ArticlePDF Available

Integrated source rock evaluation along the maturation gradient. Application to the Vaca Muerta Formation, Neuquén Basin of Argentina

Authors:

Abstract

The Vaca Muerta Formation (Tithonian–early Valanginian) is the main source rock in the Neuquén Basin and the most important unconventional shale resource in South America. In the present study, organic geochemistry, electron microscopy and basin and petroleum system modelling (BPSM) were combined to evaluate source rock properties and related processes along a transect from the early oil (east) to the dry gas (west) window. The unit is characterised by high present‐day (1–8% average) and original (2–16% average) total organic carbon contents, which increase toward the base of the unit and toward basinal (west) settings. Scanning electron microscopy shows that organic pores derived from the transformation of type II kerogen. Isolated bubble pores are typical of the oil window, whereas bubble and densely distributed spongy pores occur in the gas stage, indicating that the maturity gradient exerts strong control on organic porosity. Organic geochemistry, pressure and porosity data were incorporated into a 2‐D basin petroleum system model that includes the sequential restoration of tectonic events and calculation of compaction trends, kerogen transformation, hydrocarbon generation and estimation of pore pressure through geologic time. The W–E regional model extends from the Agrio Fold and Thrust Belts to the basin border and allows us to evaluate the relationship between thermal maturity and timing of hydrocarbon generation from highly deformed (west) to undeformed (east) regions. Modelling results show a clear decrease of maturity and OM transformation toward the eastern basin margin. Maximum hydrocarbon generation occurred in the inner sectors of the belt, at ~120 Ma; long before the first Andean compression phase, which started during the Late Cretaceous (~70 Ma). Miocene compression (15 to 7 Ma) promoted tectonic uplift of the inner and outer sectors of the belt associated with a reduction of thermal stress and kerogen cracking, as well as massive loss of retained fluids and a decrease in pore pressure. The organic matter transformation impacted (1) the magnitude of effective porosity associated with organic porosity development, and (2) the magnitude and distribution of pore pressure within the unit controlled by hydrocarbon generation and compaction disequilibrium. BPSM shows a progressive increase in effective porosity from the top to the base and toward the west region related to the original organic carbon content and maturity increasing along the same trend. Overpressure intervals with high organic carbon contents are the most prone to develop organic pores. The latter represent favorable sites for the storage of hydrocarbons in the Vaca Muerta Formation.
Basin Research. 2021;00:1–29.
|
1
EAGE
wileyonlinelibrary.com/journal/bre
Received: 29 March 2021
|
Revised: 8 July 2021
|
Accepted: 25 July 2021
DOI: 10.1111/bre.12599
RESEARCH ARTICLE
Integrated source rock evaluation along the maturation
gradient.Application to the Vaca Muerta Formation,
NeuquénBasin ofArgentina
J. B.Spacapan1
|
M.Comerio1
|
I.Brisson2
|
E.Rocha3
|
M.Cipollone1
|
J.C.Hidalgo4
© 2021 International Association of Sedimentologists and European Association of Geoscientists and Engineers and John Wiley & Sons Ltd
1YPF- Tecnología S.A. (Y- TEC- CONICET),
Buenos Aires, Argentina
2YPF S.A., Buenos Aires, Argentina
3Pluspetrol S.A., Buenos Aires, Argentina
4Schulmberger GmbH, Aachen, Germany
Correspondence
J. B. Spacapan, YPF- Tecnología S.A.
(Y- TEC- CONICET), Av. del Petróleo
Argentino (RP10) S/N, Berisso (CP 1923),
Buenos Aires, Argentina.
Email: juan.b.spacapan@ypftecnologia.com
Abstract
The Vaca Muerta Formation (Tithonian– early Valanginian) is the main source rock
in the Neuquén Basin and the most important unconventional shale resource in South
America. In the present study, organic geochemistry, electron microscopy and basin
and petroleum system modelling (BPSM) were combined to evaluate source rock
properties and related processes along a transect from the early oil (east) to the dry
gas (west) window. The unit is characterized by high present- day (1%– 8% average)
and original (2%– 16% average) total organic carbon contents, which increase to-
wards the base of the unit and basinal (west) settings. Scanning electron micros-
copy shows that organic pores derived from the transformation of type II kerogen.
Isolated bubble pores are typical of the oil window, whereas bubble and densely
distributed spongy pores occur in the gas stage, indicating that the maturity gradient
exerts strong control on organic porosity. Organic geochemistry, pressure and poros-
ity data were incorporated into a 2D basin petroleum system model that includes the
sequential restoration of tectonic events and calculation of compaction trends, kero-
gen transformation, hydrocarbon generation and estimation of pore pressure through
geologic time. The W– E regional model extends from the Agrio Fold and Thrust
Belts to the basin border and allows us to evaluate the relationship between ther-
mal maturity and timing of hydrocarbon generation from highly deformed (west) to
undeformed (east) regions. Modelling results show a clear decrease in maturity and
organic matter (OM) transformation towards the eastern basin margin. Maximum
hydrocarbon generation occurred in the inner sectors of the belt, at ca. 120Ma; long
before the first Andean compression phase, which started during the Late Cretaceous
(ca. 70Ma). Miocene compression (15– 7Ma) promoted tectonic uplift of the inner
and outer sectors of the belt associated with a reduction in thermal stress and kerogen
cracking, as well as massive loss of retained fluids and a decrease in pore pressure.
The OM transformation impacted (a) the magnitude of effective porosity associated
with organic porosity development, and (b) the magnitude and distribution of pore
pressure within the unit controlled by hydrocarbon generation and compaction dis-
equilibrium. BPSM shows a progressive increase in effective porosity from the top
2
|
EAGE
SPACAPAN et Al.
1
|
INTRODUCTION
Recognition of source rocks for oil and gas is the most criti-
cal risk element in petroleum exploration activities (Curiale
& Curtis,2016; Tissot & Welte,1984). Source rock evalua-
tion consists of assessing hydrocarbon generating potential,
organic matter (OM) type, hydrocarbons that might be gen-
erated and thermal maturity (Dembicki, 2009; Katz, 1995;
Tissot & Welte, 1984). Unconventional plays are usually
analysed in terms of average total organic carbon (TOC)
content, mineralogy, porosity, permeability, geomechanical
properties and pore pressure (Ejofodomi etal., 2011; Hazra
etal.,2019; Passey etal.,2010; Peters etal.,2017; Wang &
Gale,2009). At basin- scale, however, studies that focus on
the assessment and quantification of these parameters are still
scarce (Romero- Sarmiento etal.,2013).
Analytical techniques and numerical models including
basin and petroleum system modelling (BPSM) packages
are widely used tools to evaluate conventional and uncon-
ventional resources (Al- Hajeri et al., 2009; Hantschel &
Kauerauf, 2009; McCarthy etal., 2011; Mei etal., 2021;
Peters et al., 2017; Romero- Sarmiento et al., 2013). The
analysis of unconventional resources based on estimation
of burial temperature history and variations in maturity
gradients allows simulation of a wide variety of geological
processes through time, such as: (a) compaction trends, (b)
transformation of OM by thermal maturity, (c) generation–
retention– expulsion of hydrocarbons, (d) hydrocarbon
composition, (e) overpressure and fracturing mechanisms
and (f) preservation– development of porosity (Al- Hajeri
et al., 2009; Burgreen- Chan et al., 2015; Grohmann
etal.,2021; Hantschel & Kauerauf,2009; Mei etal.,2021;
Peters etal., 2017, 2018; Romero- Sarmiento et al., 2013;
Tissot & Welte, 1984). Recent studies integrate BPSM
tools with laboratory analysis to perform semi- quantitative
descriptions of processes that control hydrocarbon gener-
ation, expulsion and retention capacity (Mei etal., 2021;
Romero- Sarmiento etal.,2013, 2014). For instance, BPSM
of the Mississippian Barnett Shale (Fort Worth Basin,
USA) provided a quantitative means to estimate the total
volume of generated hydrocarbons and the distribution of
organic porosity at basin scale assuming that overpressure
accompanied by diagenetic precipitation of cements in-
hibited the effects of compaction (Romero- Sarmiento
etal.,2013). In that unit, model results and petrographic and
analytical studies indicated that organic pores are formed
by thermal stress and represent the main sites for hydrocar-
bon trapping (Loucks etal.,2009; Reed & Loucks,2015;
Romero- Sarmiento etal.,2014).
The Meso– Cenozoic Neuquén Basin in northern Patagonia
is one of the most important hydrocarbon basins in South
America (Legarreta etal.,2008; Uliana & Legarreta,1993;
Urien & Zambrano,1994; Veiga etal.,2020) and, in recent
years, it has become well known for its unconventional oil
and gas resources (e.g. Badessich etal., 2016; Legarreta &
Villar,2011; Sagasti et al., 2014). The basin has three ma-
rine organic- rich source rocks known as the Los Molles,
Vaca Muerta and Agrio formations (Fms). In particular, the
Vaca Muerta Fm (Tithonian– lower Valanginian) constitutes
a world- class source rock and an unconventional shale re-
source for both oil and gas (Brisson etal., 2020; Legarreta
& Uliana,1991; Legarreta etal., 2008; Veiga etal., 2020).
In recent years, a great number of sedimentological, palae-
ontological, geochemical and stratigraphic works have been
to the base and towards the west region related to the original organic carbon content
and maturity increasing along the same trend. Overpressure intervals with high or-
ganic carbon contents are the most prone to develop organic pores. The latter repre-
sent favourable sites for the storage of hydrocarbons in the Vaca Muerta Formation.
KEYWORDS
hydrocarbon generation, modelling, organic porosity, pore pressure, thermal maturation, Vaca
Muerta
Highlights
Evaluation of the Vaca Muerta Fm based on or-
ganic geochemistry, petrography and petroleum
system modelling.
Timing of hydrocarbon generation linked to tec-
tonic events and burial depth.
TOC content increases towards the base of the
unit and basinal (west) settings.
Organic pores developed from the early oil to the
dry gas window derived from the transformation
of type II kerogen.
Hydrocarbon generation and compaction disequi-
librium are the main mechanisms that generate
pore pressure.
|
3
EAGE
SPACAPAN et Al.
reported for the unit (see Section 2). However, regional
basin- scale modelling integrated with scanning electron mi-
croscopy (SEM), organic geochemistry and pore pressure
measures has not been published.
In order to evaluate the source rock potential of the Vaca
Muerta Fm (VMFm) through geological time, a transect in the
central part of the Neuquén Basin (37° S) was analysed using
a regional 2D petroleum system model that extends from the
Agrio Fold and Thrust Belt in the west to the basin border in
the east (Figure1). BPSM involves sequential restoration of
the deformation, calibrated by published thermochronolog-
ical dating, which allowed us to integrate regional tectonic
events and processes that controlled the thermal maturity and
generation and expulsion of hydrocarbons. In addition, the
present work incorporated key parameters, such as original
total organic carbon (TOC0) content, porosity and pore pres-
sure conditions, in a basin simulator to test the magnitude
and distribution of organic porosity (Øorg) and evaluate the
causes of overpressure in the VMFm through the geological
evolution of the Neuquén Basin. The characterization of such
processes provides a partial evaluation of the unconventional
potential of the unit. Such characterization, integrated with
geomechanical properties, has a direct impact on the strate-
gies of hydraulic fracturing (Peters etal.,2017).
Thus, the aims of the present article are to: (a) establish
the original and present- day TOC distribution and transfor-
mation gradients for the VMFm at different basin positions
(from dry gas to the early oil window) using a robust Rock-
Eval® pyrolysis data set, (b) characterize the origin and types
of organic pores (Øorg) as function of thermal maturation via
high- resolution 2D- SEM, (c) obtain a semi- quantitative eval-
uation of the unit using 2D regional BPSM to test the timing
of hydrocarbon generation– expulsion from the Fold Belt in
the west to the basin border in the east and (d) test kerogen
transformation into hydrocarbons, porosity development and
the magnitude of pore pressure to evaluate the unconven-
tional potential of the Vaca Muerta source rock.
2
|
GEOLOGICAL SETTING
The Meso- Cenozoic Neuquén Basin (Figure1), located in the
Northern Patagonian Andes, extends between 32° and 40° S
latitude, covering an area of 120,000km2 and includes parts
FIGURE 1 The Neuquén Basin of west-
central Argentina. Satellite image showing
the internal reference areas for the VMFm
(Brisson etal.,2020) and the positions of
analysed wells within the Agrio (Agrio-
Section) fold and thrust belt
4
|
EAGE
SPACAPAN et Al.
of the Mendoza, Neuquén, Rio Negro and La Pampa prov-
inces, Argentina (Howell etal.,2005; Vergani etal.,1995).
Three stages of basin infill are summarized in Figure2. The
rift basin stage was characterized by several NW– SE rift
depocentres formed during the Late Triassic– Early Jurassic
with continental and volcanic deposits of the Precuyo Group
(Franzese & Spalletti, 2001). A long thermal subsidence,
interrupted by localized tectonic events, characterizes the
retro- arc basin stage developed from the Early Jurassic to
Early Cretaceous with thick marine and continental succes-
sions composed of carbonate, siliciclastic and evaporite fa-
cies (Howell etal.,2005; Legarreta & Uliana,1991; Vergani
et al., 1995). The lithostratigraphic framework includes
the Cuyo, Lotena, Mendoza and Bajada del Agrio Groups
(Groeber, 1946; Leanza, 2003; Stipanicic et al., 1968;
Weaver,1931), where marine organic- rich source rocks re-
lated to long- term Palaeo- Pacific transgressions are recorded:
Los Molles, Vaca Muerta and Agrio Fms (Figure2). Finally,
the foreland basin stage in Late Cretaceous– Palaeocene
time was linked to contractional conditions and associated
with the Andean uplift (Horton, 2018). Sedimentation was
dominated by synorogenic red bed successions (Neuquén
Group) and the first shallow marine units (Malargüe Group)
from the Atlantic Ocean (Aguirre- Urreta etal.,2011; Tunik
et al., 2010). This stage was characterized by progressive
growth of S– N fold and thrust belts (e.g. Agrio, Chos Malal
and Malargüe) and reactivation of older extensional faults
(Horton etal.,2016; Rojas Vera etal.,2015).
The VMFm, the focus of this study, represents a widespread
marine transgression during the retro- arc stage, mainly com-
posed of organic- rich black shales, marls and limestones reach-
ing 800m thickness (Legarreta & Uliana,1991; Weaver,1931).
Based on ammonoid biozones, partially calibrated through
high- precision U- Pb ages, the unit ranged from Tithonian–
early Valanginian (e.g. Aguirre- Urreta et al., 2007; Naipauer
et al., 2020; Vennari et al., 2014). Along the S– N fold and
thrust belts, basinal- to- middle ramp deposits are well exposed
and have been studied by detail facies analysis (e.g. Kietzmann
etal.,2016; Otharán etal.,2020; Scasso etal.,2005; Spalletti
et al., 2000). In the subsurface of the Neuquén Embayment,
well logs, cores and seismic data provided additional infor-
mation on the unit (e.g. Desjardins et al., 2016; Dominguez
& Di Benedetto, 2019; Legarreta & Uliana, 1991; Mitchum
& Uliana,1985; Sagasti etal.,2014). Basin- scale stratigraphic
architecture shows that the unit comprises three to five pro-
gradational sequences with organic- rich transgressive inter-
vals composed of siliciclastic and carbonate- rich shales that
grade upward into highstands dominated by organic- lean
marls (Capelli etal.,2021; Dominguez etal.,2016; Kietzmann
etal.,2016; Legarreta & Villar,2015).
2.1
|
Source rocks: An overview
Many studies have analysed the regional characteristics of the
petroleum systems in the Neuquén Basin (Boll etal., 2014;
Cruz et al., 1996, 2002; Karg & Littke, 2020; Kozlowski
FIGURE 2 The classic stratigraphic chart for the Neuquén Basin
(modified from Howell etal.,2005). The VMFm (Mendoza Group)
represents the main source rock/unconventional target in the basin
|
5
EAGE
SPACAPAN et Al.
etal.,1998; Legarreta etal.,2008; Mei etal.,2021; Rocha
etal.,2018; Uliana & Legarreta,1993; Vergani etal.,2011).
The three marine source units are the Los Molles, Vaca
Muerta and Agrio Fms, which differ in terms of organic
richness, OM type, maturity and distribution throughout
the basin. The oldest source rock was deposited during the
Pleinsbachian– Callovian and covers, in the same way as the
VMFm, almost the entire basin. The unit has present- day
TOC ranging from 1% to 5% with kerogen types II (algal/
amorphous macerals) and III (palynomorphs/phytoclasts
macerals), showing a transition from immature to overma-
ture stages (Jorgensen etal.,2013; Legarreta & Villar,2011;
Martínez etal.,2008). The thickness increases from approxi-
mately 500– 900 m following the main NW– SE axis of the
basin. However, important differences in thickness are re-
lated to inherited topography of the underlying rift depocent-
ers (Legarreta & Villar,2011; Olivera etal.,2020). Based on
TOC contents, the lowermost part (ca. 200m thick) shows
the most suitable characteristics for a possible unconven-
tional target (Jorgensen etal.,2013).
Based on organic geochemistry and petrological stud-
ies, the VMFm exhibits mainly amorphous (liptinitic mac-
erals) type II kerogen with TOC content between 3% and
8% (peak values of 12%– 20%) suitable for producing oil or
gas condensate, depending on position in the basin (Brisson
etal.,2020; Karg & Littke,2020; Legarreta & Villar,2011;
Sylwan,2014). The hydrocarbon yield expressed as original
hydrogen index (HI0) averages 680mgHC/gTOC based on the
analysis of immature samples (Brisson et al., 2020; Veiga
etal.,2020). This is representative of the organic- rich inter-
vals (30– 450 m thick) considered to be the unconventional
targets of the unit (Dominguez etal.,2016). Different patterns
of organic richness, hydrocarbon source quality, distribution
of free hydrocarbons and thermal maturity allowed Brisson
etal.(2020) to distinguish six reference areas for the VMFm
(Figure1). Moderate depth of 3,000m and overpressure con-
ditions also make the unit a prime target for drilling opera-
tions and stimulation treatments which favour development
and commercialization of this play (Badessich etal., 2016;
Veiga etal.,2020).
The Agrio Fm includes two organic- rich intervals of
late Valanginian and late Hauterivian age (e.g. Comerio
etal.,2018; Pazos etal.,2020; Uliana & Legarreta, 1993).
TOC content ranges from 1% to 5% with peak values of up to
16%, including type II and III kerogens (Comerio etal.,2018,
2020; Legarreta & Villar, 2012), with average measured
HI0 of 600 mgHC/gTOC for the type II intervals (Spacapan
etal.,2018). Carbonate- rich shales are more enriched in type
II kerogen than the siliciclastic shales, which are dominated
by type III kerogen (Comerio etal., 2018). Maturity maps
indicate that both intervals range in maturity from the oil
window to the gas window (Cruz etal., 1996; Legarreta &
Villar, 2012), reaching dry gas and overmature stages in areas
where intrusions are present (Spacapan etal., 2018, 2020).
Compared with VMFm, the Agrio Fm shows lower amounts
of OM, lower effective thickness (<100 m thick), and re-
duced regional distribution.
