You are on page 1of 34

Quaternary

Quaternary
Volume 3

Chapter 8

Neogene
Cenozoic Geologic Evolution of the Southern https://doi.org/10.32685/pub.esp.37.2019.08
Published online 26 November 2020

Tumaco Forearc Basin (SW Colombian Pacific)

Paleogene
Andrés PARDO–TRUJILLO 1* , Sebastián ECHEVERRI 2 , Carlos BORRERO3 , 1 andres.pardo@ucaldas.edu.co
Universidad de Caldas
Alejandro ARENAS 4 , Felipe VALLEJO 5 , Raúl TREJOS 6 , Ángelo PLATA7 , Departamento de Ciencias Geológicas

Cretaceous
José–Abel FLORES 8 , Agustín CARDONA 9, Sergio RESTREPO 10 , Instituto de Investigaciones en Estratigrafía–
IIES
Ángel BARBOSA11 , Hugo MURCIA12 , Carlos GIRALDO 13 , Sergio CELIS14 , Grupo de Investigación en Estratigrafía y
Vulcanología–GIEV
and Sergio A. LÓPEZ15 Calle 65 n.° 26–10
Manizales, Colombia
Abstract Tumaco is a forearc basin that contains insights into the Cenozoic geologic 2 sebastian.echeverri@usp.br

Jurassic
Universidad de Caldas
evolution of SW Colombia. In this region, the subduction of the Farallon and Nazca Instituto de Investigaciones en Estratigrafía–
Plates beneath the South American Plate have controlled subsidence and magmatic IIES
Grupo de Investigación en Estratigrafía y
activity during the Oligocene to recent times. A synthesis of seismic, stratigraphic, Vulcanología–GIEV
Calle 65 n.° 26–10
petrographic, geochronologic, and biostratigraphic data from outcrops and wells is
Manizales, Colombia
presented. The Tumaco onshore basin has a trough–shaped symmetric geometry lim-

Triassic
3 borrero_c@yahoo.com
ited to the east by the Western Cordillera and to the west by the Remolino Grande– Universidad de Caldas
Departamento de Ciencias Geológicas
Gorgona Structural High. ca. 8000 m of sediments were accumulated in its depocenter Instituto de Investigaciones en Estratigrafía–
IIES
during the Cenozoic. The sedimentites are composed of mudrocks, sandstones, and Grupo de Investigación en Estratigrafía y
conglomerates, which vary in their proportions over time, and were mainly accumu- Vulcanología–GIEV

Permian
Calle 65 n.° 26–10
lated in open marine and deltaic environments. Calcareous nannofossils, foraminifera, Manizales, Colombia
and palynomorphs allowed assignment of the depositional time of the sedimentary 4 alejandro.arenas@ucaldas.edu.co
Universidad de Caldas
units; however, the low abundance, preservation, and reworking of microfossils in Instituto de Investigaciones en Estratigrafía–
some intervals require the use of multi–tools to determine the age of the deposits. IIES

Carboniferous
Grupo de Investigación en Estratigrafía y
Sandstones are mainly litharenites and feldspathic litharenites, are texturally imma- Vulcanología–GIEV
Calle 65 n.° 26–10
ture, and are composed of cherts fragments, basic to intermediate volcanic fragments, Manizales, Colombia
and crystals such as feldspars (Na and K), pyroxene, amphibole, and biotite, which can 5 diego.vallejo@ucaldas.edu.co
Universidad de Caldas
be associated with basic–intermediate volcanic, plutonic, and sedimentary rocks of the Instituto de Investigaciones en Estratigrafía–
current basement of Western Cordillera. Sediment provenance analysis (detrital zircon IIES

Devonian
Grupo de Investigación en Estratigrafía y
and heavy minerals) suggests continuous volcanism from late Oligocene to Pleistocene Vulcanología–GIEV
times, the activity of which has increased since the middle Miocene. The presence Calle 65 n.° 26–10
Manizales, Colombia
of low percentages of pre–Cenozoic zircons and metamorphic rock fragments in the 6 raul.trejos@ucaldas.edu.co
Miocene units are related to reworking of ancient sedimentary units or to a partial Universidad de Caldas
Instituto de Investigaciones en Estratigrafía–
connection with the Central Cordillera basement. The study of Miocene – Pliocene

Silurian
IIES
Grupo de Investigación en Estratigrafía y
outcrops and well cores allows the interpretation of a shallowing of the basin during Vulcanología–GIEV
the Messinian – Zanclean times. Volcanoclastic fans, as well as fluvial and coastal Calle 65 n.° 26–10
Manizales, Colombia
sediments, associated with the current Patía and Mira Rivers are partially covering the
7 angeloplata@gmail.com
Miocene – Pliocene deposits.
Ordovician

Universidad de Caldas
Instituto de Investigaciones en Estratigrafía–
Keywords: Tumaco Basin, sedimentary provenance, biostratigraphy, Colombian Pacific, IIES
Cenozoic. Grupo de Investigación en Estratigrafía y
Vulcanología–GIEV
Calle 65 n.° 26–10
Manizales, Colombia
Citation: Pardo–Trujillo, A., Echeverri, S., Borrero, C., Arenas, A., Vallejo, F., Trejos, R., Plata, Á.,
Cambrian

Flores, J.A., Cardona, A., Restrepo, S., Barbosa, Á., Murcia, H., Giraldo, C., Celis, S., & López, S.A. 8 flores@usal.es
Universidad de Salamanca
2020. Cenozoic geologic evolution of the southern Tumaco Forearc Basin (SW Colombian Pacif- Grupo de Geociencias Oceánicas
ic). In: Gómez, J. & Mateus–Zabala, D. (editors), The Geology of Colombia, Volume 3 Paleogene
– Neogene. Servicio Geológico Colombiano, Publicaciones Geológicas Especiales 37, p. 215–247.
Bogotá. https://doi.org/10.32685/pub.esp.37.2019.08
Proterozoic

215
PARDO–TRUJILLO et al.

9 agcardonamo@unal.edu.co Resumen Tumaco es una cuenca de frente de arco que guarda información sobre la
Universidad Nacional de Colombia
Sede Medellín evolución geológica cenozoica del SW de Colombia. En esta región, la subducción de las
Facultad de Minas
Medellín, Colombia
placas de Farallón y de Nazca bajo la Placa de Suramérica ha controlado la subsidencia y
10 sarestrepom@unal.edu.co actividad magmática durante el Oligoceno al reciente. Se presenta una síntesis de datos
Universidad Nacional de Colombia sísmicos, estratigráficos, petrográficos, geocronológicos y bioestratigráficos obtenidos
Medellín, Colombia
11 dunito8@hotmail.com a partir de afloramientos y pozos. La Cuenca Tumaco costa adentro tiene una geome-
Universidad de Caldas tría de artesa simétrica limitada al este por la cordillera Occidental y al oeste por el
Instituto de Investigaciones en Estratigrafía–
IIES Alto Estructural Remolino Grande‒Gorgona. Alrededor de 8000 m de sedimentos fueron
Grupo de Investigación en Estratigrafía y
Vulcanología–GIEV
acumulados en su depocentro durante el Cenozoico. Las sedimentitas corresponden a
Calle 65 n.° 26–10 lodolitas, arenitas y conglomerados, que varían en su proporción a través del tiempo,
Manizales, Colombia
University of Florida y fueron principalmente acumuladas en ambientes marinos abiertos y deltaicos. Los
Gainesville, USA
nanofósiles calcáreos, foraminíferos y palinomorfos permitieron controlar la edad de
12 hugo.murcia@ucaldas.edu.co
Universidad de Caldas acumulación de las unidades sedimentarias; sin embargo, debido a su baja abundancia
Departamento de Ciencias Geológicas y preservación y al retrabajamiento de microfósiles en algunos intervalos, se requiere el
Instituto de Investigaciones en Estratigrafía–
IIES uso de herramientas múltiples para determinar la edad de los depósitos.
Grupo de Investigación en Estratigrafía y
Vulcanología–GIEV
Las arenitas son principalmente litoarenitas y litoarenitas feldespáticas, texturalmen-
Calle 65 n.° 26–10 te inmaduras y compuestas por fragmentos de chert, fragmentos volcánicos básicos
Manizales, Colombia
13 klichegiraldo@gmail.com
a intermedios y cristales, tales como feldespato (Na y K), piroxeno, anfíbol y biotita,
Universidad de Caldas los cuales pueden asociarse con las rocas volcánicas básicas‒intermedias, plutónicas
Instituto de Investigaciones en Estratigrafía–
IIES y sedimentarias del actual basamento de la cordillera Occidental. El análisis de proce-
Grupo de Investigación en Estratigrafía y dencia (circones detríticos y minerales densos) sugiere vulcanismo continuo desde el
Vulcanología–GIEV
Calle 65 n.° 26–10 Oligoceno tardío al Pleistoceno, con incremento en la actividad desde el Mioceno medio.
Manizales, Colombia
La presencia de bajos porcentajes de circones precenozoicos y fragmentos de rocas me-
14 sergio.celis@ucaldas.edu.co
Universidad de Caldas tamórficas en las unidades del Mioceno puede estar relacionada con el retrabajamiento
Instituto de Investigaciones en Estratigrafía–
IIES
de unidades sedimentarias antiguas o con una conexión parcial con el basamento de la
Grupo de Investigación en Estratigrafía y cordillera Central. El estudio de afloramientos del Mioceno‒Plioceno y núcleos de pozo
Vulcanología–GIEV
Calle 65 n.° 26–10 permite interpretar una somerización de la cuenca durante el Mesiniano‒Zancliano.
Manizales, Colombia Abanicos volcanoclásticos, así como sedimentos fluviales y costeros, asociados a los ríos
15 seadlois@gmail.com
Agencia Nacional de Hidrocarburos Patía y Mira se encuentran cubriendo parcialmente los depósitos miocenos‒pliocenos.
* Corresponding author Palabras clave: Cuenca Tumaco, procedencia sedimentaria, bioestratigrafía, Pacífico
colombiano, Cenozoico.

1. Introduction tary deposits of the basin are below the surface. Therefore, the
spatial and temporal relationships of the stratigraphic units are
The Tumaco Basin is part of a series of forearc basins located not well known. The presence of hydrocarbon shows within
in the Pacific margin of the northern Andes (SW Colombia), wells and outcrops indicates the existence of an active petro-
where the Nazca and South American Plates have interacted at leum system (Barrero et al., 2007). Thus, oil exploration studies
least since the Miocene (Aleman & Ramos, 2000; Barrero et al., are the major source of information for the geology of the basin
2007; Borrero et al., 2012; Hall & Wood, 1985; López–Ramos, (e.g., López–Ramos, 2009; Suárez, 1990, 2007). Nevertheless,
2009; Ramos, 1999). The basin axis has an ~N30°E orientation, much of this information remains private and unpublished.
which is parallel to the current arc system and the Colombian The basement of the Tumaco Basin is composed of Cre-
Pacific subduction zone (Figure 1). Tumaco Basin is divided by taceous basic igneous and sedimentary rocks, formed in the
the Remolino Grande–Gorgona Structural High into onshore eastern Pacific and accreted to the continental margin of South
and offshore basins (Figure 1), also called the Manglares Ba- America during the Late Cretaceous – Paleogene (Echeverri et
sin by some authors (e.g., López–Ramos, 2009; Marcaillou & al., 2015a; Spikings & Simpson, 2014; Villagómez et al., 2011).
Collot, 2008). This region is characterized by a low relief and It is uncomfortably covered by siliciclastic and volcanoclastic
dense rain forest that has prevented the performance of system- rocks deposited from Eocene to Pliocene times (Barrero et al.,
atic geological studies; thus, its geological history has remained 2007; Borrero et al., 2012; Echeverri, 2012; López–Ramos,
poorly understood. The available geological maps are mainly 2009; Marín–Cerón & Sierra, 2011; Suárez, 1990, 2007). These
based on geomorphology and scarce field and chronologic con- Cenozoic rocks can reach more than 8000 m in the depocenter
trol (e.g., Nivia et al., 2003). In addition, most of the sedimen- of the basin (Agencia Nacional de Hidrocarburos & Universi-

216
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
79° W

78° W

77° W

75° W

70° W
Caribbean Sea

4° N 10° N
Buenaventura Panamá

Neogene
Venezuela

?
Nazca Plate
Cali Pacific ?

Ocean
Colombia
5° N

Paleogene
?
na

?
o

Colombia
Hig org
G
h
tur e–

Gorgona 3° N
uc nd

Island
al
Str Gra
lino


mo

Ecuador
Re

Brasil
100
91–

L–1
10
–19

973
11

–87 Perú
1–

NT
TB

99

–19
92– 1
–1

200 km
Study area
TB

284
2 0
ore

1 NT–
1990
–287 Cretaceous 1 Remolinogrande–1
fsh

3 0
2° N
Of

4 5 Paleogene 2 Majagua–1
6
2 Neogene 3 Chagüi–1
Tumaco 3
Quaternary 4 Tumaco–1–ST–S
ore

4 Tumaco–1–ST–P
sh
On

Outcrops 1 Río Patía 1

Wells 2 Río Patía 2


Pasto
City, town. 3 Salahonda

1° N Volcanoes 4 Punta Cascajal


Pacific Subduction
5 La Chorrera–Curay
Zone
Seismic lines 6 El Morro

Figure 1. Location of the Tumaco Basin (yellow color) and studied localities mentioned in the text.

dad de Caldas, 2011a; Suárez, 1990, 2007). Neogene units are age model to be established and the provenance and paleoenvi-
characterized by thick fluvial and deltaic siliciclastic deposits, ronment for the sedimentary rocks of the basin to be illustrated.
which have been particularly considered for hydrocarbon ex-
plorations (Agencia Nacional de Hidrocarburos & Antek, 2013; 2. Stratigraphy of the Tumaco Basin
Cediel et al., 2009; Echeverri, 2012; García, 2012; López–Ra-
mos, 2009; Suárez, 1990, 2007). Different stratigraphic nomenclatures have been proposed for
Due to scarce information from the Pacific basins, in 2008, the sedimentary rocks of the Tumaco Basin (Figure 2). Region-
the Agencia Nacional de Hidrocarburos (ANH) initiated a cam- al studies of the Servicio Geológico Colombiano (Arango &
paign to drill wells in different sectors of these basins (e.g., Ponce, 1980) used the stratigraphic model established by van
ANH–Tumaco 1–ST–S and 1–ST–P wells and Buenaventura der Hammen (1958), which was based on previous descriptions
1–ST–P) to investigate their petroleum systems. New geological made by Oppenheim (1949). van der Hammen (1958) reported
information from wells and outcrops collected in recent multi- that the Guapi Formation (Pliocene) uncomfortably overlays
disciplinary studies carried out by the Universidad de Caldas and the Miocene Naya Formation (Figure 2). Later, using main-
the ANH (Agencia Nacional de Hidrocarburos & Universidad ly seismic and well information, Suárez (1990) proposed four
de Caldas, 2011a, 2011b) shed light on the geological history of lithostratigraphic units that have served as references for recent
the basin. A total of 1544 analyses were performed, including studies (e.g., Agencia Nacional de Hidrocarbuos & Universidad
biostratigraphy (palynology, foraminifera, and calcareous nan- de Caldas, 2011a; Borrero et al., 2012; Escovar et al., 1992).
nofossils), petrography, and geochronology (U/Pb, Ar/Ar, and K/ This author compared the sedimentary record of the Tumaco
Ar methods). The integration of these results along with unpub- Basin to the neighboring Borbón Basin in Ecuador, defining
lished reports and published data (Bedoya et al., 2013; Borrero several units from base to top: (i) 1 Sur Unit Formation (Oligo-
et al., 2012; Echeverri et al., 2015a, 2015b, 2016) allowed an cene); (ii) Cayapas, Viche, and Angostura Formations (lower

217
PARDO–TRUJILLO et al.
218

Agencia Nacional de Hidrocar-


Age van der Marcaillou & Collot López–Ramos buros & Universidad de Caldas Agencia Nacional de Hidrocar-
Oppenheim (1949) Suárez (1990, 2007) González (2008)
(Ma) Hammen (1958) (2008) (2009) (2001a); Echeverri et al. buros & Antek (2013)
(2015a, 2016)
0 Holocene Alluvial, deltaic, and Alluvion Alluvial and deltaic deposits
Q Pleistocene Recent deposits Guapi Formation tidal deposits and fluvio–volcanic fans
Pliocene
Conglomerates,
Guapi Formation Mayorquín Formation/ Raposo Pato Mayorquín Formation Cascajal Formation Cascajal Formation
sandstones, and claystones Formation
5
Shales, San Agustín Formation Guapi Formation Raposo Formation San Agustín Formation Tangareal del Mira Formation
U sandstones, Bojayá or Tumaco Chagüí Formation
10 and conglomerates Chagüí Formation Formation
Neogene

