23 - Kietzmann Et Al. 2015 - SEDGEO

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/268518225

Cyclostratigraphy of an orbitally-driven
Tithonian–Valanginian carbonate ramp
succession, Southern Mendoza, Argentina:
Implications for the Jurassic–Cretaceous
boundary in the Neuq...

ARTICLE in SEDIMENTARY GEOLOGY · JANUARY 2015


Impact Factor: 2.67 · DOI: 10.1016/j.sedgeo.2014.10.002

CITATION READS

1 129

3 AUTHORS:

Diego Kietzmann Ricardo Palma


University of Buenos Aires University of Buenos Aires
58 PUBLICATIONS 192 CITATIONS 73 PUBLICATIONS 264 CITATIONS

SEE PROFILE SEE PROFILE

Maria Paula Iglesia Llanos


National Scientific and Technical Research …
27 PUBLICATIONS 164 CITATIONS

SEE PROFILE

Available from: Diego Kietzmann


Retrieved on: 14 December 2015
Sedimentary Geology 315 (2015) 29–46

Contents lists available at ScienceDirect

Sedimentary Geology
journal homepage: www.elsevier.com/locate/sedgeo

Cyclostratigraphy of an orbitally-driven Tithonian–Valanginian


carbonate ramp succession, Southern Mendoza, Argentina: Implications
for the Jurassic–Cretaceous boundary in the Neuquén Basin
Diego A. Kietzmann a,c,d,⁎, Ricardo M. Palma a,c,d, Maria Paula Iglesia Llanos b,c,d
a
Instituto de Estudios Andinos “Don Pablo Groeber” (IDEAN), Buenos Aires, Argentina
b
Instituto de Geociencias Básicas y Aplicadas (IGEBA), Buenos Aires, Argentina
c
Departamento de Ciencias Geológicas, Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires, Buenos Aires, Argentina
d
Consejo Nacional de Invertigaciones Científicas y Técnicas (CONICET), Argentina

a r t i c l e i n f o a b s t r a c t

Article history: Detailed sedimentological, sequence stratigraphical and cyclostratigraphical analyses have been made from four
Received 29 July 2014 lower Tithonian–lower Valanginian sections of the Vaca Muerta Formation, exposed in the southern Mendoza
Received in revised form 5 October 2014 area of the Neuquén Basin, Argentina. The Vaca Muerta Formation is characterized by decimetre-scale rhythmic
Accepted 8 October 2014
alternations of marls, shales and limestones, and consists of five facies associations, which reflect different
Available online xxxx
paleoenvironmental conditions: basin to restricted outer ramp, outer ramp, and middle ramp. Vertical organiza-
Editor: B. Jones tion within the Vaca Muerta Formation shows a well-ordered hierarchy of cycles, where elementary cycles, bun-
dles and superbundles with frequencies within the Milankovitch band have been recognized. According to
Keywords: biostratigraphic data, elementary cycles have a periodicity of ~ 20 ky, which correlates with the precession
Orbital cycles cycle of Earth's axis. Spectral analysis based on series of cycle thickness allows us to identify frequencies of
Milankovitch about 400 ky and 90–120 ky, which we interpret as the modulation of the precessional cycle by the Earth's orbital
Biostratigraphy eccentricity. Cycles are probably driven by variations in carbonate exportation, as fluctuations in shallow-water
Magnetostratigraphy carbonate production involve modifications in carbonate basinward exportation. Cyclostratigraphic data allowed
Jurassic-Cretaceous boundary
us to build a floating orbital scale for the Tithonian–lower Valanginian interval in the Neuquén Basin. Correlation
between studied sections allowed us to recognize a discontinuity between the Substeueroceras koeneni and
Argentiniceras noduliferum ammonite zones in the Malargüe Anticline area. Orbital calibration of these sections
is consistent with Riccardi's biostratigraphic scheme, wich place the Jurassic-Cretaceous boundary within the
Substeueroceras koeneni ammonite Zone. On the other hand, the base of the Vaca Muerta Formation
(Virgatosphinctes mendozanus ammonite Zone) would be probably placed in the base of the middle Tithonian
rather than the lower Tithonian, which is also consistent with our preliminary palaeomagnetic data.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction et al., 1986; Strasser and Hillgärtner, 1998; Raspini, 2001; Strasser et al.,
2004). Usually, the combination of elementary cycles into bundles
Marl-limestone rythmic successions have been cited as key (groups of 4–5 elementary cycles), and these into superbundles (groups
examples of climate-forced sedimentation resulting from numerous of 3–4 bundles), may be interpreted as the superposition of repetitive
paleoenvironmental factors (Fischer, 1964; Hardie et al., 1986; Einsele high-frequency environmental fluctuations linked to perturbations in
et al., 1991; D'Argenio et al., 2004). Basic building units have been called Earth's orbital parameters or Milankovitch cycles (e.g., Goldhammer
elementary cycles by Schwarzacher (1947), and they represent the and Harris, 1989; Goldhammer et al., 1990; Strasser, 1994; Elrick,
smallest units recognizable in the field whose vertical lithofacies organi- 1995; Balog et al., 1997; D'Argenio et al., 1999, 2004; Anderson, 2004).
zation defines a cyclic environmental change (Grötzinger, 1986; Hardie Thick carbonate platform successions are commonly characterized
by a hierarchy of stratigraphic cyclicities with well-defined durations,
which has become a powerful tool to evaluate the absolute time span
⁎ Corresponding author at: Departamento de Ciencias Geológicas, Facultad de Ciencias they represent, reaching accuracy of the order of thousands of years
Exactas y Naturales, Universidad de Buenos Aires, Intendente Güiraldes 2160, Pabellón
II, Ciudad Universitaria, (C1428EHA) Ciudad Autónoma de Buenos Aires, Argentina.
(Kauffman et al., 1991; House and Gale, 1995; Schwarzacher, 2000;
E-mail addresses: diegokietzmann@gl.fcen.uba.ar (D.A. Kietzmann), palma@ Hilgen et al., 2006). These cyclostratigraphic data can be used as a tool
gl.fcen.uba.ar (R.M. Palma), mpiglesia@gl.fcen.uba.ar (M.P. Iglesia Llanos). for relative dating (floating orbital scales) or as a base for the high-

http://dx.doi.org/10.1016/j.sedgeo.2014.10.002
0037-0738/© 2014 Elsevier B.V. All rights reserved.
30 D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46

resolution numerical dating (Berger, 1978; Berger et al., 1992; Fischer high-precision correlation between four Tithonian–Valanginian carbon-
et al., 2004). In fact, the high accuracy of the orbital datings have ate ramp cyclic successions of the Vaca Muerta Formation, using
led to the construction of a timescale for the Neogene (Lourens et al., biostratigraphic data, cyclostratigraphic approach and preliminary
2004), which has been used in the definition of numerous magnetostratigraphic data, and the orbital calibration of the stratigraph-
Global Stratotype Sections and Points (GSSPs) (Gradstein and Ogg, ic succession to contribute to discussions about the Jurassic–Cretaceous
2012), such as Zanclean GSSP and Miocene-Pliocene boundary (Van boundary in the Neuquén Basin, but also to the calibration of Andean
Couvering et al., 2000), Messinian GSSP (Hilgen et al., 2000) and ammonite zones.
Tortonian GSSP (Hilgen et al., 2003; Hüsing et al., 2007).
There are few cyclostratigraphic studies involving the Tithonian– 2. Geological Setting
Valanginian interval. Among them, it is important to mention the
contributions of Molinie and Ogg (1992) for the Callovian–Valanginian 2.1. The Neuquén Basin
of North Pacific, Strasser (1994) and Strasser et al. (2004) for the
Tithonian–Berriasian of the Jura Mountains, Park and Fürsich (2001) The Neuquén Basin was a retro-arc basin developed in Mesozoic
for the Tithonian of Solnhofen (Germany), Anderson (2004) for the times in the Pacific margin of South America (Legarreta and Uliana,
Berriasian of Spain, and Husinec and Read (2007) for the Tithonian of 1991, 1996) (Fig. 1). Its stratigraphy was defined by Groeber (1946,
the Adriatic platform. 1953) and Stipanicic (1969), who recognized three sedimentary cycles,
While cyclostratigraphic studies have been frequently used in the i.e. Jurásico, Ándico and Riográndico. This scheme was updated by
Northern Hemisphere, they are rarely applied to the Southern Hemi- Legarreta and Gulisano (1989), who emphasized the importance of
sphere geologic record. The studies of Ozawa (2006), for the middle eustatic changes in the development of depositional sequences.
Miocene of Indonesia, as well as that of Husson et al. (2011), which Different tectonic regimes exerted a first-order control in basin de-
use data from both hemispheres to calibrate astronomically the velopment and sedimentary evolution (Legarreta and Uliana, 1991,
Maastrichtian stage, are some of the few examples which may be cited. 1996). An extensional regime was established during Late Triassic–
Previous studies in the Neuquén Basin include those of Sagasti Early Jurassic, which generated a series of narrow, isolated depocentres
(2000, 2005) for the upper Agrio Formation (late Valanginian–early controlled by large transcurrent fault systems filled mainly with conti-
Barremian), and Scasso et al. (2002, 2005) for a limestone-marl interval nental deposits of the Precuyo Group (Fig. 2A, Manceda and Figueroa,
restricted to the middle Tithonian of the Vaca Muerta Formation. 1993; Vergani et al., 1995; Giambiagi et al., 2008). Thermal subsidence
Kietzmann et al. (2011b) recognized the presence of precessional and with localized tectonic events characterized the Early Jurassic to Late
eccentricity orbital cycles in the Tithonian of the Arroyo Loncoche sec- Cretaceous interval (Vergani et al., 1995). Depocentres were filled by
tion, indicating a great potential of the Vaca Muerta Formation for continental and marine siliciclastic, carbonate and evaporitic sediments
cyclostratigraphic studies. (Cuyo, Lotena, and Mendoza Groups). Marine sequences developed
Many correlations along the peri-Pacific domain are based on ages throughout the basin during Late Jurassic–Early Cretaceous, are includ-
assigned to Andean ammonite zones, so their calibration has implica- ed in the Mendoza Group (Stipanicic, 1969) or Mendoza Mesosequence
tions for global correlations. The aims of this paper are to carry out a (Legarreta and Gulisano, 1989) (Fig. 2A). Legarreta and Gulisano (1989)

Fig. 1. Location map of the Neuquén Basin showing main geological features and studied localities (see numbers).
D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46 31

Fig. 2. A) Stratigraphic chart for the Neuquén Basin showing Groeber's cycles and sequences of Legarreta and Gulisano (1989). B-C) Litostratigraphic subdivision of the Lower Mendoza
Mesosequence in the Central Neuquén sector (B) and in the Southern Mendoza sector (C).

divided the Mendoza Mesosequence into three main shallowing- By contrast, in the southern Mendoza area (Fig. 2C) the Lower
upward sedimentary cycles: Lower Mendoza Mesosequence (upper Mendoza Mesosequence consists of aggradational sequences, with
Kimmerdgian–lower Valanginian), Middle Mendoza Mesosequence a maximum thickness of 500 m towards the center of the basin
(lower Valanginian), and Upper Mendoza Mesosequence (lower (Legarreta and Gulisano, 1989). It includes continental deposits of the
Valanginian–lower Barremian). Tordillo Formation (late Kimmeridgian–early Tithonian?), basinal to
A compressive deformation regime was established during the Late middle carbonate ramp deposits of the Vaca Muerta Formation (early?
Cretaceous, and continued throughout the Cenozoic, although alternat- Tithonian–early Valanginian) and middle to inner ramp oyster-
ing with extensional events (Ramos and Folguera, 2005; Ramos, 2010). deposits of the Chachao Formation (early Valanginian), which form a
This Andean deformation resulted in the development of a series of N– homoclinal carbonate ramp system (e.g. Carozzi et al., 1981; Mitchum
S-oriented fold and thrust belts (Aconcagua, Marlargüe and Agrio fold and Uliana, 1985). Estward poorly dated tidal to continental mixed de-
and thrust belts) where excellent outcrops of the Mesozoic successions posits receiving the name of Lindero de Piedra Formation (Legarreta
are exposed. et al., 1981), have been recognized and correlated with the Vaca Muerta
and Chachao Formations (Fig. 2C).
2.2. The upper Kimmeridgian–lower Valanginian Lower
Mendoza Mesosequence 2.3. Age and biostratigraphy of the Vaca Muerta Formation

