Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Journal of South American Earth Sciences 112 (2021) 103515

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


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

Sedimentary facies of a tide-dominated estuary and deltaic complex in a


tropical setting: The middle Miocene Oficina Formation of the Orinoco Oil
Belt, Venezuela
Eurídice J. Solórzano a, b, *, Luis A. Buatois a, Williams J. Rodríguez b, M. Gabriela Mángano a
a
Department of Geological Sciences, University of Saskatchewan, 114 Science Place, Saskatoon, SK S7N 5E2, Canada
b
PDVSA-INTEVEP, Urbanización Santa Rosa, Sector El Tambor, Los Teques, Estado Miranda, Apartado 76343, Caracas, 1070A, Venezuela

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

Keywords: Although the middle Miocene Oficina Formation of the Orinoco Oil Belt represents most of Venezuela’s hy­
Fluvial braided channels drocarbon resource, a comprehensive and detailed sedimentary facies model for the whole belt has never been
Meandering estuary channels put forward. Based on the analysis of cores and well logs, nine sedimentary facies (FA-I), forming five facies
Deltaic distributary channels
assemblages (FA1-5), have been characterized. Both sedimentologic and ichnologic datasets have been inte­
Ichnofacies
Marine-marginal
grated in this study. The Oficina Formation records the typical succession of fluvial incision during relative sea-
level fall, fluvial deposition during lowstand, transition from fluvial to estuary valleys during the subsequent
transgression and deltaic progradation during highstand. FA1 consists of fluvial-braided channel (FB), floodplain
(FG2), and swamp (FH1) deposits, as well as paleosols (FG3). This facies assemblage occurs in the lower member,
representing the infill of lowstand fluvial valleys. FA2 consists of meandering estuary-channel deposits (FA, FC,
FD, FE, and FI). This facies assemblage occurs in the middle member, representing the infill of tide-dominated
estuary valleys during the early stages of the Langhian transgression. FA3 consists of tidal-flat and tidal-creek
(FC, FD, FE, FF, and FG2), and swamp (FH1 and FH2) deposits, together with paleosols (FG3). This facies
assemblage is present in the middle member, revealing backstepping and retrogradation within the estuary
system during a continuing transgression. FA4 consists of outer-estuary sandbar (FC, FD, and FG1); swamp (FH2)
deposits and paleosols are present at the margins of the estuary (FG3). This facies assemblage occurs in the
uppermost part of the middle members, representing a late stage of the Langhian transgression, culminating in a
maximum flooding surface. FA5 consists of deltaic distributary channel (FC and FD), floodplain and
interdistributary-bay (FG2), and swamp (FH1) deposits. This facies assemblage occurs in the upper member,
recording sedimentation in the delta plain of a tide-dominated delta during a highstand. Freshwater conditions in
the fluvial system, as well as in the inner portions of the estuary and the delta plain are further supported by the
presence of the Scoyenia Ichnofacies, whereas brackish-water segments of the estuary are characterized by the
Skolithos and depauperate Cruziana Ichnofacies. The substrate-controlled Teredolites, and Glossifungites Ichnofa­
cies occur in connection to erosional exhumation during ravinement. The absence of fully marine ichnofaunas is
consistent with the embayed nature of the Orinoco Oil Belt. The tropical and humid character of the Oficina
depositional systems is manifested in the development of widespread wetlands in marginal-marine settings, with
formation of swamps and embayed areas, typically displaying evidence of waterlogged paleosols with pervasive
root trace fossils. The abundance of crustacean burrows, such as Ophiomorpha and Thalassinoides, is also typical of
low-latitude marginal- and shallow-marine settings. Although tidal forces tend to be weaker in lower latitudes
because Coriolis effects are stronger there, both the estuary and delta systems are regarded as tide-dominated,
most likely reflecting the embayed nature of the shoreline.

* Corresponding author. Department of Geological Sciences, University of Saskatchewan, 114 Science Place, Saskatoon, SK S7N 5E2, Canada.
E-mail addresses: Solorzanoe@pdvsa.com (E.J. Solórzano), luis.buatois@usask.ca (L.A. Buatois), rodriguezwq@pdvsa.com (W.J. Rodríguez), gabriela.mangano@
usask.ca (M.G. Mángano).

https://doi.org/10.1016/j.jsames.2021.103515
Received 18 March 2021; Received in revised form 5 August 2021; Accepted 6 August 2021
Available online 14 August 2021
0895-9811/© 2021 Elsevier Ltd. All rights reserved.
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

1. Introduction approximately 55,314 km2, being located in the southern part of the
Eastern Venezuela Basin, sub-parallel to the Orinoco River (Fig. 1A). The
The oil sandstone deposits of the Orinoco Oil Belt represent most of Hato Viejo fault system subdivides the Orinoco Oil Belt into two prov­
Venezuela’s hydrocarbon resource (Magna Reserva Project, 2012). The inces, the western and eastern provinces (Latreille et al., 1983; Aude­
middle Miocene Oficina Formation in the Orinoco Oil Belt has been mard et al., 1985). The western province consists of the Boyaca and
traditionally considered to record deposition within fluvio-deltaic sys­ Junín areas where the Cenozoic succession overlies Cretaceous and
tems (Latreille et al., 1983; Audemard et al., 1985; Toro et al., 2001; Paleozoic strata. The eastern province includes the Carabobo and Aya­
Martinius et al., 2012, 2013, 2013; Huang et al., 2020). However, work cucho areas, where the Cenozoic succession rests on top of the Pre­
restricted to the Junín and Boyacá areas suggested a more nuanced cambrian basement (Fig. 2).
depositional setting based on the fact that the deltaic interpretation is Based on an integrated study of foraminifers, calcareous nanno­
not consistent with the retrogradational stacking pattern recorded in its plankton and palynomorphs, the Oficina Formation is considered of
middle member (Rodríguez et al., 2018). In particular, deposition in middle Miocene age (Audemard et al., 1985; Solórzano et al., 2015). The
fluvial to tide-dominated estuary settings has been proposed for the Oficina Formation in the Orinoco Oil Belt was divided into three
lower and middle members, whereas deltaic deposition has been sug­ informal members: lower, middle and upper, representing fluvial, es­
gested to be restricted to the upper member (Rodríguez et al., 2018). tuary and deltaic deposits, respectively (Solórzano et al., 2017; Rodrí­
This model is consistent with ichnologic evidence obtained in the Ori­ guez et al., 2018) (Fig. 3).
noco Oil Belt, as well as in the adjacent Oritupano Field (Solórzano et al., Integration of biostratigraphic information with sedimentologic and
2017). Regrettably, despite its economic importance, a comprehensive sequence-stratigraphic data allows interpreting the Oficina Formation as
and detailed sedimentary facies model for the Oficina Formation along a 2nd-order depositional sequence (Rodríguez et al., 2018). Based on the
the whole Orinoco Oil Belt has never been put forward. This is unfor­ presence of the nanofossil Helicosphaera ampliaperta, the NN4 zone was
tunate because a proper understanding of the variability of sedimentary identified, and is regarded as containing maximum flooding surface
facies in these tide-dominated depositional environments is important (MFS-1) (Audemard et al., 1985; Solórzano et al., 2015; Solórzano and
for both exploration strategies and reservoir characterization. Charac­ Farías, 2017). This surface extends all across the Orinoco Oil Belt,
teristic sedimentary structures of the Oficina Formation, such as inclined delineating the contact between retrogradational transgressive estuary
heterolithic stratification (IHS) (Hori et al., 2001; Choi et al., 2004; deposits below and progradational highstand deltaic deposits above
Crerar and Arnott, 2007; Rodríguez et al., 2018; Gingras et al., 2016; (Figs. 3 and 15-18). MFS-1 closely matches the middle Miocene global
Solórzano et al., 2017), commonly impact on reservoir heterogeneity. eustatic curve of Haq and Schutter (2008) (Solórzano et al., 2017).
The purposes of this paper are to: (1) document the different sedimen­ MFS-1 also has been reported outside the Orinoco Oil Belt in the Anaco
tary facies and depositional environments in the Oficina Formation in area of the Eastern Venezuela basin, based on the presence of the fora­
the Orinoco Oil Belt, (2) discuss changes in thickness/general facies minifer Praorbulina glomerosa (Campos et al., 1985; Flores et al., 2001).
relationship, and (3) place facies information within a Relative sea-level changes and regional tectonics played a significant
sequence-stratigraphic framework. The widespread extension of the role on sedimentation in the Orinoco Oil Belt because they controlled
Oficina Formation in the subsurface allows us to delineate the regional erosion and infilling of the incised-valleys, as well as the history of
architecture of these marginal-marine strata, further refining facies changes in accommodation potential in the basin. Additional evidence
models of tide-influenced settings. for the tectonic control is the wedge-shaped geometry and facies ar­
chitecture of sediments (e.g. Obi and Okogbue, 2004). The Orinoco Oil
2. Geologic setting Belt comprises a prism of Cenozoic sediment wedging toward the south
(Latreille et al., 1983; Audemard et al., 1985; Isea, 1987; Parnaud et al.,
Deposition within the Eastern Venezuela Basin can be divided into 1995; Di Croce et al., 1999).
four main phases: (1) a Paleozoic pre-rift phase, (2) a rift-and-drift phase
during Jurassic and earliest Cretaceous time, (3) a Cretaceous-Paleogene 3. Material and methods
passive margin period, and (4) a Paleogene-Quaternary oblique collision
phase (transpression) between the Caribbean and South American Thirty-five cored wells from the Oficina Formation in the Orinoco Oil
plates, during which the foreland basin and the Serranía del Interior Belt were described (B1, B2, and B3 from the Boyacá area, J1, J2, J3, J4,
were formed (Eva et al., 1989; Parnaud et al., 1995; Talwani, 2002). The J5, J6, J7, J12, J13, and J15 from the Junín area, A4, A7, A8, A1, A2,
first phase is represented by Cambrian sandstone and shale deposition, A9, A10, A11, and A12 from the Ayacucho areas, and C1, C2, C3, C6, C7,
as shown by the Hato Viejo and Carrizal formations, respectively. During C8, C9, C10, C11, C12, C13, and C20 from the Carabobo area), totaling
the second phase, red sediments, affected by basaltic sills (La Quinta 3512 m (Fig. 1B). Core-based sedimentary facies were characterized by
Formation), accumulated in the Espino graben during the Jurassic. The identifying lithology, sedimentary structures, textural characteristics,
passive margin phase consists of three major transgressive episodes. The bed and bedset thicknesses, bed contacts, bioturbation index (based on
first of these transgressive episodes took place during the Cretaceous, Taylor and Goldring, 1993), trace-fossil distribution, and ichnologic
and is associated with the formation of the main hydrocarbon source suites (Figs. 4–6). Additionally, 318 wells were analyzed based on re­
rocks. The second transgressive episode took place during the sistivity, density, neutron, and gamma ray logs. These wells were
Paleocene-Eocene. The third transgressive episode of the passive margin correlated to interpret sedimentary facies distributions and to evaluate
phase took place during the early Oligocene. Finally, the collision phase changes in thickness and general facies relationships across the Orinoco
was diachronous, spanning the early-middle Eocene in the west to the Oil Belt (Fig. 1B). Core information was matched with their log
late Oligocene-middle Miocene in the east, and continues to present day. expression. The cores are oriented. Biostratigraphic and biofacies data
Therefore, the Oficina Formation was deposited during this oblique (3000 samples from 96 wells), based on micropaleontologic information
collision phase between the Caribbean and South American plates (calcareous nanoplankton, palynomorphs, foraminifers), have been
(Parnaud et al., 1995), with the area being subjected to strong tectonic analyzed to establish the correlations used in this study (Solórzano et al.,
events that controlled sedimentation (Audemard et al., 1985; Martinius 2015).
et al., 2012).
The Eastern Venezuela Foreland Basin formed on the passive margin 4. Sedimentary facies and trace fossil content
of the South American Craton during the Neogene. This basin comprises
several petroleum provinces, and one of them is represented by the oil Table 1 summarizes the main characteristics of nine sedimentary
fields of the Orinoco Oil Belt (Parnaud et al., 1995). It covers an area of facies (FA-FI) present in the Oficina Formation, forming five facies

