respectively, in equilibrium with meteoric wa- stein, 1976; Matheney and Knauth, 1993). Lowering
ter (δ18 O of -8‰ to -4‰ vs. SMOW) at surface temperature
of δ18 O values, due to the presence of clays or detrital
(25–15ºC). The temperatures and=or isotopic compositions of
the waters were calculated using the equations cited in the text. silicates, cannot be considered, since none of these
Note that isotopic values of the quartz in the silicified samples minerals have been detected. The δ18 O values of
from the Camarena Formation are widely dispersed, with most the Camarena Formation resemble those for quartz
of them lying in the meteoric water field. In contrast, the δ18 O formed at relatively higher temperatures (Kolodny
values of the calcite, which correspond to host rock relics in the
and Epstein, 1976; Murata et al., 1977; Pisciotto,
chert, are more closely grouped and are located within the ma-
rine water field. In addition, the quartz from the radiolarian-rich 1981) or from meteoric water (Gao and Land, 1991;
facies of the Charco Formation has more uniform values typical Knauth, 1994). The calcite δ18 O values are near
of marine water. equilibrium values for marine water at surface (shelf
seafloor) temperatures of 7ºC to 16ºC. The δ18 O val-
Craig, 1965; O’Neil et al., 1969). We assume that the ues for calcite and quartz of the silicified Camarena
δ18 O values for Jurassic seawater on a planet free of Formation oolitic limestones (Fig. 7), indicate that
polar ice caps would be close to -1.2‰ (Shackleton these two mineral phases formed under different
and Kennett, 1975). However, during burial diagen- diagenetic conditions. The relatively heavy carbon
esis of carbonate sediments, interstitial waters may isotope values (C1.19 < δ13 C < C2.53‰) (Fig. 8)
acquire considerably higher δ18 O values (Lawrence are typical of marine bicarbonate and was likely in-
et al., 1977; Lawrence, 1989; Moore, 1989). Mete- herited from precursor marine carbonates (Keith and
oric water has a wide range of δ18 O values, from Weber, 1964; Anderson and Arthur, 1983; Budd and
positive values in low-latitude desert areas to val- Land, 1990; Holail, 1993).
ues lower than -40‰ in polar regions (Craig, 1961). The similarity of isotopic composition for cherts
During the Jurassic and the Cretaceous this area from the Camarena Formation and for the chert
could have been close to 20º to 30º latitude (Drewry clasts from Cretaceous redeposits (Fig. 7), supports
98 M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102

our interpretation that these clasts originated from with a meteoric component, or from marine water at
the erosion of the Camarena Formation. temperatures between 20º and 45ºC (Fig. 7). The sec-
The δ18 O values of the siliceous phases from the ond possibility is rejected for the following reasons.
Charco Formation are typical of cherts from marine (1) Temperatures as high as 45ºC, with a normal
environments that were subjected to shallow-burial geothermal gradient in a typical continental mar-
diagenesis (cf. Knauth, 1992) (Fig. 7), but the δ18 O gin basin (20ºC=km; Suchecki and Hubert, 1984),
values of two analysed calcites are more negative, would require a minimum burial depth of 1500
indicating their diagenetic stabilization in meteoric m (Fig. 9). Even considering gradients as high as
water or at higher temperatures. The δ13 C values, on 35ºC=km (Choquette and James, 1987), a depth of
the other hand, are normal marine (Fig. 8), but mete- more than 1000 m would be required. Not even in
oric diagenetic systems at high rock=water ratios also the case of the most complete and thickest sequences
retain the precursor δ13 C (Allen and Matthews, 1982; studied would the sediments accumulated during the
Lohmann, 1988; Budd and Land, 1990; Holail, 1993). Bathonian–Aptian have been buried to those depths.