The Lower Mendoza Mesosequence is an upper Kimmeridgian– Jurassic–Cretaceous Andean biostratigraphy is well defined based on
lower Valanginian broad shallowing-upward sedimentary cycle ammonites (Riccardi, 2008; Aguirre-Urreta et al., 2011; Riccardi et al.,
(Fig. 2), in which most distal facies are included into the Vaca Muerta 2011), and to a lesser extent on bivalves, brachiopods, and microfossils,
Formation. This interval was studied with some detail in the southern such as foraminifera, calcareous nannofossils, radiolarians and dinofla-
sector of the basin (Neuquén Province), where it was divided into gellates (Quattrocchio et al., 2003; Ballent et al., 2004, 2011; Bown
nine depositional sequences (Mitchum and Uliana, 1985). In the central and Concheyro, 2004).
part of the Neuquén Basin (Fig. 2B), also known as Neuquén embay- We consider here a modified version of the last Andean ammonite
ment, the Lower Mendoza Mesosequence shows a general sigmoidal ge- biostratigraphic scheme proposed by Riccardi et al. (2011) (Fig. 3),
ometry, and thickness reaching 2000 m (Gulisano et al., 1984; Legarreta which was adopted also by other authors (e.g., Vennari et al., 2014). Am-
and Gulisano, 1989; Legarreta and Uliana, 1991). It starts with the con- monite data point out that the Vaca Muerta Formation was deposited
tinental deposits of the Tordillo Formation (late Kimmeridgian–early from the early?/middle Tithonian (Virgatosphinctes mendozanus Zone)
Tithonian?), which are followed by basinal deposits of the Vaca Muerta to the early Valanginian (lower part of the Olcostephanus (O.) atherstoni
Formation (early? Tithonian–late Berriasian). Eastward these facies Zone) (Fig. 3).
change to shoreface deposits of the Quintuco Formation (late First ocurrences of calcareous nannofossils Polycostella beckmanii,
Tithonian–early Valanginian), and to sabkha deposits of the Loma Eiffellithus primus, and Umbria granulosa are located within the
Montosa Formation (lower Valanginian), forming a mixed carbonate- Pseudolissoceras zitelli ammonite Zone, indicating a middle Tithonian
siliciclastic depositional system (Gulisano et al., 1984; Mitchum and age, while Polycostella senaria and Raghodiscus asper occur within the
Uliana, 1985; Carozzi et al., 1993) (Fig. 2B). Westward the Vaca Muerta Windhaunseniceras internispinousm ammonite Zone, indicating an
Formation includes slope facies (Huncal Member), and in the Chilean upper Tithonian age (Concheyro et al., 2006; Lescano and Kietzmann,
territory pass into shallow marine/volcanic deposits (Charrier, 1985; 2010; Kietzmann et al., 2011b). The presence of Cruciellipsis cuvillieri
Leanza et al., 2011; Kietzmann and Vennari, 2013; Rossel et al., 2014). and Micrantholithus hozchulzi were recognized from the upper part of
32 D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46

Fig. 3. Biostratigraphic subdivision of the Tithonian–Valanginian interval in the Neuquén Basin, showing Andean Ammonite Zones (after Riccardi et al., 2011, and Aguirre-Urreta et al.,
2011), nannofossil events, calpionellid events and zones, and Saccocoma distribution (Kietzmann and Palma, 2009b; Lescano and Kietzmann, 2010; Kietzmann et al., 2011a,b).

the Substeueroceras koeneni Zone to the Spiticeras damesi Zone, suggest- Cyclostratigraphic analysis is based on the identification of dm-scale
ing a Berriasian age. Finally, the first occurrence of Eiffellithus windii, carbonate/siliciclastic lithofacies couplets or elementary cycles, bundles
which indicates an early Valanginian age, coincides with the Neocomites and superbundles differentiated in the field. Lower frequencies are
wichmanni ammonite Zone (Lescano and Kietzmann, 2010). searched using spectral analysis with the software POWGRAF2
Calpionellids are generally poorly preserved due to recrystallization (Pardo-Igúzquiza and Rodríguez-Tovar, 2004) using Blackman-Tukey
of carbonate mud. Chitinoidellids occurs within the Virgatosphictes method. Spectral analyses are based in the following primary premises:
mendozanus to Substeueroceras koeneni ammonite Zones. Particularly,
Chittinoidella boneti is reported for the Windhauseniceras internispinosum 1. The sections consist of a succession of elementary cycles showing
Zone in Sierra de la Cara Cura (Fernández-Carmona and Riccardi, 1998; similar thickness within the range of several decimetres, and they
Kietzmann et al., 2011a). are considered of similar time duration.
Tintinnopsella, Crassicollaria and Calpionella have been found within 2. The number of cycles is large enough to obtain statistically significant
the Corongoceras alternans and the lower part of the Substeueroceras results. According to Weedon (2003) the sampling density must be
koeneni ammonite Zones (Fernández-Carmona et al., 1996), which is at least 12 times the lowest frequency sought. In this case is at least
consistet with a late Tithonian age for these levels (Fig. 3). 25 times longer than the low frequency eccentricity cycle.
Remains of saccocomids were recognized from the Virgatosphinctes 3. Vertical changes in facies and bed thickness throughout the sections
mendozanus Zone to the Corongoceras alternans ammonite Zones, with could be related with periodic climate factors or with other non-
an acme in the Pseudolissoceras zittelli and Aulacosphinctes proximus am- periodic factors as basin evolution.
monite Zones (Kietzmann and Palma, 2009b), which occurs during 4. Data set was corrected prior to spectral analysis, substracting the
middle Tithonian (Hess, 2002). mean value and trends generated by changes in sea level (composite
depositional sequences), allowing data centering and variance stabi-
lization (see Weedon, 2003).
3. Methodology 5. Poorly-cyclic intervals are not incorporated into the spectral analysis.
However, these “non-cyclic” intervals need to be also considered for
Detailed sedimentological sections were measured and described interpreting the long term variations as well as for estimations of the
bed-by-bed in 4 localities of the southern Mendoza sector of the total number of cycles in the stratigraphic sections. An estimate of
Neuquén Basin. They include the Arroyo Loncoche, and Cuesta del the number of “hidden” cycles was given by simply dividing their
Chihuido sections, within the Malargüe Anticline area, and the Bardas thickness by the average cycle thickness of the contigous intervals.
Blancas and Arroyo Rahue sections, within the Sierra Azul Anticline An error of two low-eccentricity cycles is assumed using this
area (Fig. 1). Detailed facies analysis of these sections was published methodology.
by Kietzmann et al. (2014). Carbonate and siliciclastic facies were
analyzed taking into account bed geometry, lithology, sedimentary Sequence-stratigraphic framework is based on the identification of
structures, as well as fossil content and petrographic features. Division flooding surfaces, which were identified using juxtaposition and disloca-
of sedimentary environment within the carbonate ramp is based in tion of facies. In fact, the position of the Vaca Muerta Formation in the
Burchette and Wright (1992). sedimentary system does not allow the recognition of sequence bound-
Ammonite data from 220 ammonite levels in the entire studied sed- aries as subaerial exposure surfaces in the sense of Van Wagoner et al.
imentary sections allow us to identified the eleven early?/middle (1990). Nonetheless, stacking pattern and facies arrangement are used
Tithonian–early Valanginian Andean biozones (Fig. 3). to recognize sequence-stratigraphic trends, where flooding surfaces
D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46 33

can be interpreted as correlative conformities of subaerial exposure sur- 4.1.3. Facies association 3: bioclastic middle ramp to proximal outer ramp
faces, forming coplanar surfaces (e.g. Embry and Johannessen, 1992). FA-3 consists of gradded bioclastic floatstones/rudstones, bioturbat-
Palaeomagnetic data come from the Arroyo Loncoche section, ed wackestones, laminated packstones and wackestones, and marls
where we sampled 10 sites along the Vaca Muerta Formation. Each (Fig. 4C).
sampling site contains at least 4 oriented specimens obtained with a Allochems are dominated by bivalves, rynchonellids, gastropods,
portable drill. Because nearly 40% of the collected specimens are still serpulids, ammonites, aptychi, nautiloids, belemnites, echinoderms,
being processed, the magnetic polarities we present in this study are and abundant epistominid foraminifera and pellets, which are
preliminary. chaotically distributed. Bioturbation is represented by Thalassinoides,
Diplocaterion and Rhizocorallium.
FA-3 was deposited in a moderate well-oxygenated bioclastic
4. Sedimentology and sequence stratigraphy middle ramp to proximal outer ramp. Taphonomic features of skeletal
particles, suggest reworking by storm-wave action (Kietzmann and
4.1. Facies associations Palma, 2009a). Occurrence of laminated wackestones and marls suggest
deposition by settling of fine-grained suspended material, during fair-
A detailed facies analysis of the Vaca Muerta Formation in the south- weather stages, as well as sedimentary particles transported during
ern Mendoza sector was published by Kietzmann et al. (2014). In order storm events. Trace fossils also indicate alternating energy regimes
to establish the sedimentological features of cyclic deposits described in and a relatively well oxygenated substrate.
this paper, five facies associations characterized in this areas (FA-1 to
FA-5) are briefly described here, which correspond to basinal to middle 4.1.4. Facies association 4: bioclastic outer ramp
carbonate ramp deposits. FA-4 is represented by an alternation of peloidal and intraclastic
packstones, radiolaritic wackestones and thin gradded bioclastic
rudstones, rhytmically interbedded with marls (Fig. 5A). They contains
4.1.1. Facies association 1: Oyster autoparabiostrome dominated abundant crustacean pellets, fine to medium sand-size micritic
middle ramp intraclasts, ammonites, bivalves, gastropods and foraminifera.
FA-1 consists of oyster accumulations, which form up to several Bioturbation is rare, although there may be some Planolites,
meters thick internally bedded biostromes, alternating with poorly Palaeophycus and Thalassinoides tubes.
stratified, massive, or low-angle cross-stratified bioclastic rudstones/ FA-4 deposited in a poorly-oxygenated bioclastic outer ramp.
floatstones, marls and laminated wackestones/packstones (Fig. 4A). Sedimentation during fair-weather periods is represented by marls
Fossil content includes abundant oysters (Aetostreon sp.), ammo- and laminated radiolaritic wackestones.
nites, serpulids (Rotularia sp.) and rynchonellid brachiopods, as Pelloidal accumulations probably come from the unroofing of
well as ostracods, ophiuroid ossicles, dasycladacean algae, crustacean crustacean galleries (e.g. Tedesco and Wanless, 1991). The abundance
microcoprolites, and epistominid and textularid foraminifera. of intraclasts suggests intermittent erosion of wackestones facies and
Beds are frequently bioturbated by Taenidium, Diplocaterion, transport from bottom currents associated with storm events
Rhizocorallium and/or Thalassinoides. (Schieber et al., 2010). Thin massive or gradded bioclastic rudstones
FA-1 is interpreted as deposits of a well-oxygenated middle ramp. were originated from storm-generated turbidity-like flows. Trace fossils
Laminated wackestones and marls represent fair-weather periods. are related to the Cruziana ichnofacies, which characterizes low-energy
Low-angle cross-stratified deposits are interpreted as accretionary environments (MacEachern et al., 2008).
bioclastic bars located probably within the upper part of the middle
ramp, similar to those described by Bádenas and Aurell (2001). Trace 4.1.5. Facies association 5: distal outer ramp to basin
fossil assemblage is indicative of the Cruziana ichnofacies, which charac- FA-5 consists of dark grey to black well-laminated marls, and
terizes low-energy environments (MacEachern et al., 2008). subordinately intraclastic packstones, ripple-laminated packstones,
radiolaritic wackestones, and microbialites (Fig. 5B).
Radiolarians are abundant, but disarticulated thin-shelled oysters
4.1.2. Facies association 2: HCS dominated middle ramp are also common. These facies are particularly rich in fishes, turtles,
FA-2 is composed by marls and hummocky cross-stratified ichthyosaurs and crocodiles remains. Trace fossils are represented
grainstones, bioturbated wackestones, laminated intraclastic packstones, by Chondrites, but also evidences of cryptobioturbation are found in
massive to gradded bioclastic floatstones/rudstones, and occasionally microbialite laminae.
planar microbialite beds (Fig. 4B). Early diagenetic concretions included FA-5 deposited in a poorly-oxygenated distal outer ramp to basin.
among the marls show similar textures to packstones. Sedimentation occurred mainly from suspension. Sedimentation took
In these facies bivalves, gastropods, ammonites, echinoderms, place in a restricted marine environment with suboxic bottom waters,
Saccocoma remains, lingulid brachiopods, benthic foraminifera as indicates the absence of wave-induced structures, and the passive
(Epistomina, Lenticulina and agglutinated forms), calcareous algae, accumulation of organic materials and pelagic microfossils. However,
gastropod internal molds, small bone fragments and scales, as well the presence of fine-grained ripple laminated carbonate, as well as
abundant pellets are the dominant particles. Trace fossil are represented bioclastic remains derived from shallow-water are probably associated
by Thalassinoides, Helminthopsis, Planolites, and Chondrites. with storm-generated turbidity flows.
FA-2 represents sand-sized peloidal sediments deposited in a well-
oxygenated middle ramp. Hummocky cross-stratified grainstones 4.2. Sequence stratigraphic framework
were generated from storm-related oscillatory flows with a small unidi-
rectional component. Morever, bioclastic floatstones/rudstones repre- Based on the analysis of flooding surfaces, Kietzmann et al. (2014)
sent bioclastic tempestites originated by turbidity-like gravity flows recognized two hierarchies of depositional sequences: 1) Composite
(Kietzmann and Palma, 2011). Marls and bioturbated wackestones depositional sequences (CS) for large-scale sequences, and 2) high-
were deposited by fallout of mud, whereas microbialite beds reflect frequency depositional sequences (HFS) for those of small scale. Be-
times of low sedimentation rate. cause of its average duration, composite sequences are considered to
Trace fossils represents the ichnofacies of Skolithos and Cruziana be equivalent to third order sequences, while high-frequency sequences
(Kietzmann et al., 2014), indicating an alternating energy regime and are considered as fourth order sequences (e.g., Kerans and Tinker,
a well-oxygenated substrate (Pemberton et al., 1992). 1997). In this paper we will consider only composite sequences, because
34 D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46

