You are on page 1of 18

Journal of Petroleum Science and Engineering 213 (2022) 110410

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: www.elsevier.com/locate/petrol

Geochemical evaluation of sedimentary rocks and hydrocarbon fluids in the


Amistad offshore field, Progreso Basin Province, Southwest Ecuador
M.A. Guzmán a, G. Márquez a, *, C. Boente a, C. Witt b, A. Morato c, R. Tocco d
a
Center for Research in Sustainable Chemistry (CIQSO), University of Huelva, 21006, Huelva, Spain
b
Laboratoire D’Océanologie et de Géosciences, University of Lille, CNRS, F59000, Lille, France
c
School of Engineering Sciences, State University Santa Elena Peninsula, 240204, La Libertad, Ecuador
d
Independent Petroleum Geochemistry Consultant, Madrid, 28410, Spain

A R T I C L E I N F O A B S T R A C T

Keywords: This study contains a thorough isotopic and molecular characterization of a series of thirteen (13) gas samples
Gas geochemistry and six (6) liquid hydrocarbons produced from late Miocene reservoirs in the offshore Progreso Basin (SW
C1–C4 alkanes Ecuador), along with Rock-Eval and petrographic analyses of sixty-two (62) rock samples from the area. The
CSIA-D
principal aim of this research work is to investigate the geochemistry of hydrocarbon fluids and sedimentary
Reservoir compartments
Amistad field
rocks in the area of the Amistad field to determine the origin of the studied wet gases. Potential source rocks in
the Amistad stratigraphic section are immature and have no oil-generating capability, except for the early
Miocene Dos Bocas Formation. Studied low-boiling condensates have similar gas chromatographic fingerprints
while their compound-specific isotopic analyses of lower diamondoids (CSIA-D) indicate an oil-source correla­
tion between them and Dos Bocas extracts. Sampled gases are mainly methane of biogenic origin except for that
from the Delfín B-17X well, which might represent the thermogenic end-member gas signature of the Amistad
gases. Results suggest a complex hydrocarbon filling history with multiple charges of thermogenic gas and hy­
drocarbons generated from the distal deltaic Dos Bocas source rocks followed by biodegradation, leakage and
partial displacement by biogenic methane due to the uplifting of the Amistad structure. The Amistad gases form
three homogeneous groups located in the southern, central and northern parts of the Progreso reservoir, although
the geological segmentation of it into compartments cannot be clearly identified.

CRediT author statement mechanisms as well as an abnormally high temperature gradient (Hig­
ley, 2004). The evolution of the Gulf of Guayaquil-Tumbes Basin is
Gonzalo Márquez: Conceptualization, Writing - Original Draft, linked with extensional and strike-slip processes resulting from the
Formal Analysis, Supervision, Data Curation. Carlos Boente: Software, tectonic escape of the North Andean Block (or North Andean Sliver),
Data Curation, Visualization, Methodology. Marco Antonio Guzmán: mostly a crustal fragment accreted to the continent from Late Cretaceous
Investigation, Methodology. César Witt: Writing - Review & Editing, times (e.g. Jaillard et al., 1999). The locus of the Gulf of
Formal Analysis. Antonio Morato: Funding acquisition, Validation, Guayaquil-Tumbes Basin being related with the trailing-tail of the
Formal Analysis. Rafael Tocco: Visualization, Supervision. microplate in a complex interaction with weaknesses zones resulting
from the accretional processes (Aizprua, 2021).
1. Introduction Hydrocarbon exploration in the southern coastal Ecuador was
encouraged by the presence of several oil seeps and the Ancón field
The Progreso Basin Province includes the Santa Elena Block and the discovery in 1915 (Sheppard, 1937). During the 1940s a branch of
Gulf of Guayaquil-Tumbes Basin (Higley, 2004), which comprises the Exxon attempted to extend this oilfield southward along the Santa Elena
Esperanza, Tumbes and Jambelí sub-basins (Witt et al., 2006; Aizprua Peninsula, encountering numerous oil and gas shows (Stainforth, 1948).
et al., 2019). This region has been defined as an oil petroliferous prov­ Exploration efforts by subsidiaries of Chevron and Tenneco in the 1950s
ince because of the presence of source rock intervals, trapping and 1960s, respectively, did not report important hydrocarbon

* Corresponding author.
E-mail address: gonzalo.marquez@diq.uhu.es (G. Márquez).

https://doi.org/10.1016/j.petrol.2022.110410
Received 26 October 2021; Received in revised form 9 January 2022; Accepted 15 March 2022
Available online 17 March 2022
0920-4105/© 2022 Elsevier B.V. All rights reserved.
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

accumulations in the region (Marchant, 1965). During the early 1970s, a Since exploration begun, around 8 billion m3 of natural gas have
Phillips consortium carried out extensive exploration in the Gulf of been produced from southern coastal Ecuador, with Amistad field hav­
Guayaquil. This resulted in the discovery of the Amistad gas field ing an average annual gas production of about 40 million of ft3 and
(located 65 km northwest of the city of Machala; Fig. 1a), in which, after recoverable gas reserves estimated to be in the order of one trillion ft3
a drilling program in 1969, gas was found at some wells from the (Evangelista, 2019). A previous work (Encalada, 2017) studied gas
Miocene Progreso Formation. This offshore field is an anticlinal struc­ composition and reported abundances of ca. 97% of methane in some
ture: the presence of natural gas is reported on both sides of a large Amistad offshore wells but to date, no thorough gas analyses have been
NE-SW trending fault zone, which deepens to the east. The size of the reported from the Amistad field so the origin of these gases, along with
reservoir is approximately 25 km2 and the pay zone can be as thick as 85 their potential source rocks, remain unknown.
m (Bristow, 1975). Also in the 1970s, Tenneco drilled several explora­ The molecular composition of natural gas depends on a series of
tion wells in the Peruvian sector of the Gulf of Guayaquil, which led to geological and geochemical factors, such as source kerogen type,
the discovery of the Albacora oilfield (Evans and Whittaker, 1982). maturity, thermal cracking, biodegradation, migration, and/or mixing
Exploratory efforts in the Progreso Basin Province for following decades (Behar et al., 1992; Pallasser, 2000; Gürgey et al., 2005; Tian et al.,
reported no economic results (Higley, 2004). 2009). Generally, carbon and hydrogen isotope signatures of C1–C4

Fig. 1. a) Location of the B-17X gas well, boreholes, and fields in the Progreso Basin Province; b) map showing the situation of wells and compartmentalization in the
Amistad gas field.

2
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

gaseous hydrocarbons can be used to identify the genetic types of nat­ is not clear whether connectivity exists through the fault zone or later­
ural gas (e.g., Jenden et al., 1993; Patience, 2003; Wang et al., 2018; El ally in the southern part of the structure where the fault zone terminates
Diasty et al., 2020) and to assess gas maturity within each genetic type within the closing contour indicative of the hydrocarbon/water contact,
(Clayton, 1991; Dai et al., 2005; among others). and if local faulting has affected the reservoir connectivity and has
In this study, carbon and hydrogen isotopic analyses and C1–C4 gas caused reservoir separation into smaller compartments (Barzallo and
compositions are complemented with analyses of C5–C7 light hydro­ Bemúdez, 2016).
carbons and lower diamondoids from condensate samples of two
offshore areas located in the southern Gulf of Guayaquil in order to 2.2. Stratigraphic framework
generate a comprehensive geochemical investigation of the sources and
origin of the natural gases from both areas. Also, the organic matter The stratigraphic column that is representative of the Jambelí sub-
throughout the stratigraphic section of the Amistad structure is exam­ basin comprises sedimentary rocks of Oligocene and Neogene ages
ined. Integration of geochemical data (i.e. C1–C4 compositions, molec­ (Fig. 2; after Benítez, 1995). The Delfín wells crosses the early Miocene
ular nitrogen concentration, isotopic signatures in gas samples) allowed to Pleistocene sequences including, base to top, the Heath, Zorritos,
elucidating reservoir continuity in the Amistad field. This all informa­ Cardalitos, Tumbes, Mal Pelo, La Cruz, and Tablazo formations (Fig. 2).
tion can be used to assist in guiding future drilling programs in the The basement (Early Cretaceous Piñón Formation) is more than 1000 m
Amistad field. thick and consists of pillow lavas, wehrlites, tholeiitic basalt-andesite
porphyry flows associated with hyaloclastic breccias, and cumulate
2. Geological setting gabbros intruded by massive dolerites forming a dense network of sills
and dykes (Luzieux et al., 2006).
2.1. Tectonic evolution and structural geology The Oligocene Zapotal Formation (maximum thickness of 3200 m)
consists of a sequence of imbricated conglomerates, pebbly sandstones
The early evolution of the southernmost Ecuadorian forearc is out­ with marine fossils, tuffs and coaly shales or thin coal beds deposited by
lined by several plate-scale processes that mainly include the accretion alluvial fans near the shoreline (Bristow and Hoffstetter, 1977; Benítez,
of oceanic terranes of the Caribbean Large Igneous Province and the 1995; Jaillard et al., 1995; among others). It has not been reached by any
modern tectonic escape of the North Andean Block, which includes such of the wells of the Gulf of Guayaquil-Tumbes Basin. Reworked tuffs and
oceanic terranes and others of the South American Plate (Alvarado et al., volcaniclastic deposits yielded two U–Pb zircons ages in the Progreso
2016; Aizprua et al., 2019). The Gulf of Guayaquil-Tumbes Basin is part Basin Province between 32 and 30 Ma (Witt et al., 2019). The fauna
of a series of Plio-Pleistocene forearc basins related to the northward association reported near Zapotal village is very similar to that reported
tectonic escape of the North Andean Sliver and shows a complex ar­ in the Mancora Formation in Perú (Olsson, 1931).
chitecture made up of three main depocenters: the Esperanza, Jambelí The Miocene Dos Bocas and Villingota formations have been only
and Tumbes sub-basins (Egbue and Kellogg, 2010). Sub-basins are observed in outcrops. The former comprises a sequence of dark-gray,
separated by structural highs most-likely resulting from transpression endurated, foraminifera- and diatom-rich shales intercalated with
linked to strain-partitioning along a major weakness zone inherited from sandstones and conglomerates deposited in inner to mid-shelf environ­
the accretionary processes (Witt and Bourgois, 2010; Aizprua et al., ments, with relatively restricted oceanic circulation (Benítez, 1995). The
2019), a structural style which seems to continue towards the Villingota Formation mainly consists of slightly tuffaceous, diatoma­
Peru-Ecuador border (latitude 3◦ 23′ 34′′ S; Fig. 1a). The subsidence ceous shales deposited in outer shelf to slope environments during
related to the drift of the North Andean Sliver started as early as coastal upwelling and seems to be a lateral equivalent facies of the Dos
Oligocene, in the early to middle Pleistocene subsidence took place in Bocas Formation, both correlate with the Peruvian Heath Formation
the Esperanza and Jambelí sub-basins (at least 6 km thick in the (Benítez, 1995). Upward in the stratigraphic record, the Subibaja For­
Esperanza basin) representing one of the most significant subsidence mation is a sequence of calcite-cemented sandstones, siltstones and
through the entire North Andean forearc area. A proto Amistad structure shales (Bristow and Hoffstetter, 1977) observed in the deepest parts of
separated the Miocene depocenters of the Jambelí and Esperanza Basins. the Gulf of Guayaquil-1 and Domito-1 wells (south of the tip of the
A major reaccommodation of basin depocenters occurred during the Amistad structure and western border of the basin, respectively). This
Pleistocene implied the migration of the main depocenter from the transitional marine unit represents the progradation of sand-rich lobes
southern part of the Jambelí sub-basin towards the Esperanza sub-basin; onto a muddy shelf during a regressive cycle (Benítez, 1995). The shales
a process accompanied by a change in the tectonic signature of the of the Villingota, Dos Bocas and Subibaja formations act as top seals for
Amistad structure (Deniaud et al., 1999; Witt et al., 2006). interbedded sandy reservoirs (Higley, 2004; Deckelman et al., 2008).
The Posorja and Jambelí detachment systems are also important Provenance models based mostly on U–Pb radiometric dating (Witt
tectonic features of the Progreso Basin Province, both representing half- et al., 2017, 2019) suggest that the Miocene sedimentary series in the
grabens with oppositely dipping detachments. The NE-SW trending Amistad structure (Amistad-10 well) show affinities that may allow to
Puna-Amistad transfer fault system bounds the Esperanza and Jambelí rely them to the Tumbes sub-basin, instead of the Esperanza sub-basin.
sub-basins, denoting that the evolution of they both is tightly controlled The affinity is mostly based on the ratios between zircons derived for the
by the two above-mentioned detachments at depth (Witt and Bourgois, modern arc (very significant in SW Ecuador) and those derived from the
2010). N–S trending tensional stresses characterize the Pleistocene times continental basement (very significant in NW Peru and in the Amistad
throughout. The westernmost N–S trending Domito fault system marks structure).
the limit of the Esperanza sub-basin (Calahorrano et al., 2008). From The Progreso Formation has been reached in all wells of the Gulf of
late Pleistocene to recent times, tectonic activity is restricted along the Guayaquil-Tumbes Basin with the exception of the Esperanza-1
normal faults bounding the Esperanza sub-basin (Witt et al., 2006). exploratory well that reached only Pliocene sediments. The Progreso
The Amistad structure, of N–S orientation, is most-likely related to Formation (maximum thickness of 1400 m) consists of shales inter­
strain partitioning along the major weakness zone produced during bedded with silty sandstones. The base of the Progreso Formation was
oceanic accretions (i.e. transform boundary). It consists of a large anti­ deposited in a shallow marine environment, while the top of which
cline of flower structure, extensively faulted and compartmentalized (known as the Cerro Mala Member) represents fluvio-deltaic sedimen­
(Kraemer et al., 2001). The Amistad gas field is situated on the down tation (Benítez, 1995). To the south, in the Amistad field, the Progreso
thrown east side of the main fault zone (see Fig. 1a; Aizprua, 2021). Formation contains relatively thick shales, which may behave as
Lateral trapping mechanisms are the NE-SW trending fault zone in the important pressure seals (Higley, 2004), and gas-producing sandstones
NW part of the structure, and simple dip closure in all other directions. It with porosity and permeability of up to 30% and 20 mD (Deckelman