2
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 1. Map of Venezuela with the location of the Orinoco Oil Belt and the wells studied. A: Map of Venezuela outlining the location of the Orinoco Oil Belt. Arrow
indicates North. B: Map of the Orinoco Oil Belt showing the location of wells both with and without cores and the main stratigraphic correlations.

associations, labeled FA1 to FA5. trace fossils. Facies A is present in the middle member.
Interpretation: It was formed from the collapse of the associated
muddy cut-bank deposits (Musial et al., 2012; Martinius et al., 2015;
4.1. Facies A: intraclast breccia Moreton and Carter, 2015; Hein, 2015; Gingras and Leckie, 2017;
Brekke et al., 2017). These breccias overlie the erosional bases of the
Description: Facies A consists of 10–50 cm thick, mudstone breccia, estuary channels and indicate the thalweg position of the channels.
medium-to fine-grained sandstone, and mudstone. The sandstone is Facies A delineates the base of meandering estuary channels.
poorly sorted (Fig. 7A). Beds are stacked forming facies intervals up to
1.2 m thick. Both the intraclasts and the associated deposits are barren of

3
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 2. Chronostratigraphic framework and geological section for the Orinoco Oil Belt (modified from Magna Reserva Project, 2012).

Fig. 3. Stratigraphic chart of the Oficina Formation in the Orinoco Oil Belt showing gamma-ray log (GR), member, lithology and sedimentary structures, sub­
environments, depositional environments, sequence stratigraphic surfaces, systems tracts, and 2nd- and 3rd-order sequences.

4.2. Facies B: cross-stratified very coarse-to medium-grained sandstone planar cross-stratified, very coarse-to medium-grained, gravel-rich
sandstone with dispersed granules (Fig. 7B). Subfacies B2 consists of
Description: Facies B consists of trough and planar cross-bedded very 20–70 cm thick, poorly sorted, structureless to trough cross-stratified,
coarse-to medium-grained sandstone, locally displaying a structureless coarse-to medium-grained gravel-rich sandstone.
appearance. Beds are stacked forming facies intervals up to 12 m thick. Interpretation: Facies B shows clear evidence of channel erosion and
No trace fossils occur in this facies. Facies B occurs in the lower member. deposition. These channels are characterized by two and three-
It has been subdivided into subfacies B1 and B2. dimensional dunes that migrated along the river bottoms (Miall, 2010;
Subfacies B1 consists of 10–50 cm thick, poorly sorted, massive to Brekke et al., 2017), representing subfacies B1 and B2, respectively.

4
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 4. Sedimentologic and ichnologic log for well J3, showing vertical distribution of facies assemblages and corresponding paleoenvironmental interpretations.
Location of cored well is shown in Fig. 1B.

5
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 5. Sedimentologic and ichnologic log for well A4, showing vertical distribution of facies assemblages and corresponding paleoenvironmental interpretations.
Location of cored well is shown in Fig. 1B.

6
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 6. Sedimentologic and ichnologic log for well C11, showing vertical distribution of facies assemblages and corresponding paleoenvironmental interpretations.
Location of cored well is shown in Fig. 1B.

7
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Table 1
Sedimentary Facies of the Oficina Formation in the Orinoco Oil Belt (modified from Rodríguez et al., 2018).
Facies Lithology and Dominant Ichnology Thickness Other Interpretation
Texture physical (m) characteristics
sedimentary
structures

FA Mudstone Microfaults and No trace fossils 0.1-0.5 Lag deposits, cut


Intraclast breccia breccia, planar cross bank margins of
medium- to stratification meandering
fine-grained estuary channels
sandstone and
mudstone,
poorly sorted
FB FB1 Massive to Very coarse- to Massive to planar No trace fossils 6-12 Locally parallel- Braided fluvial
Cross-Stratified Planar cross- medium- cross- laminated channels
very coarse-to stratified grained gravel- stratification mudstone,
medium- sandstone with rich sandstone mudstone and coal
grained granules with dispersed clasts, generally oil
sandstone granules, impregnated,
poorly sorted argillaceous
FB2 Massive to Coarse- to Massive to trough No trace fossils 3-6
trough cross- medium- cross-
stratified grained gravel- stratification
sandstone with rich sandstone,
pebbles and poorly sorted
mudstone
intraclasts
FC Medium- to Trough and No trace fossils 10-15 Mud drapes Tidal channels,
Cross-stratified medium- to fine-grained sandstone fine-grained planar cross tidal flat, and tidal
with mud drapes fine sandstone, stratification sandbars
well sorted
FD FD1 Very fine- Inclined Rosselia socialis and 0.01-0.15 Outer-estuary
Inclined heterolithic IHS, very fine-grained grained heterolithic Planolites montanus sandbar complex
stratified coarse- to sandstone and mudstone sandstone, well stratification BI:0-1
fine-grained sorted,
sandstone and mudstone
mudstone FD2 Coarse- to fine- Inclined Ophiomorpha 0.01-0.91 Mud drapes Estuary channel
IHS, coarse- to fine- grained heterolithic nodosa point bars
grained sandstone and sandstone, stratification BI: 0-1
mudstone poorly sorted,
mudstone
FD3 Fine- to very Inclined Beaconites 0.01- Mud drapes Distributary
IHS, fine- to very fine- fine-grained heterolithic antarcticum, 0.018 channel point bars
grained sandstone and sandstone, well stratification Taenidium isp. BI:
mudstone sorted, 4-6
mudstone Teichichnus rectus
BI: 0-1
FD4 Fine-grained Inclined Taenidium isp. (BI: 0.02-0.18 Mud drapes, Tidal flat and creek
IHS, fine-grained sandstone, well heterolithic 4-6) Mudstone complex
sandstone and mudstone sorted, stratification Thalassinoides isp. intraclasts
mudstone BI: 0-1
FE Fine- to very Convolute Scarce 0.5-1.5 Estuary channels
Convoluted fine- to very fine-grained sandstone and fine-grained lamination Ophiomorpha and tidal flats
mudstone sandstone, well nodosa
sorted BI: 0-1
FF Sandstone FF1 Parallel- Fine- to very Wavy lamination Planolites 0.3-4 Mud drapes and Tidal sand to
Interbedded dominated laminated fine-grained (rhythmic montanus, flaser bedding mixed flats
mudstone and sandstone and sandstone, well appearance) Teichichnus rectus,
middle- to very mudstone sorted, rounded Rosselia socialis,
fine-grained clasts; Skolithos linearis,
sandstone mudstone Diplocraterion
habichii, escape
trace fossils
BI: 0-1
FF2 Muddy Fine- to very Sandstone is Ophiomorpha 0.3-4 Tidal sand to
sandstone fine-grained interlaminated to nodosa (BI: 0-1), mixed flats
sandstone interbedded with Thalassinoides isp.,
mixed with light to medium undeterminate
mudstone grey mudstone bioturbation
mottling (BI: 4-5)
Mudstone FF3 Parallel- Mudstone with Sand grains Bergaueria isp., 1-12 Tidal mud flats
dominated laminated scarce very fine dispersed in Planolites
sandy sand grains, mudstone- and montanus,
mudstone and well sorted siltstone- Thalassinoides isp.
siltstone dominated BI: 2-3
intervals
0.3-7 Tidal mud flats
(continued on next page)

8
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Table 1 (continued )
Facies Lithology and Dominant Ichnology Thickness Other Interpretation
Texture physical (m) characteristics
sedimentary
structures

FF4 Silstone and Massive Teichichnus rectus,


Bioturbated mudstone appearance Thalassinoides isp.,
siltstone and indeterminate
mudstone bioturbation
mottling
BI: 4-5
FF5 Calcareous Massive Massive No trace fossils Shell remains Tidal mud flats
massive calcareous appearance
mudstone with mudstone with
scarce scarce
limestone limestone
layers layers
FG FG1 Mudstone Massive to Planolites 3-4 Syneresis cracks, Outer-estuary
Carbonaceous parallel- montanus, abundant organic margin
rooted silty laminated, locally Teichichnus rectus, debris, and scarce
mudstone and current ripples Thalassinoides isp., mudstone clasts
thinly laminated and flaser root trace fossils
mudstone bedding BI: 3-4
FG2 Mudstone and Massive to Beaconites 0.3-1.2 Siderite nodules and Floodplains,
silstone parallel- antarcticum, bands, desiccation interdistributary
laminated, white Planolites cracks bays and tidal flats
to light gray montanus,
Taenidium isp.
BI: 4-6
FG3 Mudstone Massive to Firmground Siderite nodules and Paleosols
parallel- Thalassinoides isp., bands, desiccation
laminated, white Planolites cracks
to light gray montanus, root
trace fossils
BI: 1-4
FH FH1 Coal No trace fossils 0.1-2 Fluvial, estuary
Coal and delta plain
swamps
FH2 Bioturbated Woodground Estuary swamps
Coal Thalassinoides isp.
BI: 3-5
FI FI1 Very coarse- to Massive Ophiomorpha 8-12 Sandy deposits are Meandering tidal
Cross-Stratified medium- sandstone nodosa limited in their top channels
very coarse-to grained BI: 0-1 or base by mudstone
very fine- sandstone, deposits or
grained poorly sorted heterolithic zones,
sandstone FI2 Very fine- to Massive to planar No trace fossils 0.15-0.61 containing benthic Meandering tidal
medium- cross-stratified foraminifera and channels
grained dinoflagellates.
sandstone with
shells
FI3 Very coarse- to Massive to planar No trace fossils 2-3 Meandering tidal
fine-grained cross-stratified channels
sandstone,
poorly sorted

Grain size suggests high energy conditions. The fluvial braided river discharge cannot be completely disregarded (e.g. Gugliotta et al.,
channel-fills are represented by stacked sandstone successions of mul­ 2016a,b; Botterill et al., 2016). Structures resembling mud drapes can
tiple depositional units or storeys. Facies B represents the infill of also be formed due to fluid mud deposition under a wide variety of flow
braided-fluvial channels. conditions rather than slack water (Mackay and Dalrymple, 2011). The
presence of the cross-stratal sets in the channel infills suggests migration
4.3. Facies C: cross-stratified medium-to fine-grained sandstone with mud of two and three-dimensional dunes during periods of high river
drapes discharge when the maximum tidal limit migrated seaward. FC also
records sedimentation in outer estuary sandbars, but facies units tend to
Description: Facies C consists of 10–120 cm thick, well-sorted, be thinner than those representing the meandering estuary channels. FC
trough and planar cross-stratified, medium-to fine-grained sandstone is interpreted as a subordinate component within tidal creeks. Facies C
(Fig. 7C–D), forming facies intervals up to 15 m thick. Mud drapes represents a wide variety of tide-dominated deposits, such as
mantling oblique laminae are ubiquitous. These sandstone units do not meandering estuary channels, tidal creeks, tidal sandbars and deltaic
contain trace fossils. Facies C is present in the middle member. distributary channels. The former three record sedimentation in trans­
Interpretation: Mud drapes are commonly formed on tide-dominated gressive estuary systems and the latter one records deposition in the
point bars (Choi et al., 2004; Hovikoski et al., 2008; Dalrymple et al., lower delta plain of a tide-dominated delta.
2015; Rodríguez et al., 2018; Gingras et al., 2016). However, the pos­
sibility that some of these mud drapes may reflect seasonal variations in

9
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 7. Facies A, B and C. A: Facies A (mudstone


breccias formed at the base of meandering estuary
channel deposits due to the collapse of the associated
muddy cut-bank deposits). Well C2 (Carabobo area);
depth 982.37 m. B: Facies B (braided fluvial channel
deposits consisting of cross-stratified pebbly, very
coarse-to medium-grained sandstone and recording
dune migration). Well A8 (Ayacucho area); depth
514 m. C: Facies C (meandering estuary channel de­
posits consisting of cross-stratified medium- to fine-
grained sandstone and characterized by dune migra­
tion). Wells J1 (Junín area); depth 569.06 m. D:
Facies C (meandering estuary channel deposits with
mudstone drapes, which are formed during a brief
slack-water period, providing evidence of tidal ac­
tion). Well A9 (Ayacucho area); depth 372 m. Sand­
stone is impregnated with hydrocarbon resulting in
dark color; where it is not impregnated, it is in light
orange color. Mudstone is light gray color.