In addition, based merely on a geothermal gradient, it
6. Discussion is impossible to explain the temperature differences
(derived from the isotopic composition) at which the
6.1. Silica source different samples of the same formation (Camarena
Formation) have been generated. Furthermore, con-
No direct evidence was found for an intraforma- sidering that the Charco Formation would have been
tional silica source in the cherts or in the oolitic lime- buried to depths only 20–30 m less than the Ca-
stones of the Camarena Formation. This is not the marena Formation, such a small difference would not
case for the rocks of the same age and facies from be able to account for the observed δ18 O gradient.
other zones of the Betic Cordillera (Bustillo and (2) An alternative explanation for high tempera-
Ruiz-Ortiz, 1981) nor for most Mesozoic platform tures could be the presence of thermal water, rising
sedimentary rocks with cherts (Maliva and Siever, along faults. This is not likely because: (a) there are
1989). The close link between facies rich in siliceous no indications of hydrothermal activity in the area;
organisms (Charco Formation and the underlying (b) the indicated temperature gradient (from 20ºC to
rhythmite) and the cherts in the same stratigraphic 45ºC) is not reflected in quartz textures, the latter
sections (Fig. 3) indicates silica derivation from this being very homogeneous.
biogenic source. We conclude, therefore, that the most likely expla-
nations for the genesis of the cherts is that quartz was
6.2. Timing of chertification precipitated from waters with a meteoric component
(Knauth and Epstein, 1976; Meyers and James, 1978;
The timing of the chertification process is de- Kolodny et al., 1980), and such a scenario could also
duced by: (1) the commencement of deposition of explain the great dispersion of the δ18 O values in
silica-source facies in the Bathonian; (2) the ap- the cherts of the Camarena Formation. This propo-
pearance of chertified limestones as clasts in Aptian sition is similar to that of Whittle and Alsharnhan
redeposits; and (3) the timing of the emergence pe- (1994), with the silicification of platform limestones
riod for this sector as Bathonian–earliest Tithonian. by meteoric water and after intermediate burial.
This could produce the recharge areas for meteoric
water in the regional aquifer system. Thus, the most 6.4. Palaeogeographic model
likely timing of chertification is Bathonian–earliest
Tithonian. The stratigraphic sections presented in this paper
(Fig. 3) represent sedimentary records of different
6.3. Chertification environment tectonic blocks and different bathymetry, particu-
larly visible for the interval between breaks R.2 and
Isotopic data show that the quartz in the chert could R.3. The profiles of emerged blocks (Almirez-type)
have precipitated at surface temperatures from waters pass laterally to profiles typical of storm-dominated
M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102 99

Fig. 9. Theoretical burial depths for precipitation of silica and calcite based on the δ18 O composition that assumes a geothermal gradient
of 20ºC=km, typical of an average continental margin basin (Suchecki and Hubert, 1984). The burial depths are calculated by assuming
starting temperatures at the Jurassic shelf seafloor of 7ºC and a value of δ18 O of -1.2‰ (SMOW) for water. Shifts in the δ18 O of pore
water in sediment composed of calcium carbonate of 2‰ to 3‰ are possible with increasing depth (Lawrence, 1989), making the dotted
line more probable for depths >1000 m.

platform environments (Gigante-type) and finally to the marine phreatic, is also a preferential zone for
profiles with turbidite facies, such as Rambla Seca– silicification since it is an apt geochemical environ-
Piedras Bermejas. The subaerial exposure would ment for calcite dissolution and the simultaneous
have essentially affected the Almirez-type blocks, precipitation of quartz (Knauth, 1979).
although during the R.2 break, the emergence of
the Gigante-type blocks cannot be ruled out. The 7. Conclusions
episodes of block emergence coincided with the
deposition of the rhythmite and the radiolarite fa- (1) Dogger oolitic limestones of the Camarena
cies (Charco Formation) in the most sunken fault Formation, Internal Subbetic, eastern sector of the
blocks. The emerged sectors could have thus acted Betic Cordillera, were replaced by quartz, giving rise
as recharge areas for meteoric water that was then to chert bodies in the upper part or at the top of the
laterally fed into a confined aquifer of the lowermost formation.