3
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Fig. 2. Stratigraphic columns in the southern Gulf of Guayaquil. Note: source rocks (black circle) and reservoir rocks (black square). Hatched areas show non-
depositional periods and dotted lines separate members of the same formation.

et al., 2008). A dated deposit in the Progreso Basin Province yielded an this unit mainly comprises soft, arkosic and indurated sandstones
age of ~10 Ma (Witt et al., 2019). The Puna Formation (maximum interbedded with massive siltstones and volcanic tuffs (Placer Member)
thickness of 1000 m) represents shallow shelf sedimentation and, to­ and, on top, gray and green siltstones with lenses of sandstones (Lechuza
wards the bottom, is a sequence of transitional sandstones intercalated Member; Bristow and Hoffstetter, 1977). These latter two members were
with conglomerates and massive silty rocks (Benítez, 1995). By contrast, deposited in deltaic and estuarine environments with significant input of

4
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

siliciclastic sediments by meandering rivers. Finally, the Pleistocene increased from 35 ◦ C to 320 ◦ C at a rate of 2 ◦ C/min; and maintained at
Tablazo Formation mainly consists of shallow marine calcareous sands 320 ◦ C for 30 min. Additionally, gas chromatographic-mass spectro­
(Benítez, 1995). metric (GC-MS) and compound-specific isotope (CSIA) analyses of lower
diamondoids were done on distillates and saturate fractions of rock
3. Samples and methods extracts. For these two analyses, which were performed on a TRACE™
gas chromatograph coupled to a Thermo Scientific ISQ EC mass spec­
Eleven produced gases were sampled from the Amistad field, along trometer and a Delta V Advantage mass spectrometer with GC-isolink
with one gaseous sample from a well (B-17X) in the nearby Delfín field, interphase, respectively (Esquinas et al., 2018), the saturate frac­
using sealed glass bottles by means of metal separators connected to two tions/distillates were priorly treated with ZSM-5 zeolite to remove
sealed cylinders in parallel in the well-head. Geochemical gas data from n-alkanes (He et al., 2012). Absolute concentrations were measured by
the currently non-producing Amistad-9 (A-9) well has also been used. adding two deuterated-labeled compounds (1-methyladamantane-d3
Regarding low-boiling condensates (also called “distillates”), together and 1-methyldiamantane-d3) as internal standards to hydrocarbon liq­
with one sample (B-17X) from the Zorritos unit, five samples (labeled A- uids before performing analysis by GC-MS (Dahl et al., 1999).
3, A-6, A-7, A-8A, and A-12) were taken from the Progreso reservoir.
Sixty-two core samples from several of the aforementioned wells were 4. Results and discussion
also studied. Among them, fifty-one were analysed for vitrinite reflec­
tance maturity, of which sixteen selected samples from the Neogene 4.1. Potential hydrocarbon source rocks
sequence were analysed for maceral characterization (visual kerogen
analysis). Lastly, two representative Dos Bocas rock samples from the 4.1.1. Total organic carbon and Rock-Eval data
Amistad Sur borehole (total depth of 4506 m) were collected for fluid- Table 1 shows total organic carbon (TOC) values for selected rock
source correlation purposes. Fig. 1b shows the sampled wells in the samples throughout the stratigraphic section penetrated by the A-8A,
study area. − 11, − 12, and Amistad-Sur wells. TOC is quite low in most analysed
Optical microscopy analyses were carried out on grain pellets ground samples (0.11–0.79%; averaging 0.42%), except for the rock samples
to 1 mm following ISO 7404–2:2009 for sample preparation, ISO from the Amistad Sur well sampling the Subibaja and Dos Bocas for­
7404–5:2009 for vitrinite reflectance measurements, and ISO mations. These rock samples have TOC values between 0.98 and 2.16%
7404–3:2009 for maceral analysis; the latter being performed at a (1.45% on average), with good S2 pyrolysis peaks. As for the hydrogen
maceral level for liptinite and a maceral group level for vitrinite and indices (HI; Table 1), the Progreso and younger formations exhibit
inertinite using white light complemented with fluorescence observa­ values lower than 200 mg HC/g TOC and S2/S3 below 3, indicatives of
tions. Rock-Eval/TOC analyses were performed using a Rock Eval II Plus Type III-IV precursor organic matter with little natural gas and no oil-
TOC module apparatus, Delsi-Nermag instruments, according to generating capability (Peters, 1986). Similarly, the samples from the
Espitalié et al. (1977). Subibaja Formation (HI and S2/S3 data below 100 mg HC/g TOC and 3)
Analyses of the components in gaseous samples were performed, are only capable of generating limited amounts of natural gas. Only the
following the procedures provided by Márquez et al. (2013), through an two deepest Dos Bocas samples have Rock Eval data which can rated as
Agilent 6890 series gas chromatograph (GC) with electronic pneumatic good in source quality (HI values of 244 and 256 mg HC/g TOC, with
control. Relative standard deviations were below 0.8%. Carbon and S2/S3 exceeding 3), denoting a Type II-III kerogen and gaseous/liquid
hydrogen isotope analyses of individual compounds in the gaseous hydrocarbon generation attributes, because this type of kerogen gener­
samples were measured by gas chromatography-isotope ratio mass ates mixtures of gas and secondarily oil, in contrast to Type I-II and Type
spectrometry through a GC/C III interface connected through a com­ III-IV kerogens as they tend to be oil-prone and gas-prone, respectively
bustion reactor to a Finningan MAT Delta plus XL mass spectrometer (Tissot and Welte, 1984). These results are corroborated by the HI versus
(conditions as described in Márquez et al., 2013). The 13C/12C, D/H and OI diagram (Fig. 3a) and the plot of HI versus Tmax (Fig. 3b). This latter
15
N/14N ratios are reported in “δ” notation. δ13C, δD and δ15N values eludes any variability related to the oxygen index (Cornford et al.,
refer to V-PDB, SMOW and N-SVEC standards for carbon, hydrogen and 1998).
nitrogen stable isotope measurements. Reproducibility of the measure­ Tmax data (values below 430 ◦ C; Table 1) indicate that the entire
ments of δ13C, δ15N and δD were better than ±0.3‰, ±0.5‰ and ±5‰, section analysed is thermally immature for petroleum generation even if
respectively. good quality source rocks are present. That is, the Dos Bocas and Sub­
Core samples were softly disaggregated, air-dried for 24 h, and ibaja formations have not reached here sufficient levels of maturity to
manually grounded. Previous artificial maturation, in which potential generate hydrocarbons. All samples from the Amistad field (with the
source-rock samples are heated for 6 days at 330 ◦ C (Lijmbach et al., exception of those from the Amistad Sur well), apart from being ther­
1983), subsamples were extracted with a mixture of dichloromethane: mally immature, contain only small quantities of terrigenous organic
methanol (3:1, v/v) in a Soxtherm system (Gerhardt). The extracts were matter that cannot generate sufficient petroleum to form producible
concentrated by rotary evaporation at 40 ◦ C, then filtered (0.45 μm), and deposits. Therefore, the thermal wet gas and low-boiling condensate
finally dissolved in dichloromethane. Later, an aliquot of each rock must have migrated from deeper mature source rocks. Within this
extract was fractioned into saturated and aromatic hydrocarbons, and stratigraphic section, only a very slight maturity increase with depth is
polar compounds. Asphaltene fraction was precipitated with n-heptane observed. This is not usual (Dow, 1977) and should be explained by a
in a 1:40 v/v ratio according to the method outlined by Speight (2007). very short thermal exposure time due to uplift, by a low
Then, saturates were separated from aromatic and resin fractions by paleo-geothermal gradient, or the combination of both. If the section
liquid chromatography. The columns were filled with silica gel and was rapidly buried and then promptly uplifted, there may not have
alumina. The saturated hydrocarbons were eluted with hexane, the ar­ enough heating time to mature the organic material in the rocks: this
omatics with a mixture of dichloromethane:hexane (4:1 v/v) and, could be what happened on the Amistad structure, which has repre­
finally, the resins were eluted with methanol. sented a structural high at least since the middle Miocene to the present,
Whole oil gas chromatographic analyses of light hydrocarbons in coeval with the onset of the Jambelí sub-basin depocenter, and where
condensate samples were carried out using a J&W Agilent PONA GC recent activity has been observed in high-resolution seismic character­
column (50 m × 0.2 mm i. d.; film thickness 0.25 μm; helium was the ization (Reynaud et al., 2018). Depth-projection of the maturity profile
carrier gas) in a Delta Chrom Series 9980 instrument with flame ioni­ (%Rc) suggests that if source horizons are present, petroleum generation
zation detection. The following operating conditions were used for the in the study area might have occurred at 6000 m depth or deeper (in
gas chromatograph: temperature maintained at 35 ◦ C for 15 min; concordance with Evangelista, 2019).

5
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Table 1
Measured and calculated vitrinite reflectances (Ro and Rc), TOC values, depths, and Rock-Eval data for the sampled rocks.
Sample Depth Rock unit Lithology TOC Tmax Rr Rc S1 S2 S3 HI OI

Amistad-8D/01 2436–2450 Cerro Mala Sandy claystone 0.31 417 – 0.35% 0.02 0.30 0.45 97 145
Amistad-8A/02 2520–2534 Progreso s.s. Silty claystone 0.39 418 0.30% 0.36% 0.02 0.18 0.28 46 72
Amistad-8A/03 2646–2660 Progreso s.s. Sandy claystone 0.39 419 0.30% 0.38% 0.01 0.14 0.17 36 44
Amistad-8A/04 2688–2702 Progreso s.s. Silty claystone 0.31 418 0.29% 0.37% 0.01 0.12 0.02 39 6
Amistad-8A/05 2772–2786 Progreso s.s. Silty shale 0.52 419 0.30% 0.38% 0.03 0.28 0.32 54 62
Amistad-8A/06 2856–2870 Progreso s.s. Silty shale 0.52 420 0.31% 0.39% 0.04 0.28 0.34 54 65
Amistad-8A/07 2940–2954 Progreso s.s. Silty claystone 0.58 419 0.32% 0.39% 0.05 0.32 0.16 55 28
Amistad-8A/08 3024–3038 Progreso s.s. Silty claystone 0.79 420 0.31% 0.40% 0.03 0.41 0.23 52 29
Amistad-8A/09 3108–3122 Progreso s.s. Sandy claystone 0.58 419 0.31% 0.39% 0.02 0.25 0.16 43 28
Amistad-8A/10 3192–3206 Progreso s.s. Sandy shale 0.37 419 0.31% 0.39% 0.01 0.16 0.09 43 24
Amistad-8A/11 3262–3276 Progreso s.s. Sandy shale 0.18 420 0.32% 0.40% 0.00 0.04 0.09 22 50
Amistad-8A/12 3346–3360 Progreso s.s. Sandy shale 0.41 420 0.31% 0.40% 0.02 0.17 0.17 41 41
Amistad-8A/13 3388–3402 Progreso s.s. Sandy shale 0.51 421 0.30% 0.41% 0.02 0.22 0.14 43 27
Amistad-8A/14 3430–3444 Progreso s.s. Sandy claystone 0.37 421 0.31% 0.41% 0.15 0.33 0.60 89 162