Fig. 4. A) Facies association 1 (oyster biostrome dominated middle ramp) in the upper Berriasian of Cuesta del Chihuido section (see hammer as scale). B) Facies association 2 (HCS
dominated middle ramp) in the middle Tithonian of Bardas Blancas section. C) Facies association 3 (bioclastic middle ramp to proximal outer ramp) in the middle Tithonian of the Cuesta
del Chihuido section. Facies code: (Mrh) marls, (FRbm) massive bioclastic floatstones/rudstones, (FRbl) low-angle cross-stratificatified bioclastic floatstones/rudstones, (Fbm) massive
bioclastic floatstones, (Gphcs) hummocky cross-stratified grainstones, (Piph) horizontal laminated packstones, (Wbb) bioturbated wackestones. Hammer scale ~30 cm.

it is important to eliminate these trends for spectral analysis. High- composite depositional sequence (CS-1) starts with a regional flooding
frequency sequences are too close to the superbundle hierarchy, there- surface at the boundary between the Vaca Muerta Formation and the
fore their removal could generate fictitious frequencies. continental facies of the Tordillo Formation (see Fig. 2), in the lower?
The studied interval includes 5 trasgressive-regressive composite Tithonian Virgatosphinctes mendozanus Zone (Figs. 6, 7). The transgres-
depositional sequences (Figs. 6, 7; Kietzmann et al., 2014). First sive deposits include basinal to distal outer ramp facies (FA-4 and FA-5)
D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46 35

and the regressive system tract shows the progradation of middle ramp (FA-3) of the Vaca Muerta Formation and shallower oyster-dominated
FA-2 and FA-3. The upper limit is marked by other regional flooding deposits of the Chachao Formation.
surface (SB-1) located within the lowermost upper Tithonian
Windhauseniceras internispinosum Zone (Figs. 6, 7), which is also the 5. Cyclostratigraphic analysis
base of the second composite depositional sequence (CS-2). CS-2 en-
compasses outer ramp facies (FA-4) and middle to proximal mid ramp 5.1. Hierarchy of cycles
(FA-2 and FA-3) in its regressive part. The upper limit of CS-2 (SB-2)
lies within the uppermost upper Tithonian Substeueroceras koeneni A well-ordered hierarchy of cycles, including elementary cycles,
Zone. CS-3 has a clear transgressive-regressive trend of middle (FA-3) bundles and superbundles has been recognized within the Vaca Muerta
and outer (FA-2) facies (Figs. 6, 7). Formation. Elementary cycles have a relatively regular thickness in the
Flooding surface SB-3 that separates composite depositional se- order of 20 to 40 cm, so that they can be regarded as temporarily equiv-
quences CS-3 and CS-4 coincides with the lower part of the Spiticeras alent units. Three types of elementary cycles are recognized: limestone/
damesi Zone (upper Berriasian), while SB-4 lies in the upper part of marl, marl/marl and marl/shale (Figs. 8, 9).
this zone. CS-4 mainly includes outer ramp facies (FA-2) and prograding Each elementary cycle consists of two hemicycles of similar thick-
middle ramp facies (FA-3). In the eastern part of the southern Mendoza ness. Limestone/marl cycles begin with a limestone hemicycle, wich
sector and the Malargüe Anticline (see Figs. 2C, 7). Finally, composite pass transitionaly to a marl hemicycle. On unweathered surface, lime-
depositional sequence CS-5 ends in the lower part of the Olcostephanus stones have a net basal boundary and show enrichment in clays upward,
(O.) athertoni Zone (lower Valanginian) including middle ramp facies resulting in a transitional passage to marls. The sedimentological feature

Fig. 5. A) Facies association 4 (bioclastic outer ramp) in the upper Tithonian of the Arroyo Loncoche section, showing an alternation of marls and intraclastic/peloidal packstones. B) Facies
association 5 (distal outer ramp to basin) in the Berriasian of the Arroyo Rahue section, showing an alternation of marls and radiolaritic wackestones. Facies code: (Mrh) marls, (Wrh)
laminated radiolaritic wackestones (Piph) horizontal laminated packstones. Hammer scale ~30 cm.
36 D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46

Fig. 6. Stratigraphic sections of the Vaca Muerta Formation in the Sierra Azul area. References: (SB) sequence boundary, (TST) transgressive system tract, (RST) regressive system tract,
(E) low-frequency eccentricity cycle, (e) high-frequency eccentricity cycle, (O.a.) Olcostephanus (O.) atherstoni ammonite Zone.
D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46 37

Fig. 7. Stratigraphic sections of the Vaca Muerta Formation in the Malargüe anticline. References: (SB) sequence boundary, (TST) transgressive system tract, (RST) regressive system tract,
(E) low-frequency eccentricity cycle, (e) high-frequency eccentricity cycle.
38 D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46

cycle (Fig. 9). These facies associations are common in transgressive sys-
tem tract. By contrast, facies association 3 to 1, which represent outer
ramp to middle ramp deposits, are dominated by limestone/marl ele-
mentary cycles (Fig. 8), and are common in regressive system tract.

5.2. Spectral analysis

According to traditional ammonite biostratigraphic data, the Vaca


Muerta Formation covers the early Tithonian to early Valanginian time
interval. However, it should be noted that the lower Tithonian is not
complete in the Vaca Muerta Formation (Fig. 5), because the correlation
with the H. hybonotum Standard Zone would not be represented (see
Leanza, 1981).
Absolute ages from the Vaca Muerta Formation are rare, due to the
absence of appropiate lithologies or poor preservation of tuffs. However,
Vennari et al. (2014) obtained recently a SHRIMP age of ~139.5 Ma for
the Argentiniceras noduliferum ammonite Zone. Detrital zircon data
from the Tordillo Formation, which underlies the Vaca Muerta Forma-
tion, indicate a maximum age of ~ 147 Ma (Rossel et al., 2014) and
~ 144 Ma (Naipauer et al., 2012), and the lower Hauterivian deposits
of the Agrio Formation yielded an age of ~ 132.5 Ma (Aguirre-Urreta
et al., 2008). These data point to a duration of about 10 Ma for the
Vaca Muerta Formation on the basis of the time scale of Gradstein
et al. (2012).
In the Sierra Azul area, the Vaca Muerta Formation extends from the
Virgatosphinctes mendozanus Zone to the Olcosthephanus (O) atersthoni
Zone (Fig. 6). The Bardas Blancas section contains 480 elementary cy-
cles, which are grouped in 96 bundles and 24 superbundles, while the
Arroyo Rahue section includes 460 elementary cycles, 92 bundles and
23 superbundles. Dividing the time-length of these sections by the
number of cycles, elementary cycles in this area have duration of
about 20 ky, which can be attributed to the precessional cycle of the
Earth. Data from other Jurassic examples are consistent with this aver-
age value (e.g., Hinnov and Park, 1999; Weedon et al., 1999; Husinec
Fig. 8. Hierachy and types of cycles in the Vaca Muerta Formation.
and Read, 2007).
In the Malargüe anticline, the Vaca Muerta Formation starts in the
Virgatosphinctes mendozanus Zone, but the upper boundary reach the
lower part of the Neocomites wichmanni Zone, reaching about 10 Ma
of each hemicycle varies according to the facies association in which time lengths (Fig. 7). The Arroyo Loncoche section contains 447 elemen-
occurs. Marl/marl and marl/shale elementary cycles show variations in tary cycles, which are grouped in 88 bundles and 22 superbundles,
the content of carbonate, allowing detecting them in weathering profile. while the Cuesta del Chiuhido section include 445 elementary cycles,
Elementary cycles are grouped into sets of 4–5 elementary cycles (bun- 88 bundles and 22 superbundles. Elementary cycles in this area can
dles), and these are grouped into sets of 4–5 bundles (superbundles). also be attributed to the precessional cycle.
Such stacking pattern is used by different authors as diagnostic criteria In the Sierra Azul area, Blackman-Tuckey spectra show a peak above
to identify the influence of orbital forcing (e.g., Goldhammer et al., the 95% confidence level, which corresponds to a periodicity of 89 ky in
1990; Schwarzacher, 1993; Lehmann et al., 1998; Raspini, 2001; the Arroyo Rahue section, and 111 ky in the Bardas Blancas section
Anderson, 2004; Strasser et al., 2004, among others). The 5:1 ratio (5 el- (Fig. 10). These periodicities can be attributed to the high-frequency ec-
ementary cycles per bundles) is commonly attributed to high frequency centricity cycle. Another peak near the confidence level of 95%, with a
eccentricity (95 ky and 125 ky), while the 4:1 ratio (4 bundles per frequency of 395 and 390 ky respectively, can be attributed to the
superbundles) as low frequency eccentricity (410 ky). low-frequency eccentricity cycle.
Bundles and superbundles start with a thick elemental cycle and In the Malargüe anticline, Blackman-Tuckey spectra show a peak
show higher proportion of marls towards the top. This characteristic above the 95% confidence level, and two peaks above the 99% confi-
stacking pattern is clearly shown on the field in stratigraphic sections dence level (Fig. 10). The first one has a periodicity of 410 ky in the Ar-
with low sedimentation rate, such as the Cuesta del Chihuido section royo Loncoche section and 403 ky in the Cuesta del Chihuido section,
(Fig. 9A). Bundles and superbundles can be dominated by marl/marl el- which are consistent with the low-frequency eccentricity cycle. The
ementary cycles, so that the proportion of marls exceeds the proportion other two peaks have periodicities of 118 cycles ky and 91 ky in the Ar-
of limestones, or be dominated by limestone/marl elementary cycles, royo Loncoche section, and appear as a single peak of 95 ky in the Cuesta
i.e., the proportion of limestones is greater or similar to the proportion del Chihuido section (Fig. 10). These periodicities can be also attributed
of marls (Figs. 8, 9). This descriptive division correlates with facies and to the high-frequency eccentricity cycle.
system tracts, so that variations in the proportion of marl or limestones
have no real genetic connotation respect to the mechanisms which 6. Palaeomagnetic data
transfers the orbital signal to the sedimentary record, but with the
proximal-distal trend within the carbonate ramp. Palaeomagnetic data presented here are preliminary (see for exam-
Facies association 4 and 5, which represent basinal to outer ramp de- ple, large gaps with no data in the Loncoche section) and has to be up-
posits, are dominated by marl/marl elementary cycles. Bundles or grade in the forthcoming time. Palaeomagnetic behaviours were in
superbundles can occasionally start with a limestone/marl elementary general terms, very reliable, and only a few specimens were ruled out
D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46 39