blocks, the aquifer being sealed by impermeable fa- (2) Silicification was a slow process produced
cies above the oolitic limestones (Fig. 10). Confined by solutions low in silica and cation concentration,
aquifers can act as efficient conductors for the dis- which directly precipitated megaquartz, without an
tribution of meteoric water at depth and under the opaline precursor.
continental platform (Moore, 1989; Knauth, 1994), (3) The δ18 O values of quartz in the cherts of
aided in this case by the existence of fractures that the Camarena Formation, suggest a contribution of
increased rock porosity and facilitated communica- meteoric water during chertification.
tion (Fig. 10). The zone where the waters mixed, (4) We propose a chertification environment with
at the contact between the meteoric phreatic and a confined aquifer under submerged marine areas,
100 M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102

Fig. 10. Sketch of the proposed genetic model. Not to scale. Tilted blocks were related to listric faults. The position and names of the
three stratigraphic type-sections are indicated. The Almirez area acted as a recharge zone of meteoric water which fed a continental
phreatic zone and a mixing water zone. Sunken fault blocks (Gigante) overlain by tempestites could also have been temporarily emerged.
In the lowermost sunken fault blocks (R. Seca–P. Bermejas), more continuous sedimentation occurred, with turbidites followed by
radiolarite facies. R2 and R3 refer to the sedimentary breaks noted in Fig. 2 and in the text.

that received meteoric water from the emerged parts Group RNM-200 of the Junta de Andalucı́a. We spe-
of tilted fault blocks. Faults and fractures facilitated cially thank the Sedimentary Geology reviewers R.
water circulation. Hesse and J. Veizer and editor, B. Sellwood, for their
(5) Both the biosiliceous facies of the Charco For- helpful suggestions and comments. We are grate-
mation and the underlying radiolarian-bearing marls ful to M. Castillejo, A. Viloria and M.C. Sendra of
and rhythmites were the source of silica. These sed- the administrative and technical staff of the Museo
iments accumulated on the lowermost sunken fault Nacional de Ciencias Naturales for their help with
blocks. various aspects of this work. Thanks to Matthew
(6) Silicification took place between the Bathonian Harffy for his help in correcting the English version
and the earliest Tithonian, coincident with platform of the manuscript.
sedimentation at higher fault blocks and=or with a
phase or phases of their subaerial exposure. Chert References
clasts from erosion of chert lenses in the Camarena
Formation are included in the Aptian breccia thus Aguado, R., Rey, J., 1996. Consideraciones sobre la edad del
setting an upper limit on the timing of silicification. techo de las calizas oolı́ticas del Jurásico medio del Subbético
(7) The involvement of meteoric waters in the sili- Interno oriental (Cordilleras Béticas). Geogaceta 20, 35–38.
cification process and their connection with subaerial Aguado, R., O’Dogherty, L., Rey, J., Vera, J.A., 1991. Turbid-
itas calcáreas del Cretácico al norte de Vélez Blanco (Zona
exposure comprise a new argument in favour of the Subbética): Biostratigrafı́a y génesis. Rev. Soc. Geol. Esp. 4,
existence of phases of emergence during the Jurassic 271–304.
of specific areas of the continental margin located in Allen, J.R., Matthews, R.K., 1982. Isotope signatures associated
the pelagic realm. with early meteoric diagenesis. Sedimentology 29, 797–817.
Anderson, T.F., Arthur, M.A., 1983. Stable isotopes of oxy-
gen and carbon and their application to sedimentologic
Acknowledgements and palaeoenvironment problems. In: Arthur, T.F., Anderson,
M.A., Veizer, J., Land, L.S. (Eds.), Stable Isotopes in Sedi-
This work was carried out with the support of re- mentary Geology. Short Course Notes, Soc. Econ. Paleontol.
search projects PB93-1150-C02-02 of the DGICYT, Mineral. 10, 1–151.