10
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

4.4. Facies D: inclined heterolithic stratified coarse-to fine-grained 2017; Rodríguez et al., 2018). Inclined heterolithic stratified successions
sandstone and mudstone have been regarded as indicators of seasonal cyclicity in tidally influ­
enced settings (Hubbard et al., 2011; Sisulak and Dashtgard, 2012).
Description: Facies D consists of 30–200 cm thick, coarse-to fine- Overall, the presence of trace fossils in meandering estuary-channel
grained sandstone and mudstone beds comprising IHS stacked in up to 6 deposits with IHS reflects pauses in sedimentation or slower rates of
m thick intervals (Fig. 8A–D). This facies can be unburrowed, display sedimentation during times of reduced energy conditions. The subfacies
brackish-water trace fossils, or show elements of the Scoyenia Icnofacies defined record different salinity conditions that may reflect
(Beaconites antarcticum and Taenidium isp.) overprinting elements of the proximal-distal positions within the depositional system (Solórzano
depauperate Cruziana Ichnofacies (Teichichnus rectus). Rosselia socialis et al., 2017). Channel deposits with bioturbated IHS emplaced in the
(Fig. 8A), Planolites montanus (Fig. 8A), Ophiomorpha nodosa (Fig. 8B) inner zone of the deltaic and estuary valleys display continental
and Thalassinoides isp (Fig. 8D). are the other ichnotaxa present. Based meniscate backfilled structures overprinting a brackish-water trace-­
on its the ichnological characteristics, facies D has been subdivided into fossil suite, indicating the fluvial-tidal transition zone of the deltaic and
subfacies D1, D2, D3, and D4. Subfacies D1 includes Rosselia socialis and estuary settings (Diez-Canseco et al., 2015; Solórzano et al., 2017;
Planolites montanus, which occur in interbedded, slightly bioturbated (BI Rodríguez et al., 2018). This occurrence is interpreted as the record of
0–1), very fine-grained sandstone and mudstone with IHS (Fig. 8A). temporal fluctuations between brackish-water and freshwater to
Subfacies D2 contains Ophiomorpha nodosa, which occurs in inclined terrestrial conditions (Buatois and Mángano, 2011; Diez-Canseco et al.,
heterolithic stratified, scarcely bioturbated (BI 0–1), coarse-to fine- 2015; Solórzano et al., 2017). Marine polychaetes, potential producers
grained sandstone and mudstone (Fig. 8B). Subfacies D3 consists of of some of these brackish-water trace fossils, are typically passively
Beaconites antarcticum, Taenidium isp., and Teichichnus rectus in inter­ transported by tides from adjacent marine and marginal-marine sites
bedded, intensely bioturbated (BI 4–6), fine-to very fine-grained sand­ because their larvae do not swim (Gingras et al., 2016). Trace fossil
stone and mudstone units displaying IHS (Fig. 8C). Subfacies D4 consists cross-cutting relationships in the fluvial-tidal transition area provide
of Taenidium isp. (BI 4–6) and Thalassinoides isp. (BI 0–1) in interbedded, evidence of channel abandonment, reflecting colonization during sub­
fine-grained sandstone and mudstone units displaying IHS (Fig. 8D). sequent freshwater conditions or directly during subaerial exposure
Facies D is present in the middle and upper members. (Solórzano et al., 2017). In this context, the saltwater wedge is flushed
Interpretation: IHS results from lateral accretion of point bars in downstream during periods of high river discharge allowing establish­
meandering channels (Thomas et al., 1987; Hovikoski et al., 2008; Chen ment of freshwater conditions (Hubbard et al., 2011; Sisulak and
et al., 2015; Olariu et al., 2015; Gingras et al., 2016; Jablonski and Dashtgard, 2012; Gingras et al., 2016; Melnyk and Gingras, 2019). In
Dalrymple, 2016; Melnyk and Gingras, 2020). Although more contrast, the saltwater wedge moves upstream during periods of low
commonly recorded in estuary settings, IHS may be produced in deltaic river discharge allowing the establishment of brackish-water conditions.
systems as well (Choi et al., 2004; Martinius et al., 2012; Solórzano et al., The absence of bioturbation structures in some estuary and

Fig. 8. Facies D. A: Bioturbated mudstone-dominated


deposits with IHS (outer-estuary sandbar complex).
Ichnotaxa include Rosselia socialis (Ro) and Planolites
montanus (Pl). Well J1 (Junín area); depth 434 m. B:
Bioturbated sandstone-dominated deposits with IHS
(estuary channel point bars). Deposits contain iso­
lated specimens of Ophiomorpha nodosa (Op), forming
monospecific suites. Well A8 (Ayacucho area); depth
313.02 m C: Bioturbated sandstone-dominated de­
posits with IHS (estuary channel point bars). Taeni­
dium isp. and Beaconites antarcticum overprint
specimens of Teichichnus rectus, indicating the fluvial-
tidal transition zone in the estuary. Well J15; depth
572 m. D: Bioturbated mudstone-dominated deposits
with IHS (tidal creeks in mud-flat deposits). A
monospecific ichnofabric of Taenidium isp. occurs on
top of an interval containing a specimen of Thalassi­
noides isp. indicating the fluvial-tidal transition zone
in the estuary. Sandstone is impregnated with hy­
drocarbon resulting in dark color, while mudstone is
light color. Sandstone (A and C) displays dark to light
brown color.

11
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

distributary-channel deposits of facies D is probably due to a combina­ drapes. This subfacies contains Planolites montanus, Skolithos linearis,
tion of severe brackish-water conditions (i.e. mesohaline to oligohaline), Rosselia socialis (Fig. 10B), Teichichnus rectus, Diplocraterion habichi, and
rapid migration of two and three-dimensional dunes, and high sedi­ escape trace fossils (Fig. 10A), reflecting the presence of both the Sko­
mentation rates (Solórzano et al., 2017; Rodríguez et al., 2018; Melnyk lithos and the depauperate Cruziana Ichnofacies. Subfacies F1 is present
and Gingras, 2019). FD is interpreted as a subordinate component in the middle member.
within the tidal creeks. This facies represents deposits formed in estuary Subfacies F2 (Fig. 10D–F) consists of 1–90 cm thick, interlaminated
channel point bars, tidal creeks, tidal sandbars, and distributary channel to interbedded fine-to very fine-grained sandstone with mudstone. This
point bars. The former three record sedimentation in tide-dominated subfacies contains Ophiomorpha nodosa and Thalassinoides isp. (BI 0–1)
estuary systems, and the latter one records deposition in the lower (Fig. 10E–F) overprinted to indeterminate bioturbation mottling (BI
delta plain of a tide-dominated delta. 4–5), indicative of the Skolithos and depauperate Cruziana Ichnofacies.
Subfacies F2 is present in the middle member.
Subfacies F3 (Fig. 10G–I) consists of 1–100 cm thick, mudstone with
4.5. Facies E: convoluted fine-to very fine-grained sandstone and
dispersed, scarce very fine sand grains, displaying low to moderate in­
mudstone
tensities of bioturbation (BI 2–3). This subfacies contains Bergaueria isp.,
Planolites montanus (Fig. 10I), and Thalassinoides isp. (Fig. 10I), reflect­
Description: Facies E consists of 10–50 cm thick, well-sorted, fine-to
ing the presence of the depauperate Cruziana Ichnofacies. Monospecific
very fine-grained sandstone and mudstone with convolute lamination
occurrences of Thalassinoides isp. represents the Glossifungites Ichnofa­
(Fig. 9A–B), forming facies intervals up to 1.5 m thick. These deposits
cies (Fig. 11H). Subfacies F3 is present in the middle member.
are unbioturbated to sparsely bioturbated (BI 0–1) Ophiomorpha nodosa
Subfacies F4 (Fig. 10J-M) consists of 1–180 cm thick, massive silt­
is locally present, forming monospecific trace-fossil suites suggestive of
stone and mudstone, displaying moderate to high intensities of bio­
the Skolithos Ichnofacies. Facies E is present in the middle member.
turbation (BI 4–5). This subfacies contains Teichichnus rectus,
Interpretation: Convolute lamination may have been formed as result
Thalassinoides isp. (Fig. 10J and L), Asterosoma (Fig. 10M), and inde­
of loading, rapid sedimentation or slumping (Bridge et al., 2000;
terminate bioturbation mottling, suggestive of the depauperate Cruziana
Plink-Bjorklund, 2005; Hubbard et al., 2011; Legler et al., 2013) or may
Ichnofacies. Thalassinoides isp. is filled with regular alternations of
be interpreted as seismically induced (seismites) (Törő and Pratt, 2015).
sandstone and mudstone laminae (Fig. 10L). Subfacies F4 is present in
Although abundant in a wide variety of settings, such as prodeltas (e.g.
the middle member.
Bhattacharya and MacEachern, 2009), convolute beds are common in
Subfacies F5 (Fig. 10N–Q) consists of 10–30 cm thick, massive
intertidal areas, and are typically interpreted as having been triggered
calcareous mudstone (Fig. 10C) with scarce limestone layers (Fig. 10P
by waves, during the low-tide level, accompanied by gravity-driven
and Q) and marine bivalve shell remains (Fig. 10N–O). This subfacies is
downslope movements (Choi et al., 2004; Fan et al., 2014). Facies E
barren in trace fossils. Subfacies F5 is present in the middle member.
mostly represents meandering-channel deposits within tide-dominated
Interpretation: Flaser, wavy and lenticular bedding are common in,
estuary systems and, to a lesser extent, tidal-flat deposits.
although not exclusive of, tidal-flats environments (Weimer et al., 1981;
Reineck and Wunderlich, 1968; Hovikoski et al., 2008; Sisulak and
4.6. Facies F: interbedded mudstone and medium-to very fine-grained Dashtgard, 2012; Fan, 2012; Gingras et al., 2016, 2017). Thalassinoides
sandstone isp. forms tubular tidalites, further supporting tidal influence. Tubular
tidalites of inclined laminae occur only in open framework burrows,
Description: Facies F consists of mudstone intervals or alternations of such as Thalassinoides (Gingras et al., 2012, 2015, 2015; Wetzel et al.,
mudstone and medium-to very fine-grained sandstone. Facies F forms 2014; Rodríguez-Tovar et al., 2019). Tubular tidalites have been
intervals up to 12 m thick. It has been subdivided into subfacies F1, F2, observed in subtidal point-bar, intertidal-flat, or tidal-channel-thalweg
F3, F4, and F5. Subfacies F1 (Fig. 10A–C) consists of sparsely bio­ (Gingras and Zonneveld, 2015). Thalassinoides isp. forming tubular
turbated (BI 0–1), 1–200 cm thick, fine-to very fine-grained sandstone tidalites could be misinterpreted as Teichichnus rectus. However, Tei­
and mudstone. Mud drapes and flaser, wavy and lenticular bedding are chichnus rectus displays concave-up laminae, representing spreite. A
present. In places, wave-ripple cross-lamination and low-angle cross- distinct horizontal, circular to sub-circular causative burrow is present
lamination are present. Ripple foresets are commonly mantled by mud at the upper or lower end of the laminae (Pemberton et al., 1992; 2001).
In addition, Teichichnus rectus tends to be smaller than Thalassinoides isp.
The presence of the Skolithos and depauperate Cruziana Ichnofacies
suggests that these deposits were formed under brackish-water condi­
tions (Solórzano et al., 2017). The association of limestone and bivalve
shells indicates marine influence. Overall, facies F represents deposition
in estuary tidal flats locally dissected by tidal creeks. In particular,
subfacies F1 and F2 record deposition in tidal sand to mixed flats and
subfacies F3, F4, and F5 record sedimentation in tidal mud flats.