Angosturas Formation
Miocene

Naya

M Sandstones, Naya Formation


15 Viche Formation Naya Formation Angosturas Formation
mudrocks with Naya, Angosturas, Tumaco Formation
Viche Formation
limestone and Upper Cayapas Formation and Viche Formations
Bojayá or Tumaco Cayapas Formation
calcareous (Unit 3)
20 L Formation
concretions Lower Cayapas Formation
(Unit 3)

25 Cayapas Formation ?
U
Oligocene

1S Unit/1N Unit Uva of Buenaventura


30 Pacífico Group
L Pacífico Group Pacífico Group
Shales, Lutitas de Remolino Grande
35
U sandstones, Formation or 1 Sur Unit
and conglomerates
40
Paleogene
Eocene

M
45 Suruco? Timbiquí/Macuchi
Thickness determined from Formation
seismic information, but
50
? lithology is not know
? (not drilled)
L
55
U
Diabásico Group ?
Paleocene

60 M

Diabásico Group
L
65
Cretaceous

Igneous and Remolinogrande


Upper

70 metamorphic Diabásico Group Diabásico Group Diabásico Group Diabásico Group Naranjal/Piñón Volcanosedimentary
basement Sequence

Figure 2. Chronostratigraphic chart showing the different stratigraphic nomenclatures employed for the Tumaco Basin.
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
– middle Miocene); (iii) Chagüí and San Agustín Formations the main seismic reflectors and structures were identified (e.g.,
(middle/upper Miocene – lower Pliocene); and (iv) Guapi For- faults, folds, onlaps, toplaps, progradations). These reflectors
mation (Pliocene). Subsequently, Nivia et al. (2003) separated were tied with the stratigraphic information of the Remolino
the youngest alluvial and deltaic deposits from the Guapi For- Grande–1, Majagua–1, and Tumaco 1–ST–P wells. Addition-

Neogene
mation, defining as Cascajal and Curay Members those under- ally, as far as possible, the reflectors were tied with the surface
lying Pliocene and Miocene sedimentary deposits in the Nariño geology map of Gómez et al. (2015). Subsequently, five hori-
Department. Echeverri et al. (2016) proposed the term Cascajal zons that represent chronostratigraphic boundaries were iden-
Formation for the volcanoclastic rocks of the Messinian – Zan- tified: (i) Cretaceous/Paleogene, (ii) Paleogene/lower Miocene,

Paleogene
clean age reported by Nivia et al. (2003) (Figure 2). In contrast, (iii) lower/middle Miocene, (iv) middle/upper Miocene, and
in the northern part of the Tumaco Basin, Aspden (1984) and (v) upper Miocene – Pliocene. Lower Miocene includes the
Aspden & Nivia (1985) proposed to divide the Neogene sedi- Aquitanian and Burdigalian Stages; the middle Miocene, the
mentary cover into two formations: (i) Mayorquín Formation, Langian – Serravalian; and the upper Miocene, the Tortonian
composed mainly of mudrocks, and (ii) Raposo Formation, with – Messinian.
conglomerates and sandstones. The stratigraphic and chronos-
tratigraphic relationships of these units are unclear due to the 3.2. Biostratigraphy
low–biostratigraphic resolution and the absence of continuous
outcrops. This has caused difficulties in performing regional Biostratigraphic interpretations are based on distribution pat-
correlations between southern and northern units (González, terns and semiquantitative analyses of calcareous nannofossils
2008; Instituto Colombiano de Geología y Minería & Instituto and planktonic foraminifera reported for Majagua–1, Remo-
Geográfico Agustín Codazzi, 2005; López–Ramos, 2009; Ni- lino Grande–1, Tumaco 1–ST–S, and Tumaco 1–ST–P wells
via et al., 2003). Although Borrero et al. (2012) followed the (Agencia Nacional de Hidrocarburos & Antek, 2013; Agencia
stratigraphic nomenclature proposed by Suárez (1990), they Nacional de Hidrocarburos & Universidad de Caldas, 2011a;
grouped the sedimentary record into two megasequences, the Robertson Research inc. & Empresa Colombiana de Petróleos,
Oligocene – middle Miocene sedimentary megasequence and 1981a, 1981b). The quality of the microfossil record was eval-
the late Miocene – Holocene volcaniclastic megasequence. uated, and a composite section was built in which assemblages
Finally, Agencia Nacional de Hidrocarburos & Antek (2013) and age–indicative taxa were indicated. Biostratigraphic mod-
informally proposed the names Tumaco and Tangareal del Mira els were constructed using the standard biozones proposed by
Formations for upper Oligocene and Neogene deltaic deposits Blow (1969) and Berggren et al. (1995) for planktonic fora-
associated with the paleo–Patía and Mira Rivers (Figure 2). minifera and by Martini (1971) and Okada & Bukry (1980)
In this chapter, we will use the nomenclature of Borrero et al. for calcareous nannofossils. The biochronology of calcareous
(2012), Echeverri et al. (2011), and Echeverri et al. (2016). microfossil events follows the integrated scales of Agnini et al.
(2014), Backman et al. (2012), Lourens et al. (2004), Raffi et
3. Materials and Methods al. (2006), and Wade et al. (2011). Taxonomical and biostrati-
graphic information described by Aubry (2014, 2015), Kennett
3.1. Stratigraphy and Seismic Data & Srinivasan (1983), and Perch–Nielsen (1985) were also con-
sidered in the biostratigraphic interpretations. Biochronology
The stratigraphic information was obtained from electrical logs, of calcareous microfossil events follows the integrated scales
ditch cutting, and core samples from the Remolino Grande–1 of Agnini et al. (2014), Backman et al. (2012), Lourens et al.
(total depth (TD) 9080’–2767 m), Majagua–1 (TD 14 280’– (2004), Raffi et al. (2006), and Wade et al. (2011). Taxonomi-
4352 m), and Tumaco 1–ST–S (TD 1899.6’–579 m) wells. This cal and biostratigraphic information described by Aubry (2014,
information was also integrated with data from outcrops located 2015), Kennett & Srinivasan (1983), and Perch–Nielsen (1985)
in the Tumaco Bay (e.g., Echeverri et al., 2016), with previous were also considered in the biostratigraphic models.
results reported by Suárez (1990) and López–Ramos (2009) and Most micropaleontological results were obtained from cut-
data from the ANH Tumaco 1–ST–P well (Agencia Nacional ting samples, which in some cases (e.g., Remolino Grande–1)
de Hidrocarburos & Antek, 2013). The well cores and cuttings were collected at very large spacing, making its biostratigraphic
are stored at the national core repository of Colombia (Litoteca interpretation difficult. Nevertheless, by integrating both cal-
Nacional, Piedecuesta, Santander). careous nannofossils and planktonic foraminifera, it was pos-
Two regional seismic profiles were analyzed (Figure 3): (i) sible to constrain the age from most of the wells. To achieve a
Perpendicular to the basin (dip lines) and composed of three more accurate age model, some considerations were taken in
seismic lines and (ii) more or less parallel to the strike of the account. (i) Since the examined wells are composed of cuttings
basin (strike lines) and composed of two seismic lines (Fig- and cores that showed very abundant reworked microfossils,
ures 1, 3). These data were processed with GeoGraphix, where biostratigraphic analyses did not include specimens whose oc-

219
PARDO–TRUJILLO et al.
220

a
Seismic line: Pacífico–73_L–1973–87_2001 Seismic line: Tumaco–92_NT–1992–2840_1992 Seismic line: Tumaco–90_NT–1990–2870_1990

m Remolino Grande–1 well projection Seismic line: Tumaco–91_TB–1991–1110_2002

–2500
No seismic
–5000 data
Tumaco offshore basin Coast line Remolino Grande Structural High Tumaco onshore basin
0

–2500

–5000

0 50 100 km

b
Seismic line: Tumaco–91_TB–1991–1110_2002
ANH–Tumaco–1–ST–P well Majagua–1 well Seismic line: Tumaco–90_NT–1990–2870_1990
m
0

–2500

–5000

–2500 Mud
diapir

–5000

0 50 100 150 km

Quaternary Pliocene Middle – upper Miocene Middle – upper Miocene Lower Miocene Paleogene
boundary
Cretaceous Fault Interpreted Well Coast line Onlaps, downlaps, and toplaps.

Figure 3. Geological cross–section of Tumaco, Western Cordillera, and Patía Basins based on seismic lines geologic mapping and well data. (a) Dip line. (b) Strike line. See Figure
1 for locations.
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
currence was very discontinuous or was observed in only one (2009). This age was acquired from the average of at least three
sample. (ii) Acmes or paracmes, which are regularly used as grains of zircons with concordant ages (Kochelek et al., 2011).
bioevents in studies of deep–sea sections, were not considered These results were plotted in relative probability diagrams us-
in our results. The fine–grained matrix regularly recorded in the ing Isoplot 3.00 (Ludwing, 2003). 206Pb/238U ages were used

Neogene
material provokes a dilution, preventing the quantitative signal for grains less than 1000 Ma, and 206Pb/207Pb ages were used
(Bedoya et al., 2013). (iii) Those stratigraphic intervals lacking for grains greater than 1000 Ma. Grains with more than 10%
biostratigraphic markers but bracketed by confirmed biozones of discordance were not considered in the statistical analysis.
were illustrated as intervals of indeterminate age because there

Paleogene
was no micropaleontological support. 4. Results
3.3. Sediment Provenance Analyses The seismic lines show that the Tumaco onshore basin has
a symmetric through shape limited to the E by the Western
A total of 30 grams of dried well cuttings was collected for pe- Cordillera and to the W by the Remolino Grande–Gorgona
trographic analyses. This represents a 10–feet interval (3048 m) Structural High. Subhorizontal seismic reflectors indicate lit-
for the Majagua–1 well and intervals of 10’, 20’, and 30’ for the tle deformation of the sedimentary sequence. However, in the
Remolino Grande–1 well. Wet cutting samples were washed to southern sector of the basin, there is a structure at least 10 km
remove contamination from the drilling process. Subsequently, wide that truncates the Paleogene and Neogene strata that has
a selection of very–fine sand to coarse–sand particles from cut- been interpreted as a mud diapir (Figure 3b; Cediel et al., 2009).
tings was collected and mounted on glass slides and cover slips In the depocenter, a thickness of 6805’ (2074 m) is estimated
using balsam of Canada for petrographic analyses. A total of 82 for the Paleogene beds; 5713’ (2546 m), for the lower – middle
thin sections were studied for petrography using ditch cutting Miocene; and 11986 (3653.3 m), for the upper Miocene – Plio-
samples: 26 from the well Remolino Grande–1, 38 from the cene (Figure 3a, 3b). Calcareous microfossils are moderately
well Majagua–1, and 18 from the well Chagüí–1. On average, to poorly preserved. Even though the sedimentary record was
250 and 300 grains per slide were identified and counted. These highly discontinuous, Majagua–1, Remolino Grande–1, and
data are represented in bar diagrams showing the relative per- Tumaco 1–ST–P wells covered the largest stratigraphic record
centages of the main components. Subsequently, the sandstones (Figures 4, 5). Calcareous microfossils suggest ages spanning
with a framework higher than 50% were selected to count their from Bartonian (middle Eocene) in the Remolino Grande–1
individual particles and to quantify the composition and are well up to Pleistocene in the Tumaco 1–ST–P samples (Figures
represented in Qt–F–L and Qm–F–Lt triangles of Folk (1974). 4, 5). Miocene is the most common interval drilled by the wells,
For heavy minerals and U/Pb geochronologic analyses, allowing the definition and correlation of different biostrati-
samples of ca. 1000 gr of dried cuttings were collected along graphic intervals along the basin (Figures 4, 5). According to
tens to hundreds of feet and integrated in a single sample. The the biostratigraphic data in wells and seismic lines, the Tumaco
fraction > 63 < 250 µm was concentrated with sodium poly- Basin was divided into six chronostratigraphic sequences (Fig-
tungstate (density of 2.89 g/cm³) for heavy mineral analyses. ure 3): (i) Upper Cretaceous, (ii) Paleogene (including Eocene
In each sample, a thin section was made using a resin with a and Oligocene beds), (iii) lower Miocene, (iv) middle Miocene,
refraction index similar to balsam of Canada (1.539). A count (v) upper Miocene – Pliocene, and (vi) Pliocene – Quaternary.
of ca. 400 grains per slide was performed following the Rib-
bon–Counting method (Mange & Maurer, 1992). Mineralogical 4.1. Upper Cretaceous
identification was based on Mange & Maurer (1992). The data
obtained are represented in bar and triangular diagrams. This series includes the volcano–sedimentary succession of Re-
U/Pb geochronology was performed using the laser ablation molino Grande (Turonian – Campanian) (5630–9080’; 1716–
inductively coupled plasma mass spectrometry (LA–ICPMS) 2767 m) (Figure 3). This succession may be divided into two
method at a Washington State University Lab (USA). The segments: (i) the lower segment, composed of mudrocks and fine
samples were crushed and sieved (ca. 400–μm mesh). Zircon sandstones interbedded with basalts and microgabbros, and (ii)
concentration was initially made following the gravitational the upper segment, formed by layers of basalts and microgabbros.
methods on a Wifley table, passed through a Frantz magnetic Robertson Research inc. & Empresa Colombiana de Petróleos
separator and finally separated using methylene iodide. Ap- (1981b) provided a radiometric age in the interval 7250–7260’
proximately 100 grains per sample were analyzed in agreement (2209–2212 m) of 82.2 ± 8.1 Ma (K/Ar in total rock). New Ar/
with Gehrels et al. (2006). Separated zircons were mounted Ar geochronological data (Agencia Nacional de Hidrocarburos
on an epoxy resin, manually polished and randomly studied. & Universidad de Caldas, 2011a; Echeverri et al., 2015a) yield
The maximum depositional U/Pb ages in detrital zircons were ages of 82.9 ± 4.8 Ma (6540–6550’ or 1993–1996 m) and 76.2
performed using the method described by Dickinson & Gehrels ± 1.4 Ma (5680–5690’, or 1731–1734 m) (Figure 4). These ages

221
PARDO–TRUJILLO et al.

Remolino Grande–1 Majagua–1 well Tumaco 1–ST–P well


well
Datum: sea level

1000' 1000'
1000'

2000' 2000'
2000'

Middle
Miocene

3000' 3000'
3000'

Upper Miocene
Lower Miocene
Pliocene
4000' SW 4000'
4000'

Oligocene
5000' 5000'
5000'
Eocene

76.2 ± 1.4 Ma

6000' 6000'
6000'
Upper Miocene
82.9 ± 4.8 Ma

7000' 7000'
7000'

8000'
Upper 8000'
8000'
Cretaceous

9000'
9000'
9000'
M S G

10 000'
10 000'

11 000'
Conglomerates 11 000'

Sandstones Upper Oligocene

Mudrocks
Basalts–microgabbros 12 000' 12 000'
vf f m c vc
M: mud M S G

S: sand
vf: very fine 13 000'

f: fine
m: medium ?

c: coarse 14 000'

vc: very coarse


G: gravel M S G

Figure 4. Chronostratigraphic correlation among the Remolino Grande–1, Majagua–1, and Tumaco 1–ST–P wells. Based on Agencia Na-
cional de Hidrocarburos & Universidad de Caldas (2011a) and Agencia Nacional de Hidrocarburos & Antek (2013).