3
|
METHODOLOGY
The VMFm was analysed using a W– E section (Figure 1)
along the Agrio fold and thrust belt (Agrio FTB). It includes
information from 27 wells distributed throughout the section
and available surface and seismic data. The present study
shows organic geochemistry results and scanning electron
microscopy images from cuttings and core samples from
Well A (2,312– 3,010m), Well B (2,610– 2,790m) and Well
C (2,270– 2,430 m) on the Dorso de Los Chihuidos– NE
Platform (Figure 1). We followed the stratigraphic scheme
of Legarreta and Villar (2015) that subdivided the VMFm
in lower, middle and upper VM to describe and interpret the
results.
3.1
|
Organic geochemistry
A total of 220 cuttings and core samples were analysed for
present- day total organic carbon (TOC wt%) and Rock- Eval®
programmed pyrolysis following standard procedures from
the IFP Energies Nouvelles (Espitalié etal.,1986). Samples
were solvent extracted to remove oil mud residues. The use of
oil- based mud restricts the thorough evaluation of the unit be-
cause a quantitative measurement of free hydrocarbons (e.g.
pyrolysis S1 peak, mgHC/gROCK) is impossible, and many
ratios used to interpret and evaluate the source beds cannot
be employed in the analysis (Brisson etal.,2020). There are
many equations used to calculate the original TOC (TOC0) in
source rocks (e.g. Brisson etal.,2020; Chen & Jiang,2016;
Modica & Lapierre,2012; Peters etal.,2017). This study con-
siders the equation specific for the VMFm, which was calcu-
lated based on an extensive pyrolysis data from cores and
cutting samples from nearly 900 wells (see details in Brisson
et al., 2020). These authors indicated a homogenous char-
acter of the type II kerogen and obtained an average of HI0
of 680mgHC/gTOC based on measures of immature samples
(Brisson et al., 2020). The pyrolysis S2 (mgHC/gROCK) and
S3 (
mgCO2gTOC
) peaks allowed us to corroborate kerogen
types and thermal maturity based on the Tmax (Peters,1986).
Values of vitrinite reflectance (%Ro) were measured for Well
A and Well B; however, for Well C (early oil window), the
theoretical %Ro was calculated from Tmax values (Brisson
etal., 2020) due to the lack of vitrinite particles. The equa-
tion proposed by Brisson etal.(2020) is valid mainly for Tmax
between 410 and 470 and was derived from a strong data-
set. The usage of the equation in other source rocks should
6
|
EAGE
SPACAPAN et Al.
be tested first due to the relationship between Tmax and %Ro
appears to be kerogen- type dependent (Katz & Lin,2021). In
addition, measured %Ro data from proximal wells and ma-
turity maps were also considered to corroborate the maturity
trend of Well C. Transformation ratio (TR) was calculated
following the methodology of Waples and Tobey (2015),
which expresses the converted mass fraction of the initial re-
actant, that is, the ratio of petroleum formed from kerogen
in source rock to the total amount of petroleum that could
be formed from that kerogen (Peters etal., 2017; Tissot &
Welte,1984).
3.2
|
Scanning electron microscopy analysis
To examine microtextural features and evaluate the ther-
mal transformation of OM, 22 core samples were prepared
on thin sections normal to bedding planes, mechanically
polished, ion- milled and analysed by field emission scan-
ning electronic microscope (FE- SEM) at the Laboratorio
de Analítica in YPF- Tecnología (Y- TEC) following in-
ternal procedures. Samples were coated with carbon and
studied at different magnifications (from 150× to up to
200,000×) under secondary electron (SE) and backscat-
tered diffraction (BSD) modes. Elemental composition
of mineral phases was determined by energy- dispersive
X- ray spectroscopy (EDS) microanalysis. Pore- size dis-
tribution was measured at different magnifications using
the JMicrovision© (Roduit,2008) program and consider-
ing OM pores as spherical to oval with bubble and spongy
morphologies (Ko etal.,2017; Löhr etal.,2015; Milliken
etal.,2013; Pommer & Milliken,2015). FE- SEM resolu-
tion is in the order of 10nm; and consequently, macropores
(>50nm) are well defined. However, part of the mesopores
(2– 50 nm) and micropores (<2 nm) are not measurable
(pore sizes as defined by Rouquerol etal.,1994). Elongate
pores with crack- like morphology were not considered
because of possible generation during sample preparation
(Katz & Arango,2018; Löhr etal.,2015; Schieber,2013).
4
|
PETROLEUM SYSTEM
MODELLING
Schlumberger 2- D PetroMod® software allows full integra-
tion of tectonic events, structural restoration, facies variation
and source rock maturity through time. The software was
used for sequential restoration of the Agrio regional cross
section (Figures1 and 3). To calibrate the BPSM and evalu-
ate the history of hydrocarbon generation for the analysed
source rocks, organic geochemistry and %Ro data from the
VM, Los Molles and Agrio Fms were compiled (YPF data-
base). The age, erosion thickness, lithologies, unit thickness
and petroleum system elements for Well A and Well C are
included in Supplementary Material 1.
4.1
|
Modelling input
One- dimensional (1- D) and 2D BPSMs were performed using
PetroMod® Teclink® software (Hantschel & Kauerauf,2009),
including coupled structural restoration and backstripping tools.
This technique integrates the structural and the petroleum sys-
tems model, which is essential for simulating hydrocarbon
generation, migration and accumulation in tectonically com-
plex areas. Software calculations in tectonically complex areas
are analysed in detail by Hantschel and Kauerauf (2009) and
Burgreen- Chan et al.(2015). PetroMod® software provides a
database to estimate the surface– water interface temperature
(SWIT). The upper boundary condition for temperature calcu-
lations is determined by the sediment– water surface (onshore)
temperatures (Hantschel & Kauerauf,2009; Peters etal.,2017).
In the model, SWIT data were corrected for present and past
water depth (Supplementary Material 2). It uses a global re-
construction of the palaeo- mean surface or air temperatures
and palaeobathymetry to estimate SWIT for a given latitude
(Wygrala,1989). The most important input for predicting hy-
drocarbon generation is the evolution of thermal heat flow
through geologic time, which together with the burial history,
controls the evolution of temperature and chemical reactions
during the generation, migration, accumulation and preservation
of petroleum. Thermal history of the Agrio section was recon-
structed assuming a constant heat flow of 60mW/m2 through
time (Supplementary Material 2). Heat flow was calibrated
with %Ro data and later cross- checked with Horner- corrected
bottom- hole temperature (Deming,1989). Similar constant heat
flow of 60 mW/m2 was assumed by Lampe et al.(2006) for
the Agrio FTB. The resulting temperature fields provided theo-
retical values for temperature, %Ro, TOC, Øorg and TR, and
were used to calculate the amount of cracked kerogen. Source
rock maturation assessed using the Easy %Ro equation from
Sweeney and Burnham (1990). Lithological properties were
based on Athy's compaction law (Athy,1930) and the multipoint
model (Hantschel & Kauerauf,2009). Athy's law is a traditional
porosity versus depth curve which predicts hydrostatic pressure
based on deposition of the entire column with various litholo-
gies. Additionally, the PetroMod® software allowed to simulate
effective and secondary porosity. Effective porosity is the pore
volume in a rock that contributes to the permeability and does
not include isolated vuggy porosity or water that is bound to
clay minerals (Hantschel & Kauerauf,2009). Secondary poros-
ity in a source rock is the additional porosity that can be gener-
ated by mineral transformations (inorganic secondary porosity)
or transformation of OM into hydrocarbons (organic secondary
porosity). PetroMod® software calculates the Øorg as the cumu-
lative amount of porosity generated during OM transformation.
|
7
EAGE
SPACAPAN et Al.
In addition, it predicts the proportion of gas adsorbed within
OM pores or on mineral surfaces versus free gas in pore spaces
or natural fractures (Peters etal., 2017). Furthermore, the evo-
lution of Øorg and its relationship with the burial history were
analysed through the effective porosity curve. Pore pressure pre-
diction is important for executing a safe drilling strategy and for
accurate production modelling (Couzens- Schultz et al.,2013).
Fluid pressure modelling can be used to improve pore pressure
prediction and reduce the drilling risk posed by unanticipated
overpressures (Peters etal.,2017). Diagnostic fracture injection
testing (DFIT) data obtained from the YPF database were used
to calibrate pressure conditions in the analysed wells.
4.2
|
The Agrio Fold and Thrust Belt (FTB):
Balanced structural section
The Agrio FTB extends from the 38°50′ to 37°35′ S and is
characterized by two sectors with different structural styles
(Vergani etal., 1995; Zamora Valcarce et al.,2007; Zapata
& Folguera,2005). To the west, the inner sector shows base-
ment deformation and tectonic inversion of previous exten-
sional (Triassic) structures (Ramos et al., 2011; Rojas Vera
etal., 2010). The outer sector is dominated by thin- skinned
deformation showing basal detachments within Jurassic and
Lower Cretaceous evaporites and marine shales with a gen-
eral deformation sense to the east and several west- verging
backthrusts (Rojas Vera et al., 2015; Zamora Valcarce
etal.,2011). To the east, the Dorso de Los Chihuidos (DdLC)
is defined by an antiformal structure interpreted as the result
of tectonic inversion of basement involved normal faults
(Ramos et al., 2011). Such differences control the distribu-
tion of outcrops in the central part of the Neuquén Basin with
Jurassic sedimentary units (Cuyo and Lotena Groups) mainly
exposed in the inner sector and Upper Jurassic– Cretaceous
units (Mendoza, Bajada del Agrio and Neuquén Groups) to
the east in the outer sector (Leanza & Hugo,2001; Rojas Vera
etal., 2015; Zamora Valcarce et al.,2007, 2011). Balanced
structural section published by Rocha etal.(2018) was used in
this study, which extends from the Agrio FTB (inner and outer
FIGURE 3 Tectonic evolution for
the Agrio FTB cross section from the Late
Cretaceous to present day (see also Rocha
etal.,2018)
8
|
EAGE
SPACAPAN et Al.
sectors), the DdlC (Well A), external flank of the DdLC (Well
B) to the NE Platform (Well C) (see location in Figures1 and
3). This section was sequentially restored with the aim to rep-
resent the tectonical evolution based on available information.
Initial exhumation occurred in the Late Cretaceous (Rojas
Vera etal.,2015; Tunik etal.,2010), with two peaks of reacti-
vation in the Oligocene– Miocene and middle Miocene based
on apatite fission tracks (Rocha etal.,2018). The Agrio FTB
represents the most important depocentre during the Mesozoic
sedimentation and includes the embayment sector where the
VMFm reached highest maturity (Brisson etal., 2020). The
Mesozoic succession thins northward in the domain of Chos
Malal FTB, a sector known as Chihuido– Lomita as well as
to the basin margin (NE platform) characterized by shallower
burial and lower maturity levels.
5
|
RESULTS
5.1
|
Vaca Muerta organic geochemistry
The results from programmed pyrolysis are summarized in
Table1. The TOC content in all analysed wells is high with
mean values that range from 2.38% (Well A), 5% (Well B), to
3.21% (Well C) in the Agrio section (Figure4). The lower and
middle VM contain the highest TOC contents, with highest
values towards basinal (west) settings: 3.60% and 3.76% (Well
A), 8% and 4.83% (Well B) and 3.36% and 3.05% (Well C).
Well C shows Rock Eval®Tmax values that range from 429
to 443 (mean: 435) and calculated %Ro between 0.65
and 0.90 (mean value: 0.80), which is consistent with the
early oil window (Figure5a). %Ro measures indicate a range
from 1.60– 1.65 (wet gas, Well B) to 2.10– 2.33 (dry gas, Well
A) (Figure4). The S2 peak that coincides with the remaining
potential of hydrocarbon generation decreases towards high-
maturity zones ranging from 15.44mgHC/gROCK (1.34– 43.15
minimum and maximum values) in the early oil (Well C) to
0.68mgHC/gRO CK (0.21– 1.49) in the dry gas zone (Well A).
On the contrary, Tmax displays increased thermal maturity,
reaching mean values of 475 (460– 514 minimum and
maximum values) in the wet gas window (Well B). Tmax can-
not be measured in the dry gas window (Well A) due to very
low values of S2 indicating that hydrocarbon- generative ca-
pacity of the OM has been exhausted.
The mean values of HI also decrease as maturity increases
from 470mgHC/gTOC (238– 654 minimum and maximum val-
ues) in the early oil (Well C) to 41 mgHC/gTOC (6– 107) in
the dry gas window (Well A) for the Agrio section (Table1).
The oil- prone VMFm throughout the basin is supported by
the presence of bacterial/algae OM (Brisson et al., 2020;
Petersen etal., 2020) and is consistent with low values of
oxygen index (OI) in all analysed wells (mean values between
8 and 47
mgCO2gTOC
). However, the uppermost levels of the
upper VM at Well A record an increase in OI (Figure4). The
pseudo- van Krevelen diagram (hydrogen vs. oxygen index,
Figure5b) shows that most samples lie close to the HI- axis
and reflect increasing maturity of the analysed wells with HI
depletion from the early oil to dry gas windows. This explains
the type II kerogen identified in samples from the early- to-
peak oil window. More mature samples in the wet and dry gas
windows plot near the origin of the diagram.
The transformation ratio (TR = [HI0 − HIsample]/
HI0 × 100%, sensu Waples & Tobey,2015) increases with
maturity, showing mean values from 30% (3%– 64%) at Well
C, 97% (94%– 98%) at Well B to 93% (84%– 99%) at Well
A. The equation presented by Brisson etal.(2020) was used
to restore the original total organic carbon TOC0 content
(Table1). For the lower VM, average TOC0 contents range
from: 8.24% (1.10%– 15.50%) at Well A, 18.60% (7.29–
27.41) at Well B to 6.86% (0.56– 14.35) at Well C (Table1).
5.2
|
OM and its related pores under SEM
5.2.1
|
Oil window
Well C samples correspond to the early oil window and with re-
spect to OM two forms are documented (Figure6). Some samples
exhibit filament- like OM that defines a discontinuous wavy-
parallel lamination that is disposed between silt- size detrital
grains, calcareous debris and faecal pellets mainly composed of
calcareous coccoliths (Figure6a– c). These OM domains likely
represent relicts of marine type II kerogen compressed by me-
chanical compaction between rigid detrital grains. In such lev-
els, the OM does not show evidence of pores at SEM resolution.
Conversely, other samples show OM that is intimately bound to
the inorganic matrix composed of coccolith debris, illite/mica
phyllosilicates and quartz, plagioclase and carbonate fragments
(Figure6d,e). The OM mostly fills intergranular pore space and
lacks the filament- like structure. It occupies the interstices be-
tween those detrital components and within primary pores such
as those related to pressure shadows (Schieber,2010), and it
fills microfossil- related voids (Figure6f). Isolated, bubble pores
of variable size (40– 1,000nm, mean 200nm) are documented.
OM of detrital (terrigenous) origin occurs in few samples and in
minor amounts, which is consistent with little land- derived OM
in the VMFm (Brisson etal.,2020; Małachowska etal.,2019;
Petersen et al., 2020). Nonporous OM of terrigenous origin
(e.g. vitrinite and inertinite) is present as rigid grains with
discrete arcuate margins (Milliken et al., 2013; Pommer &
Milliken,2015). Phyllosilicate framework pores with triangu-
lar to elongated geometries (200– 1,400nm wide) are less com-
mon. Microfractures (6– 15µm thick by more than 600µm in
length, Figure 6f) oriented subparallel to bedding planes and
partially filled with secondary products (solid bitumen) are also
documented.
|
9
EAGE
SPACAPAN et Al.
TABLE 1 Summary with organic geochemistry results for the lower, middle and upper informal units of the Vaca Muerta Formation in the Agrio section (general includes all samples of the well)
Vaca
Muerta
units
TOC
%
S1
mg/g
S2
mg/g
S3
mg/g
Tmax
°C
HI=S2/TOC
mg/g
OI=S3/
TOC
mg/g
TR=HI0−HI/
HI0 %
Waples and
Tobey (2015)
TOC0 %
Brisson
etal.(2020)
Well C
(early
oil)
Middle 3.05 (0.44– 7.99) 0.15 (0.11– 0.35) 15.62 (1.34– 43.15) 0.52 (0.28– 1.06) 434 (429– 441) 497 (255– 654) 25 (4– 120) 26 (3– 62) 6.06 (0.42– 17.49)
Lower 3.36 (0.57– 6.39) 0.23 (0.06– 0.91) 15.28 (2.16– 40.18) 0.48 (0.19– 0.87) 437 (429– 443) 448 (238– 628) 17 (5– 45) 34 (7– 64) 6.86 (0.56– 14.35)
General 3.21 (0.44– 7.99) 0.19 (0.06– 0.91) 15.44 (1.34– 43.15) 0.50 (0.19– 1.06) 435 (429– 443) 470 (238– 654) 21 (4– 120) 30 (3– 64) 6.49 (0.42– 17.49)
Well B
(wet gas)
Upper 3.33 (0.40– 7.44) 0.10 (0.06– 0.16) 0.63 (0.13– 1.66) 0.27 (0.21– 0.50) 480 (460– 512) 21 (12– 38) 13 (4– 56) 96 (94– 98) 7.63 (0.87– 17.11)
Middle 4.83 (2.36– 7.10) 0.10 (0.07– 0.16) 0.81 (0.43– 1.26) 0.28 (0.22– 0.34) 466 (460– 476) 17 (12– 20) 6 (4– 11) 97 (97– 98) 11.21 (5.41– 15.90)
Lower 8 (3.18– 12) 0.10 (0.07– 0.15) 1.37 (0.52– 2.67) 0.26 (0.17– 0.33) 477 (466– 514) 17 (13– 24) 4 (2– 7) 97 (96– 98) 18.60 (7.29– 27.41)
General 5 (0.40– 12) 0.10 (0.06– 0.16) 0.87 (0.13– 2.67) 0.27 (0.17– 0.50) 475 (460– 514) 18 (12– 38) 8 (2– 56) 97 (94– 98) 11.50 (0.87– 27.41)
Well A
(dry gas)
Upper 1.75 (0.46– 4.50) 0.19 (0.10– 0.39) 0.71 (0.21– 1.49) 0.76 (0.45– 2.31) Unmeasurable
Tmax
51 (6– 107) 59 (11– 142) 92 (84– 99) 3.92 (0.97– 11.38)
Middle 3.76 (1.47– 5.0) 0.27 (0.16– 0.50) 0.61 (0.35– 0.84) 0.76 (0.53– 1.10) 17 (11– 28) 23 (11– 47) 97 (95– 98) 8.62 (3.32– 11.20)
Lower 3.60 (0.50– 6.74) 0.23 (0.10– 0.36) 0.60 (0.26– 0.91) 0.58 (0.32– 0.80) 28 (11– 60) 28 (11– 72) 95 (91– 98) 8.24 (1.10– 15.50)
General 2.38 (0.46– 6.74) 0.21 (0.10– 0.50) 0.68 (0.21– 1.49) 0.74 (0.32– 2.31) 41 (6– 107) 47 (11– 142) 93 (84– 99) 5.40 (0.97– 15.50)
Note: Minimum, maximum and mean values are presented.
Abbreviations: HI, hydrogen index; HI0, original hydrogen index; OI, oxygen index; S1, S2 and S3, peaks derived from the programed pyrolysis; Tmax, temperature reached in the S2 peak; TOC, total organic carbon; TOC0,
restoration of original TOC; TR, transformation ratio.
10
|
EAGE
SPACAPAN et Al.
|
11
EAGE
SPACAPAN et Al.