4.7. Facies G: carbonaceous, rooted silty mudstone and thinly laminated


mudstone

Description: Facies G consists of massive mudstone and parallel-to


ripple cross-laminated siltstone and very fine-grained sandstone.
Facies intervals are up to 4 m thick. It has been subdivided into subfacies
G1, G2, and G3. Subfacies G1 (Fig. 11A–B) consists of 30–150 cm thick,
massive to parallel-laminated mudstone, siltstone and very fine-grained
sandstone, displaying moderate intensities of bioturbation (BI 3–4),
locally with current ripples, flaser bedding and syneresis cracks. This
Fig. 9. Facies E (tidal flat and meandering estuary channel deposits with subfacies contains Planolites montanus, Teichichnus rectus, Thalassinoides
convolute lamination). A: Mudstone. Wells A10 (Junín area); depth 410 m. B: isp., and root trace fossils, illustrative of the depauperate Cruziana Ich­
Moderately impregnated sandstone. Well J1 (Junín area); depth 538 m. nofacies. Subfacies G1 is present in the middle member.

12
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

(caption on next page)

13
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 10. Facies F. A, B, and C: Subfacies F1 (tidal flat deposits consisting of interbedded sandstone and mudstone). Wells A12 (Ayacucho area), C1 (Carabobo area),
and J1 (Junín area), respectively; depths 891 m, 614 m, and 426 m, respectively. D, E, and F: Subfacies F2 (tidal flat deposits consisting of interbedded sandstone and
mudstone, suggestive of alternating traction and suspension fallout). Wells A11 (Ayacucho area), C8 (Carabobo area), and A8 (Ayacucho area), respectively; depths
1323 m, 438 m, and 335 m, respectively. G, H, and I: Subfacies F3 (tidal flat deposits consisting of mudstone-dominated heterolithics). Wells A2 (Ayacucho area), J13
(Junín area), and J15 respectively; depths 737 m, 434 m, and 352 m, respectively. J, K, L, and M: Subfacies F4 (tidal flat deposits consisting of massive silstone and
mudstone). Wells C9 (Carabobo area), J1 (Junín area), C9, J12 (Junín area), respectively; depths 980 m, 510 m, 988 m and 518 m, respectively. N, O, P, and Q:
Subfacies F5 (tidal flat deposits consisting of calcareous mudstone with shells and thin limestone beds). Wells A11 (Ayacucho area), A1 (Ayacucho area), C1
(Carabobo area), and C2, respectively; depths 794 m, 623 m, 555 m, and 549 m, respectively. A shows escape trace fossil (Et) displaying the classic cone-in-cone
morphology, representing rapid changes in sedimentation and B shows Rosselia socialis (Ro). Ophiomorpha nodosa (Op) and Thalassinoides isp. (Th) are present in
E and F. Syneresis crack (Sy) is present in H and Thalassinoides isp. (Th) is present in I. Thalassinoides isp. (Th) is present in J and L, the latter showing rhythmic tidal
infill. Asterosoma isp. (As) is present in M. The trace-fossil assemblage illustrates the Skolithos and depauperate Cruziana Ichnofacies, which implies brackish-water
conditions. N and O represent calcareous mudstone with shell remains of marine bivalves. P and Q display limestone layers. Sandstone is impregnated with hy­
drocarbon resulting in dark color, while mudstone is light color.

Fig. 11. Facies G. A and B: Subfacies G1 (outer-estuary margin deposits consisting of massive to parallel-laminated mudstone). Wells A4 and A12 (Ayacucho area),
respectively; depths 807 m and 831 m, respectively. C, D, and E: Subfacies G2 (tidal flat deposits consisting of massive to parallel-laminated mudstone with siderite
nodules and bands, supporting fluctuating salinity). Wells A10 (Ayacucho area), J1 (Junín area) and A10 (Ayacucho area), respectively; depths 401 m, 523 m and
410 m, respectively. F: Subfacies G2 (tidal flat deposits consisting of bioturbated mudstone with Beaconites antarcticum, indicating the presence of the Scoyenia
Ichnofacies developed in a setting located between the maximum salinity limit and the maximum tidal limit). Well J12; depth 518 m. G: Subfacies G2 (floodplain
deposits consisting of bioturbated mudstone with Taenidium isp., reflecting the presence of the Scoyenia Ichnofacies and further supporting freshwater conditions).
Well J1; 724 m. H: Subfacies G3 (firmground Thalassinoides isp. (Th) penetrating into paleosol deposits from an overlying transgressive surface). Wells A4 (Ayacucho
area); depth 862 m.

Subfacies G2 (Fig. 11C–G) consists of 30–120 cm thick, massive to in the lower, middle, and upper members.
parallel-laminated mudstone and siltstone with siderite nodules and Subfacies G3 (Fig. 11H) consists of 50–200 cm thick, massive to
bands (Fig. 11C–D), and desiccation cracks, displaying moderate to high parallel-laminated mudstone with siderite nodules and bands, syneresis
intensities of bioturbation (BI 4–6). This subfacies contains Beaconites cracks, displaying low to moderate intensities of bioturbation (BI 1–4).
antarcticum (Fig. 11F), Planolites montanus, and Taenidium isp. This subfacies contains firmground Thalassinoides isp (Fig. 11H). over­
(Fig. 11G), indicative of the Scoyenia Ichnofacies. Subfacies G2 is present printed to a fabric dominated by Planolites montanus and root trace

14
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

fossils. The former reflects the presence of the Glossifungites Ichnofacies. Subfacies H2 (Fig. 12A–B) consists of 10–50 cm thick, bioturbated
Subfacies G3 is present in the lower and middle members. (BI 3–5) coal layers that are penetrated by Thalassinoides isp., repre­
Interpretation: Subfacies G1 displays sedimentologic features sug­ senting the Teredolites Ichnofacies. Subfacies H2 is present in the middle
gestive of tidal influence (e.g. flaser bedding) and salinity fluctuations. member.
In addition, the depauperate Cruziana Ichnofacies indicates that these Interpretation: The presence of coal layers indicates a high-water
deposits were formed under brackish-water conditions (Solórzano et al., table. The presence of crustacean galleries attributed to the Teredolites
2017). Rapid and extreme changes in others factors, such as tempera­ Ichnofacies delineates transgressive surfaces of erosion (Solórzano et al.,
ture, subaerial exposure, turbulence, oxygen content, and water 2017). Following erosional exhumation of the swamp deposits during
turbidity promote a depauperate assemblage. These fluctuations result ravinement, decapod crustaceans were able to penetrate the underlying
in physiologically stressful conditions for numerous organisms, resulting coal layer under marine conditions. Subfacies H1 represents swamps
in low ichnodiversity, small body sizes, high mortality rates, rapid formed in a wide variety of settings, such as fluvial, estuary and deltaic.
reproduction capacity, short life cycles, and early sexual maturity In contrast, subfacies H2 records swamps restricted to estuary systems in
(Remane and Schlieper, 1971; Pemberton and Wightman, 1992; Buatois direct association with a relative sea-level rise. Facies H records swamp
and Mángano, 2011; Sisulak and Dashtgard, 2012; Gingras et al., 2016; deposits.
Solórzano et al., 2017). The presence of root trace fossils suggests
waterlogged paleosols. Subfacies G1 represents low-energy deposits
4.9. Facies I: cross-stratified very coarse-to very fine-grained sandstone
formed along the margins of the outer estuary.
In subfacies G2 the local presence of desiccation cracks indicates that
Description: Facies I comprises, erosively based, massive to cross-
periodic subaerial exposure took place. Siderite nodules and bands, such
stratified very coarse-to very fine-grained sandstone, forming fining-
as those present in the Oficina Formation, have been recorded in envi­
and thinning-upward intervals. This facies forms intervals up to 12 m
ronments affected by fluctuating salinity (Plummer and Gostin, 1981;
thick and is restricted to the middle member. It has been subdivided into
Postma, 1982; MacEachern et al., 2005; Hovikoski et al., 2008; Buatois
subfacies I1, I2 and I3.
et al., 2012; Martinius et al., 2012). The occurrence of siderite bands is
Subfacies I1 (Fig. 13A–B) consists of sparsely bioturbated (BI 0–1),
also consistent with freshwater environments that present low chloride
50–100 cm thick, poorly sorted, massive, very coarse-to medium-
concentrations, as well as no dissolved sulfide and high ferrous iron
grained sandstone. Ophiomorpha nodosa at sandstone tops represents the
content. However, siderite is not exclusive of freshwater environments,
Skolithos Ichnofacies.
being also present in brackish or even fully marine settings because the
Subfacies I2 (Fig. 13C–D) consists of 10–60 cm thick, unbioturbated,
presence of iron remains results in complete sulfate reduction (Postma,
poorly sorted, massive to planar cross-stratified, very fine-to medium-
1982). The presence of the Scoyenia Ichnofacies indicates that the bulk
grained sandstone with shells.
of these deposits were formed under freshwater conditions. Although
Subfacies I3 consists of 50–300 cm thick, unbioturbated, poorly
this ichnofacies is common in continental settings, it also is present in
sorted, massive to planar cross-stratified, very coarse-to fine-grained
freshwater settings located between the maximum salinity and
sandstone. Subfacies I1, I2, and I3 are intercalated within mudstone
maximum tidal limit, such as the inner estuary and deltaic environments
deposits that contain foraminifers, calcareous nannoplankton and
(Buatois et al., 1997; Mángano and Buatois, 2004; Diez-Canseco et al.,
dinoflagellates.
2015, 2016; Solórzano et al., 2017; Rodríguez et al., 2018). Floodplain
Interpretation: Erosive bases and fining- and thinning-upward trends
deposits commonly contain palynomorphs. Subfacies G2 represents
support interpretation of this facies as channel-fills. Facies I is
various environments, including floodplains in the fluvial systems, tidal
flats in the fluvial-tidal transition zone of estuary, and floodplains and
interdistributary bays in the tide-dominated deltaic systems.
A wide variety of mechanisms have been proposed for the generation
of syneresis cracks (Fig. 10H), including sediment compaction and
expulsion of water (White, 1961; Burst, 1965), microbial-mat stabili­
zation (Seilacher, 1999; Seilacher et al., 2005; Harazim et al., 2013), and
clay contraction and expansion due to fluctuating salinity (Plummer and
Gostin, 1981; Hovikoski et al., 2008; Martinius et al., 2012), the latter
being the preferred interpretation in brackish-water marginal-marine
settings (MacEachern and Pemberton, 1994). Syneresis cracks also are
interpreted as seismically induced (seismites), and in this context they
are referred to as small dikes (Törő and Pratt, 2015). In places, paleosols
display the Glossifungites Ichnofacies, which represents transgressive
surfaces of erosion (Solórzano et al., 2017). This suggests that these soils
were emplaced in areas of the coastal plain that were subjected to sig­
nificant wave erosion during ravinement and subsequent colonization
under marine conditions. Subfacies G3 represents waterlogged paleosols
within fluvial and estuary systems.
Overall facies G represents a wide variety of coastal-plain sub­
environments such as floodplains, interdistributary bays, tidal flats, and
soils formed either along the margins of the estuary or in the innermost
areas of the estuary and deltaic complex.