PB95-0106-C02-01 of the DGES, and the Research Borthwick, J., Harmon, R., 1982. A note regarding CIF, as an
M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102 101

alternative to BrF5 for oxygen isotope analysis. Geochim. Jurassic palaeokarst (Subbetic, Southern Spain). Sediment.
Cosmochim. Acta 46, 1665–1668. Geol. 87, 13–24.
Budd, D.A., Land, L.S., 1990. Geochemical imprint of mete- Keith, M.L., Weber, J.N., 1964. Isotopic composition and en-
oric diagenesis in Holocene ooid sands, Schooner Cays, Ba- vironmental classification of selected limestones and fossils.
hamas: Correlation of calcite cement geochemistry with extant Geochim. Cosmochim. Acta 28, 1787–1816.
groundwaters. J. Sediment. Petrol. 60, 361–378. Kita, I., Taguchi, S., Matsubaya, O., 1985. Oxygen isotope
Bustillo, M.A., Ruiz-Ortiz, P.A., 1981. Relación entre sedi- fractionation between amorphous silica and water at 34–93ºC.
mentación y procesos de silicificación diagenéticos: Los sı́lex Nature 314, 83–84.
del Dogger y el Malm de la Unidad Intermedia del Jabaluz– Knauth, L.P., 1979. Origin of chert in limestone. Geology 7,
San Cristobal (Cordilleras Béticas). Estud. Geol. 37, 159–175. 274–277.
Bustillo, M.A., Ruiz-Ortiz, P.A., 1987. Chert occurrences in Knauth, L.P., 1992. Orogen and diagenesis of cherts: an isotopic
carbonate turbidites: examples from the Upper Jurassic of the perspective. In: Clauer, N., Chaudhuri, S. (Eds.), Isotopic
Betic mountains (southern Spain). Sedimentology 34, 611– Signatures and Sedimentary Records. Springer, Berlin, pp.
621. 123–152.
Choquette, P.W., James, N.P., 1987. Diagenesis 12: Diagenesis in Knauth, L.P., 1994. Petrogenesis of chert. In: Heane, P.J., Pre-
limestone, 3. The deep burial environment. Geosci. Can. 14, witt, C.T., Gibbs, G.V. (Eds.). Silica. Reviews in Mineralogy
3–35. 29, Mineralogical Society of America, Washington, D.C., pp.
Clayton, R.N., Mayeda, T.K., 1963. The use of bromine pentaflu- 233–258.
oride in the extraction of oxygen from oxides an silicates for Knauth, L.P., Epstein, S., 1976. Hydrogen and oxygen isotope
isotopic analysis. Geochim. Cosmochim. Acta 27, 43–52. ratios in nodular and bedded cherts. Geochim. Cosmochim.
Clayton, R.N., O’Neil, J.R., Mayeda, T.K., 1972. Oxygen isotope Acta 40, 1095–1108.
exchange between quartz and water. J. Geophys. Res. 77, Kolodny, Y., Epstein, S., 1976. Stable isotope geochemistry of
3057–3067. deep sea cherts. Geochim. Cosmochim. Acta 40, 1195–1209.
Kolodny, Y., Taraboulos, A., Frieslander, U., 1980. Participation
Craig, H., 1961. Isotopic variations in meteoric waters. Science
of fresh water in chert diagenesis: evidence from oxygen
133, 1833–1834.
isotopes and boron track mapping. Sedimentology 27, 305–
Craig, H., 1965. The measurement of oxygen isotope paleotem-
316.
peratures. Proc. Spoleto Conf. Stable Isotopes in Oceano-
Labeyrie, L., 1974. New approach to surface seawater paleotem-
graphic Studies and Paleotemperatures, Spoleto, pp. 161–182.
peratures using 18 O=16 O ratios in silica of diatom frustules.