222
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
Planktonic Calcareous
Age
Period

Epoch
(Ma) Age (Stage)
Remolino Grande–1 Majagua–1 Tumaco 1–ST–P, 1–ST–S foraminifera nannofossils
biozones biozones
0 0.126 Upper NN21 CN15 0
b NN20
Quaternary

0.78 Middle CN14


Pleistocene

1 N22 Pt1 1
Calabrian a NN19
1.80
CN13

Neogene
2 PL6 NN18 2
Gelasian NN17
2.58 PL5 Atl.
N20/ CN12 3
3 Piacenzian NN16
3.60 N21
?
Pliocene

PL3 NN15
4 ≤ 4.1 Ma MP006 (400') PL2 NN14 CN11 4
Zanclean
N18/ PL1 NN13 5
5 5.33 ca. 6000' CN10
N19 Atl. NN12

Paleogene
6 Tumaco1 (1257') M14 6
Messinian N17b
≤ 6.6 Ma M1MP–001 (110–800') ≤ 6.5 Ma FAD of C. annulatus Alt.
7 7.25 (3950') NN11a CN9 7
≤ 7.5 Ma M1MP–002 (2830–3210') ca. 3950' N17
b

8 ≤ 7.6 Ma M1MP–003 (3430–3950') M13 8


NN11
? ca. 4350'
9 NN10 CN8 9
≤ 9.3 Ma N16 a
Tortonian M1MP–009 (9100–9320')
10 ca. 8820' 10
N15 M12 NN9 CN7
≤ 10.4 Ma RGMP–004 (1210–1450') 5563–5382'
ca. 9700' NN8 CN6
11 ca. 9700'
(Giraldo–Villegas et al., 2016) N14 M11 11
11.63 ≤ 11.5 Ma RGMP–001 (84–300') ca. 10 000' NN7
ca. 7000' N13 M10
ca. 10 050'
12 ca. 7061.5' 12
CN5
Neogene

ca. 1990' b
Serravallian N12 M9 NN6
RGMP–006 (2140–2380') 13
13 ≤ 13.1 Ma a
13.82 ≤ 13.3 Ma N11 M8
RGMP–007 (2640–2760') ca. 10 290' ca. 7250'
Miocene

14 N10 M7 14
NN5 CN4
ca. 7493' N9 M6
15 Langhian 15
b
15.97 N8 M5
16 ca. 3200' ca. 11 400' ca. 8480' 16
≤ 16.1 Ma M1MP–0014 a
ca. 8490' b NN4 CN3
(11 660–11 900')
17 N7 M4 a 17
ca. 3490' ca. 9090' N6
ca. 11 990'
18 NN3 18
Burdigalian ca. 11 990' M3 CN2
19 ca. 12 840' 19
N5
ca. 13 000'
20 20
20.44 M2
ca. 9660' NN2
21 CN1 21
Aquitanian b
22 M1 22
N4
ca. 4500' ca. 13 550' ca. 10 740' a
23.03 NN1
23 23
ca. 4570'
24 O7 24
ca. 4770'
25 NP25 25
ca. 11 070' P22
Chattian
26 O6 26
ca. 4980' CP19
27 27
27.82
O5
28
Oligocene

28 P21 NP24
O4
29 29
P20 O3 CP18
30 ca. 5680' 30

Rupelian NP23 CP17


31 31
P19 O2
Paleogene

32 32
NP22
33 P18 O1 33
Biozones supported by CP16
33.9
ca. 5695' NP21
34 calcareous microfossils 34
E16
ca. 5710' P16/
35 Palynologic age control P17 35
E15 NP
36 Priabonian 19–20 36
Not enough calcareous CP15
37 microfossils information 37
P15 E14 NP18
Eocene

37.8
38
Stratigraphic range of the 38
youngest detrital zircon population
39 NP17 39
E13
Bartonian Unconformity P14
CP14
40 40
P13 E12
ca. 5740'
41 41.2 41

Figure 5. Ages of the Remolino Grande–1, Majagua–1, Tumaco 1–ST–P, and 1–ST–S wells based on calcareous microfossils, palynology,
and detrital zircons. Geologic time scale based on Cohen et al. (2013; updated 2017/02).

223
PARDO–TRUJILLO et al.

can be correlated with the 98–60 Ma ages obtained from Gorgona nannofossils such as Cyclicargolithus abisectus, Sphenolithus
Island by Serrano et al. (2011) and from the north of Ecuador ciperoensis, and Sphenolithus distentus were recorded in some
by Vallejo et al. (2009). These rocks have been correlated by samples. Oligocene/Miocene boundary markers were not found
other authors with the Diabásico Group (sensu Nelson, 1962), at the Remolino Grande–1 well (Figure 5). Nevertheless, plank-
the Volcánica Formation (sensu Aspden, 1984), and the Dagua/ tonic foraminifera bioevents such as the last occurrence of the
Piñón Formation (López–Ramos, 2009), located in the Western Paragloborotalia opima at 27.30 Ma in the biozone O6 and the
Cordillera of Colombia and Ecuador (Figure 2). first occurrence of Paragloborotalia kugleri at 23.73 Ma in the
biozone M1a according to Wade et al. (2011) support the Pa-
4.2. Paleogene leogene/Neogene boundary near 10 740’ at the Tumaco 1–ST–P
well (Figure 5). This interpretation indicates that Paleogene cal-
4.2.1. Seismic Lines careous nannofossils reported in shallower levels correspond to
reworked species. Due to the lack of diagnostic microfossils, this
The Paleogene deposits were accumulated over a relative- boundary was unclear in the Remolino Grande–1 well; therefore,
ly smooth surface of the igneous Cretaceous basement. They it was placed at 4500’ after seismic correlation with the Tumaco
onlap onto this basement to the east (Figure 3a). Paleogene 1–ST–P well (Figures 3, 4, 5). Even though Paleocene deposits
deposits outcrop on the western border of the Western Cordil- were not found in the Remolino Grande–1 well, lower Paleocene
lera (Figure 3a; Gómez et al., 2015). To the west, Paleogene deposits have been recently reported in outcrops of the Gorgona
reflectors can be identified over the Remolino Grande–Gorgona Island 85 km to the NE (Bermúdez et al., 2016). In Figure 5, the
Structural High, where they become thinner and affected by correlation of the lower part of the Remolino Grande–1 and Tu-
normal faults. In the Tumaco offshore basin, Paleogene reflec- maco 1–ST–P wells with the calcareous nannofossils and plank-
tors increase in thickness and are affected by thrust faults (Fig- tonic foraminifera standard zones is shown. This interval ranges
ure 3a). Paleogene deposits become thickened to the south of between NP19–NP25 and P16–P22, respectively.
the basin, and their reflectors locally downlap the Cretaceous
basement (Figure 3b). 4.2.4. Petrography

4.2.2. Lithology In the Remolino Grande–1 well, two ditch cutting samples were
used for petrographic analyses. They have abundant grains of
In the Remolino Grande–1 well, the Paleogene is mainly consti- sedimentary rocks (27–80 %, siltstones, sandy siltstones, and
tuted by mudrocks and sandy siltstones interbedded with some chert; Figure 6). Plagioclase, mono– and polycrystalline quartz,
sandstone beds (Figure 4). This unit can be correlated with the pyroxene, biotite, hornblende, glauconite, and opaque minerals
Lutitas de Remolino Grande Formation or 1 Sur Unit (Echeve- are present in lower proportions. Some algae and foraminifera
rri et al., 2015a; Suárez 1990, 2007), as well as the lower part of are observed. Sandstone grains were selected to know their
the Tumaco Formation of Agencia Nacional de Hidrocarburos framework composition. They were classified as poorly sorted
& Antek (2013) (Figure 2). The Tumaco 1–ST–P well reached subarkoses, with fine to coarse angular to subrounded grains.
mudrocks, sublitharenites, and litharenites, interpreted as late They are composed of monocrystalline (32–44 %) and poly-
Oligocene (Chattian) shelf–prodelta to delta front deposits. crystalline (21–30 %) quartz, plagioclase (10–23 %), interme-
Benthic foraminifera such as Osangularia sp., Gyroidinoides diate volcanic rocks (5–6 %), chert (4–15 %), biotite (1–2 %),
broeckhiana, Discorbinella sp., Anomalinoides cicatricose, muscovite (1%), and chlorite (3%). Mud matrix and calcareous
and Anomalinoides semicribata suggest sedimentation in the cement are common.
upper–middle part of the continental slope (Agencia Nacional
de Hidrocarburos & Antek, 2013). 4.2.5. Heavy Minerals

4.2.3. Biostratigraphy The analyzed samples present higher abundances of unstable


compared to stable phases. Stable phases include zircon (2–7
The lowest sample with calcareous microfossil recovery in %) and tourmaline (1%). Unstable phases correspond to apatite
Remolino Grande–1 well yielded Reticulofenestra bisecta, Re- (4–25 %), clinopyroxene (10–52 %), orthopyroxene (1–12 %),
ticulofenestra reticulata, and Paleocene and Eocene reworked hornblende (1–13 %), and minerals of the epidote group (25–40
species (Sphenolithus anarrhopus and Sphenolithus primus, %). Olivine, oxyhornblende, and biotite are present at 1% each.
Nannotetrina spp.). The young assemblage suggests the biozones In the Tumaco 1–ST–P well, the samples contain mainly unsta-
NP16 and NP19 (Bartonian – Priabonian; Agnini et al., 2014; ble phases with abundant amphibole, pyroxene and, in lower
Perch–Nielsen, 1985). For the Oligocene, diagnostic species of proportions, epidote, apatite, clinopyroxene, and orthopyrox-
planktonic foraminifera were not identified; however, calcareous ene. In this interval, detrital zircons were not obtained.

224
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
Petrography Heavy minerals

Depth (ft)
K feldspar Augite

Lithology
Volcanic fragments Zircon Hornblende

Samples
Chert fragments
Age Chlorite Apatite Hyperstene
Plagioclase Metamorphic fragments Allochemicals
Oxyhornblende Epidote Tourmaline

Neogene
Plutonic fragments Sedimentary fragments Glauconite Olivine Biotite Actinolite Zoisite/Clino.

01 01
01
02
Upper Miocene

02 03
04

Paleogene
03 05
Onset erosion of granitoids
1000 06

04 04
07
08
05 09
10
Middle Miocene

2000 11
06 06 12
13
14
15
07 07 16
17
18
3000 19
08
20
Lower Miocene

21
09 22

4000 23

10
Oligoocene

24
Eocene –

5000

5400 12 25
0% 20% 40% 60% 80% 100% 0% 20% 40% 60% 80% 100%
Upper
Cretaceous 9000

Conglomerates Heavy minerals

Sandstones U/Pb

Mudrocks Petrography

Figure 6. Stratigraphic log of the Remolino Grande–1 well, and bar diagrams representing the composition of the ditch cuttings and
heavy minerals. Allochemicals include calcareous fossil remains (mainly foraminifera, bivalves, and gastropods).

4.3. Lower Miocene (Aquitanian – Burdigalian) reflectors and seem to be in toplap against the Pliocene reflec-
tors in the westernmost part of the seismic line (Figure 3a). In
This interval has been identified in the Remolino Grande–1 the Remolino Grande–Gorgona Structural High, the lower Mio-
(3200–4500’), Majagua–1 (14280–11220), and Tumaco 1– cene beds seem to fill some valleys dissected in the Paleogene
ST–P (8480–10740’) wells. deposits. In the strike line, the lower Miocene reflectors become
thinner to the north, but they are concordant with the Paleogene
4.3.1. Seismic Lines and middle Miocene sequences.

The lower Miocene reflectors onlap onto the Paleogene deposits 4.3.2. Lithology
towards the E of the basin (Figure 3a). Westward, they become
thinner and pinch–out in the Remolino Grande–Gorgona Struc- In the Remolino Grande–1 well, there is a dominance of clay-
tural High. In the offshore Tumaco Basin, the lower Miocene stones and sandy siltstones interbedded with some sandstone
reflectors are concordant with Paleogene and middle Miocene beds, which change abruptly to conglomerates and sandstones

225
PARDO–TRUJILLO et al.

in the upper part of the succession (Figure 4). This interval % polycrystalline) and lithic fragments (1–39 %). Lithic frag-
has abundant fossils of bivalves, gastropods, foraminifera, ments are mainly volcanic of intermediate composition (6–39
and carbonized organic matter. In the Majagua–1 well, lower %), chert (2–15 %), and low proportions of plutonic rocks,
Miocene deposits are composed of mudrocks interbedded with sandstones, micaceous and graphitic schists (≤ 1%) (Figures 7,
fine–medium sandstones and sandy mudstones. Foraminifera 9). The accessory minerals are composed of hornblende (5%),
and mollusk shells are abundant. In the Tumaco 1–ST–P, the epidote (2–3 %), glauconite (1%), biotite (1–3 %), chlorite
rocks of this period are characterized by an alternation of highly (2–6 %), and muscovite (< 1%). In the Tumaco 1–ST–P well,
bioturbated mudrocks, massive and laminated sublitharenites, the sandstones include litharenites and lithic greywackes, with
feldspathic litharenites, litharenites, and some beds of polymic- angular fragments of sedimentary, volcanic, and metamorphic
tic conglomerates. Paleontological content includes shark teeth rocks (Agencia Nacional de Hidrocarburos & Antek, 2013).
(Agencia Nacional de Hidrocarburos & Antek, 2013).
4.3.5. Heavy Minerals
4.3.3. Biostratigraphy
Two samples from the Remolino Grande–1 well and four from
Calcareous nannofossil events are the most efficient biostrati- the Majagua–1 well were analyzed (Figures 6, 7, 10). In the
graphic indicator at the Tumaco 1–ST–P and Majagua–1 wells Remolino Grande–1 well, stable phases are almost absent,
in the lower Miocene deposits. The appearances of Discoaster with only 1% zircon. Unstable phases include apatite (47%),
druggi, Heliscoaphaera ampliaperta, and Sphenolithus hetero- clinopyroxene (4–10 %), hornblende (2–5 %), and minerals of
morphus and the extinction of Triquetrorhabdulus carinatus in the epidote group (36–74 %). Oxyhornblende, actinolite, and
these wells mark the beginning of the Miocene (Aubry, 2014, chlorite are present in low proportions (≤ 1%). A higher pro-
2015; Backman et al., 2012; Perch–Nielsen, 1985). This is in portion of unstable with respect to the ultrastable phases was
agreement with the recovery of the short–lived Sphenolithus found in the Majagua–1 well (Figures 7, 10). Ultrastable phases
belemnos in the Majagua–1 well. For this interval, the Remoli- include zircon (4–26 %) and traces of tourmaline (≤ 1%). Un-
no Grande–1 well is characterized by the last occurrence of H. stable phases include apatite (51–75 %), minerals of the epi-
ampliaperta and S. heteromorphus and the occurrence of Globi- dote group (1–12 %), biotite (2–4 %), clinopyroxene (2–10
gerinoides diminutus, whose biostratigraphic range is described %), ortopyroxene (≤ 1%), chlorite (≤ 2%), hornblende (≤ 2%),
for the late Burdigalian – Langhian (Aubry, 2014; Backman et and traces of oxyhornblende and glauconite (Figures 7, 10).
al., 2012; Kennett & Srinivasan, 1983). This indicates that Re- In the Tumaco 1–ST–P well, a greater proportion of unstable
molino Grande–1 encompasses a long–lasting gap (ca. 5 Ma), minerals compared to stable ones was also observed (Agencia
which includes the Aquitanian and the lower–middle part of the Nacional de Hidrocarburos & Antek, 2013). Ultrastable phases
Burdigalian. In Figure 5, the stratigraphic equivalence of the include zircon and tourmaline (≤ 1%). Unstable phases include
Remolino Grande–1, Majagua–1, and Tumaco 1–ST–P wells hornblende (31–56 %), minerals of the epidote group (1–41
with the calcareous nannofossils and planktonic foraminifera %), apatite (< 1–3 %), clinopyroxene (6–28 %), biotite (≤ 3%),
standard zones is represented. This portion represents an inter- chlorite (1–2 %), sphene (≤ 1%), and garnet (< 1–4 %).
val between NN2 to NN6 and N4 to N12, respectively.
4.3.6. Detrital Geochronology
4.3.4. Petrography
In the Majagua–1 well, the sample M1MP–014 (depth 11660–
The ditch cutting samples from the Majagua–1 and Remoli- 11900’) was analyzed (Figure 11). A maximum depositional age
no Grande–1 wells are composed mainly of sedimentary rocks of 16.1 Ma (Burdigalian) was found. This sample shows four
(56–90 %; mudrocks, cherts, calcite, sandy shales, sandstones) main zircon populations: (i) 18.9 Ma, (ii) 45.0 Ma, (iii) 51.3 Ma,
and, in lower proportions, plagioclase, volcanic and plutonic and (iv) 20.7 Ma. Oligocene (28.6 Ma), Late Cretaceous (68.6
lithics. Fragments of quartz (mono– and polycrystalline), bio- Ma, 74.4 Ma, 74.9 Ma, 74.9 Ma, 76.9 Ma, and 83.2 Ma), and
tite, hornblende, pyroxene, glauconite, and bioclasts (foramin- Precambrian (556.7 Ma, 557.4 Ma, 844.4 Ma, and 1657.0 Ma)
ifera, algae, echinoderms) occur in lower proportions. zircons are present, but they are not abundant enough to form a
The sandstone grains of the Majagua–1 (Figure 7) and Re- population.
molino Grande–1 wells were selected to study the composition
of their framework. They are classified as feldspathic lithar- 4.4. Middle Miocene (Langhian – Serravallian)
enites and arkoses (Figures 8, 9). The sandstones are poorly
sorted, fine to medium size, with rounded to subangular grains. Rocks of this period were identified in the Remolino Grande–1
They have a high content of plagioclase (34–68 %), with low- (1450–3200’), Majagua–1 (10500–11400’), and Tumaco 1–
er proportions of quartz (3–32 % monocrystalline and 2–21 ST–P (7000–8480’) wells.

226
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
Petrography Heavy minerals
Depth (ft) K feldspar Augite

Lithology
Volcanic fragments Zircon Hornblende

Samples
Chert fragments
Age Chlorite Apatite Hyperstene
Plagioclase Metamorphic fragments Allochemicals

Neogene
Oxyhornblende Epidote Tourmaline
Plutonic fragments Sedimentary fragments Glauconite Olivine Biotite Actinolite Zoisite/Clino.