Amistad-11/01 876–960 Lechuza Shaly sandstone 0.22 – 0.37% – 0.02 0.08 0.14 36 64
Amistad-11/02 1212–1296 Lechuza Shaly sandstone 0.57 421 0.39% 0.41% 0.03 0.32 0.38 56 67
Amistad-11/03 1380–1464 Placer Shaly claystone 0.28 421 0.39% 0.41% 0.01 0.11 0.06 39 21
Amistad-11/04 1548–1632 Placer Clayey shale 0.22 421 0.39% 0.42% 0.03 0.31 0.14 141 64
Amistad-11/05 1800–1884 Puná s.s. Claystone 0.73 421 0.38% 0.41% 0.11 1.42 0.62 195 85
Amistad-11/06 2069–2153 Puná s.s. Claystone 0.22 422 0.39% 0.42% 0.01 0.17 0.11 77 50
Amistad-11/07 2237–2321 Cerro Mala Silty claystone 0.41 420 0.38% 0.40% 0.02 0.28 0.10 68 24
Amistad-11/08 2405–2489 Progreso s.s. Shale 0.67 421 0.38% 0.41% 0.03 0.37 0.35 55 52
Amistad-11/09 2489–2573 Progreso s.s. Shale 0.47 421 – 0.41% 0.02 0.37 0.21 79 45
Amistad-11/10 2574–2618 Progreso s.s. Shale 0.69 422 0.38% 0.42% 0.01 0.55 0.23 80 33
Amistad-11/11 2632–2646 Progreso s.s. Shale 0.60 422 0.38% 0.42% 0.01 0.49 0.14 82 23
Amistad-11/12 2716–2730 Progreso s.s. Sandy shale 0.25 421 0.38% 0.41% 0.01 0.10 0.13 40 52
Amistad-11/13 2814–2828 Progreso s.s. Sandy shale 0.25 420 0.37% 0.40% 0.01 0.13 0.13 52 52
Amistad-11/14 2884–2898 Progreso s.s. Shale 0.49 422 0.40% 0.42% 0.01 0.26 0.22 53 45
Amistad-11/15 2954–2968 Progreso s.s. Sandy shale 0.23 422 0.39% 0.42% 0.01 0.23 0.05 100 22

Amistad-12/01 448–532 Tablazo Sandy siltstone 0.33 414 0.28% 0.29% 0.05 0.13 0.37 39 112
Amistad-12/02 700–784 Lechuza Shaly sandstone 0.26 415 0.29% 0.30% 0.05 0.09 0.14 35 54
Amistad-12/03 952–1036 Lechuza Shaly sandstone 0.32 415 – 0.31% 0.01 0.06 0.15 19 47
Amistad-12/04 1204–1288 Placer Shaly claystone 0.27 415 0.30% 0.32% 0.03 0.15 0.72 56 267
Amistad-12/05 1456–1540 Placer Clayey shale 0.57 416 0.31% 0.32% 0.02 0.30 0.19 53 33
Amistad-12/06 1624–1708 Placer Clayey shale 0.46 416 0.31% 0.32% 0.04 0.37 0.18 80 39
Amistad-12/07 1792–1876 Puná s.s. Claystone 0.46 416 0.32% 0.32% 0.03 0.24 0.20 52 43
Amistad-12/08 2044–2128 Puná s.s. Claystone 0.37 416 0.32% 0.32% 0.03 0.28 0.14 76 38
Amistad-12/09 2296–2380 Cerro Mala Silty claystone 0.34 417 0.32% 0.33% 0.03 0.23 0.10 68 29
Amistad-12/10 2464–2478 Progreso s.s. Sandy claystone 0.47 417 0.32% 0.33% 0.02 0.22 0.13 47 28
Amistad-12/11 2534–2548 Progreso s.s. Silty claystone 0.24 416 0.32% 0.32% 0.02 0.11 0.16 46 67
Amistad-12/12 2590–2604 Progreso s.s. Sandy claystone 0.54 417 – 0.33% 0.02 0.27 0.18 50 33
Amistad-12/13 2646–2660 Progreso s.s. Silty claystone 0.71 418 0.31% 0.35% 0.02 0.30 0.18 42 25
Amistad-12/14 2702–2716 Progreso s.s. Silty shale 0.78 418 0.32% 0.35% 0.02 0.48 0.32 62 41
Amistad-12/15 2744–2758 Progreso s.s. Silty shale 0.56 417 0.32% 0.33% 0.01 0.32 0.09 57 16
Amistad-12/16 2800–2814 Progreso s.s. Shale 0.12 – 0.33% – 0.01 0.03 0.07 25 58
Amistad-12/17 2856–2870 Progreso s.s. Shale 0.30 417 0.32% 0.33% 0.01 0.12 0.10 40 33
Amistad-12/18 2912–2926 Progreso s.s. Sandy shale 0.70 417 0.32% 0.34% 0.02 0.38 0.19 54 27
Amistad-12/19 2982–2996 Progreso s.s. Sandy shale 0.50 418 0.33% 0.35% 0.02 0.24 0.26 48 52
Amistad-12/20 3108–3122 Progreso s.s. Sandy shale 0.11 – 0.32% – 0.01 0.03 0.10 27 91
Amistad-12/21 3122–3136 Progreso s.s. Sandy shale 0.14 417 0.32% 0.34% 0.01 0.02 0.18 14 127
Amistad-12/22 3136–3150 Progreso s.s. Sandy shale 0.29 417 0.33% 0.34% 0.83 0.27 0.16 92 67

Amistad-Sur/01 4144–4150 Subibaja Claystone 1.77 425 0.40% 0.49% 0.07 0.95 3.84 54 217
Amistad-Sur/02 4161–4167 Subibaja Claystone 1.69 425 – 0.49% 0.05 0.88 2.79 52 165
Amistad-Sur/03 4183–4189 Subibaja Claystone 1.18 425 – 0.49% 0.25 0.65 2.54 55 215
Amistad-Sur/04 4239–4245 Subibaja Claystone 1.12 426 0.41% 0.50% 0.16 1.12 1.26 95 112
Amistad-Sur/05 4250–4256 Subibaja Claystone 2.16 425 – 0.49% 0.22 1.55 4.13 72 191
Amistad-Sur/06 4317–4323 Dos Bocas Sandy Claystone 1.17 427 – 0.51% 0.07 0.61 3.88 52 332
Amistad-Sur/07 4342–4348 Dos Bocas Sandy Claystone 0.98 427 0.42% 0.52% 0.04 0.32 1.63 33 166
Amistad-Sur/08 4384–4390 Dos Bocas Sandy Claystone 0.98 428 – 0.53% 0.08 0.49 2.95 50 301
Amistad-Sur/09 4429–4435 Dos Bocas Sandy Claystone 1.18 428 – 0.53% 0.07 1.11 2.44 94 207
Amistad-Sur/10 4452–4458 Dos Bocas Sandy Claystone 1.70 428 – 0.53% 0.08 4.34 1.20 256 71
Amistad-Sur/11 4474–4480 Dos Bocas Sandy Claystone 2.03 429 0.43% 0.54% 0.08 4.95 1.39 244 69

Notes: TOC data in wt.%; S1and S2 in mg HC/g rock, while S3 in mg CO2/g rock; HI–
– S2/TOC; OI–
–S3/TOC; HI in mg HC/g TOC; depth in m; Tmax in ◦ C; IO in mg CO2/
g TOC. %Rc = 0.018⋅Tmax-7.16 (Jarvie et al., 2001).

4.1.2. Kerogen composition and maturity measurements humic debris. Herbaceous plant debris including traces of possible oil-
All the selected cutting samples for petrographic studies from the A- prone sporinite (Mukhopadhyay et al., 1991) and resinite are also
8A, A-11, A-12, and Amistad Sur wells consist mainly of vitrinite/ observed in the deepest Dos Bocas rock samples. Based on petroleum
huminite (58–80%) with minor quantities of inertinite (2–10%) and potential, exinite and vitrinite macerals are light oil/condensate- and
exinites not exceeding 10% (Table 2); although some amorphous mostly gas-prone, respectively, while inertinite macerals only have a
organic material is also observed, it is difficult to identify as algal or little gas potential (Stach et al., 1982). With exception of the probable

6
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Fig. 3. a) and b), respectively, HI versus OI diagram and plot of HI against Tmax for rock samples.

Table 2
Petrographic composition (vol%) for the some rock samples.
Sample Depth Rock unit Vitrinite/huminite Dark vitrinite Amorphous Exinite Inertinite

Amistad-8A/03 2646–2660 Progreso s.s. 80% 65% 10% 4% 6%


Amistad-8A/08 3024–3038 Progreso s.s. 75% 60% 15% 5% 5%
Amistad-8A/13 3388–3402 Progreso s.s. 80% 65% 10% 5% 5%

Amistad-11/02 1212–1296 Lechuza 70% 60% 20% 4% 6%


Amistad-11/05 1800–1884 Puná s.s. 58% 40% 35% 2% 5%
Amistad-11/08 2405–2489 Progreso s.s. 50% 40% 40% 5% 5%
Amistad-11/11 2632–2646 Progreso s.s. 65% 45% 20% 6% 9%
Amistad-11/14 2884–2898 Progreso s.s. 70% 65% 15% 9% 6%

Amistad-12/05 1456–1540 Placer 65% 10% 25% 5% 5%


Amistad-12/10 2296–2380 Cerro Mala 60% 10% 30% 8% 2%
Amistad-12/14 2702–2716 Progreso s.s. 65% 35% 25% 7% 3%
Amistad-12/18 2912–2926 Progreso s.s. 70% 40% 20% 5% 5%

Amistad-Sur/01 4144–4150 Subibaja 70% 55% 15% 9% 6%


Amistad-Sur/04 4239–4245 Subibaja 70% 55% 10% 10% 10%
Amistad-Sur/07 4342–4348 Dos Bocas 55% 40% 30% 10% 5%
Amistad-Sur/11 4474–4480 Dos Bocas 60% 45% 25% 10% 5%

Notes: Petrographic data expressed in a mineral-matter-free basis. Terminology according to ICCP in 2001.

Dos Bocas source rock organofacies corresponding to a prodeltaic relatively common at low maturity levels, probably resulting in sup­
depositional environment, under oxygen-deficient conditions pressed reflectance (Kalkreuth, 1982). Even though coexistence with
(Montenegro and Benítez, 1991), these kerogen assemblages indicate liptinite, whose bituminous substances can permeate through the vitri­
deposition under relatively oxic conditions, and could be characteristic nite, could often explain the presence of dark (perhydrous) vitrinite
of transitional-fluvial settings (Stach et al., 1982). The maceral analyses (Iglesias et al., 2000), in this case this option is unlikely given the low
also showed that there are two kinds of vitrinite, a dark and a light exinite content of most samples, excepting the deepest cuttings.
variety, in the cutting samples. Most samples comprise predominately Palaeo-desiccation of woody components at the time of deposition under
lipid-rich vitrinite in dark gray-black color: the vitrinitic fraction ap­ specific conditions might be another explanation for dark vitrinite
pears to be predominantly normal at the shallower samples from Cerro occurrence in cutting samples (Taylor et al., 1989), but these processes
Mala to younger strata, and then changes to a mixture of light and are generally related to environments with contrasted seasons and a
mostly dark vitrinite at deeper samples from the A-12 well (Table 2). The high inertinite content (Diessel, 2010). Alternately, associated to tissues
existence of a dark and light vitrinite in terrigenous organic facies is and resinous fillings, the vitrinite can become darker (Suarez-Ruiz et al.,