Drewry, G.E., Ramsay, A.T.S., Smith, A.G., 1974. Climatically
Nature 248, 40–41.
controlled sediments, the geomagnetic field, and trade wind
Lawrence, J.R., 1989. The stable isotope geochemistry of deep-
belts in Phanerozoic time. J. Geol. 82, 531–553.
sea pore water. In: Fritz, P., Fontes, J.Ch. (Eds.), Handbook of
Epstein, S., Buchsbraum, R., Lowenstam, H.A., Urey, H.C.,
Environmental Isotope Geochemistry 3, Elsevier, Amsterdam,
1953. Revised carbonate–water isotopic temperature scale.
pp. 317–356.
Bull. Geol. Soc. Am. 64, 1315–1326. Lawrence, J.R., Gieskes, J., Anderson, T.F., 1977. Oxygen iso-
Gao, G., Land, L.S., 1991. Nodular chert from the Arbucle group tope material balance calculations, Leg 35. Init. Rep. DSDP
Slick Hills, SW Oklahoma, a combined field, petrographic and 35, 507–512.
isotopic study. Sedimentology 38, 857–870. Lohmann, K.C., 1988. Geochemical patterns of meteoric diage-
Garcı́a-Hernández, M., Martı́n-Algarra, A., Molina, J.M., Ruiz- netic systems and their applications to paleokarst. In: Cho-
Ortiz, P.A., Vera, J.A., 1988. Umbrales pelágicos: Metodologı́a quette, P.W., James, N.P. (Eds.), Paleokarst. Springer, New
de estudio y significado de las facies. 2nd Congr. Geol. España York, pp. 58–80.
(Granada), Simposio, pp. 231–241. Maliva, R.G., Siever, R., 1988. Pre-Cenozoic nodular cherts: ev-
Garcı́a-Hernández, M., López-Garrido, A.C., Martı́n-Algarra, A., idence for opal-CT precursors and direct quartz replacement.
Molina, J.M., Ruiz-Ortiz, P.A., Vera, J.A., 1989. Las discon- Am. J. Sci. 288, 799–809.
tinuidades mayores del Jurásico de las Zonas Externas de Maliva, R.G., Siever, R., 1989. Chertification histories of some
las Cordilleras Béticas: análisis e interpretación de los ciclos Late Mesozoic and Middle Mesozoic platform carbonates.
sedimentarios. Cuad. Geol. Ibér. 13, 35–52. Sedimentology 36, 907–926.
Grigor’ev, D.P., 1965. Ontogeny of Mineral. Translated from Martı́n-Algarra, A., Vera, J.A., 1995. Neptunian dykes and asso-
Russian. Israel Program for Scientific Translation, Jerusalem, ciated features in southern Spain: mechanics of formation and
250 pp. tectonic implications. Discussion. Sedimentology 42, 960–
Hesse, R., 1987. Selective and reversible carbonate–silica re- 963.
placements in Lower Cretaceous carbonate bearing turbidites Matheney, R.K., Knauth, L.P., 1993. New isotopic temperature
of the Eastern Alps. Sedimentology 34, 1055–1079. estimates for early silica diagenesis in bedded cherts. Geology
Holail, H., 1993. Diagenetic trends of the Pleistocene calcareous 21, 519–522.
ridges, Mersa Matruh area, Egypt. Chem. Geol. 106, 375–388. McCrea, J.M., 1950. On the isotopic chemistry of carbonates and
Jiménez de Cisneros, C., Molina, J.M., Nieto, L.M., Ruiz-Ortiz, a paleotemperature scale. J. Chem. Phys. 18, 849–857.
P.A., Vera, J.A., 1993. Calcretes from a palaeosinkhole in Meyers, W.J., James, A.T., 1978. Stable isotopes of cherts and
102 M.A. Bustillo et al. / Sedimentary Geology 119 (1998) 85–102

carbonate cements in the lake Valley Formation (Mississip- Sediment. Geol. 95, 57–68.
pian), Sacramento Mts, New Mexico. Sedimentology 25, 105– Rozanski, K., Araguás, L., Gonfiantini, R., 1993. Isotopic pat-
124. terns in modern global precipitation. Climatic Change in Con-
Millot, G., 1970. Geology of Clays. Springer, New York. tinental Isotopic Records, AGU, Geophys. Monogr. 78, pp.