Fig. 9. A) Outcrops example of a cyclic interval dominated by limestone/marl elementary cycles in Cuesta del Chihuido section (upper Tithonian). B) Outcrops example of a cyclic interval
dominated by marl/marl elementary cycles in Arroyo Rahue section (upper Berriasian). References: (P) elementary cycles, attributed to precession, (e) bundles, attributed to high-fre-
quency eccentricity cycles, (E) superbundles, attributed to low-frequency eccentricity cycles.

from this study. From each site, we calculated mean directions that were normal polarity N1 which would extend until the boundary between
converted to virtual geomagnetic poles (Fig. 11A). By averaging such the Corongoceras alternans Zone and Substeueroceras koeneni Zone. On
poles, we obtained a palaeomagnetic pole that falls very consistently the top of the section, we find the N2 polarity spanning probably part
with reliable Upper Jurassic–Lower Cretaceous palaeomagnetic pole of the Argentiniceras noduliferum Zone and the Spiticeras damesi Zone.
from stable South America, indicating the primary origin of the isolated
magnetisations. 7. Discussion
Results reveal that both normal and reverse polarities that pass the
reversal test have been isolated in the Vaca Muerta Formation, indicat- 7.1. Origin of elementary cycles
ing that the antipodal directions belong to the same population. From
base to top, we find in the Arroyo Loncoche section (Fig. 11A) the re- Rhythmic calcareous successions have been cited as key examples of
verse polarity R1, located in the Virgatosphinctes mendozanus Zone, climate-forced sedimentation resulting from environmental signals or
followed by the R2 polarity that probably spans from the top of the created by diagenesis (Einsele and Ricken, 1991). Criteria for identifying
Pseudolissoceras zitteli Zone to the base of Windhauseniceras sequences of primary origin from those of diagenetic origin were enu-
internispinosum Zone. Here, there is a clear change to the dominant merated by Einsele (1982) and Einsele and Ricken (1991), which are
40 D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46

based on simple geological criteria. Recently, Westphal et al. (2010)


reviewed the use of these criteria, dividing them into “criteria of limited
use” and “unequivocal criteria”. The former include: 1) lateral continu-
ity of individual beds, 2) presence of aragonitic fossils, and 3) the use
of stable isotopes in whole rock. These criteria are highly questionable,
since there are numerous examples where these characteristics are
met both sequences of primary or secondary origin (see reference in
Westphal et al., 2010).
Among the unequivocal criteria may be mentioned: 1) presence of
palynomorphs, 2) presence of calcite fossils in both calcareous and
terrigenous hemicycles, 3) presence of sedimentary structures and tex-
tures, 4) variations in clay mineralogy, and 5) use of stable isotopes in
calcite stable components. The abundant paleontological content of
the Vaca Muerta Formation, which includes several groups of planktonic
and benthic invertebrates whith aragonitic, calcitic and organic compo-
sition, is present in both hemicycles. Furthermore, although there is ev-
idence of diagenetic overprint, taphonomic, sedimentary structures,
and petrographic evidences are irrefutable evidence that support the
primary origin of rhythmicity in the Vaca Muerta Formation (see
Kietzmann et al., 2008, 2010, 2011b, 2014; Kietzmann and Palma,
2009a,b, 2010, 2011, 2014; Martín-Chivelet et al., 2011).
Primary environmental signals result from fluctuations in the supply
of calcareous biogenic sediments produced by plankton (productivity
cycles: Einsele and Ricken, 1991) and fluctuations in the input of fine
terrigenous sediments (dilution cycles: Arthur et al., 1984; Einsele and
Ricken, 1991). Additional potential mechanisms include, fluctuations
in superficial fertility (fertility cycles: Premoli Silva et al., 1989), and
fluctuations in carbonate mud exportation from shallow areas (carbon-
ate dilution cycles: Pittet and Strasser, 1998).
Strictly, productivity cycles reflect fluctuations in the productivity of
the water surface, related with changes in the surface temperature or in
the availability of nutrients (e.g., Seibold, 1952; Prell and Hays, 1976;
Einsele, 1982). A special case of “productivity cycle” is related to varia-
tions in carbonate factory production in shallow areas associated to fluc-
tuations in sea level or changes in the water temperature, which
translate in variations in carbonate exportation. This mechanism has
been called carbonate dilution cycles (Pittet and Strasser, 1998;
Colombié and Strasser, 2003) or exportation cycles (Bádenas et al.,
2003).
The sedimentary features of the Vaca Muerta Formation in the
Mendoza sector of the Neuquén Basin restrict significantly the range
of possible mechanisms to explain the transfer of the orbital signal to
the sedimentary record. Productivity (sensu Einsele, 1982) and fertility
(sensu Premoli Silva et al., 1989) mechanisms are not adequate to ex-
plain the cyclicity in the Vaca Muerta Formation, since sedimentation
should have strong pelagic imprints. By contrast, the abundance of skel-
etal particles with signs of transport, the presence of particles derived
from shallow-water areas, such as dasycladacean algae and agluttinated
foraminifera, and the abundant crustacean microcoprolites in both
carbonate and marly facies, are strong evidences that sedimentation
was controlled by hydrodynamic factors. Therefore, cyclicity can be
interpreted in terms of “productivity” related with variations in carbon-
ate exportation, similarly to carbonate dilution cycles of Pittet and
Strasser (1998) or carbonate exportation cycles of Bádenas et al. (2003).
This mechanism would be controlled by variations in shallow-water
productivity in response to surface temperature fluctuations and/or
high frequency sea-level changes. In fact, cyclic patterns observed in
marl-dominated bundles and superbundles are typicaly transgressive;
while carbonate dominated bundles and superbundles coincide with
progradation of carbonate ramp. The decrease in shallow-water accom-
modation space may have promoted the exportation of carbonate
sediments towards distal areas of the ramp, while increase in the ac-
commodation space due to increasing water depth, promotes retention

Fig. 10. Blackman–Tuckey spectrum of studied localities, showing the presence of the low
frequency (E) and high frequency (e) eccentricity cycle.
D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46 41

Fig. 11. A) Distribution of palaeomagnetic sampling sites in the Vaca Muerta Formation at Arroyo Loncoche section, and relation with the Andean ammonite zones. Latitude of virtual geo-
magnetic poles (VGP lat.) defines the sequence of isolated polarities (R1–N2). Polarities: white = reverse, black = normal, grey = intervals with no data. B) Correlation between the local
magnetostratigraphic scale and the international geomagnetic polarity time scale (Ogg and Hinnov, 2012a,b). Solid line indicates reliable equivalence whilst dashed line represents a more
uncertain correspondence.

of carbonate in shallow-water areas and marl accumulation in deeper obliquity cycle. The obliquity signal comes and goes erratically in middle
areas. and low latitudes, but would have remained constant in some
hemipelagic sequences (e.g., Fischer et al., 2004). Therefore, the eccen-
7.2. Orbital calibration of Tithonian – Valanginian Neuquén tricity cycle has been used to build astronomical time scales (Hilgen
basin succession et al., 2000, 2003; Van Couvering et al., 2000; Lourens et al., 2004;
Hinnov and Oggs, 2007; Hüsing et al., 2007).
Milankovitch cycles vary throughout geological time (e.g., Berger Using high and low frequency eccentricity cycles identified in the
et al., 1992; Schwarzacher, 2000; Hinnov and Oggs, 2007). Berger four studied stratigraphic sections, we built a floating astronomical
et al. (1989) estimated that obliquity cycles increased a 45% from the Si- scale for the lower? Tithonian–lower Valanginian of the Neuquén
lurian, while precession cycles about 20 %. However, the periodicity of Basin, which allowed us to calibrate ammonite zones (Figs. 6, 7, 12).
the eccentricity cycle can be considered more or less constant over the There are some differences in the number of high-frequency eccentric-
last 500 Ma. According to calculations of Berger et al. (1992), during ity cycles between sections, so longer duration of biozones of the four
Tithonian–Valanginian times the primary precessional cycle of ~23 ky sections is used for building the astronomical timescale (Fig. 12). In
would have been about 21 ky, while the secondary cycle of 19 ky fact, seismic data shows the existence of discontinuities, which allow
about 18 ky, so the precessional values calculated in this study are con- to recognize depositional sequences (e.g., Mitchum and Uliana, 1985).
sistent with those estimated by this author. However, having sufficiently long time series reduces errors caused
It is also known that variations in the obliquity parameter are impor- by the presence of these minor discontinuities in spectral analysis
tant at high latitudes, while precession variations affect middle and (e.g., Weedon, 2003).
lower latitudes (Berger and Loutre, 1994). These observations led The most important difference in the thickness and number of cycles
Fischer et al. (2004) to define two categories, one related to the between ammonite biozones is shown in the Argentiniceras noduliferum
precession-eccentricity syndrome (PES) and other related to the Zone, since in the Malargüe anticline it has the scale of two high-
42 D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46

Fig. 12. C) Orbital calibration of Andean ammonite zones and correlation with standard zones. The lower boundary of the Windhauseniceras internispinosum Zone (red dotted line) is
choosen as datum to fit the orbital floating scale. Note that the second datum (base of the Neocomites wichmanni Andean Zone) fits quite well to the ICS 2004 scale.