7
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

1994). A last scenario requires that a distinct density of plants gener­ amounts of higher hydrocarbons; whereas thermogenic gas is suggested
ating dark vitrinites or different chemistry in the vitrinites may explain by methane/ethane ratios typically below 100 (Jackson et al., 2014).
the darker appearance (Petersen and Rosenberg, 1998). Gas dryness usually increases with increasing gas maturation
As shown in Table 1, vitrinite reflectance values (%Rr) near or below (Prinzhofer et al., 2000), and it is considered that values exceeding 0.95
0.4% confirm that maturity is low throughout the Neogene sequence generally indicate microbial processes (Dai, 1992). Thus, bacteriogenic
analysed. Noteworthy is also an appreciable difference between %Rc hydrocarbons have a dominant contribution to gaseous samples A-4,
and %Rr values (Table 1). This feature might be partially caused by the A-6, A-7, A-11, and A-12 with C1/C2 gas ratios clearly above 100; while
problem of reflectance suppression, which in turn can be attributed to the other Amistad gases, with C1/C2 ratios near o below 100, have an
various reasons such as overpressure in sedimentary basins and/or the appreciable thermogenic gas input. The B-17X gas contains small, but
fact that Tmax could evolve differently for Types II and III organic significant contents of ethane and other paraffins, which might reflect
matter at low maturities (Espitalié et al., 1985; Carr, 2000); in this associated oil accumulations (Dai, 1992).
context, the most reliable explanation would be given by the occurrence The molecular C1/C2 and C2/C3 ratios can be used to distinguish
of low-reflecting vitrinite. between kerogen- and oil-cracking gases by using in the ln (C1/C2)
against ln (C2/C3) graph, since the first/second ratio increases notably/
remains constant during primary cracking and vice versa during sec­
4.2. Geochemical characteristics of petroleum samples
ondary cracking (Prinzhofer and Huc, 1995). None of the Amistad gases
plot along the evolution lines of Type-II (sapropelic) or Type-III (humic)
4.2.1. Natural gas components
kerogen. Instead, the Delfín gas plotts near the evolution line of Type-II
Molecular concentrations of methane, ethane, propane, n-butane,
kerogen (Fig. 4a), with lower C1/C2 values, which are typical of primary
iso-butane, n-pentane, and iso-pentane for the natural gases throughout
cracking processes.
the studied area are shown in Table 3. Methane is the main component
Based on the ratios of iso-butane to n-butane (i-C4/n-C4) and ethane
of the sampled gases, ranging from 96.36 to 99.54%. Several groups of
to propane (C2/C3) it is possible to determine the extent of biodegra­
samples can be defined based on these results. First, the samples with
dation (as opposed to migration and other secondary processes) in the
highest molar concentrations of methane (around 99.5%) are A-4, A-6,
gas samples (e.g., Márquez et al., 2013). Relatively high values
A-7, A-11, and A-12, which also show the lowest molar concentrations of
(approaching or exceeding 3) of both ratios are observed in all the gases
heavier alkanes (Table 3). Second, samples A-1, A-5, A-8A, and A-9 have
produced from the Progreso reservoir, while the A-8D and B-17X gases
similar C1–C3 alkanes patters, with average contents of methane and
show the lowest C2/C3 and i-C4/n-C4 values (Table 4 and Fig. 4b). Since
ethane approaching 98% and 1%. Third, A-2 and A-3 samples, nearly
propane is preferentially attacked by microbes (James and Burns, 1984)
identical, have slightly lower and higher values of methane (96.94% on
and n-butane is selectively biodegraded relative to its isomeric
average) and ethane (averaging 1.611%) contents, respectively, than the
iso-alkane (Katz et al., 2002; Larter and di Primio, 2005), it is likely that
two previous groups of Amistad gaseous samples. Four, the lowest
all the analysed gases (except A-8D and B-17X) can be affected by
abundance of methane (96.39%) and highest concentration of heavier
slightly biodegradation rather than migration.
alkanes (Table 3) are noticed in the B-17X gaseous sample, excluding the
Non-hydrocarbon components are also present in most analysed
A-8D sampled gas. This latter natural gas from the Cerro Mala interval
natural gases. In the Amistad samples, CO2 and N2 range from 0.010 to
exhibited an abnormal low molar concentration of methane (68.34%),
0.033% and 0.14%–0.32%, respectively (Table 3). Delfín B-17X natural
which would suggest air contamination during sampling and/or the fact
gas (sampling the Zorritos interval) has CO2 and N2 contents of
that vertical composition is not equilibrated along the study strati­
approximately 0.2%. A-8D gaseous sample from the Cerro Mala reser­
graphic section because of barriers to molecular diffusion, as discussed
voir has 0.27% of CO2 and 30.38% of N2, values notably higher than
below.
those in the other reservoirs (see Table 3). In particular, the excep­
The gas dryness (defined as C1/ΣC1-5, which evaluates the proportion
tionally high molar nitrogen content in the A-8D gas, up to almost three
of hydrocarbon compounds in the samples) lies in the 0.963–0.998
hundred times higher than those from the other wells, seems to reflect
range for the studied samples (Table 4), averaging 0.988, with A-8D gas
air contamination during sampling, which would be supported by N2/Ar
having a value of 0.993. The molar concentration of methane in gaseous
of 90 and δ15N of 0‰ (Zhu et al., 2000); however, molecular oxygen
samples (excluding A-8D) increases with gas dryness (Pearson’s coeffi­
would have been almost consumed by reaction with the steel cylinder or
cient of about 0.99). Methane abundance in natural gas can be governed
by other feasible means (Jenden et al., 1988).
by numerous and complex factors including genetic type, gas maturity,
and geochromatography during secondary migration, among others
4.2.2. Isotopic compositions of gas hydrocarbons
(Tissot and Welte, 1984; Behar et al., 1992; Chen et al., 2000). For
The carbon isotopic compositions of the C1–C4 n-alkanes and
instance, biogenic gas is mostly methane, with no or very scarce

Table 3
Average aliphatic hydrocarbons (C1–C6+), CO2, and N2 for the gases from Amistad field and Delfín B-17X well.
Well Reservoir Latitude Longitude CH4 C2H6 C3H8 n-C4 i-C4 n-C5 i-C5 C6+ CO2 N2

Amistad-1 Progreso 3◦ 15′ 41′′ S 80◦ 27′ 36′′ W 98.50 0.708 0.218 0.028 0.089 0.006 0.030 0.046 0.030 0.32
Amistad-2 Progreso 3◦ 13′ 47′′ S 80◦ 27′ 00′′ W 97.97 0.835 0.270 0.023 0.065 0.006 0.020 0.056 0.011 0.30
Amistad-3 Progreso 3◦ 14′ 33′′ S 80◦ 28′ 06′′ W 97.83 0.842 0.271 0.022 0.063 0.006 0.021 0.053 0.010 0.31
Amistad-4 Progreso 3◦ 17′ 36′′ S 80◦ 28′ 21′′ W 99.47 0.205 0.058 0.006 0.022 0.001 0.006 0.030 0.016 0.16
Amistad-5 Progreso 3◦ 16′ 09′′ S 80◦ 28′ 20′′ W 98.53 0.700 0.212 0.028 0.088 0.006 0.027 0.044 0.033 0.31
Amistad-6 Progreso 3◦ 16′ 42′′ S 80◦ 28′ 19′′ W 99.54 0.193 0.056 0.006 0.020 0.001 0.006 0.021 0.015 0.14
Amistad-7 Progreso 3◦ 16′ 53′′ S 80◦ 27′ 55′′ W 99.51 0.202 0.059 0.006 0.021 0.001 0.006 0.028 0.012 0.15
Amistad-8D Cerro Mala 3◦ 15′ 26′′ S 80◦ 27′ 25′′ W 68.34 0.235 0.106 0.035 0.044 0.026 0.050 0.102 0.270 30.38
Amistad-8A Progreso 3◦ 15′ 36′′ S 80◦ 27′ 32′′ W 98.60 0.696 0.215 0.028 0.089 0.006 0.025 0.045 0.032 0.31
Amistad-9 Progreso 3◦ 16′ 16′′ S 80◦ 27′ 49′′ W 98.62 0.698 0.216 0.027 0.087 0.007 0.025 0.046 0.032 0.31
Amistad-11 Progreso 3◦ 17′ 00′′ S 80◦ 28′ 17′′ W 99.49 0.200 0.060 0.007 0.022 0.001 0.005 0.029 0.010 0.14
Amistad-12 Progreso 3◦ 17′ 48′′ S 80◦ 28′ 34′′ W 99.48 0.199 0.057 0.006 0.021 0.001 0.006 0.031 0.015 0.16
Delfín B-17X Zorritos 3◦ 36′ 50′′ S 80◦ 48′ 30′′ W 96.06 1.955 1.116 0.217 0.142 0.091 0.080 0.105 0.110 0.12
Amistad Sur – 3◦ 18′ 40′′ S 80◦ 28′ 19′′ W – – – – – – – – – –

Note: all data in molar percentages.

8
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Table 4
δD (‰) in methane, δ13C values (‰) in carbon dioxide and C1–C4 alkanes, as well as light hydrocarbon parameters for the gaseous samples.
Well Net pay Depth C1/ΣC1-5 i-C4/n-C4 C1/C2-3 C2/C3 C1/C2 C2/C1 δ13C1 δ13C2 δ13C3 δ13nC4 δ13iC4 δD1 δ13CCO2

A-1 3145–3215 4777 0.990 3.18 106.37 3.25 139.12 0.007 − 56.30 − 32.42 − 27.32 − 27.41 − 28.35 − 203.9 –
A-2 2585–2850 3164 0.984 2.81 88.66 3.09 117.33 0.008 − 56.26 − 31.56 − 27.48 − 27.56 − 28.44 − 205.6 − 16.45
A-3 2296–2324 2475 0.984 2.85 87.89 3.10 116.20 0.008 − 56.24 − 31.33 − 27.41 − 27.45 − 28.29 − 205.8 –
A-4 2492–2525 2944 0.997 3.66 378.21 3.53 485.22 0.002 − 61.68 − 43.34 − 30.55 − 27.29 − 28.14 − 207.0 − 18.65
A-5 2520–2530 2994 0.989 3.14 108.03 3.30 141.44 0.007 − 56.18 − 32.91 − 27.54 − 27.55 − 28.38 − 204.7 − 18.08
A-6 2492–2567 2919 0.998 3.33 399.75 3.44 515.75 0.002 − 61.39 − 43.27 − 30.45 − 27.46 − 28.43 − 210.9 − 16.96
A-7 2836–2916 3498 0.997 3.50 381.26 3.42 492.62 0.002 − 61.62 − 43.78 − 30.71 − 27.29 − 28.12 − 209.3 − 19.22
A-8D 2178–2294 2952 0.993 1.26 200.41 2.22 290.81 0.004 − 59.47 − 36.62 − 27.59 − 27.51 − 28.38 − 208.1 –
A-8A 3327–3376 4024 0.990 3.18 108.23 3.24 141.41 0.007 − 56.10 − 32.67 − 27.43 − 27.50 − 28.41 − 206.1 − 17.34
A-9 2873 3330 0.990 3.22 107.89 3.23 141.61 0.007 – – – – – – –
A-11 2733–2799 3232 0.998 3.14 382.65 3.33 497.45 0.002 − 61.37 − 43.91 − 30.80 − 27.33 − 28.17 − 211.3 –
A-12 3001–3032 3428 0.997 3.50 388.59 3.49 499.89 0.002 − 61.76 − 43.45 − 30.12 − 27.38 − 28.30 − 210.1 − 17.87
B-17X 1756 2634 0.963 0.65 31.27 1.74 49.13 0.020 − 38.45 − 30.48 − 27.96 − 27.22 − 28.09 − 180.1 –

Note: Net pay data and depths in meters.