Molina, J.M., Ruiz-Ortiz, P.A., Vera, J.A., 1995. Neptunian 1–36.
dykes and associated features in southern Spain: mechanics Ruiz-Ortiz, P.A., Bustillo, M.A., Molina, J.M., 1989. Radiolarite
of formation and tectonic implications. Discussion. Sedimen- sequences of the Subbetic, Betic Cordillera, Southern Spain.
tology 42, 957–960. In: Hein, J.R., Obradović, J. (Eds.), Siliceous Deposits of the
Moore, C.H., 1989. Carbonate Diagenesis and Porosity. Devel- Tethys and Pacific Regions. Springer, New York, pp. 107–127.
opments in Sedimentology 46, Elsevier, Amsterdam. Shackleton, N.J., Kennett, J., 1975. Paleotemperature history of
Murata, K.J., Norman, M.B., 1976. An index of crystallinity for the Cenozoic and the initiation of Antarctic glaciation: oxygen
quartz. Am. J. Sci. 276, 1120–1130. and carbon isotope analyses in DSDP Sites 227, 279 and 281.
Murata, K.J., Friedman, I., Gleason, J.D., 1977. Oxygen iso- Init. Rep. DSDP 29, 743–755.
tope relations between diagenetic silica minerals in Monterey Suchecki, R.K., Hubert, J.F., 1984. Stable isotopic and elemental
Shale, Temblor Range, California. Am. J. Sci. 277, 259–272. relationships of ancient shallow marine and slope carbonates,
O’Neil, J.R., Clayton, R.N., Mayeda, T.K., 1969. Oxygen isotope Cambro-Ordovician Cow Head Group, Newfoundland: impli-
fractionation in divalent metal carbonates. J. Chem. Phys. 51, cations for fluid flux. J. Sediment. Petrol. 54, 1062–1080.
5547–5558. Vera, J.A., Ruiz-Ortiz, P.A., Garcı́a-Hernández, M., Molina, J.M.,
Pisciotto, K.A., 1981. Diagenetic trends in siliceous facies of 1988. Paleokarst and related pelagic sediments in the Juras-
the Monterey Shale in the Santa Maria region, California. sic of the Subbetic Zone, Southern Spain. In: James, N.P.,
Sedimentology 28, 547–571. Choquette, P.W. (Eds.), Paleokarst. Springer, New York, pp.
Price, G.D., Sellwood, B.W., Valdes, P.J., 1995. Sedimentolog- 364–384.
ical evaluation of general circulation model simulations for Whittle, G.L., Alsharnhan, A.S., 1994. Dolomitization and cher-
the greenhouse Earth: Cretaceous and Jurassic case studies. tification of the Early Eocene Rus Formation in Abu Dhabi,
Sediment. Geol. 100, 159–180. United Arab Emirates. Sediment. Geol. 92, 273–285.
Rey, J., 1993. Análisis de la Cuenca Subbética durante el Ju- Wilson, R.C.L., 1966. Silica diagenesis in Upper Jurassic lime-
rásico y el Cretácico en la transversal Caravaca Vélez-Rubio. stones of southern England. J. Sediment. Petrol. 36, 1036–
Ph.D. Thesis, Univ. Granada. 1049.
Rey, J., 1995. Tectonic control in the boundaries of the genetic Winterer, E.L., Sarti, M., 1994. Neptunian dykes and associated
units: an example in the Dogger of the External Zone of the features in southern Spain: mechanisms of formation and
Betic Cordillera (Province of Murcia and Almerı́a; Spain). tectonic implications. Sedimentology 41, 1109–1132.

Você também pode gostar