frequency eccentricity cycles, while in the Sierra Azul area up to eigth correlate with the Durangites Standard Zone and the lower part of
cycles can be identified. Additionally, some missing cycles are observed the Occitanica Standard Zone. The Argentiniceras noduliferum Zone
in the upper part of the Substeueroceras koeneni Zone in the Malargüe (~ 0.8 Ma) correlates with the upper part of the Occitanica Standard
anticline. In the rest of the biozones the difference in the number of cy- Zone, and the Spiticeras damesi Zone (~1.7 Ma) with the Boisieri Stan-
cles is not greater than 2 high frequency eccentricity cycles. According to dard Zone. The lower Valanginian Neocomites wichmanni and Lissonia
Mitchum and Uliana (1985) seismic stratigraphic scheme, a significant riveroi Zones contain 2.5 low-frequency eccentricity cycles, so that its
basinward facies-belt migration is observed during the Berriasian, that duration is about 1 Ma.
could means bypass and/or erosion stages in the Malargüe anticline po- It is important to keep in mind that the number of cycles within each
sition, while in the Sierra Azul area normal depositional conditions ammonite zone may be conditioned by the facies in which they are rep-
prevailed. resented. Particularly, the basal biozones of the Tithonian, which are
There seems to be consensus in the position of lower Valanginian am- found in distal facies of carbonate ramp system, may contain minor dis-
monite zones, where nannofossils get better resolution in the Neuquén continuities that are not detectable by biostratigraphic methods or field
Basin (see Fig. 3). Similarly, different lines of evidence indicate that the evidences, and thus may have longer durations.
Windhauseniceras internispinosum Zone is located at the base of upper The polarities sequence we obtained in the Loncoche section
Tithonian, including the presence of ammonite genus Simplisphinctes shows an evident controversy between the reverse polarity of the
(Zeiss and Leanza, 2008), the first occurrence of the nannofossil Virgatosphinctes mendozanus Zone and the international geomagnetic
Polycostella senaria (Lescano and Kietzmann, 2010; Kietzmann et al., polarity time scale (GPTS), since in the GPTS the uppermost lower
2011b), the first occurrence of the preacalpionellid Chitinoidella boneti Tithonian is represented by a normal polarity spanning the Darwini
(Fernández-Carmona and Riccardi, 1998; Kietzmann et al., 2011a), and and Semiforme Standard Zones. For this reason, in concordance
a change from a reverse to a normal polarity zone (Iglesia Llanos et al., with cyclostratigraphic data, we propose that R1 in the Virgatosphinctes
2013). These two levels are used in this paper as a datum to tie our float- mendozanus Zone could be correlated with the dominant reverse polar-
ing astronomical scale. ity (M21r) of the GPTS located in the base of the Fallauxi Standard Zone
According to our cyclostratigraphic data the minimum durations of (Fig. 11B).
ammonite zones within the Vaca Muerta Formation are summarized On the other hand, the reverse polarity R2 could be correlated with
in Fig. 12. The Virgatosphinctes mendozanus Zone contains two low- the reverse polarity spanning the top of the Fallauxi and part of the
frequency eccentricity cycles, so that its duration is about 0.8 Ma. Fol- Ponti/Peroni Standard Zones in the GPTS (M20r). The R2-N1 change is
lowing the orbital calibration proposed by Gradstein et al. (2004) and probably related with the reverse-normal polarities change that occurs
Graindstein et al. (2012), it would be located at the base of the middle in the Ponti/Peroni Standard Zones in the GPTS, which also converges
Tithonian (upper part of the S. semiforme to lower part of the S. fallauxi with the cyclostratigraphic results. N1 is thus correlated with the dom-
Standard Zone). The Pesudolissoceras zittelli Zone (~ 0.6 Ma) and inant normal polarity located within the Microcanthum Standard Zone
Aulacosphinctes proximus Zone (~0.4 Ma) are located within the middle extending maybe into the Durangites Zone in the GPTS (M20-M19?).
Tithonian (upper S. semiforme to Ponti Standard Zones). The Following N1, we find a large interval with no palaeomagnetic data in
Windhauseniceras internispinosum Zone (~0.7 Ma) would be located at which the Jurassic-Cretaceous boundary would be represented and is
the base of the upper Tithonian (M. microcanthum Standard Zone), and our most urgent target. Finally, N2 is here correlated with the dominant
the Corongoceras alternans Zone (~ 1 Ma) would coincide with the normal polarity zone that spans the Boissieri Standard Zone in the GPTS
upper Tithonian (M. microcanthum Standard Zone and lower part of (M16n-M15n?). Although the polarities sequence we obtained in the
the Durangites Standard Zone). Arroyo Loncoche section is preliminary, it is clearly very consistent
In our astronomical scale the Jurassic–Cretaceous boundary is with the polarities pattern found particularly in the northern hemi-
located within the Substeueroceras koeneni Zone (~2.9 Ma), that would sphere, and is a strong incentive for our present and future work.
D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46 43

In recent years detrital zircon datings (Naipauer et al., 2012, 2014; part of the Berriasian (upper S. occitanica to S. boissieri Standard
Rossel et al., 2014; Vennari et al., 2014) have questioned the age of the Zones), since the biostratigraphic data presented by these authors are
Upper Jurassic and Lower Cretaceous Andean successions, and particular- not conclusive for the position of the Jurassic–Cretaceous boundary,
ly the Tithonian–Berriasian age of the Vaca Muerta Formation and equiv- and the A. noduliferum Zone is probably younger than early Berriasian
alent units. Regardless of the absolute ages, several lines of evidence, (see Fig. 12).
particularly ammonite, nannofossil, calpionellid and dynoflagellate bio- It is hoped in the future to have new absolute age data, which to-
stratigraphy, support a Tithonian–Valanginian age for these units, so it gether with magnetostratigraphic studies and systematic studies of cal-
would be negligent to assume a very different age. Nevertheless, there careous microfossils, and cyclostratigraphic analysis allow a better
are clear disagreements between the absolute ages and the accepted geo- calibration of the Tithonian–Berriasian, and providing more clarity to
logic time scale. this interesting interval in the Neuquén Basin.
Likewise, the duration of the Tithonian stage is still debated. In
the timescale proposed by Gradstein et al. (1994) the Tithonian has a
8. Conclusions
duration of 6.5 Ma, whereas that Pálfy et al. (2000) estimate a duration
of 8.7 Ma. Ogg (2004) proposes a duration of 5.3 Ma, and in the
The Vaca Muerta Formation is characterized by decimetre-scale
latest proposal 7.1 Ma (Ogg and Hinnov, 2012a,b). The duration of
rhythmic alternations of marls, shales and limestones, and consists of
8.7 Ma is regarded with skepticism, since the Late Jurassic isotopic
five facies associations, which represent basinal to middle carbonate
data are still scarce and chronological estimates were made using
ramp settings.
magnetostratigraphic data interpolations (Pálfy et al., 2000).
The Vaca Muerta Formation has a well-ordered hierarchy of cycles,
Considering the duration of the Tithonian between 5.3 and 7.1 Ma,
including elementary cycles (20–40 cm), bundles (3–10 m) and
this stage should contain between 252 and 338 elementary cycles,
superbundles (8–20 m). Based on spectral analysis, elementary cycles
53 to 71 bundles, and 13 to 17 superbundles. Therefore, the Jurassic–
have a periodicity of 21 ky, which correlates with the precession cycle
Cretaceous boundary in the Neuquén Basin would be within the
of Earth's axis; bundles and superbundles have preriodicities of
Substeueroceras koeneni Zone, regardless of the accepted time-length
90–120 ky and 400 ky, which we interpret as the modulation of the pre-
(Fig. 12). These results are consistent with the biostratigraphic scheme
cessional cycle by the Earth's orbital eccentricity.
proposed by Riccardi (2008) and Riccardi et al. (2000, 2011). On the
Cycles are probably driven by change in carbonate exportation, as
other hand, our floating orbital scale shows some differences in the cor-
changes in shallow-water carbonate production involve changes in car-
relation with standard ammonite biozones according to ICS-2012 and
bonate basinward exportation. Productivity and fertility mechanisms
ICS-2004 scales. Using the base of the Windhauseniceras internispinosum
are not adequate to explain the cyclicity in the Vaca Muerta Formation,
and Neocomites wichmanni ammonite Andean zones as datum levels
since sedimentation shows no strong pelagic imprints.
to tie our orbital scale, we show a better correlation with the scale
Cyclostratigraphic data allowed us to build a floating orbital scale for
ICS-2004 (Fig. 12).
the lower? Tithonian–lower Valanginian interval in the Neuquén Basin,
The floating astronomical scale presented in this paper is consistent
which in conjunction with preliminary paleomagnetic data from the
with: 1) The lack of marine lowermost Tithonian; 2) The middle
Arroyo Loncoche section, allowed us to obtain the following correlation
Tithonian position of the Pseudolissoceras zitteli and Aulacosphinctes
of ammonite zones: The Virgatosphinctes mendozanus Ammonite Zone
proximus Zones; 3) The presence of Saccocoma acme (middle Tithonian)
is probably located at the base of the middle Tithonian (upper part of
in the Pseudolissoceras zittelli and Aulacosphinctes proximus Zones, and its
the Semiforme to Fallauxi Standard Zone). The Pesudolissoceras zittelli
last occurrence within the Corongoceras alternans Zone; 3) The presence
Zone and Aulacosphinctes proximus Zone are located within the middle
of Dobeniella and Longicollaria in the base of the Vaca Muerta Formation
Tithonian (Fallauxi to Ponti Standard Zones). The Windhauseniceras
and Chitinoidella bonetti in the Windhauseniceras internispinosum Zone,
internispinosum Zone would be located at the base of the upper
as well as the presence of large forms of Calpionela alpina Lorenz,
Tithonian (Microcanthum Standard Zone), and the Corongoceras
Crassicollaria sp. and Tintinnopsella sp. within the Corongoceras alternans
alternans Zone coincides with the upper Tithonian (Microcanthum
Zone and the lowermost part of the Substeueroceras koeneni Zones.
Standard Zone and lower part of the Durangites Standard Zone).
In particular, the orbital calibration of the Tihtonian–Valanginian
The Jurassic–Cretaceous boundary is located within the
Andean succesion has implications on two key points: 1) A younger
Substeueroceras koeneni Zone that would correlate with the Durangites
age is assigned to the base of the Vaca Muerta Formation, and 2) the
Standard Zone and the lower part of the Occitanica Standard Zone. The
position of the Jurassic–Cretaceous is discussed.
Argentiniceras noduliferum Zone correlates with the upper part of the
With respect to the base of the Vaca Muerta Formation
Occitanica Standard Zone, and the Spiticeras damesi Zone with the
(V. mendozanus Zone), it appears to be younger than traditionally ac-
Boisieri Standard Zone.
cepted in previous works (e.g., Leanza, 1981, 1996; Riccardi, 1984,
Orbital calibration of these sections is consistent with biostrati-
1988, 2008; Riccardi et al., 2000, 2011; Vennari et al., 2014). A middle
graphic scheme of Riccardi et al. (2000, 2011), wich place the Jurassic-
Tithonian age is assigned rather than an early Tithonian age; these re-
Cretaceous boundary within the Substeueroceras koeneni ammonite
sults are also consistent with the lower Tithonian age for the upper
Zone.
Tordillo Formation suggested by Leanza (1993), so the correlation of
ammonite zones along the peri-Pacific region should be analyzed with
new and detailed studies. Acknowledgements
With respect to the Jurassic-Cretaceous boundary in the Neuquén
Basin, the astronomical calibration of Andean ammonite zones allows The authors are especially grateful to Alberto C. Riccardi
us to place this boundary at the base of the Substeueroceras koeneni (Universidad Nacional de La Plata y Museo, Argentina) for ammonite
Zone, which is consistent with the biostratigraphic scheme of Riccardi identification, and also to Susana Damborenea (Universidad Nacional
et al. (2000, 2011), and coincide with Salazar Soto (2012), who studied de La Plata y Museo, Argentina) for bivalve identification and Miguel
the Tithonian-Valanginian of Chile, and placed the S. koeneni Zone in the Manceñido (Universidad Nacional de La Plata y Museo, Argentina) for
lower Berriasian. In this context, the absolute age of ~139.58 Ma provid- brachiopod identification. We thank Beatriz Bádenas and Roberto A.
ed by Vennari et al. (2014) does not imply that the Jurassic–Cretaceous Scasso for their comments and suggestions that improved significantly
boundary is 5 Ma younger than accepted by the International Commis- the manuscript. This research was supported by the Consejo Nacional
sion on Stratigraphy. The tuff level dated at Las Loicas section (~30 km de Investigaciones Científicas y Técnicas (PIP-0546), the Agencia
east of the Bardas Blancas section) could corresponds to the upper Nacional de Promoción Científica y Tecnológica (PICT-2542) and the
44 D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46