hydrogen signature of methane for the gaseous samples are shown in (Table 4), which is consistent with a kerogen cracking origin of this
Table 4. Biogenic gas can be identified by methane-based carbon iso­ gaseous sample. In Amistad, some samples also follow a positive trend
topic signatures (δ13C1) below − 55‰; while δ13C1 values of thermo­ whilst samples A-1, A-2, A-3, A-5, and A-8A show a partial reversal of the
genic gas usually range from − 20 to − 55‰ (Dai, 1992). As for δ13C series: δ13C1<δ13C2<δ13C3>δ13C4 (see Table 4). Although the
depositional environment and organic matter type, it is considered that higher carbon isotopic values of propane compared to n-butane might be
δD values for methane (δD1) up to around − 190‰ are typical of explained by mixing of gases (Tao et al., 2016), biodegradation seems to
humic-type natural gases (Schoell, 1980), along with δ13C2 and δ13C3 be the more probable explanation alongside other evidences, so that C3
exceeding − 27.5‰ and − 25.5‰; instead, δ13C2 and δ13C3 values below molecules composed of the light isotope 12C are preferentially metabo­
− 29‰ and − 27‰ are typical of sapropelic-type gases, while gaseous lized by microbes with respect to 13C-containing C3 molecules (Strąpoć
samples with δ13C2 and δ13C3 between aforementioned values are either et al., 2010) and, as a result, the residual propane becomes richer in 13C
mixtures of the two end-member types or generated from type II-III (Strąpoć et al., 2007).
kerogen (Dai et al., 2005). Also, δD1 values from − 190‰ to − 180‰ The Lorant et al. (1998) diagram, C2/C3 versus δ13C2-δ13C3, can be
correspond to paralic brackish environment range (Liu et al., 2008). The used to distinguish between primary cracking gases (those generated
δ13C1 and δD1 values in Amistad gaseous samples range from − 57.26‰ directly from kerogen) and secondary cracking gases (those derived
to − 60.7‰ and − 203.5‰ to − 211.3‰ (Fig. 5a), thus indicating mixed from oil/wet gas) and to characterize oil-cracking gases of different
biogenic-thermogenic origin and mostly humic-type precursor material maturity stages (e.g., Tissot and Welte, 1984; Prinzhofer et al., 2010). In
for these gaseous samples. By contrast, δ13C1, δ13C2, δ13C3 and δD1 the particular case of the thermogenic gas from Delfín B-17X well, re­
values for the Delfín B-17X gas (Table 4 and Fig. 5a) suggest a ther­ sults suggest primary cracking and thermal maturity within the oil
mogenic input probably generated from mostly marine source rocks window. Lastly, it is possible to determine the maturity level of
deposited in paralic brackish settings. sapropelic-type gases through the methane carbon isotopic ratio (Dai,
Fig. 5b shows a plot of the C1/C2+3 ratio versus methane carbon 1992): δ13C1 from − 40.2 to − 37.4‰, from − 45 to − 40.2‰, higher than
isotopic composition (also known as Bernard diagram), a useful tool to − 37.4‰, and lower than − 45‰ usually indicate highly maturity,
differentiate the genetic type of gas (Bernard et al., 1978; Galimov, normal maturity, overmaturity, and immaturity to low maturity stages,
2006). In this diagram, two groups of Amistad samples are recognised: a respectively. Following this, the B-17X thermogenic gas (δ13C1 =
first group (A-4, A-6, A-7, A-11, and A-12) is characterised by a domi­ − 38.45‰) would be very mature.
nantly bacteriogenic gas contribution, with minor thermogenic gas
input; the other Amistad gases (A-1, A-5, A-8A, and A-8D) plot in the 4.2.3. Light hydrocarbons in condensates
“mixing” zone (Fig. 5b), but with a less dominant bacteriogenic input. Light hydrocarbon compositions of distillates, which can be used to
The B-17X natural gas sample has δ13C1 < − 40‰ and C1/C2+3 about 30 indicate source and maturity of natural gases (Thompson, 1983; Mango,
(Table 4), which denotes a thermogenic origin. The modified 1990), are shown for the study samples in Table 5. The condensates from
δ13C1-δ13C2-δ13C3 plot diagram (Fig. 6) corroborates a dominantly the Amistad Field are very similar among them and have API gravities of
biogenic origin for Amistad gaseous samples and a ca. 60◦ . In the ternary diagram of methylcyclohexane (MCH), n-heptane,
thermogenic-sapropelic origin for B-17X gas (Dai et al., 2014). Delfín and dimethylcyclopentanes (ΣDMCP), which derive mainly from
B-17X gaseous sample show a near-linear negative correlation between land-plant, algal-bacterial, and aquatic sources, respectively (Hu et al.,
δ13C values for individual C1–C4 n-alkanes and reciprocal carbon num­ 2007, 2010; Thompson, 1979), all the analysed condensates plot in the
ber, as opposed to Amistad gases as their δ13C1 values are depleted, “sapropelic” zone (Fig. 7a), where hydrocarbon liquids usually show
thereby supporting the aforementioned outcomes (Dai, 1992). Among n-C7>30% and MCH<70% (Hu et al., 1990). The Amistad distillate
other factors accounting for this latter feature (e.g., biodegradation), a samples are also characterised by low n-C7/MCH ratios (paraffinicity)
substantial biogenic signature to the Amistad gases might have added and isoheptane values (Table 5), generally related to biodegradation
enough bacteriogenic ethane to shift the thermogenic C2 isotopic char­ (Thompson, 1987). The ternary diagram of the C5–C7 normal alkanes,
acter of these samples to lighter values. isoparaffins, and cyclanes is also used here for constraining the source of
The carbon isotopic series of C1–C4 n-alkanes of the primary biotic the study samples (Cao et al., 2012; Chen et al., 2014). B-17X sample
gases usually follows a positive trend (δ13C1<δ13C2<δ13C3<δ13C4; plots clearly in the “sapropelic” zonation. Instead, all the Amistad
Boreham and Edwards, 2008). However, partial reversals of the δ13C gas condensate samples plot close to the separation line between the “sap­
series can be caused by various processes such as mixing (e.g., Fuex, ropelic” and “humic” zones (Fig. 7b) and are characterised by abundant
1977; Cai et al., 2001; Hao et al., 2008; Yang et al., 2008), biodegra­ cyclic alkanes, which is a feature consistent with biodegradation
dation (Zumberge et al., 2009; Xia et al., 2013), overmaturity (Zum­ processes.
berge et al., 2012; Gao et al., 2015) and/or thermochemical sulfate Heptane and isoheptane values are typically used to denote gas
reduction (Hao et al., 2008). Delfín B-17X follows a positive trend maturity (e.g., Chen et al., 1987). Based on the heptane value, it is

9
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Fig. 4. a) Molecular ratios C2/C3 against C1/C2 (in natural logarithmic scales) for kerogen- and mixed sampled gases analysed; b) plot of C2/C3 versus C2/iC4 for
gaseous samples denoting biodegradation and other secondary processes.

possible to define the level of gas maturity and if biodegradation hydrocarbons reflect thermal maturation of light and heavy gas frac­
occurred: percentages exceeding 30%, 18–30%, 12–18%, and tions, respectively, disparity between both data does not preclude
approaching 0% suggest highly mature, normally mature, low mature, multi-stage gas accumulation (Chen et al., 2010).
and biodegraded hydrocarbons, respectively (Thompson, 1983). The
Amistad condensates have heptane values of approximately 2%
(Table 5), which might indicate biodegradation processes. As whole oil 4.3. Oil-oil and oil-source rock correlation
gas chromatograms of the Amistad condensates, these show a relative
high abundance of n-C5 to n-C7 hydrocarbons (Fig. 8a), non-typical of Previous information from some boreholes drilled in the offshore
biodegraded oils, this could indicate that at least one late oil charge Amistad field suggested that only the Dos Bocas Formation has gaseous/
would have occurred and mixed with the original altered oil. The Delfín liquid hydrocarbon source potential in the study area (Evangelista,
B-17X distillate (value of ~21%) would be normally mature and this 2019). Therefore, a geochemical evaluation was carried out to correlate
does not necessarily contradict previously discussed δ13C1 data, as the analysed condensates with the probable Dos Bocas source rocks. As
generally maturity derived from light hydrocarbons is slightly lower shown in Table 5, Amistad and Delfín B-17X distillate samples show
than that derived from methane carbon isotopes (Sun et al., 2016); also pristane-to-phytane ratios of about 2, also similar to the potentially
since maturity levels inferred from carbon isotopic data and light parent Dos Bocas extracts, suggesting suboxic/dysoxic bottom-water
conditions during deposition. Moreover, the ratios of pristane and

10
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Fig. 5. a) Whiticar diagram for the elucidation genesis mechanisms of sampled gases; b) Bernard diagram for the gaseous samples analysed, indicating genetic types
of the sampled gases.

phytane to their adjacent normal alkanes, n-C17 and n-C18 (Table 5), and the sum of 1-, 3- and 4-MD), suggesting a type II-III precursor
would indicate that the unaltered B-17X low-boiling condensate was organic material (Schulz et al., 2001) and a maturity level close to peak
derived from a Type II/III kerogen (Peters et al., 2005), suggesting a oil-generation (Chen et al., 1996). Moreover, these non-cracked liquid
possible correlation between this sample and the Dos Bocas rock extract samples exhibit similar absolute concentrations of 3- and 4-methyldia­
with similar Pr/n-C17 and Ph/n-C18 values (1.15 and 0.79). mantane (diamondoid baseline of about 4.7 ppm), which could indi­
Additionally, as the lower diamondoids are highly thermally stable cate a common source. As expected, the evaporative fractionated
and relatively recalcitrant, they can be successfully applied in oil-oil and Amistad hydrocarbon liquids show lower values (approximately 25) of
oil-source correlation in low-boiling condensates and other liquids the ratio between 1- plus 2-methyladamantane versus 3- plus 4-methyl­
where the classical biomarkers are very scarce or absent (Schulz et al., diamantane (Moldowan et al., 2015) when compared to the B-17X
2001; Wei et al., 2007). Despite the fact that diamondoid-based source sample and Dos Bocas extracts (around 30). The compound-specific
and maturity ratios should be used with caution when evaporative isotopic signatures of the lower diamondoids (CSIA-D) provide a use­
fractionation occurs (Chakhmakhchev et al., 2017), the condensate ful tool for oil-oil and oil-source correlation in this case: the study
samples and the Dos Bocas extracts show similar values (around 75% condensates match almost perfectly with one another and with the Dos
and 22%) of dimethyl diamantane index 1 (ratio of 3,4-DMD to 4,9-DMD Bocas extracts (Fig. 8b), indicating that Amistad distillates probably
plus 3,4-DMD) and methyl diamantane parameter (ratio between 4-MD derive from the early Miocene Dos Bocas source rocks. Previous 1D

11
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Fig. 6. Plot of δ13C2-δ13C2-δ13C3 diagram for the gaseous samples analysed, suggesting genetic types of the sampled gases. Notes: I (sapropelic-type gas), II (mixed-
type), III (humic-type gas), and IV (biogenic).

mainly biogenic methane with minor quantities of thermally generated


Table 5
gas and hydrocarbon liquid. Presence of biogenic gas at great depths is
Main light hydrocarbon parameters of representative condensates in the Amis­
usually a result of rapid trap closure in actively subsiding basins (Pieri,
tad and Delfín fields.
2001). When combined δ13C1 with δD1 data (Whiticar diagram; Fig. 5a),
Parameter A-3 A-6 A-7 A-12 A-8A B-17X isotopic compositions of methane from some Amistad gases suggest that
%n-C7 2.13 2.32 2.36 2.44 2.21 25.50 the bacterial source generation mechanism followed a CO2 reduction
%MCH 45.11 44.49 42.96 43.93 43.25 37.13 pathway and not a fermentation pathway (the latest restricted to
%ΣDMCP 52.76 53.19 54.40 53.86 54.59 37.37
shallow sediments and the prior extending up to 2000 m below sediment
i-Heptane value 0.08 0.08 0.09 0.09 0.08 0.92
n-Heptane value (%) 2.05 2.12 2.15 2.18 2.07 20.66 surface; Whiticar, 1999). Some gas samples exhibit carbon isotopic
Tol/n-C7 4.12 3.41 3.33 3.77 3.90 0.59 signatures of CO2 from − 16.45‰ to − 19.22‰ (Table 4), which are
n-C7/MCH 0.05 0.05 0.05 0.05 0.05 0.68 consistent with gases formed by organic genesis (Dai et al., 1996). The
n-C6/n-C7 0.56 0.67 0.69 0.71 0.62 1.02 C2/C1 versus δ13C1 graph can be used to estimate the isotopic value of
%C5-7 n-Alkanes 4.23 3.68 4.19 3.92 3.50 29.04
%C5-7 i-Alkanes 4.98 4.73 5.13 4.99 4.26 30.46
the biogenic gas end-member: biogenic gases have extremely small or no
%C5-7 Cycloalkanes 90.79 91.59 90.68 91.09 92.24 40.50 amounts of ethane, and thus, among the studied samples, A-7, A-8D and
Pristane/Phytane 2.15 2.29 2.27 1.97 2.04 1.87 A-8A allow to extrapolate the δ13C1 value (near − 64‰; see Fig. 9a) of
Pristane/n-C17 8.66 8.22 9.14 8.73 9.46 1.22 such biogenic end-member.
Phytane/n-C18 3.23 3.56 3.56 3.19 3.45 0.78
Regarding the Delfín B-17X gas, assuming that the carbon isotope
value of the Dos Bocas kerogen is nearly − 23‰ (Montenegro and
geochemical modeling approach in the exploratory well Amistad Sur, Benítez, 1991) and using the model reported by Chung et al. (1988), the
which reported vitrinite reflectance data and penetrated the whole plot of reciprocal carbon number (1/n) versus δ13Cn (r2 > 0.999) in
sequence, suggested that the Dos Bocas Formation reached the main Fig. 9b confirms that this gaseous sample appears completely or almost
phase of oil generation during the Neogene and can still be within the oil entirely to have a thermogenic origin (Rangel et al., 2003). In the C2/C1
window in the southern Progreso Basin (Evangelista, 2019). versus δ13C1 graph (Fig. 9a), where all the gaseous samples plot
approximately along a straight line, the sample B-17X would represent
the thermogenic end-member gas signature of the Amistad gases. On the
4.4. Origin and accumulation of hydrocarbons in the study area basis of methane isotopic data and according to Schoell (1984), the A-7,
A-8D, and A-8A gases reveal predominantly bacteriogenic methane
Producible hydrocarbons in reservoir sands of the Amistad field are

12
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Fig. 7. a) and b), respectively, ternary diagrams of C7 and C5-7 light hydrocarbons in sampled gases from Amistad field and Delfín B-17X well.