01
01 01
02

Paleogene
03
1000

04
2000
05
06

3000 02 02 07

08

03 03 09
4000
11
Uppeer Miocene

04 12
5000 13

14
15
6000
05 16

17

06 18
7000
Onset of erosion of granitoids
19

07 20
8000 21
22
23
9000 24
09 09
25
26
10
10 000 27
11
28
Miocene
Middle

12 29
11 000
30
13
31

12 000 14 14
Lower Miocene

33
16 34
13 000
35
36
18 37
38
14 000
19 39
40
0% 20% 40% 60% 80% 100% 0% 20% 40% 60% 80% 100%

Conglomerates Heavy minerals

Sandstones U/Pb

Mudrocks Petrography

Figure 7. Stratigraphic log of the Majagua–1 well, and bar diagrams representing the composition of the dish cuttings and heavy minerals.
Allochemicals include calcareous fossil remains (mainly foraminifera, bivalves, and gastropods).

227
PARDO–TRUJILLO et al.

Q
1. Quartzarenite Remolino Grande–1
2. Subarkose 1
Tumaco–1–ST–S
3. Sublitharenite
Tumaco–1–ST–P
2 3 Majagua–1

Lithic Feldspathic
Arkose Litharenite
arkose litharenite

F L

Figure 8. Classification of sandstones (Folk, 1974) in cores from the Tumaco 1–ST–S and Tumaco 1–ST–P wells, and sandstone fragments
of the ditch cuttings from the Majagua–1, Remolino Grande–1, and Tumaco 1–ST–P wells. Based on Cortés et al. (2019), and Agencia
Nacional de Hidrocarburos & Universidad de Caldas (2011a, 2011b).

4.4.1. Seismic Lines litharenites, litharenites, and conglomerates. Fossils of scapho-


pods, pelecypods, ambulacres, gastropods, and echinoderms are
The middle Miocene reflectors onlap onto the lower Miocene present in this interval. This succession is interpreted as accu-
deposits towards the E of the basin (Figure 3a). They became mulated in different environments of a deltaic system (prodelta,
thinner in the Remolino Grande–Gorgona Structural High to estuarine bars, river mouth bars, and lagoons; Agencia Nacional
the west. In the Tumaco offshore basin, the middle Miocene de Hidrocarburos & Antek, 2013).
reflectors are in toplap against Pliocene reflectors (Figure 3a).
In the strike seismic line, they are conformable with the lower 4.4.3. Biostratigraphy
and upper Miocene reflectors (Figure 3b). It is not clear that
the middle Miocene reflectors were affected by the mud diapir; The Remolino Grande–1 well contains useful biostratigraphic
therefore, they are marked with a dotted line. events such as the last occurrences of Globorotalia peripher-
oronda, H. ampliaperta, and S. heteromorphus, which were
4.4.2. Lithology found together with sporadic abundances of Fohsella fohsi, Glo-
bigerinoides sicanus, and Orbulina universa. Microfossils of
In the Remolino Grande–1 well, the middle Miocene deposits this well do not support an age younger than the nanoplankton
consist of amalgamated sandstones interbedded with conglom- biozone NN6 (late Serravallian at 1990’, Figure 5). Neverthe-
erates; in lower proportions, sandy siltstones and calcareous less, this information linked to the 11.5 Ma maximum depo-
sandstones are present. In the Majagua–1 well, they mainly sitional age of detrital zircon at ca. 1450’ allows the location
consist of sandstones and conglomerates interbedded with of the middle – upper Miocene limit between ca. 1450–1990’
mudrocks (Figure 4), with abundant foraminifera and mollusk (Figures 4, 5). Biostratigraphic records of Tumaco 1–ST–P and
shells. These rocks are interpreted as delta front deposits prob- Majagua–1 wells have discontinuous abundance patterns of mi-
ably related to a relative shallowing of the depositional environ- crofossils and very abundant reworked assemblages (Figure 5).
ment (Agencia Nacional de Hidrocarburos & Universidad de In the case of the Tumaco 1–ST–P well, reworked species of
Caldas, 2011a). In the Tumaco 1–ST–P well, middle Miocene Cyclicargolithus floridanus and S. heteromorphus were report-
rocks are composed of mudrocks, sublitharenites, feldspathic ed within younger microfossils of Discoaster kugleri. Figure 5

228
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
a b
Lv Pl Fo
Lv Pl
Pl

Fo

Neogene
Pl

Qzp Lv Qzm OxHb


Lv
Lv Lv

Paleogene
Pl Pl
Pl
Pl Px
Pl

Pl Lv
500 µm 500 µm
Lv
Qz

c d
Qzm

Ch
Lv
Pl Qzm
Pl
Lv
Lv
Ch

Pl
Ls Ch
Pl

Pl
Lv Lv
Lv

Hb Ls
Pl 100 µm 500 µm

Figure 9. Microphotographs of the Majagua–1 well. (a) Sample MIPP–01: Feldspathic litharenite with plagioclase, volcanic lithic, pyroxene,
and calcareous cement. Crossed nicols. (b) MIPP–13: Feldspathic litharenite with volcanic fragments, plagioclase (zoned and twined),
quartz, oxyhornblende, and foraminifera. Crossed nicols. (c) MIPP–18: Feldspathic litharenite with plagioclase, volcanic and sedimentary
lithics. Crossed nicols. (d) MIPP–28: Feldspathic litharenites with volcanic fragments, plagioclase, quartz, chert, and calcareous cement.
Crossed nicols. Abbreviations: (Lv) volcanic lithic; (Ls) sedimentary lithic; (Ch) chert; (Pl) plagioclase; (Qz) quartz; (Qzp) polycrystalline
quartz; (Qzm) monocrystalline quartz; (Hb) hornblende; (OxHb) oxyhornblende; (Px) pyroxene; (Fo) foraminifera.

shows the stratigraphic equivalence of the Remolino Grande–1, stones, amphibole, biotite, chlorite, pyroxene, glauconite,
Majagua–1, and Tumaco 1–ST–P wells, with the calcareous opaque, and bioclasts (foraminifera). The sandstone grains of
nannofossils and planktonic foraminifera standard biozones. the Majagua–1 and Remolino Grande–1 wells are classified
This range covers the interval between zones NN4 to NN6 and as lithic arkoses, arkoses, and lower proportions of feldspathic
N8 to N12, respectively. litharenites (Figures 7, 8) and are composed of fine to coarse,
angular to rounded grains, with a high content of plagioclase
4.4.4. Petrography (23–68 %), rock fragments (8–47 %), and a minor proportion
of quartz (2–23 %). The lithic fragments are mainly volcanic
Ditch cutting samples from the Majagua–1 and Remolino of intermediate composition (8–39 %), sandstones (2–3 %),
Grande–1 wells are composed mainly by siltstones, sandy shales (1%), chert (3–8 %), and graphite schists (1–2 %) (Fig-
siltstones, sandstones, volcanic rocks, chert, and plagioclase ure 7). Accessory minerals as oxyhornblende and hornblende
(Figures 7, 8), with lower proportions of quartz grains (mono– (1–5 %), epidote (1–3 %), muscovite (< 1%), biotite (1–3
and polycrystalline), plutonic rocks, potassium feldspar, lime- %), and chlorite (1–2 %) are present. In the Tumaco 1–ST–P

229
PARDO–TRUJILLO et al.

a b c d

10 µm 10 µm

20 µm 10 µm

e f g h i

10 µm 5 µm
10 µm
5 µm

10 µm

Cpx + Opx + Ol + Spl


j k Mafic and ultramafic
Remolino Grande–1 welll

Tur Majagua–1 well

Ap Tumaco–1–ST–S well

Tumaco–1–ST–P well
Tur
Tumaco Bay outcrops

Bt OxHb Bt Pliocene

Tur Upper Miocene


Ep Ep Middle Miocene

Lower Miocene
Ep Bt Oligocene
Bt
100 µm
Ap OxHb + Hb + Bt Rt + Tur + Grt + Gln
Magmatic arcs Metamorphic

Figure 10. Some heavy minerals found in the sandstones of the Tumaco Basin. (a) Zircon. (b) Apatite. (c) Green tourmaline. (d) Epidote.
(e) Augite. (f) Hypersthene. (g) Hornblende. (h) Oxyhornblende. (i) Biotite. (j) Association of ultrastable and unstable phases: (Tur)
Tourmaline + (Ap) Apatite + (Bt) Biotite + (Ep) epidote + (OxHb) Oxyhornblende. (k) Triangular diagram with groups of heavy minerals
and their source rocks.

230
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
Majagua–1 well 0.12
M1MP–001
35
8.4 Ma
M1MP–001
a 650 b
30 0.10 M1MP–002
n: 93

550 3400
25 M1MP–009

Neogene
0.08
Number of analysis

20 18.2 Ma 25.2 Ma 450 M1MP–014

Pb/238U
700
15 0.06
350

206
3000
10 1400

Paleogene
75.6 Ma 0.04 250
500
5

0 150
0.02 2600 1000

70 8.4 Ma M1MP–002 50
n: 100 2200 300
0.00
60 1800

50 600
Number of analysis

100
40
200
30
11.1 Ma
20 0.0 0.2 0.4 0.6 0.8 1.0
12.3 Ma 207
Pb/235U
10

0 c M1MP–001 M1MP–002
8.5 Ma
90 M1MP–003 16.1 Ma 1705 Ma 13.2 Ma 11.4 Ma 9.1 Ma
57.4 Ma
n: 96
80
24.8 Ma 8.2 Ma 8 Ma
70 9 Ma 7.2 Ma 41.7 Ma 10.6 Ma
9.1 Ma
Number of analysis

90.3 Ma
60 8.3 Ma 247.8 Ma
50 8 Ma
100 µm 79.9 Ma 100 µm
40 18 Ma 13.1 Ma
30
20
M1MP–003
10
44.8 Ma 8.4 Ma 9.1 Ma 9.1 Ma 8.6 Ma 8.5 Ma
0 100 µm
8.6 Ma 8.1 Ma
11.6 Ma M1MP–009
50 8.8 Ma 8.5 Ma
n: 75 44.8 Ma 9.2 Ma
11.6 Ma 8.1 Ma
8 Ma
40
Number of analysis

8.6 Ma 9.4 Ma 8.2 Ma


10 Ma
30 9.4 Ma
12.6 Ma 9.3 Ma 8.4 Ma
20 9.4 Ma 8.8 Ma 9.2 Ma
100 µm
16.4 Ma 10.6 Ma 8.9 Ma 8.9 Ma 8.4 Ma 11.5 Ma 8.5 Ma

10
44.8 Ma
M1MP–014
0
18.9 Ma 19.2 Ma 49.8 Ma 44.2 Ma 74.9 Ma
M1MP–014 17.5 Ma
35
n: 100
46 Ma 54.6 Ma
30
47.2 Ma
17.7 Ma
Number of analysis

25 17.9 Ma 18.9 Ma 18.7 Ma


18.5 Ma
19.8 Ma 19.1 Ma 22.2 Ma
20 844 Ma
44.9 Ma
100 µm 19.1 Ma 100 µm
15 48.3 Ma 22.3 Ma

10 45 Ma

5
51.3 Ma U–Pb detrital zircons geochronology
20.7 Ma
0 0–30 Ma (Cenozoic 30–60 Ma (Timbiquí–Macuchi
65–100 Ma (Western Cordillera)
0 10 20 30 40 50 60 70 80 90 100 magmatic arc–Western Cordillera) magmatic arc)
Age (Ma)

Figure 11. (a) Relative probability diagrams of U/Pb ages of detrital zircons from the Majagua–1 well. (b) Concordia diagrams of the sam-
ples (Tera–Wasserburg diagrams, Ludwing, 2003). (c) Cathodoluminiscence of some zircon crystals. The presence of zircons with volcanic
morphology (euhedral grains with length/width ratios >3) and U/Pb ages between 12 and 8 Ma is noteworthy (from Echeverri et al., 2015b).

231
PARDO–TRUJILLO et al.

well, sandstones are mainly litharenites and lithic greywackes. 4.5.1. Seismic Lines
Rock fragments are mainly sedimentary, metamorphic, volca-
nic and plutonic mafic and felsic (Agencia Nacional de Hidro- To the east of the basin, the upper Miocene reflectors onlap over
carburos & Antek, 2013). the lower – middle Miocene and the basement rocks (Figure
3a). Towards the west of the basin, some reflectors come to
4.4.5. Heavy Minerals the surface in the Remolino Grande–Gorgona Structural High,
forming gentle hills, which are currently subject to erosion. In
Two samples from the Remolino Grande–1 well and three the offshore basin, the upper Miocene reflectors toplap against
from the Majagua–1 well were analyzed (Figures 6, 7). In the the Pliocene (Figure 3a). On the strike line, both the upper Mio-
Remolino Grande–1, stable phases are almost absent, with cene and the Pliocene reflectors become thinner towards the
only 1% zircon. The unstable phases include apatite (13–23 south of the studied area (Figure 3b).
%), clinopyroxene (12 %), orthopyroxene (1–2 %), horn- Pliocene reflectors can be divided into three sets. (i) The
blende (2–9 %), and minerals of the epidote group (59–61 %). lower set rests on the Cretaceous basement to the east, and on-
Oxyhornblende, biotite, chlorite, and olivine are present in lap onto the Paleogene (Figure 3a). Towards the west, it de-
low proportions (≤ 2%). In the Majagua–1 well, a higher pro- creases in thickness, and is truncated by the reflectors of the
portion of unstable with respect to the ultrastable phases was middle set. (ii) The middle set shows a clear progradation of
found (Figure 7). The ultrastable phases include zircon (7–21 the eastern and western reflectors towards the depocenter of
%) and traces of tourmaline (< 1%). The unstable phases in- the basin showing downlap over the set 1 (red arrows in Fig-
clude apatite (29–75 %), minerals of the epidote group (6–29 ure 3a) and a decrease in their thickness in the same direction.
%), biotite (2–5 %), clinopyroxene (1–25 %), chlorite (1–3 (iii) The upper set truncates the reflectors of the middle set and
%), hornblende (≤ 1%), actinolite (≤ 1%), and traces of oxy- pinches out towards the E and W borders of the basin (Figure
hornblende and ortopyroxene. The unstable phases such as 3a). In the Tumaco offshore basin, Pliocene beds are covering
apatite, pyroxene, and epidote are dominating and increase the middle and upper deposits, forming an erosional truncation
with respect to the Oligocene beds in the Tumaco 1–ST–P (toplap) (Figure 3a). In the Remolino Grande–1 well, there is
well (Agencia Nacional de Hidrocarburos & Antek, 2013). no Pliocene record, probably related to erosion or no deposition
In the Tumaco 1–ST–P well, a greater proportion of unstable during the Remolino Grande–Gorgona Structural High uplift.
phases compared to stable was also observed. The ultrastable
phases include zircon (1%) and traces of tourmaline (< 1%). 4.5.2. Lithology
The unstable phases include hornblende (48–68 %), minerals
of the epidote group (6–26 %), apatite (< 1–4 %), clinopyrox- In the Remolino Grande–1 well, upper Miocene beds are com-
ene (2–28 %), ortopyroxene (< 1–18 %), chlorite (1–3 %), and posed by thick beds of sandstones and sandy siltstones inter-
garnet (< 1–2 %). layed with thin beds of mudrocks. Above 760’, mudrocks with
thin sandstone beds are dominant (Figure 4). In the Majagua–1
4.4.6. Detrital Geochronology well, the lower part of the beds is composed of sandstones inter-
bedded with sandy mudstones, mudrocks and, in lower propor-
Two samples from the Remolino Grande–1 well were ana- tions, conglomerates. In the upper part (above ca. 8000’), there
lyzed (RG–MP–007: depth 2640–2760’; RG–MP–006: depth is a sudden increase in thick sandstone beds (Chagüí Formation;
2140–2380’). These samples have maximum depositional Figures 2, 4). The sandstones in some cases can be calcareous.
ages of 13.3 Ma and 13.1 Ma (Serravallian), respectively. Remnants of mollusks and carbonized organic matter are local-
They show similar patterns of zircon distributions: the most ly abundant. The thickness and frequency of sandstones vs. fine
abundant population is 12–14 Ma, followed by 22–24 Ma and grained sedimentites change through the time; for this reason,
45 Ma. Very low percentages of Late Cretaceous, Triassic, several units have been proposed (e.g., Angostura, Chagüí, and
early Paleozoic, and Precambrian zircons were recognized San Agustín Formations; Agencia Nacional de Hidrocarburos
(Figure 12). & Universidad de Caldas, 2011a; Suárez, 1990). These deposits
can be associated with deltaic systems. Good exposures of these
4.5. Upper Miocene – Pliocene rocks can be studied in the coastal cliffs of the Tumaco Bay
(Figure 13). Normal faults, slumped beds, and clastic dykes
This time interval has been partially identified in the Remo- are common in these units (Figure 13). In the Tumaco 1–ST–P
lino Grande–1 (0–1990’), Majagua–1 (0?–10500’), Tumaco well, mudrocks, calcareous mudrocks, laminated sandstones,
1–ST–P (32–7000’), Tumaco 1–ST–S (0?–1899.6’), and Tu- and matrix–supported conglomerates can be observed; they are
maco Bay outcrops. generally bioturbated. An increase upward in volcanic materi-