4.8. Facies H: coal

Description: Facies H represents coal layers, forming intervals up to Fig. 12. Facies H. A and B: Subfacies H2 (woodground Thalassinoides isp. (Th)
2 m thick. This facies has been subdivided into subfacies H1 and H2. penetrating from an overlying transgressive surface into a coal layer of swamp
Subfacies H1 consists of 1–60 cm thick, coal layers without trace fossils. origin). Well A8 (Ayacucho area) for both images; depths 326.44 m and 301 m,
Subfacies H1 is present in the lower, middle and upper members. respectively. Coal is dark gray color.

15
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 13. Facies I. A and B: Subfacies I1 (meandering


estuary channels with Ophiomorpha nodosa (Op)
indicating marine influence). Wells A9 (Ayacucho
area) and C10 (Carabobo area), respectively; depths
376.73 m and 908 m, respectively. C and D: Subfacies
I2 (meandering estuary channel deposits with
calcareous sandstone and shells of marine bivalves).
Well A2 (Ayacucho area) for both images; depths
698.60 m and 706 m, respectively. All photos repre­
sent sandstone, impregnation with hydrocarbon
resulting in dark color. Sandstone (A) is moderately
impregnated with hydrocarbon, and displays dark to
light orange color.

distinguished from similar deposits formed in the fluvial segment of the paleosols suggest that the areas between the channels may have been
Oficina Formation based on paleontologic evidence. In Subfacies I1, the characterized by some ponded and vegetated areas (Plink-Bjorklund,
presence of Ophiomorpha nodosa at the top of channel fill units indicates 2005). The presence of swamp deposits is also consistent with a high
that these channels were marine influenced during their abandonment, water table in the alluvial plain.
most likely as a result of a relative sea-level rise. In Subfacies I2, these In these fluvial deposits, the Scoyenia Ichnofacies occurs in flood­
channels were characterized by the migration of two-dimensional dunes plains located above fluvial braided-channel packages, further sup­
during periods of high river discharge when the maximum tidal limit porting freshwater conditions. These floodplain deposits are strongly
migrated seaward. Marine influence is evidenced by the occurrence of bioturbated, reflecting long colonization windows in overbank settings
remains of shells associated to transgressive events. In subfacies I3, as in (Buatois and Mángano, 2004). The presence of palynomorphs and the
the case of subfacies I2, migration of two-dimensional dunes is apparent. absence of foraminifers and calcareous nannoplankton (Suarez and
The key evidence to detect marine influence is the presence of marine Solórzano, 2014; Solórzano et al., 2015, 2017) are consistent with
microfauna and plankton in interbedded mudstone deposits. Overall, freshwater conditions. FA1 is restricted to paleotopographic lows where
facies I represents meandering channels within estuary systems. it represents the infill of incised fluvial valleys, recording the LST of the
Oficina Formation 2nd-order depositional sequence. 3D seismic avail­
5. Facies associations and depositional environments able for the Junin area further support an incised valley model (Marti­
nius et al., 2012).
The sedimentary facies (FA-I) of the Oficina Formation along the
Orinoco Oil Belt allow recognition of five facies associations, labeled
FA1 to FA5. These are stacked forming a single second-order deposi­ 5.2. FA2: meandering estuary channels
tional sequence (Rodríguez et al., 2018).
FA2 (Fig. 14A) consists of meandering estuary channels (FA, FC, FD,
FE, and FI), and is present in the middle member. Estuary channels from
5.1. FA1: fluvial braided channels the Oficina Formation are characterized by IHS resulting from lateral
accretion of point bars (Thomas et al., 1987; Plink-Bjorklund, 2005;
FA1 (Fig. 14A) consists of fluvial braided channels (FB), floodplains Hovikoski et al., 2008; Gingras et al., 2016, 2017, 2017; Ekwenye et al.,
(FG2), swamps (FH1), and paleosols (FG3), and is present in the lower 2017). Associated mud drapes suggest tidal action (Plink-Bjorklund,
member. The fluvial braided channel-fills are represented by high- 2005; Lettley et al., 2009; Musial et al., 2011; Hovikoski et al., 2008;
energy stacked sandstone successions of multiple depositional units. Hubbard et al., 2011; Legler et al., 2013; Gingras et al., 2016; Solórzano
Scarce mudstone and siltstone layers indicate limited development of et al., 2017; Rodríguez et al., 2018). Overall, the marked dominance of
floodplain settings, suggesting fluvial channels of low sinuosity. The lack tide-generated structures through the channel-fill succession suggests
of lateral accretion beds and the coarse grain size of the channel fills also that the system was tide-dominated and fluvial-influenced, the latter
suggest that the rivers were of relatively low sinuosity (Bridge et al., processes being particularly effective at times of high river discharge (Tf
2000; Plink-Bjorklund, 2005; Ekwenye et al., 2017). Pebbly of Ainsworth et al., 2011). Soft-sediment deformation was locally
mid-channel bars and channel bifurcation may have been dominant in important. In places, the bases of these channels are mantled by brec­
these fluvial systems, most likely resulting from a local decrease in flow cias, which represent lag deposits. Amalgamation of sandstone indicates
velocity or a change of slope (Rodríguez et al., 2018). Channel bars are multiple stacked channels.
represented by thick accumulations of cross-stratified sandstone formed Channel deposits with IHS emplaced in the inner zone of the estuary
mostly by frontal accretion due to unidirectional currents. Although display continental meniscate backfilled structures overprinting a
swamps deposits are more typical in high-sinuosity rivers, they may be brackish-water trace-fossil suite. Also, estuarine-channel deposits are
present in low-sinuosity systems as well, albeit subordinately (e.g. commonly intercalated with mudstone layers containing a marine
Sarma, 2014) as is the case of the lower member of the Oficina For­ fauna, which provides further evidence of a basinwide sea-level rise. The
mation. No evidence of tidal or wave action has been recorded in this stratigraphic position of FA2 resting on top of FA1 indicates a retro­
facies association, which is dominated by purely fluvial processes (F in gradational stacking pattern, signaling marine flooding of incised val­
the scheme of Ainsworth et al., 2011). Amalgamation of sandstone in­ leys and their transitioning into transgressive estuary systems. Similar
dicates multy-storey channels. Root trace fossils and waterlogged backstepping stacking patterns have been extensively documented in

16
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 14. Paleogeographic maps illustrating the geologic evolution of the Oficina Formation in the Orinoco Oil Belt and in the Oritupano Field. A: Lower to middle
member of the Oficina Formation. The Orinoco Oil Belt consists of incised fluvial valleys formed during a relative sea-level fall and incised tide-dominated estuary
valleys formed during the subsequent transgression. The estuary systems were formed in a brackish-water embayment except in the Carabobo area, which was
directly connected with the open sea. B: Upper member of the Oficina Formation. The Orinoco Oil Belt consists of highstand tide-dominated deltaic deposits formed
in a brackish-water embayment except in the Carabobo area, whose distributary channels flowed directly into the open sea. The Oritupano Field consists of regressive
wave-dominated deltaic and shoreface-offshore complex facing the open sea, and the distributary channels come from the southeastern zone of the Ayacucho area.
Additionally, El Roble Field consists of delta front, prodelta, and offshore deposits (Buatois, 2014) coming from uplifted areas located to the north. The Capiricual
Formation records deposition in fan deltas also coming from the north (Chigne et al., 2002). Sources: Di Croce et al. (1999); Solórzano et al. (2017) and unpublished
information (Chigne et al. (2002); Buatois (2014); and Parra et al. (2018).