5.2.2
|
Gas window
Samples of gas window maturity are dominated by OM of
secondary origin (see Mastalerz et al., 2018; Pommer &
Milliken,2015), which were identified as solid bitumen, the
dominant organic component in thermally mature samples
of the VMFm (Cavelan et al., 2019; Petersen et al., 2020;
Romero- Sarmiento et al., 2017). As mentioned previously,
OM is associated with the mineral matrix (intergranu-
lar space) and fills cavities within bioclasts and fractures
(Figures7 and 8). Compared with the oil window, however,
samples from the gas window show a larger range of pore
sizes. Wet gas samples (Well B) record large (2– 4µm) OM-
hosted bubble pores protected by rigid mineral phases as well
as calcite bioclasts (Figure7a– c). Nevertheless, some elon-
gated bubble pores suggest that compaction impacted OM
pores (Figure7e). OM pervasively fills intergranular spaces
and contains both spongy and bubble pores (Figure 7d–
f), which is typical of the gas window (Ko et al., 2017).
Equivalent samples from the gas window in the VMFm ana-
lysed by Cavelan etal.(2019) also show secondary solid bi-
tumen with bubble and spongy morphologies. In the present
study, SEM reveals that bubble and spongy morphologies
co- exist (Figure 7d– f), between diameters of 4µm– 800nm
(mean 500nm) and 10– 90nm (mean 55nm), respectively.
Samples from the dry gas window (Well A) also show
secondary OM products (solid bitumen) with spongy and
bubble pores (Figure8a– d). However, very small and widely
distributed spongy pores dominate (Figure8e,f). The OM is
confined within matrix cavities and is associated with both
detrital and authigenic clays as well as microcrystalline quartz
and calcite cement (Figure8b– e). In the case of the VMFm,
FIGURE 4 Organic geochemical logs for the analysed wells in the Agrio section. Blue stars (Wells A and B) represent vitrinite reflectance
(%Ro) values and pink stars (Well C) are theoretical %Ro values based on Brisson etal.(2020). TR and TOC0 were calculated following the
methodology of Waples and Tobey (2015) and Brisson etal.(2020)
FIGURE 5 Organic geochemical results for the lower, middle and upper VM (informal stratigraphic units defined by Legarreta &
Villar,2015) in the analysed wells along the Agrio (Well A, Well B and Well C) sections. (a) Hydrogen index (HI) versus calculated and measured
%Ro values. Calculated %Ro based on the equation of Brisson etal.(2020) for Well C (early oil window). Measured %Ro values are given for Well
A and Well B (gas window) and indicated with star markers. (b) HI versus oxygen index (OI) based on the pseudo- van Krevelen diagram
12
|
EAGE
SPACAPAN et Al.
FIGURE 6 SEM images of low- maturity samples (Well C). (a– c) Examples of nonporous organic matter (OM) (marine type II kerogen in the
VMFm). Note compactional features among rigid detrital grains, which in some cases are enclosed by a thin envelope of OM. Faecal pellets composed
of coccoliths and bioclasts (bc) filled with OM. (d, e) General and detail with the secondary organic matter (SOM) that hosts isolated bubble pores.
OM is distributed between detrital grains: Qz (quartz), illite/mica (red arrows) and carbonates (Cal). (f) SOM- filling microfractures and bioclasts (bc)
associated with spar calcite (Cal- s). The insert shows a detail image (scale bar=20µm)
FIGURE 7 SEM images of wet gas window samples (Well B). General (a) and detailed images (b, c) showing secondary organic matter
(SOM) filling matrix pores and bioclasts. The excellent preservation of bubble pores is linked with quartz (Qz) and calcite (Cal) cements and
calcite bioclasts (bc). (d) SOM shows both bubble and spongy pores. Note expanded mica (Mi), detrital (red arrows) and authigenic (yellow arrow)
illite. The insert shows detail of spongy pores (scale bar=2µm). (e) Elongate bubble pores (white arrows). (f) SOM- hosted spongy pores among
diagenetic microcrystalline quartz (Qz), pyrite framboids (Py) and calcite (Cal)
|
13
EAGE
SPACAPAN et Al.
siliceous and calcareous biogenic components are interpreted
to contribute to the precipitation of microquartz and calcite
(Milliken et al., 2019). There is also nonporous, elongate
OM interpreted of terrigenous origin (Figure 8a). Pores as-
sociated with the phyllosilicate framework are represented by
triangular- shaped openings (70– 1,200 nm wide, Figure 8b)
as was documented in overmature samples of the Middle
Devonian Geneseo Fm (Northern Appalachian Basin) by
Wilson and Schieber (2016). In the case of OM- hosted pores,
samples of the VMFm that experienced high- thermal maturity
range in size from 15 to 80nm (mean 40nm) for the spongy
and 100 to 700nm (mean 200nm) for the bubble pores.
6
|
BASIN AND PETROLEUM
SYSTEM MODELLING (BPSM)
The quantification of source rock maturity, hydrocarbon gen-
eration and expulsion, pore pressure distribution and porosity
(primary and secondary) was analysed in the regional 2- D
BPSM (Agrio section). BPSM fully integrated the main rock
parameters, such as kerogen type, TOC0, Øorg, TR, maturity
(%Ro) and tectonic events. The 2D modelling was conducted
for the three effective source rocks in the basin (Los Molles,
Vaca Muerta and Agrio Fms). In particular, this study pre-
sents data used to calibrate those parameters in the VMFm
because the study focuses on characterization of processes
within this unit through geological time.
6.1
|
Well calibration in the Agrio model
The conditions of temperature, OM content, maturity, po-
rosity and pore pressure were calibrated using a robust data
set from wells along the Agrio section (Figures 3 and 9).
Data presented correspond to Well A (DdLC) and Well C
(NE Platform), which represent different geological scenar-
ios, because they are located in different basin regions (see
FIGURE 8 SEM images of dry gas window samples (Well A). (a) Secondary organic matter (SOM)- hosted bubble pores and nonporous
terrigenous OM. (b, c) General and detailed images with SOM- filling pores between detrital (Pl=plagioclase; Qz- d=quartz) and diagenetic
quartz. Phyllosilicate pores (PF- pores) are also documented. In (c) note diagenetic quartz (Qz) and illite of detrital (red arrows) and diagenetic
(yellow arrow) origin. (d) SOM with bubble and spongy pores. The OM is associated with clay minerals of both detrital (red arrow) and diagenetic
(yellow arrows) origins. The insert shows spongy pores (scale bar=500nm). (e) SOM with spongy pores filling a cavity partially cemented
by calcite (Cal) and microcrystalline diagenetic quartz (Qz). There are illite/mica domains of detrital origin (red arrows) and authigenic crystals
(yellow arrows) that grew from detrital crystals. (f) Spongy pores. Note detail in the insert image (scale bar=200nm)
14
|
EAGE
SPACAPAN et Al.
Figure3). The thermal evolution of the model was calibrated
using Horner- corrected borehole temperature data (BHT),
whereas the TOC0 was calculated from pyrolysis data and
used to adjust the evolution of TOC0 for the lower, middle
and upper VM in the model (Figure9). Model and pyrolysis
results show an increase in the TOC0 content from the top
to the base of the unit. %Ro measures were used to calibrate
the thermal maturity, which indicates that the VMFm reached
the dry gas window (Well A, western flank of the Chihuidos
High) and the early oil window (Well C, NE Platform). In
order to calibrate pore pressure, we use data from DFIT.
DFIT is used as an indirect method to estimate pore pres-
sure in unconventional reservoirs (Bakar,2018). According
to DFIT data (red crosses in Figure9), overpressure increases
towards the foreland (west) region. A high pore pressure of
ca. 57MPa is registered at ca. 2,600m depth in Well A and
it is close to the lithostatic pressure, as was also documented
by Berthelon etal.(2021). Well C has a pore pressure of ca.
48 MPa at ca. 2,200 m depth. The effective porosity was
calibrated by using both well log (black crosses in Figure9)
and gas- filled porosity (red dots in Figure9) data (YPF data-
base). Results are consistent with previous studies (Askenazi
etal.,2013; Cuervo etal.,2016; Ortiz etal.,2020) and show
that the effective porosity increases towards the lower VM
(Figure 9). In Well A, this parameter ranges from 5.5% to
12.16% (9.1% average), whereas in Well C effective porosity
ranges from 3.24% to 8.44% (6.25% average).
6.2
|
Hydrocarbon generation in the
Agrio model
Evolution of the TR through time for the VMFm was com-
pared with the Los Molles and Agrio Fms and analysed in
four stages of sequential restoration (Figure 10). During
the Late Cretaceous (ca. 100Ma), TR of 95% suggests that
organic- rich deposits of the Los Molles Fm in the inner sector
reached the advanced dry gas window (Figure10a). From the
inner sector to the eastern flank of the DdLC, the lower VM
would have reached a TR average value of 95%. Expulsion
FIGURE 9 Well calibration of different parameters in the Agrio section for Well C (NE Platform) in (a) and Well A (DdLC) in (b). From left
to right: calculated temperature log (blue line). Black markers are Hornet- corrected borehole temperatures. TOC0 log showing the simulated organic
matter concentration (black curve). Green circles represent TOC0 values obtained from present- day TOC contents. Vitrinite log calculated with the
equation of Sweeney and Burnham (1990) Easy %Ro kinetic algorithm. %Ro vitrinite values (red markers) were used to calibrate the calculated
%Ro curve (black curve). Vertical profile showing the calculated pore pressure magnitude (black line), red crosses are the pore pressure determined
from DFIT tests. Calculated porosity (black curve) was calibrated using measures obtained from well logging (black cross) and gas- filled porosity
(red markers) from core samples. Depth is expressed in depth subsea
|
15
EAGE
SPACAPAN et Al.
started during the Early Cretaceous (ca. 130Ma) at a thermal
maturity of 0.82– 0.90 %Ro. Likewise, the western organic-
rich deposits of the Agrio Fm reached the early oil window,
with a TR of 15% approximately (Figure 10a). The model
predicts that large volumes of hydrocarbons were generated
and expelled from the Los Molles Fm and lower VM. Such
processes were synchronous with the deposition of the Bajada
del Agrio Group and before the formation of the main tectonic
structures in the inner sector of the belt. A marked thickness
increase in the Los Molles and VMFms was predicted in west-
ern areas where the beginning of hydrocarbon generation took
place between the Late Jurassic and Early Cretaceous. From
90 to 60Ma, more than 1,000m of sediments were deposited
(Neuquén and Malargüe Groups), which exhausted the gen-
eration capacity of the Los Molles and VMFms from the inner
sector to the DdLC. These results suggest that gas generation
related to secondary cracking reactions was a relevant factor
in the early development of overpressure in the VMFm.
The first compressional tectonic event started between 70
and 60Ma (Late Cretaceous– Palaeocene) and triggered base-
ment fault inversion in the Agrio FTB and uplifting of the
Cerro Mocho anticline in the inner (west) sector (Figure10b).
Shortening between the inner and outer sectors was accommo-
dated through a basal detachment in the Auquilco evaporites.
During this stage, the western organic- rich deposits of Los
Molles Fm and lower VM were thermally overmature, whereas
the kitchen of Agrio Fm would have reached a 45% transforma-
tion (Figure10b). To the NE platform, the base of the VMFm
would have reached more than 30% transformation (Figure10b).
During the middle Miocene (14 Ma), the model pre-
dicts that kerogens of the Los Molles Fm and lower VM
were fully transformed in the inner and outer sectors of
the belt (Figure10c). In those areas, the Los Molles source
rock reached the advanced dry gas window to overmature
(>4 %Ro). The VMFm source rock was in the gas win-
dow, whereas the Agrio Fm reached an average TR of 85%
(Figure10c). Miocene compression resulted in uplift of the
inner sector and the structural duplication of sedimentary
successions in the outer sector. The Miocene deformation
phase promoted uplift and cooling of the main source rocks
in the inner sector. Tectonic uplift temporally stopped ther-
mal maturation and secondary cracking, in conjunction with
FIGURE 10 Modelling of thermal
maturity based on transformation ratio
(TR%) in the Agrio cross section for the
three source rocks in the basin. Four stages
of sequential restoration are considered. (a)
Late Cretaceous, (b) Palaeocene, (c) middle
Miocene and (d) present day
16
|
EAGE
SPACAPAN et Al.
decreasing temperature and pressure. This process likely pro-
moted a reduction in pore pressure associated with the ex-
pulsion of large amounts of gas from the Jurassic and Lower
Cretaceous source rocks. The structural duplication in the
outer sector resulted in a component of structural thicken-
ing, which promoted thermal reactivation and, consequently,
increase in secondary cracking reactions and pore pressure.
Towards the structural heights of DdLC, the Los Molles
source rock was also overmature (>4 %Ro), and the lower
VM reached the gas window (2 %Ro).
Present- day maturity shows that the Los Molles and
VMFms are completely exhausted from the inner sector to the
eastern flank of the DdLC (Figure10d). Towards the DdLC
area, the calculated maturity trend for Well A shows that the
lower VM reached the dry gas window (Figures9b and 10d).
Furthermore, the model predicts that in the NE platform area,
the lower VM is in the early- to- peak oil window and reaches
a transformation ratio of 50% (Figure10d). Simulated and
calculated vitrinite reflectance values show a clear decrease
from the Agrio FTB to the foreland (east) region of the basin.
6.3
|
Time extractions for the lower Vaca
Muerta in the DdLC and NE Platform areas
The interplay of key processes related to regional tectonic
events and hydrocarbon generation, thermal stress, pore
pressure and porosity development was analysed in time
FIGURE 11 Time extractions for the lower VM in the Agrio model. (a) DdLC and (b) NE Platform areas. Curve references: Transformation
ratio (red), burial depth (brown), vitrinite reflectance (violet), total porosity (black), TOC (green), pore pressure (light blue) and secondary cracking
mass (dashed red). Present- day vitrinite reflectance values (black circles), pressure data derived from DFIT tests (blue stars) and the onset of
organic porosity (Øorg) development are included in both extractions
|
17
EAGE
SPACAPAN et Al.
extractions for the lower VM in the DdLC and NE Platform
areas of the Agrio section (Figure11).
The DdLC area shows that during the Early Cretaceous
the lower VM underwent a first phase of burial compaction
(brown curve in Figure11a), which reduced the effective
porosity from 60% to 13% (black curve in Figure11a). Pore
pressure increased up to 28MPa (blue curve in Figure11a)
interpreted to result from incomplete drainage of generated
fluids. From ca. 120 Ma, the unit reached a TR of 88%
(red curve in Figure11a), which induced an abrupt increase
in the porosity (black curve in Figure11a) related to the
growth of Øorg. The model also shows that the develop-
ment of Øorg added ca. 11% to the porosity and reduced
the TOC content to ca. 5% (green curve in Figure 11a).
Furthermore, continuous thermal cracking increased pore
pressure, which induced a first pulse of petroleum ex-
pulsion at thermal maturity of 0.82 %Ro (violet curve in
Figure 11a). A second phase of mechanical compaction
started at ca. 100 Ma and caused a reduction in porosity,
likely including both matrix and organic porosity (black
curve in Figure11a). At ca. 65Ma, the unit reached max-
imum burial (brown curve in Figure11a) and the effective
porosity was reduced to 10%. Secondary cracking reac-
tions, accompanied by gas generation (dashed red curve
in Figure11a), caused an increase in pore pressure of ca.
87MPa. From ca. 50Ma, progressive uplift of the DdLC
anticline started and continued during Palaeocene, Eocene
and Miocene times. Uplift and consequent erosion reduced
vertical stress, thereby reducing the required pore pressure
to maintain equilibrium with the overlying column. This
is demonstrated in the model between 50Ma and present-
day when about 500m of section was eroded, which led to
a decline in pore pressure (blue curve in Figure11a). By
present- day the pore pressure dropped to 57MPa, which is
consistent with DFIT values obtained in Well A (violet star
in Figure11a).
In the NE Platform, the lower VM underwent a contin-
uous decrease in effective porosity from 58% to 7% (black
curve in Figure11b) due to mechanical compaction that took
place between 150 and 60Ma. At 60Ma, TR of 45% (red
curve in Figure11b) shows that most of the unit was in the
main oil window during the maximum burial (brown curve
in Figure11b). The model shows that increased burial depth
and hydrocarbon generation increased the pore pressure
(blue curve in Figure11b) to ca. 62MPa. From ca. 22Ma to
present day, progressive uplift and erosion reduced the pore
pressure from 60 to 48MPa, which is comparable with DFIT
measures for Well C (violet star in Figure11b). Progressive
transformation of OM occurred in parallel with the reduction
in TOC contents (green curve in Figure11b) and increased
porosity (black curve in Figure11b), which is interpreted as
a product of Øorg development. The Øorg added 2.2% to the
effective porosity in the lower VM. The simulations allow us
to recognize significant differences in Øorg development be-
tween the DdLC (11%) and NE Platform (2.2%), which cor-
relate with TOC0 variations and transformation gradient. In
the DdLC area, the high TOC0 concentration associated with
a high transformation ratio promoted an increase in Øorg de-
velopment compared with the NE Platform.
7
|
DISCUSSION
7.1
|
Organic geochemical pattern in the
VMFm
Regional studies of the VMFm have examined the OM dis-
tribution, including richness (quantity), free hydrocarbons
FIGURE 12 Schematic evolution of
organic matter pore types with increased
thermal maturity based on SEM and
pyrolysis data for the VMFm. References:
np- OM (nonporous organic matter),
bc (bioclasts), fr (fracture) and SOM
(secondary organic matter)
18
|
EAGE
SPACAPAN et Al.
and thermal maturity (e.g. TR, %Ro) of the unit (e.g. Brisson
etal.,2020; Cruz etal.,1996, 2002; Legarreta & Villar,2011,
2015; Urien & Zambrano,1994; Veiga etal.,2020). In addi-
tion, there are many local contributions from outcrop sections
and the subsurface that include TOC and Rock- Eval® pyrolysis
analyses (e.g. Askenazi etal.,2013; Boll etal.,2014; Capelli
etal.,2021; Cavelan etal.,2019; Karg & Littke,2020; Krim
etal.,2019; Małachowska etal.,2019; Milliken etal.,2019;
Petersen etal.,2020; Romero- Sarmiento etal.,2017; Scasso
etal.,2005; Sylwan,2014). All those studies and the present
work indicate high present- day TOC values ranging from 1%
to 12% with maximum values towards the base of the unit in
the lower VM (Figure4). However, present- day TOC con-
tent may not reflect the loss of organic carbon during thermal
maturation. Thus, the original TOC0 based on the reconstruc-
tion of Brisson et al. (2020) is a suitable parameter to ex-
amine vertical and lateral richness distribution (Figure 4),
and it is applicable for the VMFm. In the Agrio section,
the lower VM shows an increase in TOC0 values from 6%
(east) to 8% (west). This observation is consistent with the
westward facies deepening of classic palaeogeographic
maps accompanied by restricted settings under low- oxygen
water conditions that favour preservation of OM (Legarreta
& Uliana,1991). The increase in TOC values from east to
west is associated with thickening of the VMFm between ca.