232
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
Remolino Grande–1 well
6 RG1–MP–001
12 Ma a b
RG–MP–001 0.28
5 n: 17 RG1–MP–004
1400 1800
0.24

Neogene
4 1800
RG1–MP–006
Number of analysis

0.20
1000

Pb/238U
1400
3 0.16 RG1–MP–007
1500
1400
0.12

206
2 28 Ma 600 1000 1300
7.4 Ma
0.08

Paleogene
1 1100
1000
0.04 200
600 900
0
0.00
700
200 600
45
RG–MP–004
500
40 12 Ma n: 73
300
35
Number of analysis

30
25 0 1 2 3
207
20 23.2 Ma Pb/235U
15
c RG–MP–04
10
5 75.1 Ma 23.2 Ma 75.3 Ma 10.7 Ma 25 Ma 12.8 Ma 23.9 Ma
49.5 Ma 42.7 Ma
0
13.6 Ma
13.6 Ma
RG–MP–006
25 n: 34

20
Number of analysis

10.9 Ma 11.5 Ma 11.7 Ma 22.5 Ma 23 Ma 10.2 Ma


15
100 µm 13.6 Ma 83.3 Ma 11.7 Ma
10.9 Ma
10

5 24.8 Ma
RG–MP–06 RG–MP–04
13.1 Ma 13.1 Ma 24.5 Ma 583.3 Ma 83.3 Ma
0
13.9 Ma
16 Ma 12.6 Ma
11.8 Ma
16 13.5 Ma RG–MP–007 12.5 Ma
22.5 Ma n: 81
14 17.2 Ma
13 Ma 1495.3 Ma
46.7 Ma
12 13.7 Ma
Number of analysis

10 23.8 Ma 11.4 Ma 100 µm


13.1 Ma 13.2 Ma 100 µm 12.3 Ma 12.8 Ma
8 10.9 Ma
6

2 U–Pb detrital zircons geochronology


0 0–30 Ma (Cenozoic 30–60 Ma (Timbiquí–Macuchi
65–100 Ma (Western Cordillera)
0 20 40 60 80 100 magmatic arc–Western Cordillera) magmatic arc)
Age (Ma)

Figure 12. (a) Relative probability diagrams of U/Pb ages of detrital zircons from the Remolino Grande–1 well. (b) Concordia diagrams
of the samples. (c) Cathodoluminiscence of some zircon crystals (from Echeverri et al., 2015b).

als is notorious. They were interpreted as fluvial and deltaic minifera, echinoderms, crustaceans, and well–preserved plant re-
deposits (Agencia Nacional de Hidrocarburos & Antek, 2013). mains were found. The unit was accumulated in a deltaic system
During the latest Miocene – Pliocene (Messinian – Zanclean), influenced by volcanism (Echeverri et al., 2016).
in the southern part of Tumaco Basin, more than 1300’ (400 m) of
thick lenticular layers of sandstones and conglomerates, with an 4.5.3. Biostratigraphy
important volcanic input (Cascajal Formation of Echeverri et al.,
2016), were accumulated and interlayered with some mudrocks Tumaco 1–ST–P and Majagua–1 are the only wells from which
and muddy sandstone beds. Locally, bivalves, gastropods, fora- middle Miocene microfossils were recovered. However, our in-

233
PARDO–TRUJILLO et al.

m A’
40 Ox
Fe a w E
36

32

28

24

20
0 20 40 m
A’
16

12

b NW SE

0 10 20 m

Mudrocks Conglomeratic Slumped beds


Fault Bivalves Logs Muddy intraclasts
sandstones

Sandstones Vegetation Ox
Colluvium Gastropods Leaves Iron oxides Clastic diques
Fe

Figure 13. Panoramic pictures of Tumaco Basin outcrops (San Agustín Formation, Tortonian – Messinian). (a) Curay. (b) La Chorrera.
The facies are mainly mudrocks interbedded with sandstone beds. Notice the presence of slumped beds and clastic dykes (see Figure
1 for location).

terpretation suggests that these wells record Tortonian sedimen- in the Tumaco 1–ST–P well according to the last occurrence of
tation. This is supported by the occurrence of Catinaster coalitus, E. kugleri (Figure 5). After this, we observed several bioevents of
Discoaster bellus, and D. hamatus in both sites. Detailed studies the calcareous nannofossils C. coalitus, D. hamatus, and plank-
conducted on the Tumaco 1–ST–P well have demonstrated that tonic foraminifera Neogloboquadrina acostaensis and Paraglo-
this assemblage is very abundant and highly resistant to disso- borotalia mayeri, as well as sporadic nannofossils D. bellus, D.
lution, making them useful in biostratigraphy. The Serravallian/ bollii, D. brouweri, and S. abies. An increase of the accumulation
Tortonian (middle/upper Miocene) boundary was placed at 7000’ rate is inferred during the Tortonian at Tumaco 1–ST–P well, as

234
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
these taxa are characterized by a short life span and they were glauconite (1–2 %), biotite (1–6 %), chlorite (1–4 %), and
found successively in the well. Alternatively, the last consistent epidote (1–6 %). In the Tumaco 1–ST–P well, sandstones are
biostratigraphic marker in the Majagua–1 well was the last oc- litharenites, feldspathic litharenites, and lower proportions
currence of C. coalitus at 8820’. of lithic arkose. Constituents are mostly volcanic. Lithics are

Neogene
As shown in Figure 5, in the Majagua–1 and Tumaco 1– mainly mudrocks, andesites, and diorites (Agencia Nacional
ST–P wells, the equivalence with the calcareous nannofossils de Hidrocarburos & Antek, 2013).
and planktonic foraminifera standard zones represents an inter-
val from NN7 to NN10 and N14 to N16, respectively. 4.5.5. Heavy Minerals

Paleogene
Diagnostic species such as D. neohamatus, D. berggrenii, D.
quinqueramus, D. asymmetricus, and Globorotalia tumida were Five samples from the Remolino Grande–1 well and ten from
identified in the upper part of the Tumaco 1–ST–P and Maja- the Majagua–1 well were analyzed (Figures 6, 7). In the Remo-
gua–1 wells, occasionally associated with abundant reworked lino Grande–1, stable phases are scarce with tourmaline (2–10
microfossils (mainly Cretaceous and Paleogene). Although it %) and zircon (1%). The unstable phases are dominant and
is difficult to identify a succession of standard zonations, the include apatite (3–20 %), clinopyroxene (9–33 %), orthopy-
microfossil assemblages allow the designation of a Messinian – roxene (1–4 %), hornblende (1–9 %), oxyhornblende (1–3 %),
Piacenzian age. Palynological data in the Tumaco 1–ST–P well biotite (1–10 %), and minerals of the epidote group (31–81 %).
show the first occurrence of Cyatheacidites annulatus at 3950’ Chlorite, olivine, and actinolite are present in low proportions
(ditch cutting sample) (Figure 5), considered a late Miocene (≤ 1%). In the Majagua–1 well, a higher proportion of unstable
biostratigraphic marker in the Llanos Basin (first occurrence with respect to the ultrastable phases was found (Figure 7).
datum at 7.1 Ma) (Jaramillo et al., 2011). The ultrastable phases include zircon (1–23 %) and tourmaline
(1–33 %). The unstable phases include apatite (4–69 %), min-
4.5.4. Petrography erals of the epidote group (4–38 %), biotite (2–6 %), musco-
vite (1 %), clinopyroxene (2–19 %), ortopyroxene (< 1–5 %),
Upper Miocene ditch cutting samples from the Majagua–1 and olivine (1%), chlorite (1–3 %), oxyhornblende (< 1–7 %), and
Remolino Grande–1 wells (upper Viche and Angostura Forma- hornblende (1–19 %). In the Tumaco 1–ST–P well, a greater
tions) are mainly composed of mudrocks, volcanic rocks, and proportion of unstable phases compared to stable phases was
plagioclase. Chert, quartz (mono and polycrystalline), potas- also observed (Agencia Nacional de Hidrocarburos & Antek,
sium feldspar, amphibole, biotite, pyroxene, chlorite, glauco- 2013). The ultrastable phases only include zircon (1%), while
nite, and bioclasts (foraminifera, algae, bivalves) are present unstable phases are dominant and include hornblende (9–72 %),
in a lower proportion. In the middle part (Chagüí Formation), minerals of the epidote group (< 1–19 %), apatite (< 1–3 %),
the samples have mainly volcanic fragments and plagioclase, clinopyroxene (4–3 6%), ortopyroxene (4–63 %), biotite (1–3
as well as sedimentary rocks (shales, sandstones, and cherts) %), chlorite (≤ 1%), and garnet (1%).
(Figure 7). In some levels, pyroxene, amphibole, and biotite
are frequent. In addition, plutonic grains, potassium feldspar, 4.5.6. Detrital Geochronology
quartz (mono and polycrystalline), tourmaline, apatite, calcite,
chlorite, epidote, glauconite, and foraminifera are present in Two samples for the Remolino Grande–1 well (RG–MP–004;
lower proportions. In the top of the Majagua–1 well, the sam- depth 1210–1450’ and RG–MP–001; depth 84–300’) were an-
ples mainly have volcanic grains, plagioclase, and sedimentary alyzed. They have 10.4 Ma and 11.5 Ma of maximum deposi-
rock fragments (chert, shale, limestone, and bioclasts). In lower tional ages, respectively. These data allow the lower – middle
proportions, schists, intermediate plutonics, potassic feldspars, Miocene boundary to be constrained between ca. 1450–1990’
quartz (mono– and polycrystalline), pyroxenes, amphibole, depth based on geochronologic and biostratigraphic data (Fig-
tourmaline, chlorite, epidote, glauconite, and bioclasts (fora- ure 5). They show similar patterns of zircon distributions: The
minifera) were identified. most abundant population is 12–14 Ma, followed by 22–24 Ma
Sandstone grains of the Majagua–1 well are classified as and 45 Ma. Very low percentages of Late Cretaceous, Trias-
lithic arkoses and feldspathic litharenites (Figure 8). They sic, early Paleozoic, and Precambrian zircons were recognized
are fine–grained, poorly sorted, and with angular to round- (Figure 12).
ed grains (Figures 7, 8), composed mainly by plagioclase In the Majagua–1 well, four samples (M1MP–009, depth
(12–54 %), quartz (1–25 %), and intermediate volcanic rock 9100–9320’; M1MP–003, depth 3430–3950’; M1MP–002,
fragments (4–54 %). Sedimentary lithics are composed by depth 2830–3210’; and M1MP–001, depth 110–800’) were
sandstones (1–8 %), mudrocks (1–14 %), and cherts (1–5 %) analyzed. They have 9.3 Ma, 7.6 Ma, 7.5 Ma, and 6.6 Ma max-
(Figure 7). Among accessory minerals were identified oxy- imum depositional ages, respectively. Three main populations
hornblende and hornblende (1–16 %), pyroxene (1–5 %), of zircons are recognized: 7–12 Ma, 16–25 Ma (especially to

235
PARDO–TRUJILLO et al.

the top of the well), and 45–50 Ma (Figure 11). In lower propor- sequence of (ca. 32’; 9.8 m) sands and muds was identified.
tions are late Oligocene – Miocene, Paleocene, and Late Creta- There were two 14C AMS ages in organic sediments, 4360 ± 30
ceous and Jurassic zircons, which are not abundant enough to BP and 4150 ± 30 BP (Holocene). These sediments belong to
constitute a population. the deltaic plain of the Mira River (López et al., 2012).
Two samples from the Tumaco 1–ST–S well, two from out-
crops of the Tumaco Bay, three in the Pleistocene fans and one 5. Interpretation
in recent sedimentites were also analyzed. In the Pliocene beds,
similar populations of detrital zircons ages are observed: the most 5.1. Cretaceous – Paleogene
abundant is 4.1–7.6 Ma, followed by 6.5–7.6 Ma, 9.0–13.0 Ma,
and 19.0–23.0 Ma. Very low percentages of Late Cretaceous, Tri- The Upper Cretaceous volcano–sedimentary succession
assic, early Paleozoic, and Precambrian zircons were recognized. of Remolino Grande probably originated in an oceanic arc
The maximum depositional ages in two samples of the Tumaco during the Late Cretaceous (Echeverri et al., 2015a). This
1–ST–S well were 6.5 Ma (1257’) and 4.1 Ma (400’) (Figure 5). unit and its basement collided obliquely against the western
margin of Ecuador and Colombia during the Late Cretaceous
4.6. Pliocene – Holocene – Paleogene, which generated thrusting, folding, and clastic
sedimentation (Barrero et al., 2006; Moreno–Sánchez & Par-
Pliocene – Holocene volcaniclastic fans were identified in the do–Trujillo, 2003; Pindell & Kennan, 2009; Villagómez et al.,
eastern border of the basin (Figure 14). The volcanic source 2011). The sedimentary record of the Remolino Grande–1 well
probably came from an old Cumbal Volcano and from the Azu- shows a Upper Cretaceous (Maastrichtian) – Eocene (ca. 20
fral volcanic activity in the Western Cordillera. Three alluvial Ma) unconformity, which could be associated with this tectonic
fans can be differentiated, which prograde in the basin through event. In northern Ecuador, the accretion of the Piñon–Naran-
time (Figure 14): (i) The oldest fan (Pliocene?), 1560 km2, is jal Terrane occurred at ca. 58 Ma (Paleocene) (Jaillard et al.,
strongly dissected and controlled by the Junín Sambianí Fault, 2009). The presence of Eocene plutonic and volcanic rocks in
and it is formed by lahar and debris flow deposits mainly com- the Western Cordillera (Agencia Nacional de Hidrocarburos
posed of dacitic and andesitic rock fragments. (ii) A younger & Geología Regional y Prospección, 2011; Barbosa–Espitia et
fan, 160 km2, formed by lahar and debris flow deposits, is main- al., 2016) suggests that an Eocene volcanic arc was developed
ly composed of andesitic rock fragments. (iii) The youngest fan after the collision of the Tumaco Basin basement with the con-
covers 2030 km2. It is formed by lahar, debris flow, and stream tinental margin (Arco de Ricaute of Spadea & Espinosa, 1996).
flow deposits and is mainly composed of dacitic pumice frag- Over these rocks, upper Eocene – lower Miocene fine–grained
ments. This deposit partially overlays the Cascajal Formation clastic sedimentary successions were discordantly accumulat-
in unconformity, and they are interlayered with recent littoral ed in shelf and prodelta environments (Lutitas de Remolino
sediments along the Pacific coast. Grande Formation or 1 Sur Unit; Figure 2).
The youngest volcanoclastic fan has 2.04 Ma as its max-
imum depositional age (Figure 14). In the Pleistocene fans 5.2. Early – Middle Miocene
and the recent sediments, detrital zircon populations of 26–20
Ma, 14–11 Ma, 9–6 Ma, and 4–3 Ma were identified. These The facies recorded in the wells suggest sedimentation in shelf,
detrital populations are in concordance with those present in prodelta, and deltaic environments (Angostura, Viche, and Ca-
the underlying stratigraphic sequence. However, there is a sig- yapas Formations; Figure 2; Agencia Nacional de Hidrocarburos
nificant presence of zircons between 3–1.5 Ma, which records & Universidad de Caldas, 2011a; Agencia Nacional de Hidro-
the most recent activity of the magmatic arc (Figure 14). In carburos & Antek, 2013). The common presence of muddy ma-
the Tumaco Bay outcrops, five samples were analyzed in the trix and the angularity and poor sorting of grains in sandstones
Pliocene – Pleistocene deposits (Figure 1). In general, stable and conglomerates suggest first–cycle sedimentation and rapid
phases occur in minor proportions compared with unstable burial. During the early Miocene, an ca. 5 Ma unconformity
ones. The stable phases are constituted by zircon (10–21 %), was identified in the Remolino Grande–1 well (Figure 5). This
tourmaline (5–9 %), and rutile (2%). The unstable phases are unconformity could be related to an exhumation pulse of the
dominant and include apatite (3–7 %), pyroxene (10–16 %), Western Cordillera and the Remolino Grande–Gorgona Struc-
hornblende (24–33 %), biotite (24–29 %), and olivine (3–5 tural High, linked to an increase in the orthogonal convergence
%). Chlorite and minerals of the epidote group are present in rates of the Nazca–South American Plates (cf. Pardo–Casas &
very low proportions (≤ 1%). Molnar, 1987; Somoza & Ghidella, 2012). Based on thermo-
Holocene deposits are mainly composed by coastal and flu- chronologic analyses of the Gorgona Island, the Tumaco Basin,
vial sediments from the Patía and Mira Rivers. To the upper part and the southern part of the Western Cordillera, Barbosa–Espitia
of the Tumaco 1–ST–S well, an unconsolidated sedimentary et al. (2013a) interpreted a progressive and generalized exhuma-

236
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
a b

79° 00' W

78° 30' W

78° 30' W
60 MB–11
2.04 Ma
n: 55
50

Neogene
Relative probability
Number of analisys
40
2° 00' N

30

Pacic Ocean 20

Paleogene
10

0
Tumaco Majagua–1 0 25 50 75 100
Age (Ma)

ANH–Tumaco–1–ST–P

1° 30' N

Guayacana
MB–11 Llorente

Junín
Wells

Sample
Ricaurte
City, town.