17
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

tide-dominated incised valleys elsewhere (Zaitlin et al., 1994; Plink-B­ mouth (Uncles et al., 2006; Dalrymple et al., 2012). Notably,
jorklund, 2005; Musial et al., 2012; Gingras et al., 2016; Ekwenye et al., wave-generated sedimentary structures are notoriously scarce in this
2017). Biostratigraphic evidence indicates that these estuarine channels outer segment of the Oficina estuary, which is characterized by an
were formed during the Langhian transgression (Audemard et al., 1985; abundance of tide-generated features and, therefore, it is considered as
Solórzano et al., 2015), and are included in the early TST of the Oficina tide-dominated, wave-influenced and fluvial-affected (Twf in the
Formation depositional sequence. scheme of Ainsworth et al., 2011). The outer estuary was dissected by
ebb/flood-dominated channels and flanked by swamps and waterlogged
5.3. FA3: tidal flat and creek complex soils developed in the adjacent coastal plain.
In places, these deposits display Thalassinoides in both firmgrounds
FA3 (Fig. 14A) represents tidal flats and tidal creeks (FC, FD, FE, FF, and woodgrounds, evidencing erosional exhumation of the substrate
and FG2), swamps (FH1 and FH2), and paleosols (FG3), and is present in during wave ravinement. Although at the seaward end of the estuary,
the middle member. The tidal flats were widely developed along the these deposits record brackish-water conditions as revealed by a low
length of the estuary system and were dissected by tidal creeks, as is diversity of trace fossils; this is consistent with the overall embayed
commonly the case in tide-dominated estuary (e.g. Ekwenye et al., nature of the Oficina Formation in the Orinoco Belt (Solórzano et al.,
2017). The local presence of IHS in the tidal-creek deposits indicates 2017). In addition to salinity, other factors such as high bedforms
minor participation of lateral accretion and limited formation of point migration rates may have been detrimental to infaunal colonization.
bars (Thomas et al., 1987). The fact that IHS is relatively rare in these FA4 represents part of the late TST of the Oficina Formation depositional
deposits indicates dominance of stable tidal creeks, which in turn sequence. Establishment of the estuary sandbar complex took place
allowed the preservation of horizontally bedded, associated tidal-flat during the latest stage of the Langhian transgression (Audemard et al.,
deposits (Dashtgard et al., 2014; Rodríguez et al., 2018). Tidal-flat de­ 1985; Solórzano et al., 2015, 2017).
posits are characteristic of modern tide-dominated macrotidal estuary
(Hamilton, 1979; Lambiase, 1980; Dalrymple et al., 1990; Dalrymple 5.5. FA5: delta plain of a tide-dominated delta
et al., 1992). The local presence of symmetric ripples in
sandstone-dominated heterolithic facies suggests that in places inter­ FA5 (Fig. 14B) consists of distributary channels (FC and FD), flood­
tidal areas were characterized by the establishment of wave-dominated plains and interdistributary bays (FG2), and swamps (FH1), and is pre­
tidal flats similar to those that occur along the Korean coast (Yang et al., sent in the upper member. The delta-plain deposits display tidal
2005, 2006, 2008). Although this segment of the Oficina estuary was influence as indicated by the presence of bioturbated IHS and mud
overwhelmingly dominated by tidal processes, tidal flats formed in areas drapes in the distributary channels (Legler et al., 2013). These channels
that were less protected from the open ocean may have been are sparse, thin, rarely amalgamated, and are separated by widespread
wave-influenced and even wave-dominated at times. Channelized areas floodplains and interdistributary bays, suggesting extensive wetland
may have been affected by fluvial processes. Accordingly, this segment development in the lower delta plain. As in the case of the estuary de­
of the Oficina estuary is thought to reflect a more complex spatial and posits, FA5 also is affected by fluctuating salinity and periodic subaerial
temporal pattern of interplay of depositional processes, encompassing exposure. The abundance of tide-generated structures suggests that the
mostly tide-dominated, wave-influenced and fluvial-affected (Twf), but system was tide-dominated and fluvial-influenced (Tf of Ainsworth
also wave-dominated, tide-influenced and fluvial-affected (Wtf) (Ains­ et al., 2011).
worth et al., 2011). The presence of meniscate trace-fossils of the Scoyenia Ichnofacies
These deposits also display the Skolithos and depauperate Cruziana within floodplain deposits indicates the establishment of a continental
Ichnofacies, which indicate brackish-water conditions, mostly reflecting invertebrate fauna in the delta-plain deposits (Solórzano et al., 2017;
the influence of tidal currents in marginal-marine, restricted settings Rodríguez et al., 2018). The Orinoco Oil Belt delta complex was
(Solórzano et al., 2017). In addition, high sedimentation rates and water emplaced within a brackish-water embayment rather than in the open
turbidity, indirectly controlled by the tides, may have limited the di­ sea. Coeval open-marine deposits of the Oficina Formation occur in the
versity of infauna in these tidal flats (Dashtgard et al., 2014). In places, Oritupano field, outside the Orinoco Oil Belt (Solórzano et al., 2017)
the Glossifungites and Teredolites Ichnofacies are present in tidal flat and (Fig. 14B). The stratigraphic occurrence of FA5 resting on top of FA4
swamp deposits, respectively, indicating erosional exhumation during indicates a progradational stacking pattern (HST).
ravinement, favoring colonization by organisms before deposition of the
overlying marine sediment. Dinoflagellates, foraminifers, calcareous 6. Regional changes in stratal architecture
nannoplankton, and bivalve shells provide further evidence of marine
influence. The tidal-flat deposits were formed during the Langhian West to east stratigraphic cross-sections along the Orinoco Oil Belt
transgression (Audemard et al., 1985; Solórzano et al., 2015), and are show that the fluvial units extend from the Junín to Carabobo areas,
included as part of the early TST. being thicker in the Junin and Boyaca areas (Fig. 15). Based on north to
south stratigraphic cross-sections and the isopach maps, the thicknesses
5.4. FA4: outer-estuary sandbar complex of the fluvial deposits range from 0 to 305 m, increasing towards the
west (Figs. 1 and 16–189A). In the Junín area, from north to south, the
FA4 (Fig. 14A) consists of outer estuarine sandbars (FC, FD, FG1), fluvial deposits decrease in thickness (30–243 m) (Fig. 19A), resting on
paleosols (FG3), and swamps (FH2), and is present in the uppermost top of Cretaceous or pre-Cretaceous strata (Fig. 16). In the Ayacucho and
interval of the middle member. The occurrence of thinner IHS strata Carabobo areas, the fluvial deposits overlie the metamorphic-igneous
suggests that laterally accreted, free-standing tidal bars may have been basement, which could have controlled sedimentation in these fluvial
dominant in the outer zone of the Oficina valley (Rodríguez et al., 2018). systems by reducing accommodation potential (Figs. 17 and 18). In fact,
Sandbars were emplaced in the estuary mouth close to the zone of fluvial deposits are absent in some zones of the Ayacucho area (e.g. A1).
maximum turbidity as indicated by the abundance of mudstone units Therefore, the estuary system in this well directly overlies the
(Jouanneau and Latouche, 1981; Dalrymple et al., 1990; Allen, 1991; metamorphic-igneous basement. In areas of reduced accommodation
Dalrymple et al., 1992; Brownfield et al., 1998; Plink-Bjorklund, 2005; space, not all the units are preserved (Hein et al., 2013). Fluvial units
Legler et al., 2013; La Croix and Dashtgard, 2014; Ekwenye et al., 2017). thin southward, eventually disappearing altogether (Audemard et al.,
Overall, the maximum turbidity zone is near the tip of the saltwater 1985). In the Ayacucho (0–91 m) and Carabobo (30–91 m) areas, the
wedge, but high turbidity zone and suspended-sediment concentration fluvial units decrease in thickness from north to south (Fig. 19A). Fluvial
may extend from the fluvial-tidal transition zone to beyond the estuary deposits reach their maximum thicknesses westward in the Junín and

18
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 15. West to east stratigraphic correlation panel showing the paleoenvironmental and sequence-stratigraphic interpretation along the Orinoco Oil Belt.
Geophysical data shown in the gamma-ray log (GR), from 0 to 200 API units, in the neutron log (NPHI), from 0 to 0.6 v/v units, and in the density log (RHOB), from
1.65 to 2.65 g/cm3 units. Stratigraphic correlation is shown in Fig. 1B (#4).

Fig. 16. North to south stratigraphic correlation panel showing the paleoenvironmental and sequence-stratigraphic interpretation in the western part of the Orinoco
Oil Belt (Junín area). Geophysical data shown in the gamma-ray log (GR), from 0 to 200 API units, in the neutron log (NPHI), from 0 to 0.6 v/v units and in the
density log (RHOB), from 1.65 to 2.65 g/cm3 units. Stratigraphic correlation is shown in Fig. 1B (#1).

Boyacá areas, because of increased subsidence caused by the compaction sediments could have come from the nearby Guayana Shield in the
of the pre-Cretaceous sediments, particularly north of these two areas south. On a regional scale, the lower member shows south to north
(Audemard et al., 1985). Deposition of the lower part of the Oficina trending fluvial channels. The fluvial units represent the infill of
Formation was mainly controlled by compressional tectonic activity incised-valley systems, which were formed during a relative sea-level
(Martinius et al., 2012), and is generally restricted to paleotopographic fall.
lows on the Cretaceous or pre-Cretaceous unconformity. The influx of West to east stratigraphic cross-sections along the Orinoco Oil Belt

19
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 17. North to south stratigraphic correlation panel showing the paleoenvironmental and sequence-stratigraphic interpretation in the eastern part of the Orinoco
Oil Belt (Ayacucho area). Geophysical data shown in the gamma-ray log (GR), from 0 to 200 API units, in the neutron log (NPHI), from 0 to 0.6 v/v units and in the
density log (RHOB), from 1.65 to 2.65 g/cm3 units. Stratigraphic correlation is shown in Fig. 1B (#2).

Fig. 18. North to south stratigraphic correlation panel showing the paleoenvironmental and sequence-stratigraphic interpretation in the eastern part of the Orinoco
Oil Belt (Carabobo area). Geophysical data shown in the gamma-ray log (GR), from 0 to 200 API units, in the neutron log (NPHI), from 0 to 0.6 v/v units and in the
density log (RHOB), from 1.65 to 2.65 g/cm3 units. Stratigraphic correlation is shown in Fig. 1B (#3).

indicate that the estuary units extend from the Boyaca to Carabobo estuary interval is greater than that of the fluvial systems in the Aya­
areas, being thicker in the Ayacucho and Carabobo areas (Figs. 15 and cucho and Carabobo areas, probably due to the space available to
19). Based on north to south stratigraphic cross-sections and the isopach accommodate sediments, while in the Junín area the thickness of the
maps, the thicknesses of the estuary deposits range from 30 to 549 m, fluvial interval is greater than estuarine units (Fig. 19B). Estuary facies
increasing towards the north (Figs. 1 and 16–189B). The thickness of the are thicker in the Ayacucho (30–549 m) and Carabobo (30–366 m) areas

20
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Fig. 19. Isopach maps of the fluvial, estuary, and deltaic environments from Boyaca to Carabobo areas in the Orinoco Oil Belt. A: Fluvial system isopach map for the
lower member of the Oficina Formation. B: Estuary system isopach map for the middle member of the Oficina Formation. C: Deltaic system isopach map for the upper
member of the Oficina Formation.

than in the Junin (30–274 m) area (Fig. 19B). The middle member 244 m, increasing towards the north (Figs. 1 and 16–189C). Deltaic
displays transgressive estuary channels that are oriented similar to the strata displays the following thickness: Carabobo (61–244 m), Junín
underlying fluvial channels and that display increased marine influence (152–244 m), Ayacucho (91–244 m), and Boyaca (152-122 m)
towards the north-northeast (Solórzano et al., 2017). (Fig. 19C). On a regional scale, the upper member shows south to north
West to east stratigraphic cross-sections through the Orinoco Oil Belt trending deltaic channels. The deltaic units represent progradation into
demonstrate that the deltaic units extend from the Boyaca to Carabobo a restricted embayment during highstand except in the Carabobo area,
areas (Fig. 15). Based on north to south stratigraphic cross-sections and which is located in closer proximity to the open sea (Fig. 14B). The
the isopach maps, the thickness of the deltaic deposits ranges from 61 to adjacent Oritupano Oil Field, located northeast of the Orinoco Oil Belt,