400 and 800m in the same trend (Dominguez et al.,2016),
which means a higher volume of generated hydrocarbons for
the same gradient of TR. Notably, Well B shows the high-
est present- day (mean 8%) and original TOC (mean 18% for
the lower VM) contents, which suggests that local effects
could have influenced these positive anomalies. As analysed
by Dominguez et al. (2016), organic- rich intervals in the
VMFm correspond to bottomset– distal foreset segments of
prograding clinoforms, conditioned by stratigraphic and in
some cases tectonic (differential subsidence and palaeo- arch
uplift) controls. The lower VM exhibits mostly stratigraphic
controls (Dominguez etal., 2016), and therefore, a position
close to the distal foreset is envisaged where optimal con-
ditions for the production and preservation of OM occurred
(Passey etal.,2010).
OM in the unit shows negligible compositional differ-
ences, both geographically and stratigraphically, and is
typically described as unstructured algal material (type II
kerogen) with very scarce land- derived (e.g. vitrinite) parti-
cles that yielded an estimated HI0 of 680mgHC/gTOC (Brisson
etal.,2020; Petersen etal.,2020; Veiga etal.,2020). As re-
corded in the Agrio section, the basin shows an E– W ther-
mal gradient, which is consistent with the decrease in HI
(Figure5) and increase in TR values to the west. The sample
set exhibits a gradual decrease in HI with increasing maturity
gradient from 470mgHC/gTOC (general average for the early
oil window, Well C) to 41mgHC/gTOC (general average for the
early oil window, Well A). This denotes the transformation
of type II kerogen into hydrocarbons and the concomitant re-
duction in HI (Hazra etal.,2019; Peters & Cassa,1994).
Some samples from the upper VM show an increase in
OI, which could suggest some terrigenous OM input; how-
ever, it may be better to interpret it as the result of OM ox-
idation during transport and deposition (Peters, 1986). In
the case of the upper VM, that increase is accompanied by
a decline in TOC content related to the shallowing upward
tendency of this interval. As the upper VM shows an increase
in the carbonate content (Capelli et al., 2021; Legarreta &
Villar,2015), impure calcite could also generate some CO2
upon pyrolysis and, as a consequence, an increase in the S3
peak and OI parameter (Peters,1986).
7.2
|
Evolution of OM across the
maturation gradient
The importance of shale reservoirs as sources of oil and gas
has motivated numerous studies that focus on parameters that
control the development and preservation of pores (Katz &
Arango,2018; Ko etal.,2017; Liu etal.,2017; Löhr etal.,2015;
Milliken et al., 2013; Reed & Loucks, 2015). The regional
analysis of the VMFm presented in this work demonstrates that
maturity is an important factor that impacts the development
of Øorg. The early and rapid changes in TR that characterize
type II kerogen from the VMFm (Brisson et al., 2020) sug-
gest that OM- hosted pores are associated with partially and ex-
hausted OM generated during primary and secondary cracking
reactions. As also documented by Tomassini etal.(2019), the
general high TOC content of the unit indicates that OM- hosted
pores are the main storage of potential fluids.
In the oil window and based on TR and SEM data, the OM
that hosts pores is partially transformed and occupies spaces
between detrital grains, the internal structure of bioclasts and
fractures (Figure12). Interestingly, some samples from the
early oil window (Well C) lack OM- hosted pores and have
the lowest values of TR (mean 30%, Table1). Filament- like
organic particles resemble those liptinite macerals identified
under optical petrography in immature to low mature sam-
ples (Brisson etal.,2020; Małachowska etal.,2019; Petersen
etal., 2020). The type II sapropelic kerogen that prevailed
during sedimentation of the VMFm (Brisson et al., 2020;
Legarreta & Villar, 2015) is prone to ductile compaction
(Figure6a– c) and loss of primary porosity (Löhr etal.,2015;
Milliken et al., 2014). In analogous shale units at low ther-
mal maturity, type II kerogen is nonporous and highly af-
fected by mechanical compaction (Comerio et al., 2020;
Schieber,2013). This observation indicates that the poten-
tial infill of pores in this early- generation stage with heavy,
bitumen- like petroleum (Petersen etal.,2020) could explain
why liptinite macerals do not show evidence of pores at
SEM resolution (Löhr etal.,2015). However, at some rock
|
19
EAGE
SPACAPAN et Al.
intervals in the early oil window, isolated bubble pores occur
in secondary OM products (e.g. solid bitumen), suggesting
that other processes were involved in the development of
Øorg. As was pointed out by Brisson etal.(2020), variations
in depositional conditions could have influenced the final
composition of oil- prone organofacies (e.g. sulphur- rich vs.
sulphur- depleted). These diagenetic differences could affect
the onset of primary cracking reactions and the generation of
liquid hydrocarbons and bitumen (Hackley & Cardott,2016).
Thus, decomposition of kinetically distinct kerogens at dif-
ferent rates might explain why that some strata at the same
thermal maturity preserve nonporous liptinite macerals,
whereas in others Øorg is associated with secondary OM
products such as the solid bitumen.
OM- hosted pores in the gas window are interpreted to
develop in secondary, highly transformed (TR> 90%) OM
and show both the bubble and spongy morphologies (Ko
et al., 2017; Milliken et al., 2013). The wet gas window
(Well B) exhibits the largest bubble pores (2– 4 µm) asso-
ciated with development of highly concentrated spongy
pores that are absent in the early oil window but become
more abundant in the dry gas window (Figure 12). Large
bubble pores occur in ‘shelter’ pores (Figure7a– c), such as
in the interior of bioclasts and around diagenetic products
(calcite cement) indicating that some lithologies (Milliken
etal.,2019) favoured their preservation. They likely formed
in the oil window as was interpreted for the Eagle Ford Fm
(Cenomanian– Turonian, Maverick Basin of south Texas),
where bubble pores include hydrocarbon liquids that were left
behind or migrated following petroleum generation (Schieber
etal., 2016). Consequently, we interpret that spongy pores
started to grow during secondary cracking due to decom-
position of partially transformed OM by thermal maturity.
Some bubble pores appear to form due to coalescence of
spongy pores (Figure8c,d) and were, therefore, also gener-
ated during the dry gas window (Cavelan etal.,2019). The
bubble geometry is mainly represented by macropores (pore
sizes >50nm) and well documented with SEM resolution.
On the contrary, spongy pores include a part of mesopores
(2– 50 nm) and presumably micropores (<2 nm), the latter
measured through nitrogen adsorption analysis by Cavelan
etal.(2019) for the VMFm. For terrigenous OM, SEM does
FIGURE 13 (A) Present- day maturity (%Ro) of the VMFm along the Agrio section. (B) Thermal maturity based on transformation ratio
(TR%) for the lower VM is presented in four areas: (a) Cerro Mocho, (b) DdLC, (c) East of DdLC and (d) NE Platform
20
|
EAGE
SPACAPAN et Al.
not show pores at any maturity, the same observations apply
for the New Albany Shale (Devonian– Mississippian, Illinois
Basin at USA), where vitrinite and inertinite macerals may
not contribute to additional Øorg (Liu etal.,2017).
SEM images do not show clear evidence of deformed
organic pores; except for some examples in Well B where
OM- hosted pores within a clay- rich matrix exhibit elongated
(flattened) forms interpreted as the product of mechanical
compaction (Figure7e). Previous studies documented that or-
ganic pores can survive mechanical compaction but preserve
features of OM deformation (Wang,2020). As documented
for the VMFm, most organic pores are preserved showing
circular forms, which implies that rigid components such as
microfossils and diagenetic products (microcrystalline quartz
and calcite cement) inhibit the effects of compaction. In
addition, the analysed samples are far from intensively de-
formed areas (e.g. outcrops in the internal belts) where OM-
hosted pores could be deformed. The OM- hosted pores are
interpreted to result from kerogen and bitumen cracking to
oil and gas, presumably in equilibrium between lithostatic
pressure and pore pressure within the unit (Ko etal.,2017).
Accordingly, overpressure- related processes could represent
important mechanisms that inhibit the effects of compaction
and contributed to preservation of Øorg (see Section7.3.3).
7.3
|
BPSM in the Agrio cross section
The 2D modelled section presented in this work reproduces
the decrease in maturity and OM transformation from west
(fold and thrust belt) to east (basin margin) for the main
source rocks in the Neuquén Basin (see Section6). For the
VMFm, the models predict temporal and spatial variations in
the timing of hydrocarbon generation– expulsion controlled
by tectonic events, burial depth and differences in thick-
ness of the Mesozoic sedimentary overburden. Model results
show that OM transformation impacted on the magnitude of
(a) the effective porosity associated with Øorg development,
and (b) the distribution of pore pressure within the unit.
7.3.1
|
Timing of hydrocarbon generation
Based on present- day %Ro values, maturity windows of the
VMFm along the Agrio section are presented in Figure 13.
Temporal and spatial variations in the timing of kerogen trans-
formation are documented between the inner sector and the
NE Platform. From the fold belt (Cerro Mocho– Pichi Mula)
to the DdLC structure, the lower VM reached the dry gas win-
dow, and according to the modelling, the maximum transfor-
mation occurred at ca. 120Ma in the inner sector (A- curve in
Figure13b) before the Late Cretaceous Andean deformation
phase that affected the western part of the belt (Cobbold &
Rossello,2003; Rojas Vera etal.,2015; Sánchez etal., 2018;
Zamora Valcarce et al., 2011; Zapata & Folguera, 2005).
Modelling results indicate that hydrocarbon generation and ex-
pulsion preceded the growth of main structures in the inner and
outer sectors of the Agrio FTB and would have been favoured
by the increase in the Mesozoic sedimentary cover towards the
inner and outer sectors (ca. 4– 5km burial depth). Nevertheless,
tectonic uplift and cooling of both sectors correlate with the
Miocene contractional pulse. As a consequence, thermal stress
and secondary cracking stopped, and pore pressure decreased,
since remnant fluids (mainly gas) were expelled from the unit.
Such interpretations were previously inferred by Gómez Omil
et al. (2014), indicating that intensely deformed areas in the
Neuquén Basin are at present day inactive and incapable of
generating hydrocarbons.
For the DdLC structure, TR of 99% shows that the lower
VM was already in the dry gas window at ca. 98Ma (B-
curve in Figure13b), whereas in the external flank of this
structure full transformation conditions were reached at ca.
80 Ma (C- curve in Figure 13b). Compressive deformation
and uplift/regional erosion promoted the exhumation of the
unit associated with expulsion of fluids as well as hydrocar-
bon migration and trapping in previously generated struc-
tures (Rocha etal.,2018). This observation is consistent with
our modelling because the dissipation of pore pressure (blue
curve in Figure11a) and fluid expulsion are linked to exhu-
mation events that occurred mainly between Palaeocene and
Miocene time. The NE Platform shows that the unit reached
the early- to- peak oil window with a TR of 40%– 50%, indi-
cating that hydrocarbons generation started from ca. 60Ma
(D- curve in Figure13b).
The model indicates that hydrocarbon generation and
expulsion started in the western VMFm before the Late
Cretaceous Andean deformation, which controlled the uplift
of inner sectors within the Agrio FTB. In the undeformed
scenario (Figure10a), hydrocarbons were able to migrate up
to 70 km through the main carrier beds in the basin (Cruz
etal., 2002; Rocha etal.,2018). At present day, the VMFm
is totally exhausted in the intensely deformed regions of the
belt, with full kerogen transformation into hydrocarbons at ca.
120Ma. At the basin scale, the Agrio model shows a clear
eastward decrease in maturity and transformation gradient as-
sociated with changes in the sedimentary column thickness
that led to differential maximum burial before the Miocene
deformation pulse. Similar interpretations were documented
along the Chos Malal FTB, where thermal maturity and
transformation also decrease eastward linked to differences
in thickness of the sedimentary overburden (Cruz etal.,1996;
Karg & Littke,2020). In the inner sector of the Agrio FTB,
the Miocene uplift caused present- day source rocks to show
relict maturity reached at deeper burial depths in the past. In
addition, there is an increase in the thermal stress and trans-
formation of the OM from north to south, mainly controlled
|
21
EAGE
SPACAPAN et Al.
increased sedimentary thickness in the same trend (Gómez
Omil etal.,2014; Legarreta etal.,2008). These observations
indicate that the effects of thermal maturity follow a N– S trend
controlled by a general thickness increase in the source rocks
of the Mendoza Group towards the area of the Agrio FTB, to
the north of the Huincul Arch (see location in Figure1).
7.3.2
|
Organic porosity estimation
The integrated modelling results and SEM images indicate
that the development of Øorg is a typical feature of the
VMFm. Petromod® software calculates Øorg as the result of
OM transformation from solid immature kerogen to less dense
fluid hydrocarbons during thermal maturation (Hantschel &
Kauerauf, 2009). Time extractions show an increase in ef-
fective porosity, interpreted as the result of Øorg linked to
the transformation of OM into hydrocarbons (Figure 14a).
The simulations are comparable with those of Modica and
Lapierre (2012): Øorg tends to increase with burial depth and
thermal transformation. Accordingly, Øorg differences for
the lower VM between the DdLC (11%) and NE Platform
(2.2%) correspond to an increase in TR values to the west in
the basin. Such observations are consistent with basin- scale
modelling of Øorg in the Mississippian Barnett Shale that al-
lowed Romero- Sarmiento etal.(2013) to estimate from 0%
(immature zones) to 4% Øorg (mature zones) of rock volume.
However, as for the Upper Devonian Duvernay Formation
in the Western Canada Basin (Chen & Jiang,2016), the de-
velopment of Øorg is also a function of TOC0 and kerogen
type, since type I kerogen generates more Øorg than type III
kerogen (Chen etal.,2015).
Time extractions show variations in the effective po-
rosity through time for the lower, middle and upper VM in
Well A (Figure14a). From 120 Ma, the lower VM (TOC0
11% average) increased its porosity by ca. 11%, which is
interpreted as a response to Øorg development. The curve
obtained by Mei etal. (2021) for the VMFm also shows a
considerable increase in porosity due to Øorg development
associated with the onset of hydrocarbon generation. To the
contrary, the upper VM (TOC0 2% average) would have de-
veloped only 1.95% of Øorg at full transformation, a value
that does not considerably affect the evolution of the poros-
ity curve (Figure14a). Time extractions indicate progressive
increase in porosity from the top to the base of the unit re-
lated to TOC0 and thermal maturity and, therefore, with the
FIGURE 14 (a) Effective porosity
curves through geologic time for the lower,
middle and upper VM in DdLC (Well A).
Simulated porosity increases as a response
to organic porosity (Øorg) development and
the porosity difference between upper and
lower VM is expressed as ΔØ. Modelling
indicates a progressive increase in porosity
from the top to the base of the unit related
to TOC0 and thermal maturity increase
and, therefore, with increase in Øorg
in the same trend. The insert shows the
calculated porosity (black curve) and data
obtained from well logging (black cross)
and gas- filled porosity (red circles) from
core samples. (b) Distribution of present- day
pore pressure in the VMFm. Pore pressure
modelling for the lower VM in Well A with
(black curve) and without hydrocarbon
generation (dashed black curve)
22
|
EAGE
SPACAPAN et Al.
increased in Øorg in the same trend. This observation agrees
with SEM analysis, which found that OM- hosted pores are
dominant in organic- rich intervals of the unit (Tomassini
etal., 2019). Although it is beyond the scope of this study
to quantify the porosity, our results suggest that organic- rich
intervals of the VMFm and associated OM pores control the
hydrocarbon storage capacity. Similar results have been doc-
umented in other worldwide class unconventional resources,
as in the Barnett (Texas) and Haynesville (Texas, Arkansas,
Louisiana) Fms, where much of the petroleum is trapped
in organic pores (Passey et al., 2010; Peters et al., 2017;
Romero- Sarmiento etal.,2013).
It is important to highlight that the constructed model
incorporates bulk kinetics specific for the VMFm (Brisson
etal., 2020), which provided accurate information on ther-
mal evolution of the OM, Øorg and timing of hydrocarbon
generation. Future investigations should incorporate detailed
compositional kinetic schemes to optimize the evolution of
TOC, Øorg and hydrocarbon retention and expulsion in the
VMFm (see Mei et al., 2021).
7.3.3
|
Pore pressure related to
hydrocarbon generation
2D basin modelling suggests that hydrocarbon genera-
tion represents an important mechanism that controls the
magnitude and distribution of pore pressure in the VMFm.
Simulations and DFIT data indicate that the unit is an over-
pressured system at different basin positions. Previous studies
proposed that the transformation of kerogen into oil and gas
represents the main mechanism for overpressure generation
in the VMFm (Badessich etal.,2016; Cobbold etal.,2013;
Rocha etal.,2018; Rodrigues etal.,2009; Veiga etal.,2020;
Zanella etal.,2015). The volume expansion associated with
hydrocarbon generation has been implicated as a cause of
overpressure in different sedimentary basins (Osborne &
Swarbrick,1997; Swarbrick etal.,2002). Hydrocarbon gen-
eration can be a cause for overpressure if the rate of volume
increase due to kerogen transformation exceeds the rate
of volume loss by fluid expulsion and migration (Berg &
Gangi,1999). Bredehoeft et al. (1994) proposed that if the
liquid generated during thermal cracking is less dense than
the solid kerogen, then there will be more liquid than created
pores and, thus, the pore pressure will increase. This over-
pressure mechanism depends upon kerogen type, abundance
of OM, thermal history and rock permeability (Osborne &
Swarbrick, 1997). For the VMFm, modelling results show
that pore pressure magnitude was influenced by abundance
of OM and generated fluids, both parameters controlled by
the position in the basin and thermal history. The highest
pore pressures coincide with the western part of the basin,
where the VMFm shows highest TOC0 contents and where
the unit reached the gas window. Osborne and Swarbrick
(1997) suggested that changes in overpressure gradients re-
lated to volume expansion after hydrocarbon generation are
controlled by two main reactions: (1) low- temperature reac-
tion rate for primary cracking of kerogen into oil and gas, and
(2) high- temperature reaction rate for secondary cracking of
oil and bitumen into gas. In particular, secondary cracking
reactions induce a significant overpressure within low per-
meability source rocks (Osborne & Swarbrick,1997). High
pressures are more readily developed by gas generation than
by oil generation because of the much lower density of gas
(Barker, 1990). Different case studies documented that the
generation of gas resulted in conditions of overpressure within
source rocks (Carcione & Gangi, 2000; Gao et al., 2019;
Hao et al., 1996; Hunt,1990; McPeek, 1981; Nunn, 2012;
Ramdhan & Goulty, 2010; Swarbrick et al., 2002; Tingay
etal.,2009). Cracking of 1 vol% of oil would generate over-
pressures close to the magnitude of the lithostatic gradient
in confined rocks (Barker,1990). The relationship between
pore pressure and maturity windows for the VMFm along the
Agrio FTB is shown in Figure14b. Well A (DdLC) in the gas
window at present day reaches 57MPa, whereas towards the
oil window (Well C), the pore pressure reaches 48MPa. The
volume expansion related to primary and secondary cracking
would have controlled such differences in pore pressure of
the VMFm between different maturity stages. These results
indicate that secondary cracking of retained hydrocarbons in
VMFm produced a significant increase in pore pressure in
western areas of the Neuquén Basin.