Universidad de Buenos Aires (UBACyT 01/XC28). This is the contribu- Embry, A.F., Johannessen, E.P., 1992. T–R sequence stratigraphy, facies analysis and
reservoir distribution in the uppermost Triassic-Lower Jurassic succession, western
tion R-145 of the Instituto de Estudios Andinos “Don Pablo Groeber”. Sverdrup Basin, Arctic Canada. In: Vorren, T.O., Bergsager, E., Dahl-Stamnes, O.A.,
Holter, E., Johansen, B., Lie, E., Lund, T.B. (Eds.), Arctic geology and petroleum poten-
tial. Norwegian Petroleum Society, Special Publication 2, pp. 121–146.
References Fernández-Carmona, J., Riccardi, A.C., 1998. First record of Chitinoidella Doben in the
Tithonian of Argentina. 10 Congreso Latinoamericano de Geología y 6 Congreso
Aguirre-Urreta, M.P., Pazos, P.J., Fanning, M., Litvak, V., 2008. First U-Pb SHRIMP age of Nacional de Geología Económica, Actas 1, Buenos Aires, p. 292.
the Hauterivian stage, Neuquén Basin, Argentina. Journal of South American Earth Fernández-Carmona, J., Álvarez, P.P., Aguirre-Urreta, M.B., 1996. Calpionélidos calcáreos y
Sciences 26, 91–99. grupos incertae sedis en la Formación Vaca Muerta (Tithoniano superior), alta cordi-
Aguirre-Urreta, B., Lazo, D.G., Griffin, M., Vennari, V.V., Parras, A.M., Cataldo, C., llera mendocina, Argentina. 13° Congreso Geológico Argentino y 3° Congreso de
Garberoglio, R., Luci, L., 2011. Megainvertebrados del Cretácico y su importancia Exploración de Hidrocarburos, Actas 5, Mendoza, p. 225.
bioestratigráfica. In: Leanza, H.A., Arregui, C., Carbone, O., Danieli, J.C., Vallés, J.M. Fischer, A.G., 1964. The Loffer cyclothems of the Alpine Triassic. Geological Survey of
(Eds.), Geología y Recursos Naturales de la Provincia del Neuquén. Asociación Kansas Bulletin 169, 107–149.
Geológica Argentina, Buenos Aires, pp. 465–488. Fischer, A.G., D'Argenio, B., Premoli Silva, I., Weissert, H., Ferreri, V., 2004. Cyclostratigraphic
Anderson, E.J., 2004. The cyclic hierarchy of the “Purbeckian” Sierra del Pozo Section, approach to Earth's history: An introduction. In: D'Argenio, B., Fischer, A.G., Premoli
Lower Cretaceous (Berriasian), southern Spain. Sedimentology 51, 455–477. Silva, I., Weissert, H., Ferreri, V. (Eds.), Cyclostratigraphy: Approaches and case histories.
Arthur, M.A., Dean, W.E., Stow, D.A.V., 1984. Models for the deposition of Mesozoic fine- Society of Economic Paleontologists and Mineralogists, Special Publication 81, pp. 5–16.
grained organic-carbon-rich sediment in the deep sea. In: Stow, D.A.V., Piper, D.J.W. Giambiagi, L., Bechis, F., Lanés, S., Tunik, M., García, V., Suriano, J., Mescua, J., 2008.
(Eds.), Fine-grained sediments: deep-water processes and facies. Geological Society Formación y evolución triásico-jurásica del Depocentro Atuel, Cuenca Neuquina,
of London, Special Publication 15, pp. 527–560. provincia de Mendoza. Revista de la Asociación Geológica Argentina 63,
Bádenas, B., Aurell, M., 2001. Proximal-distal facies relationships and sedimentary pro- 520–533.
cesses in a storm dominated carbonate ramp (Kimmeridgian, northwest of the Iberi- Goldhammer, R.K., Harris, M.T., 1989. Eustatic controls on the stratigraphy and geometry
an Range, Spain). Sedimentary Geology 139, 319–340. of the Latemar buildup (Middle Triassic), the Dolomites of northern Italy. In: Crevello,
Bádenas, B., Aurell, M., Rodríguez-Tovar, F.J., Pardo-Igúzquiza, E., 2003. Sequences stratig- P.D., Wilson, J.L., Sarg, J.F., Read, J.F. (Eds.), Controls on carbonate platform and basin
raphy and bedding rhythms of an outer ramp limestones succession (Late development. Society of Economic Paleontologists and Mineralogists, Special Publica-
Kimmeridgian, Northeast Spain). Sedimentary Geology 161, 153–174. tion 44, pp. 323–338.
Ballent, S.C., Ronchi, D.I., Angelozzi, G.N., 2004. Microfósiles calcáreos tithonianos Goldhammer, R.K., Dunn, P.A., Hardie, L.A., 1990. Depositional cycles, composite sea-level
(Jurásico superior) en el sector oriental de la cuenca Neuquina, Argentina. changes, cycle stacking patterns, and the hierarchy of stratigraphic forcing. Geological
Ameghiniana 41, 13–24. Society of America Bulletin 102, 535–562.
Ballent, S., Concheyro, A., Náñez, C., Pujana, I., Lescano, M., Carignano, A.P., Caramés, A., Gradstein, F.M., Ogg, J.G., 2012. The chronostratigraphic scale. In: Gradstein, F.M., Ogg, J.G.,
Angelozzi, G., Ronchi, D., 2011. Microfósiles mesozoicos y cenozoicos. In: Leanza, Schmitz, M.D., Ogg, G.M. (Eds.), The geologic time scale. Elsevier, Oxford, pp. 31–42.
H.A., Arregui, C., Carbone, O., Danieli, J.C., Vallés, J.M. (Eds.), Geología y Recursos Gradstein, F.M., Agterberg, F.P., Ogg, J.G., Hardenbol, J., Van Veen, P., Thierry, J., Huang, Z.,
Naturales de la Provincia del Neuquén. Asociación Geológica Argentina, Buenos 1994. A Mesozoic time scale. Journal of Geophysical Research 99, 24051–24074.
Aires, pp. 489–528. Gradstein, F.M., Ogg, J.G., Smith, A.G., 2004. A geologic time scale. Cambridge University
Balog, A., Read, J.F., Haas, J., 1997. Late Triassic Milankovitch cycle record of a Hungarian Press, Cambridge, p. 589.
marine carbonate platform compared with record from Italian Alps and United Gradstein, F.M., Ogg, J.G., Schmitz, M.D., Ogg, G.M., 2012. The geologic time scale. Elsevier,
States rift basins. American Association of Petroleum Geologists Annual Convention, Oxford, p. 1144.
Houston. Groeber, P., 1946. Observaciones geológicas a lo largo del meridiano 70. Hoja Chos Malal.
Berger, A., 1978. Long-term variations of daily insolation and Quaternary climatic chang- Revista Sociedad Geológica Argentina 1, 178–208.
es. Journal of the Atmospheric Sciences 35, 2362–2367. Groeber, P., 1953. Ándico. In: Groeber, P., Stipanicic, P.N., Mingramm, A. (Eds.), Geografía
Berger, A., Loutre, M.F., 1994. Astronomical forcing through geological time. In: de Boer, de la República Argentina. Sociedad Argentina de Estudios Geográficos 2, Buenos
P.L., Smith, D.G. (Eds.), Orbital forcing and cyclic sequences. International Association Aires, pp. 349–351.
of Sedimentologists, Special Publication 19, pp. 15–24. Grötzinger, J.P., 1986. Cyclicity and paleoenvironment dynamics, Rocknest platform,
Berger, A., Loutre, M.F., Dehant, V., 1989. Astronomical frequencies for pre-Quaternary Rocknest Formation, Wopmay, Oregon, Northwest Territories, Canada. GSA Bulletin
paleoclimate studies. Terra Nova 1, 474–479. 97, 1208–1231.
Berger, A., Loutre, M.F., Laskar, J., 1992. Stability of the astronomical frequencies over the Gulisano, C., Gutierrez Pleimling, R., Digregorio, R.E., 1984. Análisis estratigráfico del
Earth's history for paleoclimate studies. Science 255, 560–566. intervalo Tithoniano-Valanginiano (Formaciones Vaca Muerta, Quintuco y
Bown, P., Concheyro, A., 2004. Lower Cretaceous calcareous nannoplankton from the Mulichinco) en el suroeste de la provincia del Neuquén. 9° Congreso Geologico
Neuquén Basin, Argentina. Marine Micropaleontology 52, 51–84. Argentino, Actas 1, Bariloche, pp. 221–235.
Burchette, T.P., Wright, V.P., 1992. Carbonate ramp depositional systems. Sedimentary Hardie, L.A., Bosellini, A., Goldhammer, R.K., 1986. Repeated subaerial exposure of subtidal
Geology 79, 3–57. carbonate platforms, Triassic, northern Italy: evidence for high-frequency sea-level
Carozzi, A.V., Bercowski, F., Rodriguez, M., Sanchez, M., Vonesch, T., 1981. Estudio de oscillations on a 104 year scale. Paleoceanography 1, 447–457.
microfacies de la Formación Chachao (Valanginiano), Provincia de Mendoza. 8 Hess, H., 2002. Remains of Saccocomids (Crinoidea: Echinodermata) from the
Congreso Geológico Argentino, Actas 2, San Luis, pp. 545–565. Upper Jurassic of southern Germany. Stuttgarter Beiträge zur Naturkunde B 329,
Carozzi, A.V., Orchuela, I.A., Rodriguez Schelotto, M.L., 1993. Depositional models of the 1–57.
Lower Cretaceous Quintuco-Loma Montosa Formation, Neuquén Basin, Argentina. Hilgen, F.J., Iaccarino, S., Krijgsman, W., Villa, G., Langereis, C.G., Zachariasse, W.J., 2000.
Journal of Petroleum Geology 16, 421–450. The global boundary stratotype and point (GSSP) of the Messinian stage (uppermost
Charrier, R., 1985. Estratigrafía, evolución tectónica y significado de las discordancias de Miocene). Episodes 23, 1–6.
los Andes chilenos entre 32°S y 36°S durante el Mesozoico y Cenozoico. In: Frutos, Hilgen, F.J., Abdul Aziz, H., Krijgsman, W., Raffi, I., Turco, E., 2003. Integrated stratigraphy
J., Oyarzún, R., Pincheira, M. (Eds.), Geologia y Recursos Minerales de Chile. and astronomical tuning of the Serravallian and lower Tortonian at Monte dei Corvi
Universidad de Concepción, Concepción, pp. 101–133. (Middle-Upper Miocene, northern Italy). Palaeogeography, Palaeoclimatology,
Colombié, C., Strasser, A., 2003. Depositional sequences in the Kimmeridgian of the Palaeoecology 199, 229–264.
Vocontian Basin (France) controlled by carbonate export from shallow-water plat- Hilgen, F.J., Brinkhuis, H., Zachariasse, W.J., 2006. Unit stratotypes for global stages: the
forms. Geobios 36, 675–683. Neogene perspective. Earth-Science Reviews 74, 113–125.
Concheyro, A., Palma, R.M., Lescano, M., López Gómez, J., Chivelet, J.M., Kietzmann, D.A., Hinnov, L.A., Oggs, J.G., 2007. Cyclostratigraphy and the astronomical time scale. Stratigra-
2006. Nanofósiles calcáreos en los episodios de productividad y dilución de la phy 4, 239–251.
Formación Vaca Muerta. 9° Congreso Argentino de Paleontología y Bioestratigrafía, Hinnov, L.A., Park, J., 1999. Strategies for assessing Early-Middle (Pliensbachian-Aalenian) Ju-
Córdoba, p. 218. rassic cyclochronologies. In: Shackleton, N.J., Mc Cave, I.N., Weedon, G.P. (Eds.),
D'Argenio, B., Ferreri, V., Iorio, M., Raspini, A., Tarling, D.H., 1999. Diagenesis and rema- A Discussion: Astronomical (Milankovitch) Calibration of the Geological Timescale, Phil-
nence acquisition in the Cretaceous carbonates of Monte Raggeto Southern Italy. In: osophical Transactions of the Royal Society, London, Series A 357, pp. 1831–1859.
Tarling, D.H., Turner, P. (Eds.), Palaeomagnetism Diagenesis in Sediments. House, M.R., Gale, A.S., 1995. Orbital forcing timescales and cyclostratigraphy. Geological
Geolological Society of London, Special Publication 151, pp. 147–156. Society of London, Special Publication 85, 1–210.
D'Argenio, B., Fischer, A.G., Premoli Silva, I., Weissert, H., Ferreri, V., 2004. Husinec, A., Read, J.F., 2007. The Late Jurassic Tithonian, a greenhouse phase in the Middle
Cyclostratigraphy: Approaches and case histories. Society of Economic Paleontolo- Jurassic–Early Cretaceous “cool” mode: Evidence from the cyclic Adriatic platform,
gists and Mineralogists. Special Publication 81, 1–311. Croatia. Sedimentology 54, 317–337.
Einsele, G., 1982. Limestone–Marl cycles (periodicities): diagnosis, significance, causes – Hüsing, S.K., Hilgen, F.J., Abdul Aziz, H., Krijgsman, W., 2007. Completing the Neogene
A review. In: Einsele, G., Seilacher, A. (Eds.), Cyclic and event stratification. Springer, geological time scale between 8.5 and 12.5 Ma. Earth and Planetary Science Letters
Berlin, pp. 8–53. 253, 340–358.
Einsele, G., Ricken, W., 1991. Limestone-Marl alternation – an overview. In: Einsele, G., Husson, D., Galbrun, B., Laskar, J., Hinnov, L.A., Thibault, N., Gardin, S., Locklair, R.E., 2011.
Ricken, W., Seilacher, A. (Eds.), Cycles and events in stratigraphy. Springer, Berlin, Astronomical calibration of the Maastrichtian (Late Cretaceous). Earth and Planetary
Heidelberg, New York, pp. 23–47. Science Letters 305, 328–340.
Einsele, G., Ricken, W., Seilacher, A., 1991. Cycles and events in stratigraphy. Springer, Iglesia Llanos, M.P., Palma, R.M., Kietzmann, D.A., 2013. First magnetostratigraphic
Berlin, Heidelberg, New York, pp. 1–955. results of the Upper Jurassic-Lower Cretaceous Vaca Muerta Formation, Neuquén
Elrick, M., 1995. Cyclostratigraphy of Middle Devonian carbonates in the eastern Great Basin, Argentina. Latinmag Letters 3. Latin American Association of Paleomagne-
Basin. Journal of Sedimentary Research B65, 61–79. tism and Geomagnetism, Montevideo, OB022, pp. 1–2.
D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46 45