Road
Piedraancha
0 10 20 km

Figure 14. (a) Distribution of Pliocene – Pleistocene volcanoclastic fans in the southern Tumaco Basin. In green: the oldest fan; purple:
the younger fan; brown: the youngest fan (see explanation in the text). (b) Relative probability diagrams of U/Pb ages of detrital zircons
from the MB–11 sample (color conventions as in Figure 11).

tion cooling in the area between 24–22 and 20–16 Ma. Thermo- Three compositional associations in lower – middle Miocene
chronologic analyses of the Western and Eastern Cordilleras of sandstones are indicative of sediment sources: (i) Abundance of
Ecuador also recorded an exhumation event between 25–22 Ma, plagioclase and intermediate volcanic fragments can be asso-
which suggests a regional event linked to plate–scale kinematic ciated with volcanic igneous rock sources. (ii) Olivine and py-
changes (Spikings & Crowhurst, 2004; Spikings et al., 2005). roxenes (mainly augite) are related to mafic–ultramafic igneous
In Ecuador, a middle Miocene cooling event has been attributed rocks (Figure 10). (iii) Potassium feldspar, quartz, amphibole,
to the increase in compressive stress during the collision of the and occasionally biotite are related to acid and/or intermedi-
Carnegie Ridge with the South American margin, which started ate plutonic rocks. This can be compared to the present–day
at ca. 15 Ma (Spikings et al., 2010). According to López–Ramos basement of the Western Cordillera, where upper Eocene and
(2009) the division of the Tumaco Borbón and Manglares (Tu- Oligocene granitoids intrude the Cretaceous oceanic sedimen-
maco offshore) Basins occurred in the middle Miocene time and tary and basic igneous rocks (Figure 15; Barbosa–Espitia et
was related to the subduction of the young and hot Nazca Plate al., 2016; Gómez et al., 2015). In the Majagua–1 well, there is
and the decrease in the convergence rate. more than 50% sedimentary lithics with respect to igneous frag-
ments and feldspars. In contrast, the Remolino Grande–1 well

237
PARDO–TRUJILLO et al.

shows an increasing tendency in the content of volcanic, pluton- López–Ramos, 2009) and to the increase in the volcanic ac-
ic fragments, and feldspars in the middle Miocene (Figures 6, tivity of the magmatic arc. López–Ramos (2009) suggests that
7). The analysis of heavy minerals in the Remolino Grande–1 the considerable sediment accumulation in the Tumaco Basin
well shows a domain of unstable phases (mainly epidote, augite, would have resulted from crustal buckling due to horizon-
and hornblende), and the Majagua–1 well presents a domain of tal stress transfer into the overriding plate and the erosion of
ultrastable phases (mainly zircon and apatite). These compo- the Western Cordillera. He also indicates that the Remolino
sitional contrasts are probably related to different source areas Grande Gorgona High was uplifted and allowed sediments
or disconnection in the paleo–drainage systems or may also be to dam in the Tumaco onshore basin. Late Miocene – Plio-
influenced by the formation of topographic highs that controlled cene uplift pulses (ca. 14–10 and ca. 6–4 Ma) recognized in
the drainages and acted as sedimentation barriers. southwestern Colombia (Barbosa–Espitia et al., 2013b) could
The 24–18 Ma and 12–14 Ma most frequent populations be related to the following: (i) Subduction of the young oce-
of detrital zircons indicate a magmatic activity during these anic crust with change in the subduction angle between Naz-
periods in the vicinity of the basin, which can be associated ca and the South American Plate (Echeverri et al., 2015b),
with Western Cordillera plutonic bodies that intrude the Cre- (ii) orthogonal convergence of the Nazca Plate (Pardo–Casas
taceous oceanic sedimentary and basic igneous rocks of the & Molnar, 1987; Somoza, 1998), or (iii) the collision of a
Western Cordillera (Agencia Nacional de Hidrocarburos & buoyant slab segment, derived from the Nazca Plate, such
Geología Regional y Prospección, 2011; Barbosa–Espitia et as a small ridge or seamount, similar to, or being part of, the
al., 2016; Echeverri et al., 2015b). Magmatic activity close to Carnegie Ridge, and (iv) the collision of the Panamá–Chocó
the source is also consistent with the preservation of feldspars Block (Barbosa–Espitia et al., 2013a).
and volcanic fragments and is associated with the erosion of López–Ramos (2009) indicated a upper Miocene (Torto-
middle–upper crustal levels in orogenic zones with magmatic nian) unconformity (called U2) based on seismic reflectors in
activity. The occurrence of chert and other sedimentary lithics the Tumaco Basin. It can be associated with a facies change
in the sandstones, as well as Mesozoic, Paleozoic, and Pre- in the Majagua–1 (8000’; base of the Chagüí Formation) and
cambrian detrital zircons, suggests reworking of pre–Miocene the Tumaco 1–ST–P wells (ca. 5400’; base of the Tangareal
sedimentary sequences of the Western Cordillera or, alterna- del Mira Formation of Agencia Nacional de Hidrocarburos &
tively, a connection with the Central Cordillera basement. This Antek, 2013) (Figure 5). This unconformity can be related with
connection could explain the presence of metamorphic frag- an eustatic sea level drop (10.5 Ma; Haq et al., 1987; López–
ments in very low percentages. Additionally, marine fossils Ramos, 2009). Based on thermochronological constraints,
(mollusks and foraminifera) have been reported in the rocks Barbosa–Espitia et al. (2013b) proposed that the Remolino
of the Cauca–Patía Basin for the Oligocene – Miocene inter- Grande–Gorgona Structural High was uplifted between 14–10
val (León et al., 1973). This basin is located to the east of the Ma. This is also supported by the common presence of early
Tumaco Basin between the Central and Western Cordilleras, Miocene reworked microfossils (Figure 15).
which could indicate that the Tumaco and Cauca–Patía Basins During the late Miocene, the upward increase in the pro-
were connected and that some sediments came directly from portion of unstable minerals such as plagioclase, pyroxene,
erosion of the Central Cordillera basement. oxyhornblende, hornblende, and volcanic fragments indicates
an intensification of magmatic activity (Figure 7). This is in
5.3. Late Miocene – Pliocene agreement with a dominant 12–8 Ma zircon population. The
presence of microcline and biotite, as well as the occurrence of
In general, the presence of sandstones and conglomerates in the 25–18 Ma and 50–30 Ma zircon populations, suggest the
the Majagua–1 and Remolino Grande–1 wells, as well as the onset of the erosion of granitoids (Figures 6, 7), which could
abundance of mollusks and carbonized organic matter, suggest be associated with the Western Cordillera basement (e.g., Pie-
deposition in the delta plain and nearshore environments (San drancha Granodiotrite, Nulpi Gabronorite, and ca. 44 Ma dikes
Agustín and Chagüí Formations; Figure 2). Fine–grained inter- intruding the Timbiquí Formation; Figure 15; Agencia Nacional
vals with abundant mollusks and foraminifera could be asso- de Hidrocarburos & Geología Regional y Prospección, 2011a;
ciated with the lower delta front–prodelta transition. Agencia Agencia Nacional de Hidrocarburos & Universidad de Caldas,
Nacional de Hidrocarburos & Antek (2013) proposed sedi- 2011a; Barbosa–Espitia et al., 2016; Echeverri et al., 2015b).
mentary environments varying from fluvial channels, estuarine The presence of chert, sandstone, and shale fragments, as well
and mouth bars, delta front and prodelta environments in the as Mesozoic, Paleozoic, and Precambrian zircons, suggests
Tumaco 1–ST–P well. During this period, an increase in the reworking of the pre–Miocene sedimentary cover or partial
sedimentation rate is notorious (Figures 5, 15). connection with the Central Cordillera basement. The good
The late Miocene increase in the sedimentation rate may preservation of plagioclase and volcanic lithics suggests short
be related to high subsidence in the basin (Echeverri, 2012; transport, rapid burial, and negligible diagenesis effects.

238
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
Eocene – Oligocene (ca. 41–23 Ma)

Tumaco offshore Remolino Grande–Gorgona


Structural High Tumaco onshore basin
basin

Neogene
Magmatism
Upper Cretaceous – Eocene unconformity

Early Miocene First cycle of sediments

Paleogene
Rapid burial Volcanic, ultramafic,
acid, and intemediate
plutonic sources.

Exhumation pulse based on thermochronologic analyses

Middle – late Miocene


Increase in the
subsidence of the basin
Miocene reworked
Increase in the sedimentation rate
microfossils
Abundance of mollusk, foraminifera,
and carbonized organic matter.

Pliocene

Miocene reworked
Volcanic fans
microfossils

Pliocene – Holocene

Coast Remolino Grande–1


line well proyection

0 50 100 150 km

Quaternary Pliocene Middle – upper Lower Miocene Paleogene Cretaceous


Miocene
Middle – upper Subsidence and Onlaps, downlaps,
Fault Miocene boundary and toplaps.
exhumation
Magma chamber Pluton
Well Pyroclastic numb
(active volcanism) (ancient magmatism)

Figure 15. Eocene to Holocene geologic evolution of the southern Tumaco Basin. Not palinspatically restored. The thick red arrows
indicate the direction of sediment transport. See the explanation in the text.

239
PARDO–TRUJILLO et al.

The Miocene – Pliocene boundary is difficult to locate age model of the sedimentary deposits. The sequence starts
with the available information. Based on paleontological data, in the Paleogene (NP19 and P16), and the upper part of the
Agencia Nacional de Hidrocarburos & Antek (2013) indicat- record is consistent with a Messinian – Piacenzian age. The
ed the First Appearance Datum (FAD) of Cyatheacidites an- information to establish the age of the upper Miocene – Plio-
nulatus in the Tumaco 1–ST–P well at 3950’ (ditch cutting cene deposits is still limited.
sample). The FAD of this species is located approximately at The Neogene sandstones of the Tumaco Basin are main-
the Tortonian – Messinian boundary (7.1 Ma, Jaramillo et al., ly related to nearby intermediate to mafic magmatic sources
2011). The correlation of seismic reflectors with the Maja- similar to the present–day Western Cordillera basement. Eo-
gua–1 well (Figure 3b) allowed the establishment of a depth cene and Cretaceous zircon populations can be related to the
of ca. 6000’ for this boundary. A 4.1 Ma (Zanclean) mini- erosion of the Western Cordillera plutonic and volcanic rocks.
mal age of detrital zircons from the Tumaco 1–ST–S at 400’ The presence of Jurassic and older sources would be linked to
indicates that the Miocene – Pliocene boundary is between the reworking coming from units of the Western Cordillera,
3950–400’ (Figure 5). These difficulties show the importance although it is possible that in some areas, a direct connection
of using multi–tools for chronostratigraphic interpretation, as with the Central Cordillera basement existed. Differences in
well as the need to acquire more geological information in petrography and heavy mineral associations for sediments ac-
other places of the basin. cumulated at the same time could be related to different source
areas or disconnection in the paleo–drainage systems related
5.4. Pliocene – Pleistocene to topographic highs.
Detrital zircons and petrographic data indicate that sedimen-
During the Pliocene – Pleistocene, volcaniclastic fans were tation was contemporary with magmatic activity, which started
identified at the SE border of the basin (Figure 14). A pro- at ca. 26 Ma (late Oligocene) until today, and recorded an in-
gradation of seismic reflectors to the west (Figure 3a) can creasing activity since the middle Miocene.
probably be associated with the increase in the erosion rates The seismic, biostratigraphic, and geochronological data
and/or the volcanic activity. Seismic information (e.g., onlap enabled the identification and quantification of the duration of
surfaces and eastward prograding reflectors in the western some unconformities in the Tumaco Basin (e.g., Late Creta-
border of the onshore basin) also shows an influence of the ceous – Eocene, early Miocene), which seem to be related to
Remolino Grande–Gorgona Structural High in the sedimen- regional and global tectonic processes, such as changes in the
tation (Figures 3a, 15). direction and convergence rates of tectonic plates, subduction
López–Ramos (2009) identified an important Pliocene (Zan- of seamounts, and eustasy.
clean) unconformity (U3) marked by a deep erosional surface
and an abrupt facies change (sandstones and conglomerates of Acknowledgements
the Cascajal Formation; Echeverri et al., 2016) and probably
related to a regional Andean orogenic event (e.g., van der Ham- We thank the Agencia Nacional de Hidrocarburos and the
men et al., 1973). Thermochronological data obtained for the Vicerrectoría de Investigaciones y Posgrados (Universidad
Western Cordillera (Piedrancha Pluton) and in the sedimentary de Caldas) for their financial support (contract 092 of 2009,
fill of the Tumaco Basin recorded an ca. 4 Ma exhumation event Universidad de Caldas–Agencia Nacional de Hidrocarburos).
(Barbosa–Espitia et al., 2013a). This time period also coincides Special thanks are due to the Instituto de Investigaciones en
with a marine eustatic drop (Haq et al., 1987). Estratigrafía (IIES) staff for their logistic support. People of the
communities of Tumaco, Salahonda, Cascajal, and La Chorrera
6. Conclusions are greatly acknowledged for their hospitality and help during
the field work. We give special thanks to many people that par-
The Tumaco Basin has a through symmetric shape with ca. 26 ticipate in the 092 contract: Clemencia GÓMEZ, Jairo Alon-
000’ (ca. 8000 m) of sediments in its depocenter. These deposits so OSORIO, Elvira C. RUÍZ, Gustavo HINCAPIÉ, Sebastián
accumulated as a response to the subduction of the Farallon and ROSERO, Marcelo JARAMILLO, Juan Carlos PATIÑO, Ale-
Nazca Plates beneath the South American Plate, controlling sub- jandro PINILLA, César MADRID, Rosa Ester NAVARRETE,
sidence, magmatic activity, and sedimentation rates. Francisco SIERRO, Alejandra MEJÍA, Erika BEDOYA, Di-
The sedimentary fill is mainly composed of mudrocks, ana OCHOA, Silane Da SILVA, Fatima LEITE, Millerlandy
sandstones, and conglomerates that varied in proportion ROMERO, Jenny GARCÍA, and José María JARAMILLO,
through time and accumulated in open marine to deltaic envi- among others. We also thank Jorge GÓMEZ TAPIAS (Servi-
ronments. The integration of biostratigraphic data obtained by cio Geológico Colombiano) for his help and patience during the
the analysis of calcareous nannofossils, planktonic foramin- editing of this paper. Fundación Carolina supported AP–T with
ifera, palynology, and detrital zircons allows us to produce an a postdoctoral stay in Salamanca (Spain) for the finalization of

240
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
this document. The critical revision of the manuscript by Eti- Aubry, M.P. 2014. Cenozoic coccolithophores: Discoasterales (CC–B).
enne JAILLARD and Mario SUÁREZ is greatly appreciated. Atlas of micropaleontology series, part 1 of 4. The Micropa-
We thank Daniela MATEUS ZABALA and the editorial board leontology Press Foundation, 104 p. New York.
of the Servicio Geológico Colombiano for helpful comments Aubry, M. 2015. Cenozoic coccolithophores: Discoasterales (CC–D).