The Petromod® software calculates the pore pressure
caused by primary and secondary cracking reactions consid-
ering different parameters such as kerogen reduction, poros-
ity, density of hydrocarbons and compressibility (Hantschel
& Kauerauf,2009). The magnitude of fluid expansion on pore
pressure was also evaluated in a model excluding hydrocar-
bon generation in pore pressure calculations for the VMFm
in the DdLC area (Figure14b). Interestingly, the two models
show discrepancies in the magnitude of pore pressure, con-
firming that hydrocarbon generation, especially gas, plays a
significant role in pressure generation. In the second model,
pore pressure generation was related to disequilibrium com-
paction and pore pressure magnitude is lower than the model
with hydrocarbon generation. In particular, during the peak
of gas generation (ca. 50Ma), the pore pressure difference is
ca. 30% on average between both models (Figure14b). Our
results suggest that both disequilibrium compaction and hy-
drocarbon generation could act as coupled processes contrib-
uting to pore pressure generation within the VMFm.
On the other hand, the present- day distribution of pore
pressure (Figure14b) depends on the magnitude of different
processes that reduce pore pressure such as fracturing, uplift
and erosion (Doré & Jensen,1996; Law & Dickinson,1985;
Law & Spencer, 1998; Neuzil & Pollock, 1983). Such
|
23
EAGE
SPACAPAN et Al.
processes reduce vertical stress, thereby reducing the re-
quired pore pressure to maintain equilibrium with the over-
lying column (Burgreen- Chan et al., 2015). Pore pressure
reduction is recorded in the time extraction for DdLC when
ca. 1,000m of sedimentary rocks were eroded from 65Ma
to present day. The pore pressure in the unit decreases from
87MPa (50Ma) to 57MPa (present day), demonstrating the
capacity of fracturing, uplift and erosion to reduce pressure
and promote late expulsion and migration of hydrocarbons
into potential traps. The amount of pore pressure reduction
appears to be a consequence of the permeability increase
triggered by fracturing, uplift and erosion associated with
a decrease in lithostatic load. In the same way, the Miocene
tectonic uplift stopped the thermal stress and secondary
cracking, thus promoting a substantial pore pressure reduc-
tion in the Agrio Fold belt. The reduction in pore pressure in
the VMFm associated with those processes is in agreement
with observations and modelling from other studies in the
unit (Mei etal.,2021). We consider that tectonic uplift could
be an important process for redistribution of hydrocarbons
within the basin and individual traps because of changes
in structural configuration (e.g. Shanley & Cluff,2015). In
the model, the faults were assumed to be closed (i.e. im-
permeable) for the simulated steps. Compartmentalization
due to faults is a common feature in petroleum systems and,
in many cases, is identified by laterally different pressure
regimens (Borge, 2002; Borge & Sylta, 1998; Gibson &
Bentham,2003; Williams & Madatov,2005). This is doc-
umented in the Pichi Mula triangular zone, where closed
faults inhibited the drainage of fluids and resulted in high
present- day pressure (Figure14b).
Finally, 2D basin modelling by Berthelon et al. (2021)
suggested that horizontal (tectonic) compression related
to Andean deformation could increase overpressure in the
VMFm. However, the authors mentioned that their proposed
model does not integrate horizontal stress with hydrocarbon
generation in the calculation of the overpressure. The model-
ling in our study shows that generation of hydrocarbons and
compaction disequilibrium represents the main mechanisms
controlling overpressure in the VMFm. However, we do not
rule out the possibility that compressional stress related to
Andean deformation may have also influenced the distribu-
tion and magnitude of pore pressure. This proposal should be
modelled in detail considering the magnitudes and temporal
relationships among maximum burial, hydrocarbon genera-
tion and horizontal stress.
8
|
CONCLUSIONS
This study analyses the unconventional petroleum potential
of the VMFm source rock at different thermal maturities
based on organic geochemistry, electron microscopy and
a regional BPSM from the Agrio FTB to the basin border.
A robust data set was incorporated into the model with the
purpose of integrating regional tectonic events and processes
that controlled the generation and expulsion of hydrocarbons,
organic porosity development and pore pressure mechanisms
through geologic time.
In the Agrio FTB, the lower Vaca Muerta shows an
increase in TOC0 towards the west from 4% to 8% mean
values. The increase in TOC0 from east to west is associ-
ated with thickening of the unit, which suggests the po-
tential larger volumes of generated hydrocarbons for the
same thermal gradient. Thermal maturity based on the
transformation ratio (TR) controlled the development of
organic pores. TR increases westward, indicating that in
conjunction with increased TOC0, organic pores represent
the main control on total porosity in organic- rich intervals
of the unit. Along the Agrio FTB, ca. 11% of organic po-
rosity was reached in the DdLC (gas window), whereas
only ca. 2% was reached in the NE Platform (early oil) for
the lower Vaca Muerta. Organic geochemistry and elec-
tron microscopy demonstrate that OM- hosted pores are
related to thermal degradation of oil- prone type II kerogen
into partially (TR ca. 3%– 60%) nearly completely (TR ca.
80%– 99%) transformed OM. Based on SEM images, iso-
lated bubble pores are typical of the oil window, whereas
bubble and densely distributed spongy pores occur in the
gas window.
BPSM shows that temporal and spatial variations in the
timing of hydrocarbon generation are linked to tectonic
events, burial depth and differences in thickness of the
Mesozoic sedimentary overburden. The model shows a clear
decrease in maturity and OM transformation to the east (basin
margin) for the main source rocks in the Neuquén Basin. 2D
modelling predicts that large volumes of hydrocarbons were
generated and expelled from VMFm during the Early and
Late Cretaceous in western sectors of the basin.
The VMFm is an overpressure cell at different basin po-
sitions. BPSM indicates that hydrocarbon generation and
compaction disequilibrium were main mechanisms that
controlled magnitude and distribution of pore pressure.
The volume expansion associated with hydrocarbon gen-
eration reached maximum values from secondary cracking
(transformation of oil and bitumen into gas) during the Late
Cretaceous– Palaeocene.
The distribution of OM and mechanisms that control pres-
sure and porosity along the maturity gradient represent key
parameters to evaluate unconventional plays in the context
of basin evolution. In the Vaca Muerta Formation, overpres-
sure intervals with high organic carbon contents are the most
prone to develop organic pores, which represent favourable
sites for the storage of hydrocarbons.
24
|
EAGE
SPACAPAN et Al.
ACKNOWLEDGEMENTS
This study is published with the permission of YPF and
YTEC. We are especially grateful to L. Monti, G. Sagasti
and R. Manoni (YPF), C. Smal and J.P. Alvarez (Y- TEC)
for allowing us to publish the present data. To M. Fasola,
for his review and constructive comments. Special thanks
to A. Floridia, N. Jausoro and F. Medina (Laboratorio de
Microscopía in Y- TEC) for SEM analysis. H. Villar (GeoLab
Sur) is thanked for pyrolysis analysis and vitrinite determina-
tions. Special thanks to M.F. Romero- Sarmiento and B. Katz
who provided thoughtful comments and further improved
this article. We also express our gratitude to K. Peters for
his useful edits, suggestions and ideas. We acknowledge C.
Jhonson (Asociate Editor of BR) for constructive comments
on the original version of the manuscript.
PEER REVIEW
The peer review history for this article is available at https://
publo ns.com/publo n/10.1111/bre.12599.
DATA AVAILABILITY STATEMENT
The data that support the findings of this study are available
from the corresponding author upon reasonable request.
REFERENCES
Aguirre- Urreta, B., Tunik, M., Naipauer, M., Pazos, P., Ottone,
E., Fanning, M., & Ramos, V. A. (2011). Malargüe Group
(Maastrichtian– Danian) deposits in the Neuquén Andes, Argentina:
Implications for the onset of the first Atlantic transgression related
to Western Gondwana break- up. Gondwana Research, 19, 482– 494.
https://doi.org/10.1016/j.gr.2010.06.008
Aguirre- Urreta, M. B., Mourgues, F. A., Rawson, P. F., Bulot, L. G., &
Jaillard, E. (2007). The Lower Cretaceous Chañarcillo and Neuquén
Andean basins: Ammonoid biostratigraphy and correlations.
Geological Journal, 42, 143– 173. https://doi.org/10.1002/gj.1068
Al- Hajeri, M. M., Saeed, M. A., Derks, J., Fuchs, T., Hantschel, T.,
Kauerauf, A., Neumaier, M., Schenk, O., Swientek, O., Tessen, N.,
Welte, D., Wygrala, B., Kornpihl, D., & Peters, K. (2009). Basin and
petroleum system modeling. Oilfield Review, 21, 14– 29.
Askenazi, A., Biscayart, P., Cáneva, M., Montenegro, S., & Moreno, M.
(2013). Analogía entre la Formación Vaca Muerta y shale gas/oil
plays de EEUU. Society of Petroleum Engeneers (SPE).
Athy, L. F. (1930). Density, porosity, and compaction of sedimentary
rocks. AAPG Bulletin, 14, 1– 24.
Badessich, M. F., Hryb, D. E., Suarez, M., Mosse, L., Palermo, N.,
Pichon, S., & Reynolds, L. (2016). Vaca Muerta shale— Taming a
giant. Oilfield Review, 28, 26– 39.
Bakar, R. (2018). Modeling and analysis of diagnostic fracture injection
tests (DFITs) (M.Sc. Thesis). Department of Petroleum Engineering,
Colorado School of Mines.
Barker, C. (1990). Calculated volume and pressure changes during the
thermal cracking of oil to gas in reservoirs. AAPG Bulletin, 74,
1254– 1261.
Berg, R. R., & Gangi, A. F. (1999). Primary migration by oil- generation
microfracturing in low- permeability source rocks: Application to
the Austin Chalk, Texas. AAPG Bulletin, 83, 727– 756.
Berthelon, J., Brüch, A., Colombo, D., Frey, J., Traby, R., Bouziat,
A., Cacas- Stentz, M. C., & Cornu, T. (2021). Impact of tectonic
shortening on fluid overpressure in petroleum system modelling:
Insights from the Neuquén basin, Argentina. Marine and Petroleum
Geology, 127, 104933.
Boll, A., Alonso, A., Fuentes, F., Vergara, M., Laffitte, G., & Villar, H.
J. (2014). Proceedings of IX Congreso de Exploración y Desarrollo
de Hidrocarburos, Mendoza, Argentina, Actas (pp. 3– 44).
Borge, H. (2002). Modelling generation and dissipation of overpressure
in sedimentary basins: An example from the Halten Terrace, off-
shore Norway. Marine and Petroleum Geology, 19, 377– 388. https://
doi.org/10.1016/S0264 - 8172(02)00023 - 5
Borge, H., & Sylta, Ø. (1998). 3D modelling of fault bounded pressure
compartments in the North Viking Graben. Energy Exploration and
Exploitation, 16(4), 301– 323.
Bredehoeft, J. D., Wesley, J. B., & Fouch, T. D. (1994). Simulations
of the origin of fluid pressure, fracture generation, and the move-
ment of fluids in the Uinta Basin, Utah. AAPG Bulletin, 78(11),
1729– 1747.
Brisson, I. E., Fasola, M. E., & Villar, H. J. (2020). Organic geochemi-
cal patterns of the Vaca Muerta Shale. In D. Minisini, M. Fantin, I.
Lanusse, & H. Leanza (Eds.), Integrated geology of unconvention-
als: The case of the Vaca Muerta play, Argentina. AAPG Memoir
120, Chapter 11.
Burgreen- Chan, B., Meisling, K. E., & Graham, S. (2015). Basin and
petroleum system modelling of the East Coast Basin, New Zealand:
A test of overpressure scenarios in a convergent margin. Basin
Research, 28, 536– 567. https://doi.org/10.1111/bre.12121
Capelli, I. A., Scasso, R. A., Spangenberg, J. E., Kietzmann, D. A.,
Cravero, F., Duperron, M., & Adatte, T. (2021). Mineralogy and
geochemistry of deeply- buried marine sediments of the Vaca
Muerta- Quintuco system in the Neuquén Basin (Chacay Melehue
section), Argentina: Paleoclimatic and paleoenvironmental implica-
tions for the global Tithonian- Valanginian reconstructions. Journal
of South American Earth Sciences, 107, 103103.
Carcione, J. M., & Gangi, A. F. (2000). Gas generation and overpres-
sure: Effects on seismic attributes. Geophysics, 65, 1769– 1779.
https://doi.org/10.1190/1.1444861
Cavelan, A., Boussafir, M., Rozenbaum, O., & Laggoun- Défarge, F.
(2019). Organic petrography and pore structure characterization
of low- mature and gas- mature marine organic- rich mudstones:
Insights into porosity controls in gas shale systems. Marine and
Petroleum Geology, 103, 331– 350. https://doi.org/10.1016/j.marpe
tgeo.2019.02.027
Chen, Z., & Jiang, C. (2016). A revised method for organic porosity
estimation in shale reservoirs using Rock- Eval data: Example from
Duvernay Formation in the Western Canada Sedimentary Basin.
AAPG Bulletin, 100, 405– 422. https://doi.org/10.1306/08261
514173
Chen, Z., Wang, T., Liu, Q., Zhang, S., & Zhang, L. (2015). Quantitative
evaluation of potential organic- matter porosity and hydrocarbon
generation and expulsion from mudstone in continental lake ba-
sins: A case study of Dongying sag, eastern China. Marine and
Petroleum Geology, 66, 906– 924. https://doi.org/10.1016/j.marpe
tgeo.2015.07.027
Cobbold, P. R., & Rossello, E. A. (2003). Aptian to recent compressional
deformation, foothills of the Neuquén Basin, Argentina. Marine and
Petroleum Geology, 20, 429– 443. https://doi.org/10.1016/S0264
- 8172(03)00077 - 1
|
25
EAGE
SPACAPAN et Al.
Cobbold, P. R., Zanella, A., Rodrigues, N., & Løseth, H. (2013).
Bedding- parallel fibrous veins (beef and cone- in- cone): Worldwide
occurrence and possible significance in terms of fluid overpressure,
hydrocarbon generation and mineralization. Marine and Petroleum
Geology, 43, 1– 20. https://doi.org/10.1016/j.marpe tgeo.2013.01.010
Comerio, M., Fernández, D. E., & Pazos, P. J. (2018). Sedimentological
and ichnological characterization of muddy storm related deposits:
The upper Hauterivian ramp of the Agrio Formation in the Neuquén
Basin, Argentina. Cretaceous Research, 85, 78– 94. https://doi.
org/10.1016/j.cretr es.2017.11.024
Comerio, M., Fernández, D. E., Rendtorff, N., Cipollone, M., Zalba, P. E.,
& Pazos, P. J. (2020). Depositional and postdepositional processes of
an oil- shale analog at the microstructure scale: The Lower Cretaceous
Agrio Formation, Neuquén Basin, northern Patagonia. AAPG
Bulletin, 104, 1679– 1705. https://doi.org/10.1306/04082 017419
Couzens- Schultz, B., Axon, A., & Azbel, K. (2013). Pore pressure pre-
diction in unconventional resources. In IPTC Paper 16849 presented
at the International Petroleum Technology Conference, Beijing,
China, 26– 28 March.
Cruz, C. E., Boll, A., Gómez Omil, R., Martínez, E. A., Arregui, C.,
Gulisano, C., Laffitte, G., & Villar, H. J. (2002). Hábitat de hi-
drocarburos y sistemas de carga Los Molles y Vaca Muerta en el
sector central de la Cuenca Neuquina, Argentina. In Congreso de
Exploración y Desarrollo de Hidrocarburos (No. 5).
Cruz, C. E., Villar, H. J., & Muñoz, N. G. (1996). Los sistemas petrole-
ros del Grupo Mendoza en la Fosa de Chos Malal. Cuenca Neuquina,
Argentina. In 13° Congreso Geológico Argentino y 3° Congreso de
Exploración de Hidrocarburos, Actas 1, AGA– IAPG, Buenos Aires,
Argentina, 13– 18 October (pp. 45– 60).
Cuervo, S., Lombardo, E., Vallejo, D., Crousse, L., Hernandez, C., &
Mosse, L. (2016). Towards a simplified petrophysical model for the
Vaca Muerta Formation. In Unconventional Resources Technology
Conference (URTeC), San Antonio, Texas, 1– 3 August 2016 (pp.
778– 796).
Curiale, J. A., & Curtis, J. B. (2016). Organic geochemical applications
to the exploration for source- rock reservoirs– A review. Journal
of Unconventional Oil and Gas Resources, 13, 1– 31. https://doi.
org/10.1016/j.juogr.2015.10.001
Dembicki, H. Jr (2009). Three common source rock evaluation er-
rors made by geologists during prospect or play appraisals. AAPG
Bulletin, 93, 341– 356. https://doi.org/10.1306/10230 808076
Deming, D. (1989). Application of bottom- hole temperature correc-
tions in geothermal studies. Geothermics, 18, 775– 786. https://doi.
org/10.1016/0375- 6505(89)90106 - 5
Desjardins, P., Fantín, M., González Tomassini, F., Reijenstein,
H., Sattler, F., Dominguez, R. F., Kietzmann, D., Leanza, H. A.,
Bande, A., Benoit, S., Borgnia, M., Vittore, F., Gil, G., Simo, T.,
& Minisini, D. (2016). Capítulo 2: Estratigrafía Sísmica Regional.
In G. González (Ed.), Transecta Regional de la Formación Vaca
Muerta (IAPG), Buenos Aires.
Dominguez, F., Noguera, I. L., Continanzia, M. J., Mykietiuk, K., Ponce, C.,
Pérez, G., Guerello, R., Caneva, M., Di Benedetto, M., & Cristallini, E.
(2016). Organic- rich stratigraphic units in the Vaca Muerta formation, and
their distribution and characterization in the Neuquén Basin (Argentina).
Unconventional Resources Technology Conference (URTEC).
Dominguez, R. F., & Di Benedetto, M. (2019). Understanding 3- D
distribution of organic- rich units in the Vaca Muerta Formation.
In Unconventional Resources Technology Conference, Denver,
Colorado, 22– 24 July 2019 (pp. 5114– 5128).
Doré, A. G., & Jensen, L. N. (1996). The impact of late Cenozoic up-
lift and erosion on hydrocarbon exploration: Offshore Norway and
some other uplifted basins. Global and Planetary Change, 12, 415–
436. https://doi.org/10.1016/0921- 8181(95)00031 - 3
Ejofodomi, E., Baihly, J. D., Malpani, R., Altman, R. M., Huchton, T.
J., Welch, D., & Zieche, J. (2011). Integrating all available data
to improve production in the Marcellus Shale. In North American
Unconventional Gas Conference and Exhibition. Society of
Petroleum Engineers.
Espitalié, J., Deroo, G., & Marquis, F. (1986). La pyrolyse Rock- Eval
et ses applications. Troisième partie. Revue de l'institut Français du
Pétrole, 41(1), 73– 89. https://doi.org/10.2516/ogst:1986003
Franzese, J. R., & Spalletti, L. A. (2001). Late Triassic continental-
extension in southwestern Gondwana: Tectonic segmentation and
pre- breakup rifting. Journal of South American Earth Sciences, 14,
257– 270.
Gao, J., Zhang, J.- K., He, S., Zhao, J.- X., He, Z.- L., Wo, Y.- J., Feng, Y.-
X., & Li, W. (2019). Overpressure generation and evolution in Lower
Paleozoic gas shales of the Jiaoshiba region, China: Implications for
shale gas accumulation. Marine and Petroleum Geology, 102, 844–
859. https://doi.org/10.1016/j.marpe tgeo.2019.01.032
Gibson, R. G., & Bentham, P. A. (2003). Use of fault- seal analysis in
understanding petroleum migration in a complexly faulted anti-
clinal trap, Columbus Basin, offshore Trinidad. AAPG Bulletin, 87,
465– 478.