Fig. 8. a) example of whole-oil gas chromatogram for


a representative gas condensate sample; b) CSIA-D for
oil-oil and oil-source correlations between represen­
tative condensate samples and with an extract ob­
tained from Dos Bocas source rocks. Note the
identification of the lower diamondoids 1 to 12 in the
order given: adamantane; 1-methyladamantane; 1,3-
dimethyladamantane; 1,3,5-trimethyladamantane; 2-
methyladamantane; cis-1,4-dimethyladamantane;
trans-1,4-dimethyladamantane; 1,3,6-trimethylada­
mantane; 1,2-dimethyladamantane; cis-1,3,4-trime­
thyladamantane; trans-1,3,4-trimethyladamantane,
and 1,2,5,7-tetramethyladamantane.

associated with a minor thermogenic input (<10%, ~17% and ~30%, condensates with ratios markedly below one), also indicative of partial
respectively), assuming that the δ13C1 value for each sample is equal to vaporization (Mango, 1997). Based on geochemical data and oil-source
the weighted average between the δ13C1 signatures of biogenic (− 64‰) rock correlations, the Delfín B-17X thermogenic gas and liquid could be
and thermogenic (− 38.45‰) end-members. associated with normally mature oil, which originated from the deeply
Amistad condensate samples have a series of light hydrocarbons in buried distal/deltaic Dos Bocas source rocks and possibly migrated
the n-C5 to n-C7 range showing a distribution with disproportionate vertically along faults.
amounts of toluene (Tol/n-C7 or aromaticity exceeding 3; Table 5), A probable hydrocarbon filling scenario for the Amistad structure
which would be indicative of partial vaporization. This feature is implies that thermal gas and liquid hydrocarbons might be related to a
coherent with variations in the volatile-sensitive n-C6/n-C7 ratio (note light crude oil, which nowadays could be present in reservoirs below the
that in B-17X distillate the ratio is about one as opposed to Amistad current pay zones to depths of up to about 6000 m (Evangelista, 2019).

13
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Fig. 9. a) Plot of methane carbon isotope ratio versus C2/C1 molecular ratio for sampled gases. This plot is used to estimate the characteristic isotopic value that may
be assigned to the locally sourced bacterial gas end-member present in the Amistad region. Data from this plot may be also used to accurately determine the bacterial
and thermogenic inputs into Amistad gaseous samples; b) plot of δ13C versus the inverse C number of individual C1–C4 n-alkanes.

Later, much of the original thermogenic hydrocarbons may have been remaining gases from the Amistad field, although the Gastar diagrams in
lost and/or phase-separated prior to in-reservoir biodegradation and Figs. 10b and c do not reveal large compositional differences between
biogenic gas charge. Dry gas of biogenic origin may have partitioned the group of gases from the southern part of the Progreso reservoir (A-4,
into an oil and gas phases. Thus, only residual quantities of altered hy­ A-6, A-7, A-11, and A-12) and those from the central part (A-1, A-5, and
drocarbon liquid and thermal gas in the Amistad reservoirs escaped A-8A), the E-W normal F1 fault could act as a intra-reservoir barrier
leakage and biogenic gas displacement. The effect of this mixing would between at least two separate reservoir bodies with partial or no con­
be that the thermogenic component of the resultant gas phase is diluted nectivity between them (Fig. 1b), which would be in agreement with
to such an extent that the Amistad gases appear completely to have a prior literature (Barzallo and Bemúdez, 2016). In addition, even though
bacteriogenic origin. The Amistad field has a complex history of fault­ the differences between the studied gases in each group are close to
ing, burial, uplift and erosion, therefore multiple charges of thermogenic analytical accuracy, baffles and barriers to diffusion within the central
hydrocarbons may have taken place, probably at different time intervals and southern parts of the Progreso reservoir cannot be completely
in the geologic past due to their distinct levels of maturity. Also, sedi­ dismissed.
ment reburial may have caused most of the biogenic gas to go into water The small differences of A-1, A-5, and A-8A gaseous samples with
solution and subsequent overburden removal resulting in fizz gas. respect to A-2 and A-3 gases (Fig. 10c and d) apparently discard the
notion of compartmentalization in the north-central area in the Amistad
field, where the E-W normal F2 fault might act as a partial seal (Fig. 1b).
4.5. Compartmentalization in Amistad structure
However, minimal differences between the latter two gaseous samples
were within experimental error range, suggesting that both can be
Comparison of chemical and isotopic data for Amistad gas samples
grouped, representing a separate compartment or sub-compartment in
was made to allow delineation of reservoir compartmentalization in the
the northern part of the Progreso reservoir. The relatively low C1/C2
Amistad Field. For this, natural gas data from the Amistad structure is
values for A-2 and A-3 gases indicate that they might be slightly affected
studied with Gastar-diagrams (Márquez et al., 2013). A series of 11
by the preferential loss of methane by cap rock leakage. This feature is
normalized parameters (C1/C2, C2/C3, iC4/nC4, δ13C1, δ13C2, δ13C3,
consistent with previous results by Barzallo and Bemúdez (2016).
δ13iC4, δ13nC4, δ13C3-δ13C2, δ13nC4-δ13iC4, and δ13C2-δ13C1) were
examined following the method proposed by Prinzhofer et al. (2000).
5. Conclusions
These parameters are related with the main processes that affect the
chemical and isotopic compositions of natural gases: maturity (all of
Organic richness and thermal maturity are low in most rock samples
them), efficiency of accumulation (the three isotopic ratios differences),
throughout the Neogene sequence in the Amistad area. Only several Dos
segregative migration (those involving methane) and biodegradation
Bocas rock samples contain enough organic matter to be considered as
(C2/C3 and iC4/nC4). Fig. 10 shows the mentioned-above data plotted in
potential source rocks, but even in these latter samples thermal matu­
four Gastar-diagrams.
ration is low. The vitrinite/huminite in most rock samples depicts
The A-8D gas from the Cerro Mala reservoir exhibits a markedly
slightly suppressed reflectance. All the analysed condensates show
different Gastar-diagram (Fig. 10a). In spite of contamination of this
similar gas chromatographic fingerprints. Molecular data and
sample during collecting, as discussed above, the significant variations
compound-specific isotopic analyses of lower diamondoids (CSIA-D)
in C1–C4 alkane chemical and isotopic data between this gaseous sample
indicate oil-source correlation between such hydrocarbon liquids and
and the other Amistad gases would indicate a lack of vertical commu­
the Dos Bocas rock extracts.
nication between the Progreso and Cerro Mala reservoirs. As for the

14
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Fig. 10. a), b), c) and d), four normalized Gastar diagrams for the groups of Amistad gaseous samples coming from the studied reservoirs. The scales are the same for
all 11 axes of each diagram.

Producible gases in reservoir sands on the Amistad structure and Declaration of competing interest
Delfín B-17X well are mainly composed of hydrocarbons dominated by
biogenic and thermogenic methane, respectively. The latter gas might The authors declare that they have no known competing financial
represent the thermogenic end-member gas signature of the Amistad interests or personal relationships that could have appeared to influence
gases. Results, combined with geological data available for the study the work reported in this paper.
area, lead to the conclusion that the Amistad area has experienced a
complex geological history in which liquid thermal hydrocarbons and Acknowledgments
gas were generated in deeply buried distal/deltaic Dos Bocas source
rocks and migrated along deep faults to late Miocene reservoir sands; Authors are grateful to Professors Galo Montenegro (ESPOL) and
later, thermal petroleum were leaked out and partially displaced by Marcos Escobar†(LUZ) for access to samples and scientific assistance.
subsequent biogenic methane during uplift events. Therefore, petroleum Carlos Boente obtained a post-doctoral contract within the program
could be present in deeper reservoirs and consideration should be given PAIDI 2020 (Ref 707 DOC 01097), co-financed by the Junta de Anda­
to the drilling of the bore of an exploratory well to test the petroleum lucía (Spain) and the EU.
potential below the current pay zones in the Amistad structure.
Analytical results also denote that Amistad hydrocarbons show References
biodegradation and partial vaporization. Finally, the heterogeneities
identified in Gastar diagrams are not decidedly indicative of three lateral Aizprua, C.A., 2021. Forearc Crustal Structure and Controlling Factors on Basin
Formation across the Southernmost Northern Andes. Norwegian University of
compartments being connected by the F1 and F2 normal faults. In turn, a Science and Technology, Trondheim, p. 242. PhD thesis.
reasonable communication within the southern, central and northern Aizprua, C., Witt, C., Johansen, S.E., Barba, D., 2019. Cenozoic stages of forearc
parts of the Progreso reservoir in the Amistad gas field could be inferred evolution following the accretion of a sliver from the late Cretaceous-Caribbean
large Igneous province: SW Ecuador-NW Peru. Tectonics 38, 1441–1465, 10.1029/
from the C1–C4 alkane analyses, though it is impossible to conclude 2018TC005235.
complete fluid homogenization in each of them.