Kauffman, E.G., Elder, W.P., Sageman, B.B., 1991. High-resolution correlation: a new MacEachern, J.A., Kerrie, L.B., Pemberton, S.G., Gingras, M.K., 2008. The ichnofacies para-
tool in chronostratigraphy. In: Einsele, G., Ricken, W., Seilacher, A. (Eds.), Cycles digm: high-resolution paleoenvironment interpretation of the rock record. In:
and events in stratigraphy. Springer, Berlin, pp. 795–819. MacEachern, J.A., Bann, K.L., Gingras, M.K., Pemberton, S.G. (Eds.), Applied
Kerans, Ch., Tinker, S., 1997. Sequence stratigraphy and characterization of carbonate res- ichnologySociety of Economic Paleontologists and Mineralogists, Short Course
ervoirs. Society of Economic Paleontologists and Mineralogists, Short Course Notes Notes. American Association of Petroleum Geologists, Tulsa, pp. 27–64.
40, pp. 1–128. Manceda, R., Figueroa, D., 1993. La inversión del rift mesozoico de la faja fallada y plegada
Kietzmann, D.A., Palma, R.M., 2009a. Tafofacies y biofacies de Formación Vaca de Malargüe. Provincia de Mendoza. 12 Congreso Geologico Argentino y 2 Congreso
Muerta en el sector surmendocino de la Cuenca Neuquina: implicancias de Exploración de Hidrocarburos, Actas 3, Mendoza, pp. 219–232.
paleoecológicas, sedimentológicas y estratigráficas. Ameghiniana 46, 321–343. Martín-Chivelet, J., Palma, R.M., López-Gómez, J., Kietzmann, D.A., 2011. Earthquake-
Kietzmann, D.A., Palma, R.M., 2009b. Microcrinoideos saccocómidos en el Tithoniano de la induced soft-deformation structures in Upper Jurassic open-marine microbialites
Cuenca Neuquina. ¿Una presencia inesperada fuera de la región del Tethys? (Neuquén Basin, Argentina). Sedimentary Geology 235, 2010–2221.
Ameghiniana 46, 695–700. Mitchum, R.M., Uliana, M., 1985. Seismic stratigraphy of carbonate depositional se-
Kietzmann, D.A., Palma, R.M., 2010. New crustacean microcoprolites from the Lower quences, Upper Jurassic-Lower Cretaceous, Neuquén Basin, Argentina. In: Berg, B.R.,
Cretaceous (middle Berriasian - lower Valanginian) of the Neuquén Basin, south- Woolverton, D.G. (Eds.), Seismic Stratigraphy 2. An integrated approach to hydrocar-
ern Mendoza, Argentina. Journal of South American Earth Sciences 30, 58–64. bon analysis American Association of Petroleum Geologists, Memoir 39. American As-
Kietzmann, D.A., Palma, R.M., 2011. Las tempestitas peloidales de la Formación Vaca sociation of Petroleum Geologists, Tulsa, pp. 255–283.
Muerta (Tithoniano-Valanginiano) en el sector surmendocino de la Cuenca Molinie, A.J., Ogg, J.G., 1992. Milankovitch Cycles in Upper Jurassic and Lower Cretaceous
Neuquina, Argentina. Latin American Journal of Sedimentology and Basin Analysis radiolarites of the Equatorial Pacific: Spectral analysis and sedimentation rate curves.
18, 121–149. Proceedings. Ocean Drilling Program. Scientific Results 129, 529–547.
Kietzmann, D.A., Palma, R.M., 2014. Early Cretaceous crustacean microcoprolites from Naipauer, M., García Morabito, E., Marques, J.C., Tunik, M., Rojas Vera, E., Vujovich, G.I.,
Sierra de la Cara Cura, Neuquén Basin, Argentina: Taphonomy, environmental distri- Pimentel, M., Ramos, V.A., 2012. Intraplate Late Jurassic deformation and exhumation
bution, and stratigraphic correlation. Cretaceous Research 49, 214–228. in western central Argentina: constraints from surface data and U-Pb detrital zircon
Kietzmann, D.A., Vennari, V.V., 2013. Sedimentología y estratigrafía de la Formación Vaca ages. Tectonophysics 524–525, 59–75.
Muerta (Tithoniano-Berriasiano) en el área del cerro Domuyo, norte de Neuquén, Naipauer, M., Tunik, M., Marques, J., Rojas Vera, E.A., Vujovich, G., Pimentel, M., Ramos,
Argentina. Andean Geology 40, 41–65. V.A., 2014. U–Pb detrital zircon ages of Upper Jurassic continental successions: impli-
Kietzmann, D.A., Palma, R.M., Bressan, G.S., 2008. Facies y microfacies de la rampa cations for the provenance and absolute age of the Jurassic–Cretaceous boundary in
tithoniana-berriasiana de la Cuenca Neuquina (Formación Vaca Muerta) en la sección the Neuquén Basin. In: Sepúlveda, S.A., Giambiagi, L.B., Moreiras, S.M., Pinto, L.,
del arroyo Loncoche – Malargüe, provincia de Mendoza. Revista de la Asociación Tunik, M., Hoke, G.D., Farías, M. (Eds.), Geodynamic processes in the Andes of central
Geológica Argentina 63, 696–713. Chile and Argentina. Geological Society, London, Special Publications 399. http://dx.
Kietzmann, D.A., Blau, J., Fernández, D.E., Palma, R.M., 2010. Crustacean microcoprolites doi.org/10.1144/SP399.1.
from the Upper Jurassic – Lower Cretaceous of the Neuquén Basin, Argentina: Ogg, J.G., 2004. The Jurassic Period. In: Gradstein, F.M., Ogg, J.G., Smith, A.G. (Eds.), A geo-
Systematics and biostratigraphic implications. Acta Palaeontologica Polonica 55, logic time scale. Cambridge University Press, Cambridge, pp. 307–343.
277–284. Ogg, J.G., Hinnov, L., 2012a. The Jurassic Period. In: Gradstein, F.M., Ogg, J.G., Schmitz, M.,
Kietzmann, D.A., Martín-Chivelet, J., Palma, R.M., López-Gómez, J., Lescano, M., Concheyro, Ogg, G. (Eds.), The geologic time scale 2012. Elsevier, Amsterdam, pp. 731–792.
A., 2011a. Evidence of precessional and eccentricity orbital cycles in a Tithonian Ogg, J.G., Hinnov, L., 2012b. The Cretaceous Period. In: Gradstein, F.M., Ogg, J.G.,
source rock: the mid-outer carbonate ramp of the Vaca Muerta Formation, Northern Schmitz, M., Ogg, G. (Eds.), The geologic time scale 2012. Elsevier, Amsterdam,
Neuquén Basin, Argentina. American Association of Petroleum Geologists Bulletin 95, pp. 793–854.
1459–1474. Ozawa, T., 2006. Milankovitch 41000-year cycles in lithofacies and molluscan content
Kietzmann, D.A., Blau, J., Riccardi, A.C., Palma, R.M., 2011b. An interesting finding of in the tropical Middle Miocene Nyalindung Formation, Jawa, Indonesia.
chitinoidellids (Calpionellidea Bonet) in the Jurassic-Cretaceous boundary of Palaeogeography, Palaeoclimatology, Palaeoecology 235, 382–405.
the Neuquén Basin. 18 Congreso Geológico Argentino, Actas CD, Neuquén, Pálfy, J., Smith, P.L., Mortensen, J.K., 2000. A U-Pb and 40Ar/39Ar time scale for the
pp. 1480–1481. Jurassic. Canadian Journal of Earth Sciences 37, 923–944.
Kietzmann, D.A., Palma, R.M., Martín-Chivelet, J., López-Gómez, J., 2014. Sedimentol- Pardo-Igúzquiza, E., Rodríguez-Tovar, F., 2004. POWGRAF 2: a program for graphical spec-
ogy and sequence stratigraphy of a Tithonian-Valanginian carbonate ramp (Vaca tral analysis in cyclostratigraphy. Computers and Geosciences 30, 533–542.
Muerta Formation): a misunderstood exceptional source rock in the Southern Park, M.H., Fürsich, F., 2001. Cyclic nature of lamination in the Tithonian Solnhofen
Mendoza area of the Neuquén Basin, Argentina. Sedimentary Geology 302, Plattenkalk of southern Germany and its palaeoclimatic implications. International
64–86. Journal of Earth Sciences 90, 847–854.
Leanza, H.A., 1981. Faunas de ammonites del Jurásico y Cretácico inferior de América Pemberton, S.G., van Wagoner, J.C., Wach, G.D., 1992. Ichnofacies of a wave-dominated
del Sur, con especial consideración de la Argentina. In: Volkheimer, W., shoreline. In: Pemberton, S.G. (Ed.), Application of ichnology to petroleum explora-
Musacchio, E. (Eds.), Cuencas Sedimentarias de América del Sur, Buenos Aires, tion. Society of Economic Palaeontologists and Mineralogists, Core Workshop 17,
pp. 559–597. Tulsa, pp. 339–382.
Leanza, H.A., 1993. Estratigrafía del Mesozoico posterior a los Movimientos Intermálmicos Pittet, B., Strasser, A., 1998. Depositional sequences in deep-shelf environments
en la comarca del Cerro Chachil, provincia de Neuquén. Revista de la Asociación formed through carbonate mud import from the shallow platform (late
Geológica Argentina 48, 71–84. Oxfordian, German Swabian Alb and eastern Swiss Jura). Eclogae Geologicae
Leanza, H.A., 1996. Advances in the ammonite zonation around the Jurassic/Cretaceous Helvetiae 91, 149–169.
boundary in the Andean Realm and correlation with Tethys. Jost Wiedmann Sympo- Prell, W.L., Hays, J.D., 1976. Late Pleistocene faunal and temperature patterns of the
sium, Abstracts, Tübingen,pp. 215–219. Colombia Basin, Caribbean Sea. In: Cline, R.M., Hays, J.D. (Eds.), Investigation of late
Leanza, H.A., Sattler, F., Martinez, R., Carbone, O., 2011. La Formación Vaca Muerta y Quaternary peleoceanography and paleoclimatology. Geological Socity of America
equivalentes (Jurásico Tardío – Cretácico Temprano) en la Cuenca. Neuquina. In: Memoir 145, pp. 201–220.
Leanza, H.A., Arregui, C., Carbone, O., Danieli, J.C., Vallés, J.M. (Eds.), Geología y Premoli Silva, I., Erba, E., Tornaghi, I., 1989. Paleoenvironmental signals and changes in
Recursos Naturales de la Provincia del Neuquén. Asociación Geológica Argentina, surface fertility in Mid-Cretaceous Corg-rich facies of the Fucoid Marls (central
Buenos Aires, pp. 113–129. Italy). Geobios 11, 225–236.
Legarreta, L., Gulisano, C.A., 1989. Análisis estratigráfico de la Cuenca Neuquina Quattrocchio, M.E., Martínez, M.A., García, V.M., Zavala, C.A., 2003. Palinoestrtigrafía del
(Triásico Superior-Terciario Inferior). In: Chebli, G.A., Spalletti, L.A. (Eds.), Tithoniano-Hauteriviano del centro-oeste de la Cuenca Neuquina, Argentina. Revista
Cuencas Sedimentarias ArgentinasSerie Correlación Geológica 6. Universidad Española de Micropaleontología 354, 51–74.
Nacional de Tucumán, Tucumán, pp. 221–243. Ramos, V.A., 2010. The tectonic regime along the Andes: Present-day and Mesozoic re-
Legarreta, L., Uliana, M.A., 1991. Jurassic–Cretaceous Marine Oscillations and Geom- gimes. Geological Journal 45, 2–25.
etry of Back Arc Basin, Central Argentina Andes. In: McDonald, D.I.M. (Ed.), Sea Ramos, V.A., Folguera, A., 2005. Tectonic evolution of the Andes of Neuquén: constraints
level changes at active plate margins: Process and product. International Associ- derived from the magmatic arc and Foreland deformation. In: Veiga, G.D., Spalletti,
ation of Sedimentologists, Special Publication 12, pp. 429–450. L.A., Howell, J.A., Schwarz, E. (Eds.), The Neuquén Basin, Argentina: A case study in
Legarreta, L., Uliana, M.A., 1996. The Jurassic succession in west central Argentina: sequence stratigraphy and basin dynamics. Geological Society of London, Special
stratal patterns, sequences, and paleogeographic evolution. Palaeogeography, Publications 252, pp. 15–35.
Palaeoclimatology, Palaeoecology 120, 303–330. Raspini, A., 2001. Stacking pattern of cyclic carbonate platform strata: Lower Creta-
Legarreta, L., Kozlowski, E., Boll, A., 1981. Esquema estratigráfico y distribución de fa- ceous of southern Appennines, Italy. Journal of the Geological Society 158,
cies del Grupo Mendoza en el ámbito surmendocino de la cuenca neuquina. 8 353–366.
Congreso Geológico Argentino, Actas 3, San Luis, pp. 389–409. Riccardi, A.C., 1984. Las asociaciones de amonitas del Jurásico y Cretácico de la Argentina.
Lehmann, C., Osleger, D.A., Montañez, I.P., 1998. Controls on cyclostratigraphy of Lower IX Congreso Geológico Argentino, Actas 4, Buenos Aires, pp. 559–595.
Cretaceous carbonates and evaporites, Cupido and Coahuila platforms, Northeastern Riccardi, A.C., 1988. The Cretaceous System of southern South America. Geological Society
Mexico. Journal of Sedimentary Research 68, 1109–1130. of America, Memoir 168. Geological Society of America, Boulder.
Lescano, M., Kietzmann, D.A., 2010. Nanofósiles Calcáreos de la Formación Vaca Muerta Riccardi, A.C., 2008. The marine Jurassic of Argentina: a biostratigraphic framework.
(Tithoniano inferior- Valanginiano inferior) en la región sudoccidental de la Provincia Episodes 31, 326–335.
de Mendoza. 10 Congreso Argentino de Paleontología y Bioestratigrafía y 7 Congreso Riccardi, A.C., Leanza, H.A., Damborenea, S.E., Manceñido, M.O., Ballent, S.C., Zeiss, A.,
Latinoamericano de Paleontología, Actas, La Plata, p. 94. 2000. Marine Mesozoic Biostratigraphy of the Neuquén Basin. Zeitschrift für
Lourens, L., Hilgen, F., Shackleton, N.J., Laskar, J., Wilson, D., 2004. The Neogene Period. In: Angewandte Geologie, Sonderheft 1, 103–108.
Gradstein, F., Ogg, J., Smith, A. (Eds.), A geologic time scale 2004. Cambridge Univer- Riccardi, A.C., Damborenea, S.E., Manceñido, M.O., Leanza, H.A., 2011. Megainvertebrados
sity Press, Cambridge, pp. 409–440. jurásicos y su importancia geobiológica. In: Leanza, H.A., Arregui, C., Carbone, O.,
46 D.A. Kietzmann et al. / Sedimentary Geology 315 (2015) 29–46