Gómez Omil, R., Caniggia, J., & Borghi, P. (2014). La Formación Vaca
Muerta en la faja plegada de Neuquén y Mendoza. Procesos que
controlaron su madurez. IX Congreso de Exploración y Desarrollo
de Hidrocarburos. Actas DVD (pp. 71– 96).
Groeber, P. (1946). Observaciones geológicas a lo largo del meridi-
ano 70° 1. Hoja Chos Malal. Revista de la Asociación Geológica
Argentina, 1, 177– 208.
Grohmann, S., Fietz, W. S., Nader, F. H., Romero- Sarmiento, M. F.,
Baudin, F., & Littke, R. (2021). Characterization of Late Cretaceous
to Miocene source rocks in the Eastern Mediterranean Sea: An in-
tegrated numerical approach of stratigraphic forward modelling and
petroleum system modelling. Basin Research, 33, 846– 874. https://
doi.org/10.1111/bre.12497
Hackley, P. C., & Cardott, B. J. (2016). Application of organic pe-
trography in North American shale petroleum systems: A review.
International Journal of Coal Geology, 163, 8– 51. https://doi.
org/10.1016/j.coal.2016.06.010
Hantschel, T., & Kauerauf, A. I. (2009). Fundamentals of basin and
petroleum systems modeling (p. 476). Springer.
Hao, F., Li, S., Sun, Y., & Zhang, Q. (1996). Characteristics and origin
of the gas and condensate in the Yinggehai Basin, offshore South
China Sea: Evidence for effects of overpressure on petroleum gen-
eration and migration. Organic Geochemistry, 24, 363– 375. https://
doi.org/10.1016/0146- 6380(96)00009 - 5
Hazra, B., Wood, D. A., Mani, D., Singh, P. K., & Singh, A. K. (2019).
Evaluation of shale source rocks and reservoirs (p. 139). Springer.
Horton, B. K. (2018). Sedimentary record of Andean mountain build-
ing. Earth- Science Reviews, 178, 279– 309.
Horton, B. K., Fuentes, F., Boll, A., Starck, D., Ramirez, S. G., &
Stockli, D. F. (2016). Andean stratigraphic record of the transi-
tion from backarc extension to orogenic shortening: A case study
from the northern Neuquén Basin, Argentina. Journal of South
American Earth Sciences, 71, 17– 40. https://doi.org/10.1016/j.
jsames.2016.06.003
26
|
EAGE
SPACAPAN et Al.
Howell, J. A., Schwarz, E., Spalletti, L. A., & Veiga, G. D. (2005). The
Neuquén Basin: An overview. In G. D. Veiga, L. A. Spalletti, J. A.
Howell, & E. Schwarz (Eds.), The Neuquén Basin: A case study in
sequence stratigraphy and basin dynamics (Vol. 252, pp. 1– 14).
Geological Society, London, Special Publications.
Hunt, J. M. (1990). Generation and migration of petroleum from ab-
normally pressured fluid compartments. AAPG Bulletin, 74, 1– 12.
Jorgensen, L., López Pezé, A., & Pisani, F. (2013). Caracterización de
la Formación Los Molles como reservorio de tipo Shale Gas en el
ámbito Norte de la Dorsal de Huincul, Cuenca Neuquina, Argentina,
Mostrando su analogía con reservorio de Shale Gas probado en
EEUU. Society of Petroleum Engineers.
Karg, H., & Littke, R. (2020). Tectonic control on hydrocarbon genera-
tion in the northwestern Neuquén Basin, Argentina. AAPG Bulletin,
104, 2173– 2208. https://doi.org/10.1306/05082 018171
Katz, B. J. (Ed.) (1995). Petroleum source rocks— An introductory over-
view. In: Petroleum source rocks (pp. 1– 8). Springer- Verlag; Berlin.
Katz, B. J., & Arango, I. (2018). Organic porosity: A geochemist's view
of the current state of understanding. Organic Geochemistry, 123,
1– 16. https://doi.org/10.1016/j.orgge ochem.2018.05.015
Katz, B. J., & Lin, F. (2021). Consideration of the limitations of thermal
maturity with respect to vitrinite reflectance, Tmax, and other prox-
ies. AAPG Bulletin, 105(4), 695– 720.
Kietzmann, D. A., Ambrosio, A. L., Suriano, J., Alonso, M. S.,
González Tomassini, F., Depine, G., & Repol, D. (2016). The Vaca
Muerta- Quintuco system (Tithonian– Valanginian) in the Neuquén
basin, Argentina: A view from the outcrops in the Chos Malal
fold and thrust belt. AAPG Bulletin, 100, 743– 771. https://doi.
org/10.1306/02101 615121
Ko, L. T., Loucks, R. G., Ruppel, S. C., Zhang, T., & Peng, S. (2017).
Origin and characterization of Eagle Ford pore networks in the
south Texas Upper Cretaceous shelf. AAPG Bulletin, 101, 387– 418.
https://doi.org/10.1306/08051 616035
Kozlowski, S., Cruz, C. E., & Sylwan, C. (1998). Modelo explorato-
rio en la faja corrida de la Cuenca Neuquina, Argentina. Boletín de
Informaciones Petroleras, 55, 4– 23.
Krim, N., Tribovillard, N., Riboulleau, A., Bout- Roumazeilles, V.,
Bonnel, C., Imbert, P., Aubourg, C., Hoareau, G., & Fasentieux, B.
(2019). Reconstruction of palaeoenvironmental conditions of the
Vaca Muerta Formation in the southern part of the Neuquén Basin
(Tithonian- Valanginian): Evidences of initial short- lived develop-
ment of anoxia. Marine and Petroleum Geology, 103, 176– 201.
https://doi.org/10.1016/j.marpe tgeo.2019.02.011
Lampe, C., Kornpihl, K., Sciamanna, S., Zapata, T., Zamora, G., &
Varadé, R. (2006). Petroleum systems modeling in tectonically
complex areas— A 2D migration study from the Neuquen Basin,
Argentina. Journal of Geochemical Exploration, 89, 201– 204.
Law, B. E., & Dickinson, W. W. (1985). Conceptual model for origin of
abnormally pressured gas accumulations in low- permeability reser-
voirs. AAPG Bulletin, 69, 1295– 1304.
Law, B. E. & Spencer, C. W. (1998). Abnormal pressures in hydrocar-
bon environments. In: B. E. Law, G. F. Ulmishek & V. I. Slavin
(Eds.), Abnormal pressures in hydrocarbon environments (Vol. 70,
pp. 1– 11). AAPG Memoir.
Leanza, H. A. (2003). Las sedimentitas huitrinianas y rayosianas
(Cretácico inferior) en el ámbito central y meridional de la cuenca
Neuquina, Argentina. Servicio Geológico Minero Argentino, Serie
Contribuciones Técnicas, Geología 2 (p. 31).
Leanza, H. A., & Hugo, C. A. (2001). Hoja Geológica 3969- I: Zapala:
Programa Nacional de Cartas Geológicas de la República Argentina
1:250.000. Instituto de Geología y Recursos Minerales Boletín 275
(p. 128).
Legarreta, L., & Uliana, M. A. (1991). Jurassic– marine oscillations and
geometry of back- arc basin fill, Central Argentina Andes. In D. I.
M. McDonald (Ed.), Sedimentation, tectonics and eustacy: Sea level
changes at active plate margins (pp. 429– 450). Blackwell Scientific
Publications.
Legarreta, L., ViIIar, H. J., Cruz, C. E., Laffitte, G. A., & Varadé, R.
(2008). Revisión integrada de los sistemas generadores, estilos de
migración- entrampamiento, y volumetría de hidrocarburos en los
distritos productivos de la cuenca Neuquina, Argentina. In C. E.
Cruz, J. F. Rodríguez, J. J. Hechern, & H. J. Villar (Eds.), Sistemas
Petroleros de las Cuencas Andinas (pp. 79– 108). Instituto Argentino
del Petróleo y el Gas.
Legarreta, L., & Villar, H. J. (2011). Geological and geochemical keys
of the potential shale resources, Argentina basins. AAPG Search and
Discovery Article 80196. Accessed November 7, 2011. http://www.
searc handd iscov ery.com/pdfz/docum ents/2011/80196 legar reta/
ndx_legar reta.pdf.html
Legarreta, L., & Villar, H. J. (2012). Las facies generadoras de hidrocar-
buros de la Cuenca Neuquina. Petrotecnia, 14– 39.
Legarreta, L., & Villar, H. J. (2015). The Vaca Muerta Formation (Late
Jurassic- Early Cretaceous), Neuquén Basin, Argentina: Sequences,
facies and source rock characteristics. In Extended Abstracts.
Liu, B., Schieber, J., & Mastalerz, M. (2017). Combined SEM and reflected
light petrography of organic matter in the New Albany Shale (Devonian-
Mississippian) in the Illinois Basin: A perspective on organic pore
development with thermal maturation. International Journal of Coal
Geology, 184, 57– 72. https://doi.org/10.1016/j.coal.2017.11.002
Löhr, S., Baruch, E., Hall, P., & Kennedy, M. (2015). Is organic pore
development in gas shales influenced by the primary porosity
and structure of thermally immature organic matter? Organic
Geochemistry, 87, 119– 132. https://doi.org/10.1016/j.orgge
ochem.2015.07.010
Loucks, R. G., Reed, R. M., Ruppel, S. C., & Jarvie, D. M. (2009).
Morphology, genesis, and distribution of nanometer- scale pores in
siliceous mudstones of the Mississippian Barnett Shale. Journal
of Sedimentary Research, 79, 848– 861. https://doi.org/10.2110/
jsr.2009.092
Małachowska, A., Mastalerz, M., Hampton, L., Hupka, J., & Drobniak,
A. (2019). Origin of bitumen fractions in the Jurassic- early
Cretaceous Vaca Muerta Formation in Argentina: Insights from
organic petrography and geochemical techniques. International
Journal of Coal Geology, 205, 155– 165. https://doi.org/10.1016/j.
coal.2018.11.013
Martínez, M. A., Prámparo, M. B., Quattrocchio, M. E., & Zavala, C. A.
(2008). Depositional environments and hydrocarbon potential of the
Middle Jurassic Los Molles Formation, Neuquén Basin, Argentina:
Palynofacies and organic geochemical data. Andean Geology, 35,
279– 305. https://doi.org/10.5027/andge oV35n 2- a05
Mastalerz, M., Drobniak, A., & Tankiewicz, A. B. (2018). Origin, prop-
erties, and implications of solid bitumen in source- rock reservoirs:
A review. International Journal of Coal Geology, 195, 14– 36.
https://doi.org/10.1016/j.coal.2018.05.013
McCarthy, K., Rojas, K., Niemann, M., Palmowski, D., Peters, K., &
Stankiewicz, A. (2011). Basic petroleum geochemistry for source
rock evaluation. Oilfield Review, 23, 32– 43.
McPeek, L. A. (1981). Eastern Green River basin: A developing giant
gas supply from deep, overpressured Upper Cretaceous sandstones.
AAPG Bulletin, 65, 1078– 1098.
|
27
EAGE
SPACAPAN et Al.
Mei, M., Burnham, A. K., Schoellkopf, N., Wendebourg, J., & Gelin,
F. (2021). Modeling petroleum generation, retention, and expulsion
from the Vaca Muerta Formation, Neuquén Basin, Argentina: Part I.
Integrating compositional kinetics and basin modeling. Marine and
Petroleum Geology, 123, 104743.
Milliken, K. L., Ko, L. T., Pommer, M., & Marsaglia, K. M. (2014).
SEM Petrography of Eastern Mediterranean sapropels: Analogue
data for assessing organic matter in oil and gas shales. Journal of
Sedimentary Research, 84, 961– 974.
Milliken, K. L., Reed, R. M., McCarty, D. K., Bishop, J., Lipinski,
C., Fischer, T. B., Crousse, L., & Reijenstein, H. (2019). Grain as-
semblages and diagenesis in the Vaca Muerta Formation (Jurassic-
Cretaceous), Neuquén Basin, Argentina. Sedimentary Geology, 380,
45– 64. https://doi.org/10.1016/j.sedgeo.2018.11.007
Milliken, K. L., Rudnicki, M., Awwiller, D. N., & Zhang, T. (2013).
Organic matter– hosted pore system, Marcellus formation
(Devonian), Pennsylvania. AAPG Bulletin, 97, 177– 200. https://doi.
org/10.1306/07231 212048
Mitchum, R. M. Jr, & Uliana, M. A. (1985). Seismic stratigraphy of car-
bonate depositional sequences, Upper Jurassic- Lower Cretaceous,
Neuquén Basin, Argentina. In O. R. Berg, & D. G. Woolverton
(Eds.), Seismic stratigraphy II: An integrated approach to hydro-
carbon exploration (pp. 255– 274). American Association Petroleum
Geologists, Memoir 39.
Modica, C. J., & Lapierre, S. G. (2012). Estimation of kerogen porosity
in source rocks as a function of thermal transformation: Example
from the Mowry Shale in the Powder River Basin of Wyoming esti-
mation of kerogen porosity as a function of thermal transformation.
AAPG Bulletin, 96, 87– 108. https://doi.org/10.1306/04111 110201
Naipauer, M., Comerio, M., Lescano, M. A., Vennari, V. V., Aguirre-
Urreta, B., Pimentel, M. M., & Ramos, V. A. (2020). The Huncal
Member of the Vaca Muerta Formation, Neuquén Basin of Argentina:
Insight into biostratigraphy, structure, U- Pb detrital zircon ages
and provenance. Journal of South American Earth Sciences, 100,
102567. https://doi.org/10.1016/j.jsames.2020.102567
Neuzil, C. E., & Pollock, D. W. (1983). Erosional unloading and fluid
pressures in hydraulically "tight" rocks. The Journal of Geology, 91,
179– 193. https://doi.org/10.1086/628755
Nunn, J. A. (2012). Burial and thermal history of the Haynesville Shale:
Implications for overpressure, gas generation, and natural hydrof-
racture. Gulf Coast Association of Geological Society Journal, 1,
81– 96.
Olivera, D. E., Martínez, M. A., Zavala, C., Di Nardo, J. E., & Otharán,
G. (2020). New contributions to the palaeoenvironmental framework
of the Los Molles Formation (Early- to- Middle Jurassic), Neuquén
Basin, based on palynological data. Facies, 66, 1– 21. https://doi.
org/10.1007/s1034 7- 020- 00607 - 8
Ortiz, A. C., Crousse, L., Bernhardt, C., Vallejo, D., & Mosse, L.
(2020). Reservoir properties: Mineralogy, porosity, and fluid types.
In D. Minisini, M. Fantín, I. Lanusse Noguera, & H. Leanza (Eds.),
Integrated geology of unconventionals: The case of the Vaca Muerta
play, Argentina. AAPG Memoir 121 (pp. 329– 350).
Osborne, M. J., & Swarbrick, R. E. (1997). Mechanisms for generating
overpressure in sedimentary basins: A reevaluation. AAPG Bulletin,
81, 1023– 1041.
Otharán, G., Zavala, C., Arcuri, M., Di Meglio, M., Zorzano, A.,
Marchal, D., & Koller, G. (2020). Análisis de facies de fangolitas
bituminosas asociadas a flujos fluidos de fango. Sección inferior de
la Formación Vaca Muerta (Tithoniano), Cuenca Neuquina central,
Argentina. Andean Geology, 47, 384– 417.
Passey, Q. R., Bohacs, K., Esch, W. L., Klimentidis, R., & Sinha,
S. (2010). From oil- prone source rock to gas- producing shale
reservoir- geologic and petrophysical characterization of unconven-
tional shale gas reservoirs. In International Oil and Gas Conference
and Exhibition, Beijing, China, June 8– 10, 2010, SPE Paper 131350
(p. 29).
Pazos, P. J., Comerio, M., Fernández, D. E., Gutiérrez, C., Estebenet,
M. C. G., & Heredia, A. M. (2020). Sedimentology and Sequence
Stratigraphy of the Agrio Formation (Late Valanginian- Earliest
Barremian) and the Closure of the Mendoza Group to the North of
the Huincul High. In: A. Folguera & D. Kietzmann (Eds.), Opening
and Closure of the Neuquén Basin in the Southern Andes (pp. 237–
265). Springer.
Peters, K. E. (1986). Guidelines for evaluating petroleum source rock
using programmed pyrolysis. AAPG Bulletin, 70, 318– 329.
Peters, K. E., Burnham, A. K., Walters, C. C., & Schenk, O. (2018).
Guidelines for kinetic input to petroleum system models from open-
system pyrolysis. Marine and Petroleum Geology, 92, 979– 986.
https://doi.org/10.1016/j.marpe tgeo.2017.11.024
Peters, K. E. & Cassa, M. R. (1994). Applied source rock geochemistry.
In: L. B. Magoon & W. G. Dow (Eds.), The Petroleum System—
From Source to Trap (Vol. 60, pp. 93– 117). American Association
of Petroleum Geologists Memoir.
Peters, K. E., Schenk, O., Hosford Scheirer, A., Wygrala, B., & Hantschel,
T. (2017). Basin and petroleum system modeling, Chapter 11. In C.
S. Hsu & P. R. Robinson (Eds.), Handbook of petroleum technology
(2nd edn., pp. 381– 417). Springer.
Petersen, H. I., Sanei, H., Gelin, F., Loustaunau, E., & Despujols, V.
(2020). Kerogen composition and maturity assessment of a solid
bitumen- rich and vitrinite- lean shale: Insights from the Upper
Jurassic Vaca Muerta Shale, Argentina. International Journal of
Coal Geology, 229, 103575.
Pommer, M., & Milliken, K. (2015). Pore types and pore- size distribu-
tions across thermal maturity, Eagle Ford Formation, southern Texas.
AAPG Bulletin, 99, 1713– 1744. https://doi.org/10.1306/03051
514151
Ramdhan, A. M., & Goulty, N. R. (2010). Overpressure- generating mech-
anisms in the Peciko field, lower Kutai Basin, Indonesia. Petroleum
Geoscience, 16, 367– 376. https://doi.org/10.1144/1354- 07930 9- 027
Ramos, V. A., Mosquera, A., Folguera, A., & García Morabito, E.
(2011). Evolución tectónica de los Andes y del Engolfamiento
Neuquino adyacente. In Geología y Recursos Naturales de la
Provincia de Neuquén. Relatorio del VXIII Congreso Geológico
Argentino, Buenos Aires (pp. 335– 348).
Reed, R. M., & Loucks, R. G. (2015). Low- thermal- maturity (0.7%
VR) mudrock pore systems: Mississippian Barnett Shale, southern
Fort Worth Basin. Gulf Coast Association of Geological Society, 4,
15– 28.
Rocha, E., Brisson, I., & Fasola, M. (2018). Modelado de Sistemas
petroleros a lo largo de la faja plegada de cuenca neuquina. In M.
Gardini, M. L. Ayoroa, C. E. Cruz, M. Gomez, M. Limeres, P.
Malone, R. Manceda, G. Peroni, & H. Villar (Eds.), X Congreso de
Exploración y Desarrollo de Hidrocarburos, Mendoza, Argentina,
5– 9 November (pp. 301– 314).
Rodrigues, N., Cobbold, P. R., Loseth, H., & Ruffet, G. (2009).
Widespread bedding- parallel veins of fibrous calcite ('beef')
in a mature source rock (Vaca Muerta Fm, Neuquén Basin,
Argentina): Evidence for overpressure and horizontal compres-
sion. Journal of the Geological Society, 166, 695– 709. https://doi.
org/10.1144/0016- 76492 008- 111
28
|
EAGE
SPACAPAN et Al.