21
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

hosts deposits formed in strandplain environments and in deltaic sys­ In spite of potential local effects, it is unsurprising that relative sea-
tems that prograded into the open sea rather than a restricted embay­ level changes have played a key role on sedimentation of the Oficina
ment (Solórzano et al., 2017) (Fig. 14B). Progradation from the uplifted Formation and in the other units regardless of latitudinal setting. The
areas in the north may have been locally important in some of the Oficina Formation records the typical succession of fluvial incision
northern fields (e.g. El Roble Field, Fig. 14A–B). during sea-level fall, transition from fluvial to estuary valleys during the
subsequent transgression and deltaic progradation during highstand.
7. Discussion Notably, although tectonics were a significant factor at the time of
deposition, evolution of the Oficina Formation shows a close match with
7.1. Interplay of eustatic changes, depositional processes, and latitude the global sea-level curve of Haq and Schutter (2008) (Solórzano et al.,
2015). Similar depositional histories in response to relative sea-level
Ongoing research is emphasizing that climate may have been over­ changes are shared by other units elsewhere. In particular, the Creta­
looked in the generation of facies models (Martini, 2014; Martinius ceous McMurray Formation of western Canada shows a large-scale
et al., 2014). The Oficina tide-dominated marginal-marine depositional depositional evolution which is similar to that of the Oficina Forma­
systems developed in a tropical and humid climate with widespread tion, with a lowstand fluvial lower member, a transgressive estuarine
wetlands, swamps and embayed areas, displaying waterlogged paleosols middle member, and a highstand deltaic upper member (Crerar and
with pervasive root trace fossils and crustacean burrows, such as Arnott, 2007; Musial et al., 2012; Harris et al., 2016; Gingras et al.,
Ophiomorpha and Thalassinoides, which are also typical of low-latitude 2016). Similarities between the Oficina and McMurray formations are
marginal- and shallow-marine settings (Quiroz et al., 2010). The Ofi­ not restricted to stratal stacking pattern, but include sedimentary facies
cina Formation in the Orinoco Oil Belt shows marked similarities with (e.g. IHS) and trace-fossil content (low-diversity ichnofaunas) as well. In
various marginal-marine units of different ages and latitudinal settings the Aspelintoppen Formation of Arctic Norway, coastal-plain aggrada­
in terms of sedimentologic, ichnologic, and sequence-stratigraphic fea­ tion took place mostly during transgressions (Plink-Bjorklund, 2005). In
tures. The fact that these units were formed under a wide variety of the Bay of Fundy, the preservation of coastal sediments is strongly
latitudinal settings prompts us to evaluate the role of climate on depo­ influenced by transgression resulting from rapid sea-level rise over the
sition in marginal-marine tide-dominated settings. Although some past 6000 years (Amos et al., 1991; Shaw and Courtney, 2002; Dasht­
marginal-marine subenvironments are specific to certain latitudinal re­ gard and Gingras, 2005). Quaternary climatic and sea-level changes
gions (e.g. mangroves in tropical areas), others are ubiquitous with have controlled coastal zonality at different spatial and temporal scales
respect to latitude (e.g. sandy beaches, barriers, lagoons). Differences (Solomon, 2007; Kelletat et al., 2014).
are associated with the variable efficiency and intensity of the same A common characteristic between the Oficina Formation and the
physical azonal processes and locally dominant zonal processes, such as other units analyzed is the presence of a depauperate marine ichnofauna
the effects of sea-ice cover in the Arctic and carbonate reef building in reflecting brackish-water conditions in central and outer regions of the
the tropics (Martini, 2014). Zonal features are associated with latitude, estuary and a freshwater ichnofauna in the fluvial system and the most
but azonal features are not. proximal areas of the estuary and deltaic systems. Regardless of the
Overall, the characteristics of the Oficina depositional systems overall ichnologic similarities across a broad spectrum of latitudinal
resemble those of the modern Orinoco Delta, despite the differences in settings, it has been noted that the distribution of shallow-marine ich­
the degree of tidal influence in both cases, namely overall tidal domi­ nofaunas may be controlled by climate and that three climatic zones
nance in the Oficina Formation and tidal influence restricted to embayed may be recognized, namely (1) tropical and subtropical with Ophio­
areas in the Orinoco Delta (Méndez, 2000; Buatois et al., 2012). The morpha, echinoid burrows as well as other ichnotaxa, (2) temperate with
formation and overall physiography of coastal environments depend on echinoid burrows and Thalassinoides and (3) arctic with only molluscan
bedrock, glacial advance and retreat, relative sea-level changes, and and worm structures (Goldring and Cadee, 2004). Subsequent work
sediment redistribution by fluvial, coastal and marine processes (Hein extended the dominance of mollusk and worm burrows to the temperate
et al., 2014). Physical processes, such as tides, waves and winds, as well zone (Gingras et al., 2006). Incipient Ophiomorpha along the eastern
as biological and chemical factors, play a significant role across climatic coast of the Americas does not extend further into high latitudes,
belts (Kelletat et al., 2014). Tidal currents are the main hydrodynamic reaching as far as 34◦ N and 27◦ S, whereas modern examples of Tha­
agent in the systems analyzed in this paper, spanning lower- and lassinoides may extend up to 70◦ N and 50◦ S (Goldring and Cadee, 2004;
higher-latitude settings. However, tidal forces tend to be weaker in Martini, 2014). Both ichnogenera are particularly common in
higher latitudes because Coriolis effects are stronger there (Martinius marginal-marine deposits of the Oficina Formation. In general, crusta­
et al., 2014). In any case, tidal action is revealed by the overwhelming cean burrows tend to display higher diversity in tropical shallow-marine
presence of sedimentary structures, such as tubular tidalites, and mud environments of northern South America (e.g. Quiroz et al., 2010).
drapes. Also, the abundance of well-defined tidal channels dissecting However, Ophiomorpha has been mentioned in the tidal sandbars of the
tidal flats, such as those recorded in the Oficina Formation, is more high-latitude Eocene Aspelintoppen Formation of Arctic Norway
typical of low latitudes, being relatively rare in arctic to subarctic set­ (Plink-Bjorklund, 2005). This high-latitude occurrence of Ophiomorpha
tings (Martini, 2014). may simply reflect the warmer conditions of the Eocene. In contrast,
Estuary is often associated with transgressive conditions, whereas arctic tidal flats display a clear dominance of worm burrows (e.g. Nereis
delta is associated with regressions (Dalrymple et al., 2003, 2012; Dal­ divesiscolor and Arenicola marina) with some participation of
rymple and Choi, 2007). A tide-dominated estuary is a transgressive mollusk-generated structures, such as Macoma balthica (Aitken et al.,
coastal setting at the mouth of a river, receiving sediment from both 1988; Weslawski et al., 1999; Martini, 2014). In general, arctic and
river and sea, encompassing a wide spectrum of salinity levels (Pritch­ subarctic infaunal associations show remarkable similarities with cold
ard, 1967; Dalrymple et al., 1992; Dalrymple, 2006; Dalrymple et al., temperate tidal flats of the Bay of Fundy, except for the absence of
2012). A tide-dominated delta is a prograding coastal environment that Corophium volutator (Martini, 2014). Benthic activity in higher latitudes
receives clastic sediment from a river source, being mostly reworked by is controlled by temperature and insolation, as well as short runoff
tidal currents (Hori et al., 2001; Coleman, 1981). However, because seasons and strong fluvial discharge seasonality, which impart specific
shoreline trajectory is controlled by the sediment supply, the shoreline stress factors on the benthos (Martinius et al., 2014). As a result, ich­
may retrograde or prograde due to fluctuations in sediment supply at a nodiversity levels are affected negatively by a number of factors, such as
local scale. For example, the distributaries in abandoned delta lobes of ice cover, erosion by ice pushing and lifting of sediment (Martini, 2014).
the Mahakan Delta are transformed into estuaries as a result of reduced
sediment supply (e.g. Allen and Chambers, 1998).

22
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

7.2. Secular changes in bioturbation in tide-dominated marginal-marine Acknowledgments


settings
We thank Petróleos de Venezuela filial (PDVSA Intevep) for allowing
Secular changes in bioturbation by the continental and marginal- publication of this study. Financial support was provided by PDVSA
marine infauna have significantly impacted on sediment mixing (Bua­ Intevep. Genesis Rodriguez helped with the drawings. M.G.M. thanks
tois et al., 2005; Diez-Canseco et al., 2015). Although brackish-water additional funding by the George J. McLeod Enhancement Chair in
ichnofaunas, which reflect the activity of highly conservative biotas, Geology.
tend to be quite persistent through the Phanerozoic, subtle increases in
ichnodiversity and intensity of bioturbation are apparent through References
geologic time in marginal-marine environments (Buatois et al., 2005). In
this regard, the Miocene illustrates the appearance of the modern Ainsworth, R.B., Vakarelov, B.K., Nanson, R.A., 2011. Dynamic spatial and temporal
prediction of changes in depositional processes on clastic shorelines: toward
brackish-water benthos (Buatois et al., 2005) and the marginal-marine improved subsurface uncertainty reduction and management. AAPG (Am. Assoc. Pet.
ichnofaunas of the Oficina Formation favour direct comparison with Geol.) Bull. 95, 267–297.
modern analogues, particularly in tropical settings. Aitken, A.E., Risk, M.J., Howard, J.D., 1988. Animal–sediment relationships on subarctic
intertidal flat, Pangnirtung Fjord, Baffin Island, Canada. J. Sediment. Petrol. 58,
In comparison, changes in the composition of freshwater ichnofau­ 969–978.
nas through the Phanerozoic have been more remarkable. This is Allen, G.P., 1991. Sedimentary processes and facies in the Gironde estuary: a recent
particularly evident if the fluvio-estuary transition of the Oficina For­ model of macrotidal estuarine systems. In: Smith, G.D., Reinson, G.E., Zaitlin, B.A.,
Rahmani, R.A. (Eds.), Clastic Tidal Sedimentology, vol. 16. Canadian Society of
mation is compared with similar environments in the Paleozoic from an Petroleum Geologists, pp. 29–40.
ichnologic perspective. The Carboniferous freshwater to terrestrial ich­ Allen, G.P., Chambers, J.L.C., 1998. Sedimentation in the Modern and Miocene
nofaunas from the North American mid-continent are characterized by Mahakam Delta. Indonesian Petroleum Association, Jakarta, p. 231.
Amos, C.L., Tee, K.T., Zaitlin, B.A., 1991. The post-glacial evolution of Chignecto Bay,
superficial trackways and trails of the Mermia and Scoyenia Ichnofacies
Bay of Fundy, and its modern environment of deposition. In: Smith, D.G., Reinson, G.
(Buatois et al., 1997, 1998, 1998; Mángano and Buatois, 2004). In E., Zaitlin, B.A., Rahmani, R.A. (Eds.), Clastic Tidal Sedimentology, vol. 16.
contrast, post-Paleozoic freshwater to terrestrial ichnofaunas display a Canadian Society of Petroleum Geologists, pp. 59–89.
remarkable increase in burrowing depth in fluvio-tidal transitions, as is Audemard, F., Aspiritxaga, I., Baumann, P., Isea, A., Latreille, M., 1985. Marco Geológico
del Terciario de la Faja Petrolífera del Orinoco. In: Espejo, A., Rios, J.H., Bellizzia, N.
illustrated by deposits of the Oficina Formation, which are typically P. (Eds.), Memorias, VI Congreso Geologico Venezolano, vol. 1, pp. 70–108.
intensely bioturbated by meniscate burrows of the Scoyenia Ichnofacies Bhattacharya, J.P., MacEachern, J.A., 2009. Hyperpycnal rivers and prodeltaic shelves in
(Solórzano et al., 2017). Accordingly, secular increases in extent and the Cretaceous seaway of North America. J. Sediment. Res. 79, 184–209.
Botterill, S.E., Campbell, S.G., Timmer, E.R., Gingras, M.K., Hubbard, S., 2016.
depth of bioturbations have resulted in higher disturbance of the pri­ Recognition of wave-influenced deltaic and bay-margin sedimentation, Bluesky
mary sedimentary fabrics. Formation, Alberta. Bull. Can. Petrol. Geol. 64, 389–414.
In addition, the ability of the infauna to penetrate deeper into the Brekke, H., MacEachern, J.A., Roenitz, T., Dashtgard, S.E., 2017. The use of
microresistivity image logs for facies interpretations: an example in point-bar
sediment have led to common overprint of trace-fossil suites, as illus­ deposits of McMurray Formation, Alberta, Canada. AAPG (Am. Assoc. Pet. Geol.)
trated in the Oficina Formation by the common occurrence of freshwater Bull. 101, 655–682.
to terrestrial trace fossils cross-cutting previously emplaced brackish- Bridge, J.S., Jalfin, G.A., Georgieff, S.M., 2000. Geometry, lithofacies, and spatial
distribution of Cretaceous fluvial sandstone bodies, San Jorge Basin, Argentina:
water trace fossils (Diez-Canseco et al., 2015; Solórzano et al., 2017). outcrop analog for the hydrocarbon-bearing Chubut Group. J. Sediment. Res. 70,
This fact underscores the importance of performing ichnofabric analysis 341–359.
in order to make accurate use of ichnologic data to infer paleosalinity in Brownfield, R.L., Brenner, R.L., Pope, J.R., 1998. Distribution of the Bandera Shale of the
Marmaton Group, Middle Pennsylvanian of southeastern Kansas. Current Research
marginal-marine settings.
in Earth Science 241, 29–41.
Buatois, L.A., 2014. Descripción e interpretación sedimentológica icnológica del núcleo
8. Conclusions RPN-78. Informe técnico para PDVSA.
Buatois, L.A., Mángano, M.G., 2004. Animal-substrate interactions in freshwater
environments: applications of ichnology in facies and sequence stratigraphic analysis
Based on the recognition of nine sedimentary facies (FA-I) and five of fluvio-lacustrine successions. In: McIlroy, D. (Ed.), The Application of Ichnology to
facies associations (FA1-5), the Oficina Formation is interpreted as Palaeoenvironmental and Stratigraphic Analysis, vol. 228. Geological Society,
recording lowstand fluvial deposits (lower member), passing upward Special Publication, pp. 311–333.
Buatois, L.A., Mángano, M.G., Maples, C.G., Lanier, W.P., 1997. The paradox of
into transgressive estuary deposits (middle member), and highstand nonmarine ichnofaunas in tidal rhythmites: integrating sedimentologic and
lower delta-plain deposits (upper member). The abundance of mud ichnologic data from the Late Cretaceous of eastern Kansas, USA. Palaios 12,
drapes, bioturbated IHS and tubular tidalites through all the middle and 467–481.
Buatois, L.A., Mángano, M.G., Maples, C.G., Lanier, W.P., 1998a. Ichnology of an Upper
upper members suggests tidal dominance. Ichnologic evidence indicates Carboniferous fluvio-estuarine paleovalley: the Tonganoxie sandstone, Buildex
freshwater conditions in the fluvial systems, the inner part of the estuary quarry, eastern Kansas, USA. J. Paleontol. 72, 152–180.
and the delta plain, whereas brackish-water conditions dominated in the Buatois, L.A., Mángano, M.G., Genise, J.F., Taylor, T.N., 1998b. The ichnologic record of
the continental invertebrate invasion: evolutionary trends in environmental
rest of the estuary valley, including its outer region, which is consistent expansion, ecospace utilization, and behavioral complexity. Palaios 13, 217–240.
with the embayed physiography of the paleocoastline in the Orinoco Oil Buatois, L.A., Mángano, M.G., 2011. Ichnology Organism—Substrate Interactions in
Belt. Fluvial, estuary, and deltaic strata extend across the whole belt. In Space and Time (University of Saskatchewan). Cambridge University Press.
Buatois, L.A., Gingras, M.K., MacEachern, J., Mángano, M.G., Zonneveld, J.-P.,
these zones, the channels display a south-north orientation. Estuary
Pemberton, S.G., Netto, R.G., Martin, A.J., 2005. Colonization of brackish-water
valleys display increased marine influence towards the north-northeast. systems through time: evidence from the trace-fossil record. Palaios 20, 321–347.
Comparisons with other marginal-marine units under a broad spectrum Buatois, L.A., Santiago, N., Herrera, M., Plink-Bjorklund, P., Steel, R., Espin, M.,
of latitudinal settings stress the importance of tidal dominance and Parras, K., 2012. Sedimentological and ichnological signatures of changes in wave,
river and tidal influence along a Neogene tropical deltaic shoreline. Sedimentology
relative sea-level changes as main controls on sedimentation. However, 59, 1568–1612.
the establishment of extensive marginal-marine wetland systems and the Chen, S., Steel, R.J., Olariu, C., 2015. Palaeo-Orinoco (Pliocene) channels on the tide-
types of burrowing infauna that characterize the Oficina Formation are dominated Morne L’Enfer delta lobes and estuaries, SW Trinidad. In: Ashworth, P.J.,
Best, J.L., Parsons, D.R. (Eds.), Fluvial-Tidal Sedimentology. Developments in
consistent with the tropical nature of these coastal ecosystems. Sedimentology, vol. 68. Elsevier, pp. 227–281.
Chigne, N., Camposano, C.L., Rodriguez, I., Rodriguez, W.J., Castillo, M.G., Daza, J.,
Declaration of competing interest Dávila, M., Márquez, F., Di Gianni, N., Mata, L., Ruiz, M., Guerra, B., Ruiz, F.,
Arocha, B., Medina, J., 2002. Proyecto generación de prospecto Capiricual. Informe
técnico PDVSA Exploración y Producción.
The authors declare that they have no known competing financial Choi, K.S., Dalrymple, R.W., Chun, S.S., Kim, S., 2004. Sedimentology of modern,
interests or personal relationships that could have appeared to influence inclined heterolithic stratification (IHS) in the macrotidal Han River Delta, Korea.
J. Sediment. Res. 74, 677–689.
the work reported in this paper.