15
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Alvarado, A., Audin, L., Nocquet, J.M., Jaillard, E., Mothes, P., Jarrín, P., Segovia, M., Diessel, C.F.K., 2010. The stratigraphic distribution of inertinite. Int. J. Coal Geol. 81,
Rolandone, F., Cisneros, D., 2016. Partitioning of oblique convergence in the 251–268. https://doi.org/10.1016/j.coal.2009.04.004.
Northern Andes subduction zone: migration history and the present-day boundary of Dow, W.G., 1977. Kerogen studies and geological interpretations. J. Geochem. Explor. 7,
the North Andean Sliver in Ecuador. Tectonics 35, 1048–1065. https://doi.org/ 79–99. https://doi.org/10.1016/0375-6742(77)90078-4.
10.1002/2016TC004117. Egbue, O., Kellogg, J., 2010. Pleistocene to present North Andean “escape.
Barzallo, M.I., Bemúdez, D.A., 2016. Estudio de compartimentalización del Campo Tectonophysics 489, 248–257. https://doi.org/10.1016/j.tecto.2010.04.021.
Amistad. BSc thesis. Escuela Politécnica Superior del Litoral, Guayaquil, p. 81. El Diasty, W.Sh, Peters, K.E., Moldowan, J.M., Essa, G.I., Hammad, M.M., 2020. Organic
Behar, F., Kressmann, S., Rudkiewicz, J.L., Vandenbroucke, M., 1992. Experimental geochemistry of condensates and natural gases in the northwest Nile Delta offshore
simulation in a confined system and kinetic modelling of kerogen and oil cracking. Egypt. J. Petrol. Sci. Eng. 187, 106819. https://doi.org/10.1006/j.
Org. Geochem. 19, 173–189. https://doi.org/10.1016/0146-6380(92)90035-V. petrol.2019.106819.
Benítez, S., 1995. Évolution géodynamique de la Province cótière sud-équatorienne au Encalada, B.A., 2017. Análisis de alternativas para evitar el ahogamiento de los pozos de
Crétacé-supérieur-Tertiaire. These de Doctoat d’ Universite Institute Dolomieu, gas perforados en el Amistad field. BSc thesis, Facultad de Ingeniería, Universidad
Grenoble, p. 160. Tecnológica Equinoccial, Quito, p. 63.
Bernard, B.B., Brooks, J.M., Sackett, W.M., 1978. Light hydrocarbons in recent Texas Espitalié, J., Deroo, G., Marquis, F., 1985. La pyrolyse Rock-Eval et ses applications.
continental shelf and slope sediments. J. Geophys. Res. 83, 4053. https://doi.org/ Première partie. Oil Gas Sci. Technol. 40, 563–579. https://doi.org/10.2516/ogst:
10.1029/JC083iC08p04053. 1985035.
Boreham, C.J., Edwards, D.S., 2008. Abundance and carbon isotopic composition of neo- Espitalié, J., Laporte, J., Madec, M., Marquis, F., Leplat, P., Paulet, J., Boutefeu, A., 1977.
pentane in Australian natural gases. Org. Geochem. 39, 550–566. https://doi.org/ Méthode rapide de caractérisation des roches méres, de leur potential pétrolier et de
10.1016/j.orggeochem.2007.11.004. leur degré d’evolution. Oil Gas Sci. Technol. 32, 23–42.
Bristow, C.R., 1975. On the age of the Zapotal sands of southwest Ecuador. Newsl. Esquinas, N., Márquez, G., Permanyer, A., Gallego, J.R., 2018. Geochemical evaluation of
Stratigr. 4, 119–134. https://doi.org/10.1127/nos/4/1975/119. crude oils from the Caracara and Tiple areas, eastern Llanos Basin, Colombia: palaeo
Bristow, C.R., Hoffstetter, R., 1977. Fasc. 5a2, Ecuador; Centre National de la Recherche biodegradation and oil mixing. J. Petrol. Geol. 41, 113–134. https://doi.org/
Scientifique, Paris Lex. Stratigr. Int. V, 410. 10.1111/jpg.12696.
Cai, C., Hu, W., Worden, R.H., 2001. Thermochemical sulphate reduction in Evangelista, Z., 2019. Modelo 1D de historia de soterramiento de la Cuenca petrolera
Cambro–Ordovician carbonates in central Tarim. Mar. Petrol. Geol. 18, 729–741. Progreso. BSc thesis, Universidad de Huelva, Huelva (Spain), p. 63.
https://doi.org/10.1016/S0264-8172(01)00028-9. Evans, C.D.R., Whittaker, I.E., 1982. The geology of the Westem part of the Borbon basin,
Calahorrano, A., Sallares, V., Collot, J., Sage, F., Ranero, C.R., 2008. Nonlinear variations north-west Ecuador. In: Leggett, J.K. (Ed.), Trench- Forearc Geology: Sedimentation
of the physical properties along the southern Ecuador subduction channel: results and Tectonics on Modern and Ancient Active Plate Margins, 10. Geol. Soc. Lon. Sp.
from depth-migrated seismic data. Earth Planet Sci. Lett. 267, 453–467. https://doi. Pub., pp. 191–198
org/10.1016/j.epsl.2007.11.061. Fuex, A.N., 1977. The use of stable carbon isotopes in hydrocarbon exploration.
Cao, J., Wang, X., Sun, P., Zhang, Y., Tang, Y., Xiang, B., Lan, W., Wu, M., 2012. J. Geochem. Explor. 77, 155–188.
Geochemistry and origins of natural gases in the central Junggar Basin, northwest Galimov, E.M., 2006. Isotope organic geochemistry. Org. Geochem. 37, 1200–1262.
China. Org. Geochem. 53, 166–176. https://doi.org/10.1016/j. https://doi.org/10.1016/j.orggeochem.2006.04.009.
orggeochem.2012.06.009. Gao, L., Guimond, J., Thomas, E., Huang, Y., 2015. Major trends in leaf wax abundance,
Carr, A.D., 2000. Suppression and retardation of vitrinite reflectance, Part I: formation δ2H and δ13C values along leaf venation in five species of C3 plants: physiological
and significance for hydrocarbon exploration. J. Petrol. Geol. 23, 313–343. https:// and geochemical implications. Org. Geochem. 78, 144–152. https://doi.org/
doi.org/10.1111/j.1747-5457.2000.tb01022.x. 10.1016/j.orggeochem.2014.11.005.
Chakhmakhchev, A., Sanderson, J., Pearson, C., Davidson, N., 2017. Compositional Gürgey, K., Philp, R.P., Clayton, C., Emiroğlu, H., Siyako, M., 2005. Geochemical and
changes of diamondoid distributions caused by simulated evaporative fractionation. isotopic approach to maturity/source/mixing estimations for natural gas and
Org. Geochem. 113, 224–228. https://doi.org/10.1016/j.orggeochem.2017.06.016. associated condensates in the Thrace Basin, NW Turkey. Appl. Geochem. 20,
Chen, Z., Cao, Y., Ma, Z., Zhen, Y., 2014. Geochemistry and origins of natural gases in the 2017–2037. https://doi.org/10.1016/j.apgeochem.2005.07.012.
Zhongguai area of Junggar Basin, China. J. Petrol. Sci. Eng. 119, 17–27. https://doi. Hao, F., Guo, T., Zhu, Y., Cai, X., Zou, H., Li, P., 2008. Evidence for multiple stages of oil
org/10.1016/j.petrol.2014.05.007. cracking and thermochemical sulfate reduction in the Puguang gas field, Sichuan
Chen, J.F., Miao, Z.Y., Zhang, C., Chen, H.Y., 2010. Geochemical characteristics of light Basin, China. Am. Assoc. Petrol. Geol. Bull. 92, 611–637. https://doi.org/10.1306/
hydrocarbons in natural gas in the Tabei Uplift of the Tarim Basin and their 01210807090.
implications. Oil Gas Geol. 31, 271–276. He, M., Moldowan, J.M., Nemchenko-Rovenskaya, A., Peters, K.E., 2012. Oil families and
Chen, J.F., Xu, Y.C., Huang, D.F., 2000. Geochemical characteristics and origin of natural their inferred source rocks in the Barents Sea and northern Timan-Pechora Basin,
gas in Tarim Basin, China. AAPG Bull. 84, 591–606. Russia. Am. Assoc. Petrol. Geol. Bull. 96, 1121–1146. https://doi.org/10.1306/
Chen, J., Fu, J., Sheng, G., Liu, D., Zhang, J., 1996. Diamondoid hydrocarbon ratios: 10181111043.
novel maturity indices for highly mature crude oils. Org. Geochem. 25, 179–190. Higley, D.K., 2004. The Progreso Basin Province of Northwestern Peru and Southwestern
https://doi.org/10.1016/S0146-6380(96)00125-8. Ecuador: Neogene and Cretaceous-Paleogene Total Petroleum Systems. U.S.
Chen, K.M., Jin, W.M., He, Z.H., Chen, J.P., Yang, Z.F., 1987. Composition characteristics Geological Survey Bulletin 2206-B, Denver, p. 25.
of light hydrocarbons in continental oil/condensate and their geological Hu, T.L., Ge, B.X., Chang, Y.J., Liu, B., 1990. The development and application of
significance. Petrol. Explor. Dev. 14, 33–34. fingerprint parameters for hydrocarbons absorbed by source rocks and light
Chung, H.M., Gormly, J.R., Squires, R.M., 1988. Origin of gaseous hydrocarbons in hydrocarbons in natural gas. Petrol. Geol. Exp. 12, 375–394.
subsurface environments: theoretical considerations of carbon isotope distribution. Hu, G.Y., Li, J., Li, Z.S., Luo, X., Sun, Q.W., Ma, C.H., 2007. The discussion of light
Chem. Geol. 71, 97–104. https://doi.org/10.1016/0009-2541(88)90108-8. hydrocarbon for gas genetic type identification. Sci. China Earth Sci. 37, 111–117.
Clayton, C., 1991. Carbon isotope fractionation during natural gas generation from Hu, G.Y., Li, J., Li, J., Shan, X.Q., Han, Z.X., 2010. The origin of natural gas and the
kerogen. Mar. Petrol. Geol. 8, 232–240. https://doi.org/10.1016/0264-8172(91) hydrocarbon charging history of the Yulin gas field in the Ordos Basin, China. Int. J.
90010-X. Coal Geol. 81, 381–391. https://doi.org/10.1016/j.coal.2009.07.016.
Cornford, C., Gardner, P., Burgess, C., 1998. Geochemical truths in large data sets. I: Iglesias, M., Jiménez, A., del Rio, J., Suarez-Ruiz, I., 2000. Molecular characterisation of
geochemical screening data. Org. Geochem. 29, 519–530. https://doi.org/10.1016/ vitrinite in relation to natural hydrogen enrichment and depositional environment.
S0146-6380(98)00189-2. Org. Geochem. 31, 1285–1299. https://doi.org/10.1016/S0146-6380(00)00086-3.
Dai, J.X., Ni, Y., Hu, G., Huang, S., Liao, F., Yu, C., Gong, D., Wu, W., 2014. Stable carbon Jaillard, E., Laubacher, G., Bengtson, P., Dhondt, A.V., Bulot, L.G., 1999. Stratigraphy
and hydrogen isotopes of gases from the large tight gas fields in China. Sci. China and evolution of the Cretaceous forearc Celica-lancones basin of south-western
Earth Sci. 57, 88–103. https://doi.org/10.1007/s11430-013-4701-7. Ecuador. J. South Am. Earth Sci. 12, 51–68, 10 .1016/S0895 -9811 (99)00006 -1.
Dai, J.X., Qin, S.F., Tao, S.Z., Zhu, G.Y., Mi, J.K., 2005. Developing trends of natural gas Jaillard, E., Ordonez, M., Benitez, S., Berrones, G., Jimenez, N., Montenegro, G.,
industry and the significant progress on natural gas geological theories in China. Nat. Zambrano, I., 1995. Basin development in an accretionary, oceanic-floored fore-arc
Gas Geosci. 16, 127–142. setting: southern Coastal Ecuador during Late Cretaceous-Late Eocene time. In:
Dai, J.X., Song, Y., Dai, C.S., Wang, D.R., 1996. Geochemistry and accumulation of Tankard, A.J., Suarez, S.R., Welsink, H.J. (Eds.), Petroleum Basins of South America,
carbon dioxide gases in China. AAPG Bull. 80, 1615–1626. American Association of Petroleum Geologists Memoir 62, pp. 615–631.
Dai, J.X., 1992. Identification of different hydrocarbon gas. Sci. China Earth Sci. 2, Jackson, R.B., Down, A., Phillips, N.G., Ackley, R.C., Cook, C.W., Plata, D.L., Zhao, K.,
185–193. 2014. Natural gas pipeline leaks across Washington, DC. Environ. Sci. Technol. 48,
Dahl, J.E., Moldowan, J.M., Peters, K.E., Claypool, G.E., Rooney, M.A., Michael, G.E., 2051–2058. https://doi.org/10.1021/es404474x.
Mello, M.R., Kohnen, M.L., 1999. Diamondoid hydrocarbons as indicators of natural James, A.T., Burns, B.J., 1984. Microbial alteration of subsurface natural gas
oil cracking. Nature 399, 54–57. https://doi.org/10.1038/19953. accumulations. AAPG Bull. 68, 957–960.
Deckelman, J.A., Connors, F.X., Shultz, A.W., Glagola, P.A., Menard, W.M., Schwegal, S. Jarvie, D.M., Claxton, B.L., Henk, B., Breyer, J., 2001. Oil and shale gas from the Barnett
R., Shearer, J.N., 2008. Neogene oil and gas reservoirs in the Progreso Basin, shale, Fort Worth basin, Texas. In: Am. Assoc. Pet. Geol. Annual National
offshore Ecuador and Peru: implications for petroleum exploration and Convention, Denver, June 3–6, p. A100.
development. J. Petrol. Geol. 31, 43–60. https://doi.org/10.1111/j.1747- Jenden, P.D., Drazan, D.J., Kaplan, I.R., 1993. Mixing of thermogenic natural gases in
5457.2008.00406.x. northern Appalachian Basin. AAPG Bull. 77, 980–998.
Deniaud, Y., Baby, P., Basile, C., Ordonez, M., Montenegro, G., Mascle, G., 1999. Opening Jenden, P., Kaplan, I., Poreda, R., Craig, H., 1988. Origin of nitrogen-rich natural gases in
and tectonic and sedimentary evolution of the Gulf of Guayaquil: Neogene and the California Great Valley: evidence from helium, carbon and nitrogen isotope
Quaternary fore-arc basin of the south Ecuadorian Andes. CR Earth Planet. Sci. 328, ratios. Geochem. Cosmochim. Acta 52, 851–861. https://doi.org/10.1016/0016-
181–187, 10.1016/S1251-8050(99)80094-9. 7037(88)90356-0.