Roduit, N. (2008). JMICROVISION version 1.2.7: Image analysis tool-
box for measuring and quantifying components of high definition
images. Accessed November 1, 2012. http://www.jmicr ovisi on.com
Rojas Vera, E. A., Folguera, A., Zamora Valcarce, G., Giménez, M.,
Ruiz, F., Martínez, P., Bottesi, G., & Ramos, V. A. (2010). Neogene
to Quaternary extensional reactivation of a fold and thrust belt: The
Agrio belt in the Southern Central Andes and its relation to the
Loncopue trough (38°– 39° S). Tectonophysics, 492, 279– 294.
Rojas Vera, E. R., Mescua, J., Folguera, A., Becker, T., Sagripanti,
L., Fennell, L., Orts, D., & Ramos, V. A. (2015). Evolution of the
Chos Malal and Agrio fold and thrust belts, Andes of Neuquén:
Insights from structural analysis and apatite fission track dating.
Journal of South American Earth Sciences, 64, 418– 433. https://doi.
org/10.1016/j.jsames.2015.10.001
Romero- Sarmiento, M.- F., Ducros, M., Carpentier, B., Lorant, F.,
Cacas, M.- C., Pegaz- Fiornet, S., Wolf, S., Rohais, S., & Moretti, I.
(2013). Quantitative evaluation of TOC, organic porosity and gas re-
tention distribution in a gas shale play using petroleum system mod-
eling: Application to the Mississippian Barnett Shale. Marine and
Petroleum Geology, 45, 315– 330. https://doi.org/10.1016/j.marpe
tgeo.2013.04.003
Romero- Sarmiento, M. F., Ramiro- Ramirez, S., Berthe, G., Fleury,
M., & Littke, R. (2017). Geochemical and petrophysical source
rock characterization of the Vaca Muerta Formation, Argentina:
Implications for unconventional petroleum resource estimations.
International Journal of Coal Geology, 184, 27– 41. https://doi.
org/10.1016/j.coal.2017.11.004
Romero- Sarmiento, M. F., Rouzaud, J. N., Bernard, S., Deldicque,
D., Thomas, M., & Littke, R. (2014). Evolution of Barnett Shale
organic carbon structure and nanostructure with increasing matu-
ration. Organic Geochemistry, 71, 7– 16. https://doi.org/10.1016/j.
orgge ochem.2014.03.008
Rouquerol, J., Avnir, D., Fairbridge, D. H., Everett, J. H., Pernicone,
N., Ramsay, J. D. F., Sing, K. S. W., & Unger, F. (1994).
Recommendations for the characterization of porous solids. Pure
and Applied Chemistry, 68, 1739– 1758.
Sagasti, G., Foster, M., Hryb, D., Ortiz, A., & Lazzari, V. (2014).
Understanding geological heterogeneity to customize field devel-
opment: An example from the Vaca Muerta unconventional play,
Argentina. In Unconventional Resources Technology Conference,
Denver, Colorado (pp. 797– 816).
Sánchez, N. P., Coutand, I., Turienzo, M., Lebinson, F., Araujo, V., &
Dimieri, L. (2018). Tectonic evolution of the Chos Malal fold- and-
thrust belt (Neuquén Basin, Argentina) from (U- Th)/He and fission
track thermochronometry. Tectonics, 37, 1907– 1929. https://doi.
org/10.1029/2018T C004981
Scasso, R. A., Alonso, M. S., Lanes, S., Villar, H. J., & Laffitte, G.
(2005). Geochemistry and petrology of a Middle Tithonian
limestone- marl rhythmite in the Neuquén Basin, Argentina:
Depositional and burial history. In G. D. Veiga, L. A. Spalletti, J.
A. Howell, & E. Schwarz (Eds.), The Neuquén Basin, Argentina: A
case study in sequence stratigraphy and basin dynamics (Vol. 252,
pp. 207– 229). Geological Society, London, Special Publication.
Schieber, J. (2010). Common themes in the formation and preservation
of intrinsic porosity in shales and mudstones— Illustrated with ex-
amples across the Phanerozoic. In Society of Petroleum Engineers
Unconventional Gas Conference, Pittsburgh, Pennsylvania,
February 23– 25, 2010, SPE- 132370- MS (p. 10).
Schieber, J. (2013). SEM observations on ion- milled samples of
Devonian black shales from Indiana and New York: The petrographic
context of multiple pore types. In W. Camp, E. Diaz, & B. Wawak
(Eds.), Electron microscopy of shale hydrocarbon reservoirs. AAPG
Memoir 102 (pp. 153– 171).
Schieber, J., Lazar, R., Bohacs, K., Klimentidis, B., Ottmann, J., &
Dumitrescu, M. (2016). An SEM study of porosity in the Eagle
Ford Shale of Texas – Pore types and porosity distribution in a dep-
ositional and sequence stratigraphic context. AAPG Memoir, 110,
153– 172.
Shanley, K. W., & Cluff, R. M. (2015). The evolution of pore- scale
fluid- saturation in low- permeability sandstone reservoirs. AAPG
Bulletin, 99, 1957– 1990. https://doi.org/10.1306/03041 411168
Spacapan, J. B., D'Odorico, A., Palma, O., Galland, O., Vera, E. R.,
Ruiz, R., Leanza, H. A., Medialdea, A., & Manceda, R. (2020).
Igneous petroleum systems in the Malargüe fold and thrust belt,
Río Grande Valley area, Neuquén Basin, Argentina. Marine and
Petroleum Geology, 111, 309– 331. https://doi.org/10.1016/j.marpe
tgeo.2019.08.038
Spacapan, J. B., Palma, J. O., Galland, O., Manceda, R., Rocha, E.,
D'Odorico, A., & Leanza, H. A. (2018). Thermal impact of igne-
ous sill- complexes on organic- rich formations and implications for
petroleum systems: A case study in the northern Neuquén Basin,
Argentina. Marine and Petroleum Geology, 91, 519– 531. https://
doi.org/10.1016/j.marpe tgeo.2018.01.018
Spalletti, L., Franzese, J., Matheos, S., & Schwarz, E. (2000). Sequence
stratigraphy of tidally- dominated carbonate- siliciclastic ramp; the
Tithonian of the southern Neuquén Basin, Argentina. Journal of the
Geological Society of London, 157, 433– 446.
Stipanicic, P. N., Rodrigo, F., Baulies, O. L., & Martínez, C. G.
(1968). Las Formaciones presenonianas en el denominado Macizo
Nordpatagónico y regiones adyacentes. Revista de la Asociación
Geológica Argentina, 23, 67– 98.
Swarbrick, R. E., Osborne, M. J., & Yardley, G. S. (2002). Comparison
of overpressure magnitude resulting 809 from the main generating
mechanisms. In A. R. Huffman & G. L. Bowers (Eds.), Pressure
regimes in 810 sedimentary basins and their prediction. AAPG
Memoir 76 (pp. 1– 12).
Sweeney, J. J., & Burnham, A. K. (1990). Evaluation of a simple model
of vitrinite reflectance based on chemical kinetics. AAPG Bulletin,
74, 1559– 1570.
Sylwan, C. (2014). Source rock properties of Vaca Muerta
Formation, Neuquina Basin, Argentina. Simposio de Recursos No
Convencionales. Ampliando el Horizonte Energético. IX° Congreso
de Exploración y Desarrollo de Hidrocarburos (pp. 365– 386).
Tingay, M. R., Hillis, R. R., Swarbrick, R. E., Morley, C. K., & Damit,
A. R. (2009). Origin of overpressure and pore- pressure prediction in
the Baram province, Brunei. AAPG Bulletin, 93, 51– 74. https://doi.
org/10.1306/08080 808016
Tissot, B. P., & Welte, D. H. (1984). From kerogen to petroleum. In B.
P. Tissot & D. H. Welte (Eds.), Petroleum formation and occurrence
(pp. 160– 198). Springer.
Tomassini, F. G., Smith, L., Rodriguez, M. J., Kietzmann, D., Jausoro,
I., Floridia, M. A., Cipollone, M., Caneiro, A., Epele, B., Santillán,
N., & Medina, F. (2019). Semi- quantitative SEM analysis of the
Vaca Muerta Formation and its impact on reservoir characteriza-
tion, Neuquén Basin, Argentina. In Unconventional Resources
Technology Conference (URTeC), Denver, Colorado, 22– 24 July
2019 (pp. 3175– 3188).
Tunik, M. A., Folguera, A., Naipauer, M., Pimentel, M., & Ramos, V.
A. (2010). Early uplift and orogenic deformation in the Neuquén
Basin: Constraints on the Andean uplift from U- Pb and Hf isotopic
|
29
EAGE
SPACAPAN et Al.
data of detrital zircons. Tectonophysics, 489, 258– 273. https://doi.
org/10.1016/j.tecto.2010.04.017
Uliana, M., & Legarreta, L. (1993). Hydrocarbon habitat in a Triassic-
to- Cretaceous Sub- Andean setting: Neuquén Basin, Argentina.
Journal of Petroleum Geology, 16, 397– 420.
Urien, M. C., & Zambrano, J. J. (1994). Petroleum systems in the
Neuquén Basin, Argentina. In L. B. Magoon & W. G. Dow (Eds.),
The petroleum system– From source to trap. AAPG Memoir 60 (pp.
513– 534).
Veiga, R. D., Vergani, G. D., Brissón, I. E., Macellari, C. E., & Leanza,
H. A. (2020). The Neuquén Super Basin. AAPG Bulletin, 104,
2521– 2555. https://doi.org/10.1306/09092 020023
Vennari, V. V., Lescano, M., Naipauer, M., Aguirre- Urreta, M. B.,
Concheyro, A., Schaltegger, U., Armstrong, R., Pimentel, M., &
Ramos, V. A. (2014). New constraints on the Jurassic- Cretaceous
boundary in the High Andes using high- precision U- Pb data.
Gondwana Research, 26, 374– 385. https://doi.org/10.1016/j.
gr.2013.07.005
Vergani, G., Arregui, C., Carbone, O., Leanza, H. A., Danieli, J. C., &
Vallés, J. M. (2011). Sistemas petroleros y tipos de entrampamien-
tos en la Cuenca Neuquina. In Geologıa y Recursos Naturales de la
Provincia de Neuquén: XVIII Congreso Geológico Argentino (pp.
645– 656).
Vergani, G. D., Tankard, A. J., Belotti, H. J., & Welsink, H. J. (1995).
Tectonic evolution and paleogeography of the Neuquén Basin,
Argentina. In A. J. Tankard, S. R. Suárez, & H. J. Welsink (Eds.),
Petroleum Basins of South America. AAPG Memoir 62 (pp.
383– 402).
Wang, F. P., & Gale, J. F. (2009). Screening criteria for shale- gas sys-
tems. Gulf Coast Association of Geological Societies Transactions,
59, 779– 793.
Wang, G. (2020). Deformation of organic matter and its effect on
pores in mud rocks. AAPG Bulletin, 104, 21– 36. https://doi.
org/10.1306/04241 918098
Waples, D., & Tobey, M. H. (2015). Like space and time, transformation
ratio is curved. Search and Discovery Article #41713 (2015) Posted
October 26, 2015.
Weaver, C. E. (1931). Paleontology of the Jurassic and Cretaceous of
West Central Argentina. Memoirs of the University of Washington
(Vol. 1, p. 469). University of Washington Press.
Williams, K. E., & Madatov, A. G. (2005). Analysis of pore pressure
compartments in extensional basins. In 25th Annual, GCSSEPM
Foundation Annual Bob F. Perkins Research Conference
Proceedings: Petroleum Systems of Divergent Continental Margin
Basins (pp. 862– 890).
Wilson, R. D., & Schieber, J. (2016). The influence of primary and sec-
ondary sedimentary features on reservoir quality: Examples from
the Geneseo Formation of New York, U.S.A. In T. Olson (Ed.),
Imaging unconventional reservoir pore systems. AAPG Memoir
112 (pp. 167– 184).
Wygrala, B. P. (1989). Integrated study of an oil field in the southern Po
Basin, Northern Italy. Berichte Kernforschungsanlage Jülich, 2313,
217.
Zamora Valcarce, G., Rapalini, A. E., & Spagnuolo, C. M. (2007).
Reactivación de estructuras cretácicas durante la deformación mio-
cena, Faja Plegada del Agrio, Neuquén. Revista de la Asociación
Geológica de Argentina, 62, 299– 307.
Zamora Valcarce, G. Z., Zapata, T., & Ramos, V. A. (2011). La faja
plegada y corrida del Agrio. In Relatorio Geología y Recursos
Naturales de la provincia del Neuquén (pp. 367– 374).
Zanella, A., Cobbold, P. R., Ruffet, G., & Leanza, H. A. (2015).
Geological evidence for fluid overpressure, hydraulic fracturing and
strong heating during maturation and migration of hydrocarbons in
Mesozoic rocks of the northern Neuquén Basin, Mendoza Province,
Argentina. Journal of South American Earth Sciences, 62, 229– 242.
https://doi.org/10.1016/j.jsames.2015.06.006
Zapata, T., & Folguera, A. (2005). Tectonic evolution of the Andean
fold and thrust belt of the southern Neuquén Basin, Argentina. In:
G. D. Veiga, L. A. Spalletti, J. A. Howell, & E. Schwarz (Eds.), The
Neuquén Basin: A case study in sequence stratigraphy and basin
dynamics (Vol. 252, pp. 37– 56). The Geological Society, Special
Publication.
SUPPORTING INFORMATION
Additional supporting information may be found online in
the Supporting Information section.
How to cite this article: Spacapan, J. B., Comerio,
M., Brisson, I., Rocha, E., Cipollone, M., & Hidalgo,
J. C. (2021). Integrated source rock evaluation along
the maturation gradient. Application to the Vaca
Muerta Formation, Neuquén Basin of Argentina.
Basin Research, 00, 1– 29. https://doi.org/10.1111/
bre.12599
... Sample C1 was obtained from the Well C Donadelli et al. Organic Geochemistry xxx (xxxx) 104690 (early oil window), C3 from the Well B (wet gas) and samples C4 and C5 from the Well A (dry gas) analyzed by means of organic geochemistry by (Spacapan et al., 2021). Sample C2 comes from the Well D (peak of oil window) located close to the boundary between the NE platform and the Chihuido-Lomita areas (Brisson et al., 2020). ...
... Oxygen Index of Oori sample is 30 mgCO 2 /gTOC, a relatively higher value than the obtained for natural core rocks (with values between 6 and 10 mgCO 2 /gTOC). Although some grade of weathering or a contribution of organic matter richer in oxygen cannot be discounted for outcrop samples, OI values between 8 and 47 mgCO 2 /gTOC are normal for the Vaca Muerta Formation (Spacapan et al., 2021). After the exposure to temperature, OI sharply decrease to a value of 2 mgCO 2 /gTOC indicating a markedly loss of oxygen. ...
Article
Artificial maturation of kerogen is a widely used technique to assess the hydrocarbon potential of shale rocks and to observe thermal transformations of organic matter in the laboratory. However, the degree of reproducibility of natural geological transformation at the molecular level is still not fully established. In the present work, a set of kerogens isolated from Vaca Muerta Formation core samples at varying levels of thermal maturity were studied. Another set of samples was obtained by anhydrous pyrolysis of kerogen in a closed system. Molecular structures measured by solid-state techniques (13C NMR, XPS and FTIR) were compared in natural and artificially matured samples at equivalent levels of thermal maturation. We observed that heating in an anhydrous closed system accurately reproduces most of the molecular structural changes observed during natural thermal maturation, with exceptions related to the branching degree and oxygen containing groups. The evolution of some parameters, such as N and S content, are highly variable in the natural samples because of differences in their deposition environments. In these cases, artificial thermal maturation changes are smaller and follow more clearly defined trends because variables other than thermal changes are not present. This is the first work comparing natural and artificial thermal maturation in the Vaca Muerta Formation and add valuable data regarding the limitation of artificial maturation that can be extrapolated to other formations.
... Elemental composition of mineral phases was determined by EDS (energy-dispersive X-ray spectroscopy). A semi-quantitative pore-size distribution in the organic matter was measured at different magnifications with 2D-SEM images by means the JMicrovision software (Roduit, 2008) and considering spherical to oval forms as an indicative of the thermal transformation of organic matter (Spacapan et al., 2021). ...
... Other samples analyzed herein are thermally altered which prevents an ambiguous reconstruction of the kerogen type based only on rock eval data alone. Organicrich shales such as those recorded in the Mississippian Barnett (Texas), the Upper Jurassic Haynesville (Texas, Arkansas, Louisiana), and the Upper Jurassic-Lower Cretaceous Vaca Muerta formations, are characterized by organic pores which provide space for oil and gas storage (Peters et al., 2017;Brisson et al., 2020;Spacapan et al., 2021). In fact, isolated bubble pores hosted in solid bitumen (Fig. 5) reflect that at higher thermal conditions the hydrocarbon storage potential would be partially controlled by organic pores. ...
Article
The present study explores the oil and gas potential of the Huallaga-Marañon Andean retroforeland basins in Perú, considering a regional transect which extends from the Eastern Cordillera in the west trough the Huallaga fold-and-thrust belt at the Andean deformation front to the Marañon Basin in the east. The sequential restoration of main tectonic events was derived from previous stratigraphic and structural interpretations revisited with reprocessed seismic lines, and new geochemical, petrographical, geochronological and thermochronological data. The resulting framework was tested in a 2-D Basin and Petroleum System Modeling (BPSM) to check the relationship between burial history, thermal maturity evolution and timing of hydrocarbon generation, expulsion and trapping at regional scale. BPSM suggests that during the Late Cretaceous (ca. 80 Ma) the petroleum generated in the Upper Triassic-Lower Jurassic source rock known as the Aramachay Formation migrated through the Middle Upper Jurassic Sarayaquillo Formation basal sandstones up to 250 km and accumulated in the southeast of the Marañon Basin. From the Late Miocene, the reactivation of vertical faults favored the migration and accumulation of petroleum in Cretaceous siliciclastic reservoirs which is consistent with the hydrocarbon fields existence in this sector of the basin. In addition, results of modeling show that in the southwest of the Marañon Basin the development of structures during Late Miocene times postdated the generation and expulsion of hydrocarbons calculated for the Ara-machay Formation which introduces a risk of synchronization between oil migration and charge. However, part of the generated hydrocarbons during the Late Cretaceous and Cenozoic are partially preserved in Jurassic reservoirs due to the high seal capacity of mudstones of the Sarayaquillo Formation. Such hydrocarbon fluids are redistributed and migrated into structural traps during Miocene times. In addition, the Miocene gas generation and expulsion also will reduce the risk of synchronization with the development of structures. In the Eastern Cordillera, organic-rich levels of the Aramachay Formation are in the late immature to early oil window show the occurrence of organic pores adding hydrocarbon storage capacity, suggesting this unit to be considered as a potential unconventional target in future exploration activities.
... However, the leakage system of pyrolysis gas is controlled by faults, diapirs, and other deep structures. For example, the charging period of overpressure-related fractures is subjected to the generation and development of overpressure [83], and the formation of structural fractures is controlled by tectonic activities [4,84]; meanwhile, polygonal fractures are related to the mechanism implicated in the formation of local fractures [81]. Therefore, spatial configuration models of fractures and traditional transport systems, such as diapirs, fractures, sand bodies, and hydrate reservoirs, are significant in the evaluation of seepage systems. ...