Neogene
that improved the chapter presentation. Atlas of micropaleontology series. The Micropaleontology
Press Foundation, 433 p. New York.
References Backman, J., Raffi, I., Rio, D., Fornaciari, E. & Pälike, H. 2012.
Biozonation and biochronology of Miocene through Pleisto-

Paleogene
Agencia Nacional de Hidrocarburos & Antek. 2013. Estudio integrado cene calcareous nannofossils from low and middle latitudes.
de los núcleos obtenidos de los pozos perforados por la ANH y Newsletters on Stratigraphy, 45(3): 221–244. https://doi.
de las muestras de pozo (núcleos y ripios) disponibles en la Li- org/10.1127/0078-0421/2012/0022
toteca Nacional, localizados en zonas de interés para la ANH. Barbosa–Espitia, A.A., Restrepo–Moreno, S.A., Pardo–Trujillo, A.,
Pozo Tumaco–1–ST–P. Agencia Nacional de Hidrocarburos, Osorio, J.A. & Ochoa, D. 2013a. Uplift and exhumation of
unpublished report. the southernmost segment of the Western Cordillera, Colom-
Agencia Nacional de Hidrocarburos & Geología Regional y Prospec- bia, and development of the neighboring Tumaco Basin. GSA
ción. 2011. Cartografía geológica a escala 1:100 000 de las Annual Meeting 125th Anniversary of GSA. Proceedings 45(7):
planchas 340, 362, 385 y 409 localizadas en la Cuenca Tu- p. 548. Denver, USA.
maco, así como el inherente levantamiento de columnas Barbosa–Espitia, A.A., Restrepo–Moreno, S., Pardo–Trujillo, A.,
estratigráficas y muestreo litológico para los análisis petro- Ochoa, D.I. & Osorio, J.A. 2013b. Uplift and exhumation of
gráficos, bioestratigráficos, geoquímicos, petrofísicos y ra- the southernmost segment of the Western Cordillera and de-
diométricos. Contrato 084 de 2010, unpublished report. 207 p. velopment of the Tumaco Basin. XIV Congreso Colombiano
Agencia Nacional de Hidrocarburos & Universidad de Caldas. 2011a. de Geología. Abstract, p. 387. Medellín.
Estudio geológico integrado en la Cuenca Tumaco onshore. Barbosa–Espitia, A.A., Foster, D.A., Restrepo–Moreno, S.A., Par-
Síntesis cartográfica, sísmica y análisis bioestratigráfico, petro- do–Trujillo, A. & Kamenov, G. 2016. Geochemistry and geo-
gráfico, geocronológico, termocronológico y geoquímico de chronology of plutonic rocks from the Colombian Western
testigos de perforación y muestras de superficie. Contrato 092 Cordillera: New insights into the Cenozoic tectonics of the
de 2009. Unpublished report, 191 p. Manizales, Colombia. NW South American plate margin. GSA Annual Meeting. Ab-
Agencia Nacional de Hidrocarburos & Universidad de Caldas. 2011b. stracts with Programs, 48(7): Paper No. 347–27. https://doi.
Estudio integrado de los núcleos y registros obtenidos de los org/10.1130/abs/2016AM-279826
pozos someros (slim holes) perforados por la Agencia Nacio- Barrero, D., Laverde, F., Ruiz, C.C. & Alfonso, C.A. 2006. Oblique
nal de Hidrocarburos. Informe Final Contrato 093 de 2009. collision and basin formation in western Colombia: The origin,
unpublished report. evolution and petroleum potential of Cauca–Patía Basin. IX
Agnini, C., Fornaciari, E., Raffi, I., Catanzariti, R., Pälike, H., Back- Simposio Bolivariano de Exploración en Cuencas Subandinas.
man, J. & Rio, D. 2014. Biozonation and biochronology of Abstracts CD ROM. Cartagena, Colombia.
Paleogene calcareous nannofossils from low and middle lati- Barrero, D., Pardo, A., Vargas, C.A. & Martínez, J.F. 2007. Colombian
tudes. Newsletters on Stratigraphy, 47(2): 131–181. https://doi. sedimentary basins: Nomenclature, boundaries and petroleum
org/10.1127/0078-0421/2014/0042 geology, a new proposal. Agencia Nacional de Hidrocarburos,
Aleman, A. & Ramos, V.A. 2000. Northern Andes. In: Cordani, U.G., 92 p. Bogotá.
Milani, E.J., Thomaz–Filho, A. & Campos, D.A. (editors), Tec- Bedoya, E.L., Flores, J.A. & Pardo, A. 2013. Nanofósiles calcáreos
tonic evolution of South America. 31st International Geological y bioestratigrafía del Mioceno tardío del SW de la Cuenca
Congress. Proceedings, p. 453–480. Rio de Janeiro, Brazil. Tumaco onshore, Pacífico colombiano. Boletín de Geología,
Arango, J.L. & Ponce, A. 1980. Memoria explicativa: Reseña expli- 35(1): 55–66.
cativa del mapa geológico del departamento de Nariño. Scale Berggren, W.A., Kent, D.V., Swisher III, C.C. & Aubry, M.P. 1995. A
1:400 000. Ingeominas, 40 p. Bogotá. revised Cenozoic geochronology and chronostratigraphy. In:
Aspden, J. 1984. The geology of the Western Cordillera and Pacific Berggren, W.A., Kent, D.V., Aubry, M.P. & Hardenbol, J. (edi-
coastal plain in the Department of Valle del Cauca (sheets 261, tors), Geochronology, time scales and global stratigraphic cor-
278, 279, 280 and 299). Ingeominas–Misión Británica, Inter- relation. SEPM Special Publications 54: p. 129–212. https://
nal report 1959, 61 p. Cali. doi.org/10.2110/pec.95.04.0129
Aspden, J. & Nivia, A. 1985. Memoria explicativa: Reseña explicati- Bermúdez, H.D., García, J., Stinnesbeck, W., Keller, G., Rodríguez,
va del mapa geológico preliminar de la plancha 278 bahía de J.V., Hanel, M., Hopp, J., Schwarz, W., Trieloff, M., Bolívar,
Buenaventura. Scale 1:100 000. Ingeominas, 13 p. Bogotá. L. & Vega, F.J. 2016. The Cretaceous – Paleogene boundary

241
PARDO–TRUJILLO et al.

at Gorgonilla Island, Colombia, South America. Terra Nova, Echeverri, S., Cardona, A., Pardo, A., Mosalve, G., Valencia, V.A., Borre-
28(1): 83–90. https://doi.org/10.1111/ter.12196 ro, C., Rosero, S. & López, S. 2015b. Regional provenance from
Blow, W.H. 1969. Late middle Eocene to recent planktonic foramin- southwestern Colombia fore–arc and intra–arc basins: Implica-
iferal biostratigraphy. In: Bronnimann, P. & Renz, H.H. (ed- tions for middle to late Miocene orogeny in the northern Andes.
itors), Proceedings of the First International Conference on Terra Nova, 27(5): 356–363. https://doi.org/10.1111/ter.12167
Planktonic Microfossils, 1, p. 199–422. Geneva, Switzerland. Echeverri, S., Pardo–Trujillo, A., Borrero, C., Cardona, A., Rosero,
Borrero, C., Pardo, A., Jaramillo, C.M., Osorio, J.A., Cardona, A., S., Celis, S.A. & López, S.A. 2016. Estratigrafía del Neóge-
Flores, A., Echeverri, S., Rosero, S., García, J. & Casti- no superior al sur de la Cuenca Tumaco (Pacífico colom-
llo, H. 2012. Tectonostratigraphy of the Cenozoic Tumaco biano): La Formación Cascajal, propuesta de redefinición
Forearc Basin, Colombian Pacific, and its relationship with litoestratigráfica. Boletín de Geología, 38(4): 43–60. http://
the northern Andes orogenic build up. Journal of South Amer- dx.doi.org/10.18273/revbol.v38n4-2016003
ican Earth Sciences, 39: 75–92. https://doi.org/10.1016/j. Escovar, R., Gómez, L.A. & Ramírez, J.R. 1992. Interpretación de la
jsames.2012.04.004 sísmica Tumaco 90 y evaluación preliminar del área: Bogotá.
Cediel, F., Restrepo, I., Marín–Cerón, M.I., Duque–Caro, H., Cuar- Empresa Colombiana de Petróleos, unpublished report, 74
tas, C., Mora, C., Montenegro, G., García, E., Tovar, D. & p. Bogotá.
Muñoz, G. 2009. Geology and hydrocarbon potential, Atrato Folk, R. 1974. Petrology of sedimentary rocks. Hemphill Publishing
and San Juan Basins, Chocó (Panamá) Arc. Tumaco Basin Company, 182 p. Austin, USA.
(Pacific realm), Colombia. Agencia Nacional de Hidrocarbu- García, J. 2012. Procedencia de los sedimentos neógenos de Tumaco
ros and Fondo Editorial Universidad EAFIT, 172 p. Medellín, onshore y Gorgona: Una contribución a la tectónica de antear-
Colombia. co del Pacífico sur de Colombia. Master thesis, Universidad
Cohen, K.M., Finney, S.C., Gibbard, P.L. & Fan, J.X. 2013 (updated de Caldas, 82 p. Manizales, Colombia.
2017/02). The ICS International Chronostratigraphic Chart. Gehrels, G.E., Valencia, V. & Pullen, A. 2006. Detrital zircon geochro-
Episodes, 36(3): 199–204. nology by laser–ablation multicollector ICPMS at the Arizona
Cortés, J.E., Mejía−Molina, A., Morton, A., Vargas, C.A. & Cortés, LaserChron Center. In: Loszewski, T. & Huff, W. (editors),
S.I. 2019. Provenance, tectonic setting, and weathering of Geochronology: Emerging opportunities. Paleontology Soci-
sediments in Tumaco−1 ST−P well, Tumaco Forearc Basin, ety Short Course: Paleontology Society Papers, 11, 10 p.
Colombia: Insights from petrography, heavy minerals, X− Giraldo–Villegas, C.A., Celis, S.A., Rodríguez–Tovar, F.J., Pardo–Tru-
ray diffraction, and whole−rock chemostratigraphy. Journal jillo, A., Vallejo–Hincapié, D.F. & Trejos–Tamayo, R.A. 2016.
of South American Earth Sciences, 96: 102219. https://doi. Ichnological analysis of the upper Miocene in the ANH–Tuma-
org/10.1016/j.jsames.2019.102219 co–1–ST–P well: Assessing paleoenvironmental conditions at
Dickinson, W.R. & Gehrels, G.E. 2009. Use of U–Pb ages of detri- the Tumaco Basin, in the Colombian Pacific. Journal of South
tal zircons to infer maximum depositional ages of strata: A American Earth Sciences, 71: 41–53. https://doi.org/10.1016/j.
test against a Colorado Plateau Mesozoic database. Earth and jsames.2016.06.008
Planetary Science Letters, 288(1–2): 115–125. https://doi. Gómez, J., Montes, N.E., Nivia, Á. & Diederix, H., compilers. 2015.
org/10.1016/j.epsl.2009.09.013 Geological Map of Colombia 2015. Scale 1:1 000 000. Ser-
Echeverri, S. 2012. Estratigrafía y análisis de procedencia del inter- vicio Geológico Colombiano, 2 sheets. Bogotá. https://doi.
valo Eoceno – Plioceno de la cuenca de forearc de Tumaco org/10.32685/10.143.2015.936
“onshore” (SW de Colombia) y su relación con la tectónica González, H., compiler. 2008. Investigación integral del Andén Pa-
de la margen pacífica del NW de Suramérica. Master thesis, cífico colombiano. Tomo 1: Geología; Cuenca de Tumaco.
Universidad de Caldas, 127 p. Manizales. Igac–Ingeominas. p. 85–86. Bogotá.
Echeverri, S., Borrero, C., Moreno, M., Pardo–Trujillo, A. & Castillo, Hall, M.L. & Wood, C.A. 1985. Volcano–tectonic segmentation of the
H. 2011. Nomenclatura estratigráfica para la sucesión neó- northern Andes. Geology, 13(3): 203–207. https://doi.org/10.1
gena expuesta en la bahía de Tumaco, Cuenca Tumaco costa 130/0091-7613(1985)13<203:VSOTNA>2.0.CO;2
adentro, SW de Colombia: Redefinición y formalización. XIV Haq, B.U., Hardenbol, J. & Vail, P.R. 1987. Chronology of fluctuating
Congreso Latinoamericano de Geología y XIII Congreso Co- sea levels since the Triassic. Science, 235: 1156–1167.
lombiano de Geología. Abstracts, p. 364. Medellín, Colombia. Instituto Colombiano de Geología y Minería & Instituto Geográfico
Echeverri, S., Cardona, A., Pardo–Trujillo, A., Borrero, C., Rosero, S. Agustín Codazzi. 2005. Mapa geológico de las planchas 361
& López, S. 2015a. Correlación y geocronología Ar–Ar del Bis, 383 Tumaco, 384 La Chorrera, 407 Cabo Manglares y 408
basamento cretácico y el relleno sedimentario Eoceno supe- Barbacoas. Scale 1:100 000. Bogotá.
rior – Mioceno (Aquitaniano inferior) de la cuenca de antearco Jaillard, E., Lapierre, H., Ordoñez, M., Toro–Álava, J., Amórtegui, A.
de Tumaco, SW de Colombia. Revista Mexicana de Ciencias & Vanmelle, J. 2009. Accreted oceanic terranes in Ecuador:
Geológicas, 32(2): 179–189. Southern edge of the Caribbean Plate? Geological Society of

242
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
London, Special Publications 328, p. 469–485. https://doi. Nelson, H.W. 1962. Contribución al conocimiento de la cordillera
org/10.1144/SP328.19 Occidental: Sección carretera Cali–Buenaventura. Boletín
Jaramillo, C., Rueda, M. & Torres, V. 2011. A palynological zonation Geológico, 10(1–3): 81–108.
for the Cenozoic of the Llanos and Llanos Foothills of Colom- Nivia, A., Pérez, C. & Sepúlveda, J. 2003. Memoria explicativa: Geo-

Neogene
bia. Palynology, 35(1): 46–84. https://doi.org/10.1080/01916 morfología y geología de la plancha 383 Tumaco. Scale: 1:100
122.2010.515069 000. Ingeominas, 39 p. Cali, Colombia.
Kennett, J.P. & Srinivasan, M.S. 1983. Neogene planktonic foramin- Okada, H. & Bukry, D. 1980. Supplementary modification and in-
ifera: A phylogenetic atlas. Hutchinson Ross Publishing Com- troduction of code numbers to the low–latitude coccolith

Paleogene
pany. 265 p. biostratigraphic zonation (Bukry, 1973; 1975). Marine Mi-
Kochelek, E.J., Amato, J.M., Pavlis, T.L. & Clift, P.D. 2011. Flysch cropaleontology, 5: 321–325. https://doi.org/10.1016/0377-
deposition and preservation of coherent bedding in an accre- 8398(80)90016-X
tionary complex: Detrital zircon 111 ages from the Upper Cre- Oppenheim, V. 1949. Geología de la costa sur del Pacífico de Colom-
taceous Valdez Group, Chugach Terrane, Alaska. Lithosphere, bia. Boletín Instituto Geofísico de los Andes Colombianos,
3(4): 265–274. (1), Serie C: 4–23.
León, L.A., Padilla, L.E. & Marulanda, N. 1973. Geología, recursos Pardo–Casas, F. & Molnar, P. 1987. Relative motion of the Nazca
minerales y geoquímica de la parte NE del cuadrángulo 0–5 El (Farallon) and South American Plates since Late Cretaceous
Bordo, departamento del Cauca. Ingeominas, Internal report time. Tectonics, 6(3): 233–248. https://doi.org/10.1029/
1652, 235 p. Popayán. TC006i003p00233
López, P.A., Pardo, A., Ochoa, D. & Rosero, S. 2012. Palinología de Perch–Nielsen, K. 1985. Mesozoic calcareous nannofossils. In: Bol-
los sedimentos recientes del río Mira, Pacífico colombiano. li, H., Saunders, J.B. & Perch–Nielsen, K. (editors), Plankton
Geología Colombiana, 37(1): 21–22. stratigraphy. Cambridge University Press, p. 329–426.
López–Ramos, E. 2009. Evolution tectono–stratigraphique du doble Pindell, J.L. & Kennan, L. 2009. Tectonic evolution of the Gulf of
basin avant–arc de la marge convergente Sud Colombienne– Mexico, Caribbean and northern South America in the mantle
Nord Equatorienne pedant le Cenozoïque. Doctorade thesis, reference frame: An update. In: James, K.H., Lorente, M.A. &
Université de Nice Sophia Antipolis, 349 p. Nice, France. Pindell, J.L. (editors), The origin and evolution of the Carib-
Lourens, L.J., Hilgen, F.J., Shackleton, N.J., Laskar, J. & Wilson, D. bean Plate. Geological Society of London, Special Publication
2004. The Neogene Period. In: Gradstein, F.M., Ogg, J.G. & 328, p. 1–55. https://doi.org/10.1144/SP328.1
Smith, A.G. (editors), Geological time scale 2004, Cambridge Raffi, I., Backman, J., Fornaciari, E., Pälike, H., Rio, D., Lourens, L.
University Press, pp. 409–440. Cambridge, UK. & Hilgen, F. 2006. A review of calcareous nannofossil astro-
Ludwing, K.R. 2003. User´s manual for Isoplot 3.00: A geochrono- biochronology encompassing the past 25 million years. Qua-
logical toolkit for Microsoft Excel. Berkeley Geochronology ternary Science Reviews, 25(23–24): 3113–3137. https://doi.
Center, Special Publication (4), 70 p. org/10.1016/j.quascirev.2006.07.007
Mange, M.A. & Maurer, H.F.W. 1992. Heavy minerals in colour. Ramos, V.A. 1999. Plate tectonic setting of the Andean Cordillera.
Chapman & Hall, 147 p. London, UK. Episodes, 22(3): 183–190.
Marcaillou, B. & Collot, J.Y. 2008. Chronostratigraphy and tectonic Robertson Research inc. & Empresa Colombiana de Petróleos. 1981a.
deformation of the north Ecuadorian–South Colombian off- Geochemical and biostratigraphic analysis of Wainoco No.1,
shore Manglares Forearc Basin. Marine Geology, 255(1–2): Majagua, Colombia. Internal report 413, 61 p. Houston, USA.
30–44. https://doi.org/10.1016/j.margeo.2008.07.003 Robertson Research inc. & Empresa Colombiana de Petróleos. 1981b.
Marín–Cerón, M.I. & Sierra, G. 2011. Tumaco Basin. In: Cediel, F. Biostratigraphic, geochemical and petrological analysis of
(editor), Petroleum geology of Colombia: Geology and hydro- samples from Arco well Remolino Grande–1, Colombia. In-
carbon potential, 13. Agencia Nacional de Hidrocarburos and ternal report 318, 75 p. Houston, USA.
Universidad EAFIT, 77 p. Medellín. Serrano, L., Ferrari, L., López–Martínez, M., Petrone, C.M. & Jarami-
Martini, E. 1971. Standard Tertiary and Quaternary calcareous nan- llo, C. 2011. An integrative geologic, geochronologic and geo-
noplankton zonation. Proceedings 2nd Planktonic Conference, chemical study of Gorgona Island, Colombia: Implications for
Proceedings, 739–785. Roma, Italy. the formation of the Caribbean Large Igneous Province. Earth
Moreno–Sánchez, M. & Pardo–Trujillo, A. 2003. Stratigraphical and and Planetary Science Letters, 309(3–4): 324–336. https://doi.
sedimentological constraints on western Colombia: Implica- org/10.1016/j.epsl.2011.07.011
tions on the evolution of the Caribbean Plate. In: Bartolini, Somoza, R. 1998. Updated Nazca (Farallon)–South America relative
C., Buffler, R.T. & Blickwede, J. (editors), The circum–Gulf motions during the last 40 My: Implications for mountain
of Mexico and the Caribbean: Hydrocarbon habitats, basin building in the central Andean region. Journal of South Amer-
formation, and plate tectonics. American Association of Pe- ican Earth Sciences, 11(3): 211–215. https://doi.org/10.1016/
troleum Geologists Memoir 79, p. 891–924. Tulsa, USA. S0895-9811(98)00012-1