16
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Kalkreuth, W.D., 1982. Rank and petrographic composition of selected Jurassic-lower Schoell, M., 1980. The hydrogen and carbon isotopic composition of methane from
Cretaceous coals of British Columbia. Can. B. Can. Petrol. Geol. 30, 112–139. natural gases of various origins. Geochem. Cosmochim. Acta 44, 649–661. https://
https://doi.org/10.35767/gscpgbull.30.2.112. doi.org/10.1016/0016-7037(80)90155-6.
Katz, B.J., Narimanov, A., Huseinzadeh, R., 2002. Significance of microbial processes in Sheppard, G., 1937. The Geology of Southwestern Ecuador. Billing and Sons, London,
gases of the South Caspian basin. Mar. Petrol. Geol. 19, 783–796. https://doi.org/ p. 275.
10.1016/S0264-8172(02)00086-7. Schulz, L.K., Wilhelms, A., Rein, E., Steen, A.S., 2001. Application of diamondoids to
Kraemer, P.E., Weiner, A., Alvarez, P., 2001. Tectonoestratigraphic framework of a distinguish source rock facies. Org. Geochem. 32, 365–375. https://doi.org/
transtensional forearc basin - the Tumbes Basin of northwestern Peru. In: Poster 10.1016/S0146-6380(01)00003-1.
Abstract, Search and Discovery Article #90906©2001, AAPG Annual Convention. Speight, J.G., 2007. The Chemistry and Technology of Petroleum, fifth ed. Taylor and
June 3-6, 2001, Denver, Colorado. Francis, Boca Raton, Florida, p. 953.
Larter, S., di Primio, R., 2005. Effects of biodegradation on oil and gas field PVT Stach, E., Mackowsky, M.Th, Teichmüller, M., Taylor, G.H., Chandra, D., Teichmüller, R.,
properties and the origin of oil rimmed gas accumulations. Org. Geochem. 36, 1982. Textbook of Coal Petrology, third ed. Gebruder Borntraeger, Berlin, p. 535.
299–310. https://doi.org/10.1016/j.orggeochem.2004.07.015. Stainforth, R.M., 1948. Applied micropaleontology in coastal Ecuador. J. Paleontol. 22,
Lijmbach, G.W.M., van der Veen, F.M., Englehardt, E.D., 1983. Geochemical 113–151.
characterization of crude oils and source rocks using field ionization mass Strąpoć, D., Mastalerz, M., Schimmelmann, A., Drobniak, A., Hasenmueller, N.R., 2010.
spectrometry. In: Bjorøy, M., Albrecht, C., Cornford, C. (Eds.), Advances in Organic Geochemical constraints on the origin and volume of gas in the new Albany shale
Geochemistry. John Wiley and Sons, New York, pp. 788–798. (Devonian–Mississippian), eastern Illinois basin. AAPG Bull. 94, 1713–1740. https://
Liu, Q., Dai, J., Li, J., Zhou, Q., 2008. Hydrogen isotope composition of natural gases doi.org/10.1306/06301009197.
from the Tarim Basin and its indication of depositional environments of the source Strąpoć, D., Mastalerz, M., Eble, C., Schimmelmann, A., 2007. Characterization of the
rocks. Sci. China Earth Sci. 51, 300–311. https://doi.org/10.1007/s11430-008- origin of coalbed gases in southeastern Illinois Basin by compound-specific carbon
0006-7. and hydrogen stable isotope ratios. Org. Geochem. 38, 267–287. https://doi.org/
Lorant, F., Prinzhofer, A., Behar, F., Huc, A.-Y., 1998. Carbon isotopic and molecular 10.1016/j.orggeochem.2006.09.005.
constraints on the formation and the expulsion of thermogenic hydrocarbon gases. Suarez-Ruiz, I., Jimenez, A., Iglesias, M.J., Laggoun-Defarge, F., Prado, J.G., 1994.
Chem. Geol. 147, 249–264. https://doi.org/10.1016/S0009-2541(98)00017-5. Influence of resinite on huminite properties. Energy Fuel. 8, 1417–1424. https://doi.
Luzieux, L.D.A., Heller, F., Spikings, R., Vallejo, C.F., Winkler, W., 2006. Origin and org/10.1021/ef00048a033.
Cretaceous tectonic history of the coastal Ecuadorian forearc between 1◦ N and 3◦ S: Sun, P., Wang, Y., Leng, K., Li, H., Ma, W., Cao, J., 2016. Geochemistry and origin of
paleomagnetic, radiometric and fossil evidence. Earth Planet Sci. Lett. 249, 400–414. natural gas in the eastern Junggar Basin, NW China. Mar. Petrol. Geol. 75, 240–251.
https://doi.org/10.1016/j.epsl.2006.07.008. https://doi.org/10.1016/j.marpetgeo.2016.04.018.
Mango, F.D., 1997. The light hydrocarbons in petroleum: a critical review. Org. Tao, K., Cao, J., Wang, Y., Ma, W., Xiang, B., Ren, J., Zhou, N., 2016. Geochemistry and
Geochem. 26, 417–440. https://doi.org/10.1016/S0146-6380(97)00031-4. origin of natural gas in the petroliferous Mahu sag, northwestern Junggar Basin, NW
Mango, F.D., 1990. The origin of light hydrocarbons in petroleum: a kinetic test of the China: carboniferous marine and Permian lacustrine gas systems. Org. Geochem.
steady-state catalytic hypothesis. Geochem. Cosmochim. Acta 54, 1315–1323. 100, 62–79. https://doi.org/10.1016/j.orggeochem.2016.08.004.
https://doi.org/10.1016/0016-7037(90)90156-F. Taylor, G.H., Liu, S.Y., Diessel, C.F.K., 1989. The cold-climate origin of inertinite-rich
Marchant, S., 1965. Gravity slide deposits in timor and Ecuador. Geol. Mag. 102, Gondwana coals. Int. J. Coal Geol. 11, 1–22. https://doi.org/10.1016/0166-5162
464–465. https://doi.org/10.1017/S0016756800053759. (89)90110-9.
Márquez, G., Escobar, M., Lorenzo, E., Gallego, J.R., Tocco, R., 2013. Using gas Thompson, K.F.M., 1987. Fractionated aromatic petroleums and the generation of gas-
geochemistry to delineate structural compartments and assess petroleum reservoir- condensates. Org. Geochem. 11, 573–590. https://doi.org/10.1016/0146-6380(87)
filling directions: a Venezuelan case study. J. South Am. Earth Sci. 43, 1–7. https:// 90011-8.
doi.org/10.1016/j.jsames.2012.12.008. Thompson, K.F.M., 1983. Classification and thermal history of petroleum based on light
Moldowan, J.M., Dahl, J., Zinniker, D., Barbanti, S.M., 2015. Underutilized advanced hydrocarbons. Geochem. Cosmochim. Acta 47, 303–316. https://doi.org/10.1016/
geochemical technologies for oil and gas exploration and production-1. The 0016-7037(83)90143-6.
diamondoids. J. Petrol. Sci. Eng. 126, 87–96. https://doi.org/10.1016/j. Thompson, K.F.M., 1979. Light hydrocarbons in subsurface sediments. Geochem.
petrol.2014.11.010. Cosmochim. Acta 43, 657–672. https://doi.org/10.1016/0016-7037(79)90251-5.
Montenegro, G., Benítez, S., 1991. Estado actual del conocimiento geoquímico de la Tian, H., Xiao, X., Yang, L., Xiao, Z., Guo, L., Shen, J., Lu, Y., 2009. Pyrolysis of oil at high
cuencas del suroeste ecuatoriano. Bolet. Geol. Ecuat. 2 (1), 27–46. temperatures: gas potentials, chemical and carbon isotopic signatures. Sci. Bull. 54,
Mukhopadhyay, P.K., Hatcher, P.G., Calder, J.H., 1991. Hydrocarbon generation from 1217–1224. https://doi.org/10.1007/s11434-008-0590-0.
deltaic and intermontane fluviodeltaic coal and coaly shale from the Tertiary of Tissot, B.P., Welte, D.H., 1984. Petroleum Formation and Occurrence, second ed.
Texas and Carboniferous of Nova Scotia. Org. Geochem. 17, 765–783. https://doi. Springer-Verlag, New York, p. 699.
org/10.1016/0146-6380(91)90020-K. Wang, K., Pang, X., Zhang, H., Zhao, Z., Su, S., Hui, S., 2018. Characteristics and genetic
Olsson, A.A., 1931. Contributions to the Tertiary paleontology of northern Peru, part 4, types of natural gas in the northern Dongpu Depression, Bohai Bay basin, China.
the Peruvian Oligocene. Bull. Am. Paleontol. 17, 100–264. J. Petrol. Sci. Eng. 170, 453–466. https://doi.org/10.1006/j.petrol.2018.06.080.
Pallasser, R.J., 2000. Recognising biodegradation in gas/oil accumulations through the Wei, Z., Moldowan, J.M., Peters, K.E., Wang, Y., Xiang, W., 2007. The abundance and
δ13C compositions of gas components. Org. Geochem. 31, 1363–1373. https://doi. distribution of diamondoids in biodegraded oils from the San Joaquin Valley:
org/10.1016/S0146-6380(00)00101-7. implications for biodegradation of diamondoids in petroleum reservoirs. Org.
Patience, R.L., 2003. Where did all the coal gas go? Org. Geochem. 34, 375–387. https:// Geochem. 38, 1910–1926. https://doi.org/10.1016/j.orggeochem.2007.07.009.
doi.org/10.1016/S0146-6380(02)00214-0. Whiticar, M.J., 1999. Carbon and hydrogen isotope systematics of bacterial formation
Petersen, H.I., Rosenberg, P., 1998. Reflectance retardation (suppression) and source and oxidation of methane. Chem. Geol. 161, 291–314. https://doi.org/10.1016/
rock properties related to hydrogen-enriched vitrinite in Middle Jurassic coals, S0009-2541(99)00092-3.
Danish North Sea. J. Petrol. Geol. 21, 247–263. https://doi.org/10.1111/j.1747- Witt, C., Bourgois, J., 2010. Forearc basin formation in the tectonic wake of a collision-
5457.1998.tb00781.x. driven, coastwise migrating crustal block: the example of the North Andean block
Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The Biomarker Guide: Biomarkers and and the extensional Gulf of Guayaquil-Tumbes Basin (Ecuador-Peru border area).
Isotopes in Petroleum Systems and Earth History, second ed. Cambridge University Geol. Soc. Am. Bull. 122, 89–108. https://doi.org/10.1130/B26386.1.
Press, Cambrige, p. 1150. Witt, C., Bourgois, J., Michaud, C., Ordonez, M., Jimenez, N., Sosson, M., 2006.
Peters, K.E., 1986. Guidelines for evaluating petroleum source rocks using programmed Development of the Gulf of Guayaquil (Ecuador) during the quaternary as an effect
pyrolysis. AAPG Bull. 70, 318–329. of the north andean block tectonic escape. Tectonics 25. https://doi.org/10.1029/
Pieri, M., 2001. Italian petroleum geology. In: Vai, F., Martini, I.P. (Eds.), Anatomy of an 2004TC001723. TC3017, 22.
Orogen: the Apennines and Adjacent Mediterranean Basins. Springer Publishing, Witt, C., Reynaud, J.Y., Barba, D., Poujol, M., Aizprua, C., Rivadeneira, M., Amberg, C.,
New York, pp. 533–549. 2019. From accretion to forearc basin initiation: the case of SW Ecuador, Northern
Prinzhofer, A., Dos Santos Neto, E.V., Battani, A., 2010. Coupled use of carbon isotopes Andes. Sediment. Geol. 379, 138–157. https://doi.org/10.1016/j.
and noble gas isotopes in the Potiguar basin (Brazil): fluids migration and mantle sedgeo.2018.11.009.
influence. Mar. Petrol. Geol. 27, 1273–1284. https://doi.org/10.1016/j. Witt, C., Rivadeneira, M., Poujol, M., Barba, D., Beida, D., Beseme, G., Montenegro, G.,
marpetgeo.2010.03.004. 2017. Tracking ancient magmatism and Cenozoic topographic growth within the
Prinzhofer, A.A., Mello, M.R., da Sila Freitas, L.C., Takaki, T., 2000. A new geochemical Northern Andes forearc: constraints from detrital U-Pb zircon ages. Geol. Soc. Am.
characterization of natural gas and its use in oil and gas evaluation. In: Mello, M.R., Bull. 129, 415–428. https://doi.org/10.1130/B31530.1.
Katz, B.J. (Eds.), Petrol. Syst. South Atl. Margins. AAPG Bull., 70, pp. 107–119. Xia, X., Chen, J., Braun, R., Tang, Y., 2013. Isotopic reversals with respect to maturity
Prinzhofer, A.A., Huc, A.Y., 1995. Genetic and post-genetic molecular and isotopic trends due to mixing of primary and secondary products in source rocks. Chem. Geol.
fractionations in natural gases. Chem. Geol. 126, 281–290. https://doi.org/10.1016/ 339, 205–212. https://doi.org/10.1016/j.chemgeo.2012.07.025.
0009-2541(95)00123-9. Yang, C., Luo, X., Li, J., Li, Z., Liu, Q., Wang, Y., 2008. Geochemical characteristics of
Rangel, A., Katz, B., Ramirez, V., Vaz dos Santos Neto, E., 2003. Alternative pyrolysis gas from epimetamorphic rocks in the northern basement of Songliao
interpretations as to the origin of the hydrocarbons of the Guajira Basin, Colombia. Basin, Northeast China. Sci. China Earth Sci. 51, 140–147. https://doi.org/10.1007/
Mar. Petrol. Geol. 20, 129–139. https://doi.org/10.1016/S0264-8172(03)00061-8. s11430-008-5019-8.
Reynaud, J.Y., Witt, C., Pazmiño, A., Gilces, S., 2018. Tide-dominated deltas in active
margin basins: insights from the Guayas estuary, Gulf of Guayaquil, Ecuador. Mar.
Geol. 403, 165–178.
Schoell, M., 1984. Recent advances in petroleum isotope geochemistry. Org. Geochem. 6,
645–663. https://doi.org/10.1016/0146-6380(84)90086-X.

17
M.A. Guzmán et al. Journal of Petroleum Science and Engineering 213 (2022) 110410

Zhu, Y., Shi, B., Fang, C., 2000. The isotopic compositions of molecular nitrogen: Zumberge, J., Ferworn, K., Brown, S., 2012. Isotopic reversal (“rollover”) in shale gases
implications on their origins in natural gas accumulations. Chem. Geol. 164, produced from the Mississippian Barnett and Fayetteville formations. Mar. Petrol.
321–330. https://doi.org/10.1016/S0009-2541(99)00151-5. Geol. 31, 43–52. https://doi.org/10.1016/j.marpetgeo.2011.06.009.
Zumberge, J.E., Ferworn, K.A., Curtis, J.B., 2009. Gas character anomalies found in
highly productive shale gas wells. Geochem. Cosmochim. Acta 73 (Suppl. 1), A1539.

18

You might also like