Article
Full-text available
The Qiongdongnan Basin (QDNB), located in the north of the South China Sea, is a Cenozoic rift basin with abundant oil and gas resources. Large flake hydrates have been found in the core fractures of Quaternary formations in the deep-water depression of the QDNB. In order to understand the spatial distribution patterns of these fractures, their geneses in sedimentary basins, and their influences on gas migration and accumulation, such fractures have been observed using high-resolution 3D seismic images and visualization techniques. Four types of fractures and their combinations have been identified, namely bed-bounded fractures/microfaults, unbounded fractures, fracture bunches, and fracture clusters. Bed-bounded fractures/microfaults are mainly short and possess high density; they have developed in mass transport depositions (MTDs) or Meishan and Sanya Formations. The unbounded fractures/microfaults that occur in Miocene–Pliocene formations are mainly long and discrete, and are dominantly caused by strong tectonic movements, the concentration of stress, and sustained intense overpressure. The fracture bunches and fracture clusters that occur in Oligocene–Early Miocene formations have commonly developed with the accumulation of large numbers of fractures and may be related to the release of pressure, diapirs, and basement fault blocks (228.9 ± 1 Ma). In this study, six fluid charging or leakage models are proposed based on distinct fracture types, assuming the uniform conductivity of each fracture. In a 3D space view, a vertical decrease in the fracture scale (number or density) will more likely result in gas supply than dispersion, thus promoting the accumulation of gas in the reservoirs. Nevertheless, the fractures above the Bottom Simulating Reflect (BSR)/seismic anomaly are excessively developed, and bed-bounded fractures within a particular layer, such as MTDs, can easily cause seabed leakage. These results are useful for explaining the vertical migration of gas/fluids in areas and formations with less developed gas chimneys, faults, diapirs, and other structures, particularly in post-rifting basins.
... The Vaca Muerta Formation of the Neuquén Basin Argentina, is a marine source rock with excellent characteristics from the perspective of source rockunconventional plays: (1) high present-day TOC content (1 -8% with peaks of 12%); (2) original hydrogen index (HIo) estimated to be approximately 680 mg HC/g TOC; (3) moderate depths of 3.150 m; (4) overpressure distribution in the order of 47-61 MPa, (5) lack of expandable clays in mature areas; and (6) different landing zones for the production of oil, wet gas and dry gas [3,4,5,6,7,8]. ...
Article
Full-text available
The Vaca Muerta Formation (Upper Jurassic–Lower Cretaceous) of the Neuquén Basin, Argentina is a world-class source rock and renowned as an unconventional reservoir for both oil and gas. The present study examines rotary sidewall cores representative of the oil and gas windows to analyze correlations between nuclear magnetic resonance (NMR T2 and T1-T2 maps at 2 MHz), Gas Filled Porosity (GFP), Rock-Eval programmed pyrolysis and quantitative X-ray diffraction (XRD) mineralogy. Shale samples are characterized by high present-day total organic carbon contents ranging from 3 to 7.20% (mean: 4.30%) for Well A at the oil window and, 0.72 to 11.77% (mean: 5.86%) for Well B at the gas window. The analyzed set also covers a wide compositional spectrum from carbonate-rich (> 50% of calcite and minor dolomite), mixed carbonate-siliciclastic (30 to 50% of calcite), to siliciclastic-rich (< 30% of calcite) samples. Notably, the clay fraction is dominated by interstratified illite-smectite minerals with less than 10% of expandable layers. Total porosity, calculated from Gas Filled Porosity plus NMR T2 Cumulative Distribution, is in the range of 10 to 20 porosity units. In this work we also present a new NMR sequence to detect solid-like organic matter with NMR at 2 MHz in an integrated workflow to characterize petrophysical properties of the Vaca Muerta Formation.
Article
Full-text available
Fibrous calcite bed‐parallel veins (BPVs) are a typical feature of the Upper Jurassic – Lower Cretaceous Vaca Muerta Formation in the subsurface of the Neuquén Basin (Argentina). The formation is considered to be the main source rock in the basin as well as an important unconventional play. This study examines the growth of BPVs through an analysis of core from three wells located along a transect extending for some 150 km from the NE Platform near the basin margin in the east to the Agrio fold‐and‐thrust belt at the Andean deformation front in the west. The main objective is to integrate fluid inclusion data with the palaeothermal and palaeopresure evolution obtained from a regional‐scale 2D basin and petroleum systems model to examine the timing of fracture development and its relationship with hydrocarbon generation in the Vaca Muerta Formation through time. The apertures of BPVs were measured in more than 360 m of core from three wells (wells A, D and E). This data was combined with optical petrography to investigate the number of calcite cementation events, and the temperature of cement precipitation based on fluid inclusion data. The organic geochemical and mineralogical characteristics of the Vaca Muerta source rock were also analysed. The integrated results were incorporated into a poro‐elastic basin model to investigate the impact of horizontal shortening due to Andean compression on pore pressure development and fracturing in the Vaca Muerta Formation. This framework allowed the timing of BPV formation to be determined together with possible mechanisms governing overpressure conditions through time. Near the Andean deformation front in the west of the modelled section where the Vaca Muerta Formation is in the wet gas window (well D) and dry gas window (well A), BPVs are characterized by two or more generations of calcite fibres indicating multiple growth phases. Calcite which precipitated during cementation event 1 (E1) in the internal zones of BPVs consists of crystals oriented perpendicular to fracture walls, indicating perpendicular vein opening. Calcite precipitated during cementation event 2 (E2) in the outer zones of BPVs includes curved and oblique crystals. During this phase, shear occurred between the opening vein walls as a result of horizontal shortening. Cementation event 3 (E3) is characterized by an equant mosaic of calcite crystals which preserve intracrystalline porosity. E1cements formed between 110 and 90 Ma with trapping temperatures of ∼112 °C (upper Vaca Muerta, well A) and ∼125 °C (lower Vaca Muerta, well D). Fracturing resulted from disequilibrium compaction and from volumetric expansion due to primary cracking of kerogen within the oil window. E2 cements record a trapping temperature of ∼159 °C and formed between 70 and 55 Ma (lower Vaca Muerta, well D) during maximum burial of the Vaca Muerta Formation, synchronous with the secondary cracking of retained liquid hydrocarbons and the beginning of Andean compression. E3 cements (upper Vaca Muerta, well A) have a trapping temperature of ∼162 °C, and formed between 65 Ma and 53 Ma synchronous with the generation of thermogenic gas. By contrast, in the east of the modelled section in the less deformed foreland area of the Neuquén Basin where the Vaca Muerta Formation is in the early oil window (well E), BPVs are composed of a single generation of calcite fibres (E1)with a trapping temperature of ∼118 °C. The E1 cement is characterized by calcite crystals which are oriented perpendicular to fracture walls with no evidence of shearing. According to model simulations, cementation here occurred between 64 Ma and 53 Ma during maximum burial and was related to overpressures which resulted from both disequilibrium compaction and primary transformation of kerogen into oil. The data presented suggests that in some intervals of the Vaca Muerta Formation, a slight increase in TOC content is accompanied by an increase in vein thickness, with the highest number of cementation events occurring towards the Andean deformation front in the west of the study area compared to the foreland in the east.
Thesis
Diagnostic fracture injection tests (DFIT) are used as an indirect method to determine closure pressure and formation effective permeability in unconventional reservoirs as a first step in formation evaluation. The information obtained from DFIT is particularly useful because it is obtained before any production for a given well is available. In DFIT, a small fracture is created by injecting few barrels of completion fluid until formation breaks down and a fracture is initiated and propagates a short distance into the reservoir. Then, injection is stopped and the pressure decline (or falloff) is monitored. From this pressure decline, the effective permeability of the formation is estimated by Nolte’s G-function, log-log plot, or square root of time analysis. In this research, the viability of the common DFIT analysis techniques was investigated for unconventional reservoirs with and without micro-fractures by using a numerical hydraulic fracturing simulator, CFRAC. The results of numerical simulations were investigated to assess the impact of permeability, residual fracture aperture, and complex fracture networks on conventional DFIT interpretations. For the example considered in this work, the commonly used G-function analysis yielded estimates of permeability over an order of magnitude higher than the simulated matrix permeability. Error in the G-function estimates of permeability were higher for higher matrix permeability and in the existence of a fracture network. On the other hand, straight-line analysis of ∆p versus G-time yielded much closer (in the same order of magnitude) estimates of permeability.
Article
Petroleum System Modelling (PSM) is a method which reproduces the burial history of sedimentary basins, together with rock properties, thermal and stress state, fluid flow and chemical transfers. The Vaca-Muerta Formation (Fm) in the Neuquén foreland basin, Argentina, presents exceptionally high overpressure despite the latest erosional history coeval with the current basin shortening. PSM currently accounts for vertical compaction laws only, which are not sufficient to match the observed pore pressure while keeping permeability values realistic. It also prevents to discuss the relationships between natural fracturing and the basin hydrodynamic. To assess these phenomena, a code coupling a PSM and a geomechanical simulator is used, in which we consider a 3D poro-elastoplastic geomechanical framework that accounts for both burial and tectonic compaction. Using this coupled approach, we calibrate porosity and pore pressure in a 3D geological model of the Neuquén basin using successives tectonic shortening phases related to the Andean subduction. Compared to PSM results using similar parameters, this study quantifies how much Andean tectonic deformation influenced pore pressure evolution in the basin. It shows that Late Miocene to recent tectonic could explain most of the overpressure observed in the Vaca-Muerta Fm. A shear-induced fracturing index provided by the constitutive model suggests that fracturing in the Vaca-Muerta Fm is very likely to occur during one of the main Andean deformation phases. The work suggests that pore pressure prediction in regions that have been subjected to lateral tectonic loading should be handled considering a 3D geomechanical approach. Using the Neuquén basin as an example, the present study discusses the impact of tectonic in pore pressure evolution, and its role in natural fluid migration in sedimentary basins subjected to tectonic deformation.
Article
The Vaca Muerta-Quintuco (VM-Q) system of Tithonian-early Valanginian age was studied in the Chacay Melehue section of the Neuqu´en Basin (western Argentina) by means of sedimentological, mineralogical and geochemical analyses in order to determine the main driving factors that triggered the paleoenvironmental change from a carbonate ramp (Vaca Muerta Formation, VMFm) to a mixed siliciclastic/carbonatic marine environment (Quintuco Formation, QFm). The VMFm was divided into two stratigraphic intervals: Lower VMFm (LVMFm) and Upper VMFm (UVMFm), whereas the QFm is subdivided into the Puesto Barros Member (PBMb) and the Cerro La Visera Member (CVMb), which can be correlated to other sections in the basin (e.g., Puerta Curaco). Isolated, turbiditic sandstone beds, correlated to the Huncal Member, are included in the QFm. The LVMFm (Tithonian) and the UVMFm (Berriasian) are constituted by dark, well-laminated marls, mudstones, calcite concretions and tuffs. The PBMb (earlier early Valanginian) is constituted by marls and sandstones, whereas the CVMb (later early Valanginian) is constituted by marls, mudstones, siltstones, sandstones and coquinas. The LVMFm (Total organic carbon, TOC ~ 1–4 wt%) is characterized by the enrichment of redox sensitive trace elements (RSTE), where the enrichment of Ni and Cu suggest high productivity in the water column, and the enrichment of Mo, U, V points to sea bottom anoxia, with periods of increased oxygenation as deduced from higher P concentrations in marls and mudstones. The clay mineral association is constituted by mixed-layer illite/smectite formed by the transformation of smectitic layers. The predominance of smectite in coeval sucessions, less overprinted by burial diagenesis, suggests a temperate and semi-arid climate in the adjacent continent. The UVMFm (TOC ~ >1 wt%) is characterized by a gradual decrease of the RSTE pointing to a decrease in productivity and a slight increase in the oxygenation of the sea bottom. In addition, a change towards more humid conditions in the continent is inferred by both the increase in the Chemical Index of Alteration and the Al2O3/TiO2 ratio. The PBMb (TOC ~1 wt%) has even lower content of RSTE indicating diminished sea water productivity and a gradual rise of the oxygenation of the sea bottom. In this interval, increased illite contents suggest periods of enhanced physical weathering, probably related to the tectonic uplift of the Huincul Ridge. The RSTE in the CVMb (TOC <1 wt%) documents a fully oxygenated sea bottom, where the productivity of the water column was negligible. The presence of kaolinite and the increment of the detrital sedimentation in the CVMb indicate a change towards more humid conditions in the hinterlands. The enhanced runoff caused by this paleoclimatic change towards more humid conditions that started in the early Berriasian and increased during the later early Valanginian triggered the change from carbonate ramp to mixed siliciclastic/carbonatic to siliciclastic marine paleoenvironments. The organic carbon isotope composition (δ13Corg vs. VPDB) of the VMFm ranges between - 30.0 and - 23.4‰, whereas in the QFm values range between - 29.0 and - 23.9‰. Two positive carbon isotope excursions (PCIE) are recorded in the system: PCIE-A in the lower part of the VMFm (early Tithonian) with δ13Corg values ~ - 25‰, and PCIE-B in the upper part of the QFm (later early Valanginian) with δ13Corg values ~ - 24.5‰. The shift in δ13Corg at PCIE-B is up to +4.2‰ and marks the onset of the Weissert Event. This is the first time that the onset of this event is recorded in the Neuqu´en Basin, within the Lissonia riveroi ammonite zone. The results of our study confirm that clay mineralogy, trace elements and stable isotopes are valuable proxies of past ocean-climate variability even in sediment deposits that underwent nearsurface and deep burial diagenesis.
Article
With more than 6000 m of sedimentary thickness and several superimposed petroleum systems, the Neuquén Basin of Argentina has all the elements to be considered a super basin. The basin developed during the Triassic–Early Jurassic in a rift environment that generated a localized petroleum system. Continued subsidence during the Early Jurassic resulted in the first marine ingression (Los Molles Formation, total organic carbon [TOC] 1%–7%). This unit matured 120 to 55 m.y.a., before the formation of major structures, with the exception of the Huincul high. A fast marine flooding in a retroarc setting occurred during the latest Jurassic–Early Cretaceous. At its base lies the Vaca Muerta Formation, a 50- to 1000-m-thick world-class source rock, composed by calcareous shales and marls (TOC 2%–9%). This unit reached maturity at 95 Ma, when the only traps available were structures along the Huincul high or stratigraphic. A new marine ingression occurred during the Early Cretaceous, depositing bituminous shales and marls (Agrio Formation, TOC 2%–5%) that matured during the Paleogene. These superimposed petroleum systems, combined with different structural settings, gave rise to numerous plays and trapping configurations that resulted in the discovery of 14 billion BOE. In addition, the unconventional potential of the Vaca Muerta is now being developed, with estimated resources of 91.5 TCF and 14.3 billion bbl of oil. The Neuquén Basin has gone through three main phases of development: conventional, tight, and unconventional. After a century of production, the basin still shows a wide variety of opportunities for both conventional and world-class unconventional plays.
Article
Improved understanding and prediction of petroleum expulsion, retention, and producibility from source rocks (SR) are achieved by integrating 3D basin modeling with customized compositional kinetics and retention model for the lower Vaca Muerta Formation (LVM), Neuquén Basin, Argentina. Modeling results indicate basin-wide expulsion from LVM started in the late Cretaceous when petroleum generation exceeded the SR retention capacity at thermal maturity of 0.9-1.0% R o , associated with increasing pore pressure (P p), increasing then maintained porosity (ø), permeability (k), and free petroleum retained in the SR (free petroleum). Expulsion ceased around the middle Eocene when kerogen, polar components, and medium aromatics were totally cracked at about 1.1-1.3% R o. In the late Eocene, a different type of expulsion occurred in deeply buried areas at > 1.3% R o due to continued cracking of the remaining larger hydrocarbons to form smaller hydrocarbons. The petroleum cracking created volume expansion, which in turn induced rock fracturing, which then reduced P p , ø, k, and free petroleum. Then, continuous thermal cracking increased P p and maintained ø, k, and free petroleum. In the mid-Miocene, tectonic uplift stopped the thermal stress and secondary cracking, and decreased pressure and temperature. Modeled petroleum compositions and properties such as gas wetness, oil • API gravity, gas/oil ratio are calibrated to Vaca Muerta (VM) core extracts, produced oils, and mud gases at wells with maturities ranging from oil window to dry gas. Our study indicates larger and polar components are preferentially sorbed within the source rock and cracked through thermal maturation by 1.3%R o. Genetic saturates/aromatics ratio (S/A) equals about one at <0.8% R o , and gradually increases in all retained fluids due to preferential generation of saturates and preferential thermal cracking of aromatics, which are mostly consumed by 1.65%R o. The preferential sorption of aromatics causes a higher S/A in the free oil. The method and the insights of this study improve the fundamental understanding of petroleum expulsion and retention and can be applied to global plays.
Article
Assessment of thermal maturity by solid bitumen reflectance measurements in vitrinite-lean or vitrinite-barren shales is widely used in unconventional shale gas and shale oil plays but the method is commonly challenged by the complexity in the measurements. The Upper Jurassic marine Vaca Muerta shale in the Neuquén Basin in western Argentina is both a prolific conventional source rock and a primary unconventional reservoir for oil and gas. The shale is rich in solid bitumen and lean or barren in vitrinite and offers an excellent opportunity to gain insight into thermal maturity assessment of a shale with a complex organic matter composition. The kerogen composition of a set of immature to highly mature Vaca Muerta shale samples was investigated by organic petrography and reflectance measurements were taken on vitrinite and solid bitumen. The results were integrated with available Rock-Eval data and stable carbon isotope values of shale gas (ethane and propane) collected at the same depth as the rock samples. The maceral composition of immature Vaca Muerta shale is highly sapropelic and oil-prone with Hydrogen Index values of 600 mg HC/g TOC or higher. The liptinite macerals consist primarily of alginite with minor amounts of inertinite and/or oxidized vitrinite. Indigenous vitrinite is generally scarce and in some samples absent. Vitrinite reflectance values range from 0.79%Ro to more than 1.94%Ro corresponding to early peak oil to dry gas maturity. Solid bitumen is abundant in all mature samples. Two general types of solid bitumen morphology were observed: (i) BR I, a dark reflecting porous variant disseminated in the clayey mineral matrix and (ii) BR II, a lighter reflecting more ‘dense’ textured variant. Bands of denser solid bitumen were commonly seen ‘flowing’ in more porous or weak zones in the matrix and around mineral grains. Degassing pores were frequently observed. Solid bitumen reflectance measurements on BR I and BR II provided two populations with different reflectance ranges. The correlation between vitrinite and BR II reflectances is considered to have highest confidence. The correlation can be expressed by the equation VReqv = (BR II + 0.3043)/1.026. The established VReqv-BR II correlation yields results comparable to other published correlations but also demonstrates significant discrepancy to others. This warrants caution when assessing the maturity of complex unconventional shale gas and oil plays by translating solid bitumen reflectance values into vitrinite reflectance equivalents using correlations not calibrated to that specific shale play. Both Hydrogen Index values and ethane and propane stable carbon isotope values of Vaca Muerta shale gas show strong correlations to VReqv derived from BR II. Thus, in cases with absence of vitrinite and questionable solid bitumen measurements reasonable maturities (vitrinite reflectance equivalents) can be inferred from these relationships.