243
PARDO–TRUJILLO et al.

Somoza, R. & Ghidella, M.E. 2012. Late Cretaceous to recent plate Suárez, M.A. 2007. Geological framework of the Pacific coast sedi-
motions in western South America revisited. Earth and mentary basins. Geología Colombiana, 32: 47–62.
Planetary Science Letters, 331–332: 152–163. https://doi. Vallejo, C., Winkler, W., Spikings, R.A., Luzieux, L., Heller, F. &
org/10.1016/j.epsl.2012.03.003 Bussy, F. 2009. Mode and timing of terrane accretion in the
Spadea, P. & Espinosa, A. 1996. Petrology and chemistry of Late Cre- forearc of the Andes in Ecuador. In: Kay, S.M., Ramos, V.A.
taceous volcanic rocks from the southernmost segment of the & Dickinson, W.R. (editors), Backbone of the Americas: Shal-
Western Cordillera of Colombia (South America). Journal of low subduction, plateau uplift, and ridge and terrane collision.
South American Earth Sciences, 9(1–2): 79–90. https://doi. Geological Society of America, Memoirs 204, p. 197–216.
org/10.1016/0895-9811(96)00029-6 https://doi.org/10.1130/2009.1204(09)
Spikings, R. & Crowhurst, P. 2004. (U–Th)/He thermochronometric van der Hammen, T. 1958. Estratigrafía del Terciario y Maastrichtiano
constraints on the late Miocene – Pliocene tectonic develop- continentales y tectogénesis de los Andes colombianos. Bo-
ment of the northern Cordillera Real and the Interandean De- letín Geológico, 6(1–3): 67–128.
pression, Ecuador. Journal of South American Earth Sciences, van der Hammen, T., Werner, J.H. & van Dommelen, H. 1973.
17: 239–251. Palynological record of the upheaval of the northern An-
Spikings, R.A. & Simpson, G. 2014. Rock uplift and exhumation of des: A study of the Pliocene and lower Quaternary of the
continental margins by the collision, accretion, and subduction Colombian Eastern Cordillera and the early evolution of
of buoyant and topographically prominent oceanic crust. Tec- its high–Andean biota. Review of Palaeobotany and Pal-
tonics, 33(5): 635–655. https://doi.org/10.1002/2013TC003425 ynology, 16(1–2): 1–122. https://doi.org/10.1016/0034-
Spikings, R.A., Winkler, W., Hughes, R.A. & Handler, R. 2005. 6667(73)90031-6
Thermochronology of allochthonous terranes in Ecua- Villagómez, D., Spikings, R., Magna, T., Kammer, A., Winkler, W. &
dor: Unravelling the accretionary and post–accretionary Beltrán, A. 2011. Geochronology, geochemistry and tectonic
history of the northern Andes. Tectonophysics, 399(1–4): evolution of the Western and Central Cordilleras of Colom-
195–220. bia. Lithos, 125(3–4): 875–896. https://doi.org/10.1016/j.lith-
Spikings, R., Crowhurst, P.V., Winkler, W. & Villagómez, D. 2010. Syn– os.2011.05.003
and post–accretionary cooling history of the Ecuadorian Andes Wade, B.S., Pearson, P.N., Berggren, W.A. & Pälike, H. 2011. Re-
constrained by their in–situ and detrital thermochronometric view and revision of Cenozoic tropical planktonic foramin-
record. Journal of South American Earth Sciences, 30(3–4): iferal biostratigraphy and calibration to the geomagnetic
121–133. https://doi.org/10.1016/j.jsames.2010.04.002 polarity and astronomical time scale. Earth–Science Re-
Suárez, M.A. 1990. Estudio geológico regional e interpretación sís- views, 104(1–3): 111–142. https://doi.org/10.1016/j.earsci-
mica estratigráfica en la provincia sedimentaria de la costa rev.2010.09.003
Pacífica. Ecopetrol, unpublished report, 119 p. Bogotá.

Explanation of Acronyms, Abbreviations, and Symbols:


ANH Agencia Nacional de Hidrocarburos LA–ICPMS Laser ablation inductively coupled plasma mass
AP–T Andrés Pardo–Trujillo spectrometry
FAD First Appearance Datum TD Total depth

244
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
Authors’ Biographical Notes Alejandro ARENAS is a geologist of the
Departamento de Ciencias Geológicas at
Andrés PARDO–TRUJILLO is a ge- the Universidad de Caldas (Manizales,
ologist in the Departamento de Ciencias Colombia) and has an MS in earth sci-

Neogene
Geológicas at the Universidad de Cal- ences from the Universidad de Caldas
das (Manizales, 1998). Andres obtained (Manizales, Colombia). Alejandro has
his MS in vegetal micropaleontology worked in REDCOM as a profession-
in 1997 and his PhD in Science from al for Ecopetrol in the field of reservoir

Paleogene
Liège University (Belgium, 2004). He development. Alejandro is currently a
has worked since 1989 as a profes- researcher at the Instituto de Investiga-
sor in the Departamento de Ciencias ciones en Estratigrafía (IIES) and belongs to the Grupo de Investigación
Geológicas at Universidad de Caldas, en Estratigrafía y Vulcanología (GIEV) Cumanday. He has strong capa-
Colombia, in sedimentology, regional geology, palynology, and field bilities in 3D modeling and seismic interpretation.
geology. From 2006–2009, he worked as an advisor at the Agen-
cia Nacional de Hidrocarburos (ANH), where he participated in the Felipe VALLEJO is a geologist from
geological study of several Colombian sedimentary basins. Andres the Universidad de Caldas (Colombia),
is currently the director of the Instituto de Investigaciones en Es- MS and PhD student in geology at Uni-
tratigrafía (IIES) and of the Grupo de Investigación en Estratigrafía y versidad de Salamanca (Spain). Felipe
Vulcanología (GIEV) at the Universidad de Caldas. His main interest works as a researcher and consultant
is the study of the geological and biological evolution of the NW at the Instituto de Investigaciones en
corner of South America during the Cretaceous – Cenozoic. Estratigrafía (IIES) of the Universidad
de Caldas (Colombia). Felipe is also
Sebastián ECHEVERRI is a geologist part of the Grupo de Investigación en
from the Universidad de Caldas (Co- Estratigrafía y Vulcanología (GIEV)
lombia), with an MS in earth sciences Cumanday. His work has been focused on building a biostratigraphic
from the same university and a PhD in and biochronologic framework based on calcareous nannofossils for
science–geotectonics from the Univer- the Caribbean and Pacific Basins of Colombia.
sidade de São Paulo (Brasil). Sebastián
works as a researcher in the Grupo de Raúl TREJOS is a geologist from the
Investigación en Estratigrafía y Vulca- Universidad de Caldas (Colombia) and
nología (GIEV) Cumanday – Institu- has an MS in earth sciences from the
to de Investigaciones en Estratigrafía Universidad de Salamanca (Spain). He
(IIES) at Universidad de Caldas (Colombia) and is part of the EGEO is currently a PhD student in geology
research group of the Universidad Nacional (Medellín). His main at the Universidad de Salamanca. Since
research area is the regional geology and tectono–stratigraphic evo- 2013, Raúl has worked as a researcher
lution of convergent margins. at the Instituto de Investigaciones en
Estratigrafía (IIES) – Universidad de
Carlos BORRERO attained a BS in Caldas and belongs to the Grupo de In-
geology from Universidad Nacional de vestigación en Estratigrafía y Vulcanología (GIEV) Cumanday. He
Colombia (Bogotá, 1983). Stratigraphy, is recognized as a junior researcher by the Ministerio de Ciencias,
volcanology, field geology, and hazard Tecnología e Innovación in Colombia. His main experience is in mi-
and risk assessment and management cropaleontology (foraminifera).
have been the focus of his long career
in academia at Universidad de Caldas
(Manizales, 1985–2015) and as a con-
sultant geologist (1991–present). Cur-
rently, he is semiretired.

245
PARDO–TRUJILLO et al.

Ángelo PLATA obtained his bachelor’s regional tectonics of convergent margins, especially the Andean and
degree in biology from the Universi- Caribbean Orogens.
dad Industrial de Santander, Colombia,
in 2007. He worked at the Colombian Sergio RESTREPO is a geologist from
oil company Ecopetrol for 5 years as a Universidad Nacional de Colombia. He
junior biostratigrapher before going to holds a PhD in geology and geography
Spain in 2012 to obtain his Master of from the Department of Geological
Earth Science degree from the Depart- Sciences at the University of Florida
ment of Geology, Universidad de Sala- (UF, 2009) and is currently an associate
manca. After returning from Universidad professor in the Departamento de Geo-
de Salamanca, he began his current role as palynologist in chief of ciencias y Medio Ambiente at Universi-
the Palynology Lab for Instituto de Investigaciones en Estratigrafía, dad Nacional de Colombia and adjunct
Universidad de Caldas, where he studies terrestrial palynomorphs and postdoctoral researcher at University of
marine diatoms. Currently, Angelo is a candidate for his PhD at Uni- Florida. Sergio completed two postdoctoral studies under the auspices
versidad de Salamanca. of the Smithsonian Institution (2009–2010) and the National Science
Foundation (2011). During his doctoral and postdoctoral work, Sergio
José–Abel FLORES is a geologist received research support from GSA, AGI, NSF, and the COMPTON
in the Department of Geology of the Foundation, among others. Since 2010, Sergio has collaborated in
Universidad de Salamanca (Spain) and several interdisciplinary projects carried out by the Instituto de Inves-
obtained his PhD in sciences at the tigaciones en Estratigrafía (Universidad de Caldas, Colombia) while
Universidad de Salamanca in 1997. A also developing some investigations supported by Colciencias and the
specialist in calcareous nannofossils, he National Geographic Society. Sergio's work in geosciences concen-
has been a full professor of micropale- trates on the use of geochronology, thermochronology, and isotopic
ontology and oceanography since 2006 tools to understanding morphotectonic process over various spatio-
and an invited professor in several in- temporal scales, as well as on the development of educational tools to
stitutions in Europe, the Americas, and advance earth and environmental literacy among rural communities.
Japan. He centered his research in biostratigraphy and paleoceanog- Geographically, Sergio’s work focuses on the Andes, the Caribbean,
raphy, participating in five ODP/IOPD expeditions and working with and Central America.
different hydrocarbon companies (Repsol, Gas Natural, Petrobras,
Ecopetrol, Exxon). His interest and expertise cover the Cretaceous Ángel BARBOSA is a Colombian ge-
to present–day environments, combining microfossil and biogeo- ologist who obtained his bachelor’s de-
chemistry information. In recent years, he has participated in several gree from the Universidad de Caldas in
Colombian projects with the Universidad de Caldas, particularly the 2009 and an MS in geological sciences
paleoceanographic evolution of tertiary basins, the consequences of in 2012. His master’s thesis focused on
the communication between the Atlantic and the Pacific, and the defi- low–temperature thermochronology ap-
nition of a high–resolution astronomically calibrated scale based on plied to understand the thermotectonic
micro and calcareous nannofossils. coevolution of the south part of the Co-
lombian Western Cordillera and the Tu-
Agustín CARDONA is graduated maco Basin. He has worked on several
with a degree in geology from Univer- projects in Colombia involving geological mapping, sampling, and the
sidad EAFIT in 1999. He obtained MS use of analytic techniques such as mass spectrometry to obtain thermo
and PhD degrees in geochemistry and and geochronologic data. Angel is currently a PhD candidate at the
geotectonics, respectively, from the University of Florida, where he has served as a TA, teaching several
Universidade de São Paulo, Brasil. Sub- courses, including physical geology, engineering and environmental
sequently, he was a postdoctoral fellow geology, and introduction to earth science. His research interests in-
at the Smithsonian Tropical Research clude understanding the Cenozoic tectonic evolution of the NW Andes
Institute (2006–2011), working in sev- using the magmatic and sedimentary record within sedimentary basins
eral paleogeogeographic–oriented proj- and cordilleran massifs and investigating the role of sediments in the
ects in northern Colombia and Panamá. origin and evolution of continental crust.
Since 2012, he has served as full–time professor at the Universidad
Nacional de Colombia in Medellín. His main interest is related to

246
Cenozoic Geologic Evolution of the Southern Tumaco Forearc Basin (SW Colombian Pacific)

Quaternary
Hugo MURCIA is a geologist from the Sergio CELIS is a geologist from the
Universidad de Caldas, Colombia, who Universidad de Caldas (Colombia),
received his MS in earth sciences from has an MS in earth sciences from the
the Universidad Nacional Autónoma de same university, and received his PhD

Neogene
México, México, and his PhD in geol- in earth sciences from the University of
ogy from the University of Auckland, Granada (Spain). Sergio is a researcher
New Zealand. Hugo works as a pro- at the Instituto de Investigaciones en
fessor at the Departamento de Ciencias Estratigrafía (IIES) and has experience

Paleogene
Geológicas as a researcher at the Insti- as a professor in the Departamento de
tuto de Investigaciones en Estratigrafía Ciencias Geológicas at the Universidad
(IIES), both at Universidad de Caldas, Colombia. Hugo also belongs de Caldas. Sergio also belongs to the Grupo de Investigación en Es-
to the Grupo de Investigación en Estratigrafía y Vulcanología (GIEV) tratigrafía y Vulcanología (GIEV) Cumanday and is recognized as a ju-
Cumanday and is recognized as an associate researcher by the Minis- nior researcher by the Ministerio de Ciencias, Tecnología e Innovación
terio de Ciencias, Tecnología e Innovación in Colombia. He specializes in Colombia. His main experience is in ichnology and basin analysis.
in volcanology and petrology.
Sergio A. LÓPEZ is graduated with a
Carlos GIRALDO is a geologist of the bachelor's degree in geology from the
Departamento de Ciencias Geológicas Universidad de Caldas, Colombia, and
de la Universidad de Caldas (Maniza- received a master's degree in earth and
les, Colombia, 2014) and a student of environmental sciences from the Uni-
the Master of Earth Sciences program versidad EAFIT, Medellín, Colombia,
at the Universidad de Caldas. During with an emphasis on morphodynamics
2014–2015, he worked at the Instituto of tropical deltaic systems. He current-
de Investigaciones Marinas y Costeras ly holds a doctoral candidate position in
José Benito Vives de Andréis (INVE- the Department of Geological Sciences
MAR) in areas related to marine geolo- from the Universidad de Buenos Aires, Argentina. He has worked as
gy. Since 2015, he has been linked to the Instituto de Investigaciones a researcher in interdisciplinary projects on modern deltas (Univer-
en Estratigrafía (IIES) and the Grupo de Investigación en Estratigrafía sidad EAFIT, 2004–2008), active tectonics and geodynamics (Servi-
y Vulcanología (GIEV) Cumanday. His research field is related to cio Geológico Colombiano, 2008–2010), and as advisor of the new
ichnology, sedimentology, and stratigraphy, and its integration into ventures technical team from the Agencia Nacional de Hidrocarburos
sedimentary basin analysis to reconstruct ancient depositional environ- (2011–2017).
ments and interpret their possible sequence stratigraphic framework.

247

You might also like