Danieli, J.C., Vallés, J.M. (Eds.), Geología y Recursos Naturales de la Provincia del Mountains). In: de Boer, P.L., Smith, D.G. (Eds.), Orbital forcing and cyclic sequences.
Neuquén. Asociación Geológica Argentina, Buenos Aires, pp. 441–464. International Association of Sedimentologists, Special Publication 19, pp. 285–301.
Rossel, P., Oliveros, V., Mescua, J., Tapia, F., Ducea, M.N., Calderón, S., Charrier, R., Hoffman, Strasser, A., Hillgärtner, H., 1998. High-frequency sea-level fluctuation recorded on a shal-
D., 2014. The Upper Jurassic volcanism of the Río Damas-Tordillo Formation (33°-35. low carbonate platform (Berriasian and Lower Valanginian of Mount Saleve, French
5°S): Insights on petrogenesis, chronology, provenance and tectonic implications. Jura). Eclogae Geologicae Helvetiae 91, 375–390.
Andean Geology 41, 529–557. Strasser, A., Hillgärtner, H., Pasquier, J.B., 2004. Cyclostratigraphic timing of sedimentary
Sagasti, G., 2000. La sucesión rítmica de la Formación Agrio (Cretácico Inferior) en el sur processes: An example from the Berriasian of the Swiss and French Jura Mountains.
de la provincia de Mendoza, y su posible vinculación con Ciclos de Milankovitch. In: D'Argenio, B., Fischer, A.G., Premoli Silva, I., Weissert, H., Ferreri, V. (Eds.),
Revista de la Asociación Argentina de Sedimentologia 7, 1–22. Cyclostratigraphy: Approaches and case histories. Society of Economic Paleontolo-
Sagasti, G., 2005. Hemipelagic record of orbitally-induced dilution cycles in Lower Creta- gists and Mineralogists, Special Publication 81, pp. 135–151.
ceous sediments of the Neuquén Basin. In: Veiga, G.D., Spalletti, L.A., Howell, J.A., Tedesco, L.P., Wanless, H.R., 1991. Generation of sedimentary fabrics and facies by repet-
Schwarz, E. (Eds.), The Neuquén Basin, Argentina: A Case Study in Sequence Stratig- itive excavation and storm infilling of burrow networks: Holocene of south Florida
raphy and Basin Dynamics. Geological Society of London, Special Publications 252, and Caicos Platform, B.W.I. Palaios 6, 326–343.
pp. 231–250. Van Couvering, J.A., Castradori, D., Cita, M.B., Hilgen, F.J., Rio, D., 2000. The base of the
Salazar Soto, C.A., 2012. The Jurassic-Cretaceous Boundary (Tithonian - Hauterivian) in Zanclean Stage and of the Pliocene Series. Episodes 23, 179–187.
the Andean Basin of Central Chile: Ammonites, Bio- and Sequence Stratigraphy and Van Wagoner, J.C., Mitchum, R.M., Campion, K.M., Rahmanian, V.D., 1990. Siliciclastic se-
Palaeobiogeography(PhD. Thesis) Universität Heidelberg, Heidelberg. quence stratigraphy in well logs, cores, and outcrops: Concepts for high-resolution
Scasso, R.A., Alonso, S.M., Lanés, S., Villar, H.J., Lippai, H., 2002. Petrología y geoquímica de correlation of time and facies. American Association of Petroleum Geologists,
una ritmita marga-caliza del Hemisferio Austral: El Miembro Los Catutos (Formación Methods in Exploration Series 7, 1–55.
Vaca Muerta), Tithoniano medio de la Cuenca Neuquina. Revista de la Asociación Vennari, V.V., Lescano, M., Naipauer, M., Aguirre-Urreta, B., Concheyro, A., Schaltegger, U.,
Geológica Argentina 57, 143–159. Armstrong, R., Pimentel, M., Ramos, V.A., 2014. New constraints on the Jurassic-
Scasso, R.A., Alonso, S.M., Lanés, S., Villar, H.J., Lippai, H., 2005. Geochemistry and petrol- Cretaceous boundary in the High Andes using high-precision U-Pb data. Gondwana
ogy of a Middle Tithonian limestone-marl rhythmite in the Neuquén Basin, Research 26, 374–385.
Argentina: depositional and burial history. In: Veiga, G.D., Spalletti, L.A., Howell, J.A., Vergani, G.D., Tankard, A.J., Belotti, H.J., Welkink, H.J., 1995. Tectonic evolution and paleo-
Schwarz, E. (Eds.), The Neuquén Basin, Argentina: A case study in sequence geography of the Neuquén Basin, Argentina. In: Tankard, A.J., Suarez Soruco, R.,
stratigraphy and basin dynamics. Geological Society of London, Special Publication Welsink, H.J. (Eds.), Petroleum Basins of South America. American Association of Pe-
252, pp. 207–229. troleum Geologists, Memoir 62. American Association of Petroleum Geologists, Tulsa,
Schieber, J., Southard, J.B., Schimmelmann, A., 2010. Lenticular shale fabrics resulting from pp. 383–402.
intermittent erosion of water-rich muds - interpreting the rock record in the light of Weedon, G., 2003. Time–series analysis and cyclostratigraphy. Examining stratigraphic
recent flume experiments. Journal of Sedimentary Research 80, 119–128. record of environmental cycles. Cambridge University Press, New York, pp. 1–259.
Schwarzacher, W., 1947. Über die Sedimentäre Rhythmik des Dachsteinkalkes von Lofer. Weedon, G.P., Jenkyns, H.C., Coe, A.L., Hesselbo, S.P., 1999. Astronomical calibration of
Geologische Bundesanstalt, Verhandlungen H10–12, 175–188. the Jurassic time scale from cyclostratigraphy in British mudrock formations. Philo-
Schwarzacher, W., 1993. Cyclostratigraphy and the Milankovitch theory. Elsevier, sophical Transactions of the Royal Society A: Mathematical, Physical and Engineering
Amsterdam, pp. 1–224. Sciences 357, 1787–1813.
Schwarzacher, W., 2000. Repetition and cycles in stratigraphy. Earth-Science Reviews 50, Westphal, H., Hilgen, F., Munnecke, A., 2010. An assessment of the suitability of individual
51–75. rhythmic carbonate successions for astrochronological application. Earth-Science
Seibold, E., 1952. Chemische Untersuchungen zur Bankung im unteren Malm Schwabens. Reviews 99, 19–30.
Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 95, 337–370. Zeiss, A., Leanza, H.A., 2008. Interesting new ammonites from the Upper Jurassic of
Stipanicic, P.N., 1969. El avance en los conocimientos del Jurásico argentino a partir del Argentina and their correlation potential: new possibilities for global correlations at
esquema de Groeber. Revista de la Asociación Geológica Argentina 24, 367–388. the base of the Upper Tithonian by ammonites, calpionellids and other fossil groups.
Strasser, A., 1994. Milankovitch cyclicity and high-resolution sequence stratigraphy Newsletters on Stratigraphy 42, 223–247.
in lagoonal-peritidal carbonates (upper Tithonian–lower Berriasian, French Jura

You might also like