23
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Coleman, J.M., 1981. Deltas: Processes of Deposition and Models for Exploration. Management: Albuquerque, Transatlantic Arts Incorporated. Colston Paper 30,
Burgess Publishing, Minneapolis, p. 124. pp. 162–172.
Crerar, E.E., Arnott, R.W.C., 2007. Facies distribution and stratigraphic architecture of Haq, B.U., Schutter, S.R., 2008. A chronology of Paleozoic sea-level changes. Science
the Lower Cretaceous McMurray Formation, Lewis Property, northeastern Alberta. 322, 64–68.
Bull. Can. Petrol. Geol. 55, 99–124. Harris, B.S., Timmer, E.R., Ranger, M.J., Gingras, M.K., 2016. Continental ichnology of
Dalrymple, R.W., 2006. Incised valleys in space and time: an introduction to the volume the Lower McMurray Formation inclined heterolithic strata at Daphne Island,
and an examination of the controls on valley formation and filling. In: Dalrymple, R. Athabasca River, north-eastern Alberta, Canada. Bull. Can. Petrol. Geol. 64,
W., Tillman, R.W. (Eds.), Incised Valleys in Space and Time, vol. 85. Society for 218–232.
Sedimentary Geology, pp. 5–12. Hein, F.J., 2015. The Cretaceous McMurray oil sands, Alberta, Canada: a world-class,
Dalrymple, R.W., Choi, K., 2007. Morphologic and facies trends through the fluvial- tidally influenced fluvial–estuarine system—an Alberta government perspective. In:
marine transition in tide-dominated depositional systems: a schematic framework for Ashworth, P.J., Best, J.L., Parsons, D.R. (Eds.), Fluvial-Tidal Sedimentology,
environmental and sequence-stratigraphic interpretation. Earth Sci. Rev. 81, Developments in Sedimentology, vol. 68. Elsevier, pp. 561–621.
135–174. Hein, F.J., Graham, D., Fairgrieve, B., 2013. A regional geologic framework for the
Dalrymple, R.W., Knigh, R.J., Zaitlin, B.A., Middleton, G.V., 1990. Dynamics and facies Athabasca oil sands, northeastern Alberta, Canada. In: Hein, F.J., Leckie, D.,
model of a macrotidal sand-bar complex, Cobequid Bay-Salmon River Estuary (Bay Larter, S., Suter, J.R. (Eds.), Heavy-oil and Oil-Sand Petroleum Systems in Alberta
of Fundy). Sedimentology 37, 577–612. and beyond, American Association of Petroleum Geologists Studies in Geology, vol.
Dalrymple, R.W., Zaitlin, B.A., Boyd, R., 1992. Estuarine facies models: conceptual basis 64, pp. 207–250.
and stratigraphic implications. J. Sediment. Petrol. 62, 1130–1146. Hein, C.J., Fitzgerald, D.M., Buynevich, I.V., Van Heteren, S., Kelley, J.T., 2014.
Dalrymple, R.W., Baker, E.K., Harris, P.T., Hughes, M.G., 2003. Sedimentology and Evolution of paraglacial coasts in response to changes in fluvial sediment supply. In:
stratigraphy of a tide-dominated foreland-basin delta (Fly River, Papua New Martini, I.P., Wanless, H.R. (Eds.), Sedimentary Coastal Zones from High to Low
Guinea). In: Sidi, F.H., Nummedal, D., Imbert, P., Darman, H., Posamentier, H.W. Latitudes: Similarities and Differences, vol. 388. Geological Society, London, Special
(Eds.), Tropical Deltas of Southeast Asia, vol. 76. Society for Sedimentary Geology Publications, pp. 1–32.
Special Publication, pp. 147–173. Hori, K., Saito, Y., Zhoa, Q., Cheng, X., Wang, P.Y., Li, C., 2001. Sedimentary facies and
Dalrymple, R.W., Mackay, D.A., Ichaso, A.A., Choi, K., 2012. Processes, Holocene progradation rate of the Changjiang (Yangtze) delta, China.
morphodynamics, and facies of tide-dominated estuaries. In: Davis, R.A., Geomorphology 41, 233–248.
Dalrymple, R.W. (Eds.), Principles of Tidal Sedimentology. Springer, New York, Hovikoski, J., Räsänen, M., Gingras, M., Ranzi, A., 2008. Tidal and seasonal controls in
pp. 79–107. the formation of Late Miocene inclined heterolithic stratification deposits, western
Dalrymple, R.W., Kurcinka, C.E., Jablonski, B.V.J., Ichaso, A.A., Mackay, D.A., 2015. Amazonian foreland basin. Sedimentology 55, 499–530.
Deciphering the relative importance of fluvial and tidal processes in the Huang, W., Li, S., Chen, H., Fu, C., 2020. A river-dominated to tide-dominated delta
fluvial–marine transition. In: Ashworth, P.J., Best, J.L., Parsons, D.R. (Eds.), Fluvial- transition: a depositional system case study in the Orinoco heavy oil belt, Eastern
Tidal Sedimentology. Developments in Sedimentology, vol. 68. Elsevier, pp. 3–45. Venezuelan Basin. Mar. Petrol. Geol. 118, 104389.
Dashtgard, S.E., Gingras, M.K., 2005. Facies architecture and ichnology of recent salt- Hubbard, S.M., Smith, D.G., Nielsen, H., Leckie, D.A., Fustic, M., Spencer, R.J., Bloom, L.,
marsh deposits: Waterside Marsh, New Brunswick, Canada. J. Sediment. Res. 75, 2011. Seismic geomorphology and sedimentology of a tidally influenced river
596–607. deposit, Lower Cretaceous Athabasca oil sands, Alberta, Canada. AAPG (Am. Assoc.
Dashtgard, S.E., Pearson, N.J., Gingras, M.K., 2014. Sedimentology, ichnology, ecology Pet. Geol.) Bull. 95, 1123–1145.
and anthropogenic modification of muddy tidal flats in a cold-temperate Isea, A., 1987. Geological synthesis of the Orinoco oil belt, eastern Venezuela. J. Petrol.
environment: Chignecto Bay, Canada. In: Martini, I.P., Wanless, H.R. (Eds.), Geol. 10, 135–148.
Sedimentary Coastal Zones from High to Low Latitudes: Similarities and Differences, Jablonski, B.V., Dalrymple, R.W., 2016. Recognition of strong seasonality and climatic
vol. 388. Geological Society, London, Special Publications, pp. 229–245. cyclicity in an ancient, fluvially dominated, tidally influenced point bar: middle
Di Croce, J., Bally, A.W., Vail, E., 1999. In: Mann (Series Editor, R., Hsu, K.J. (Eds.), McMurray Formation, Lower Steepbank River, north-eastern Alberta, Canada.
Sequence Stratigraphy of the Eastern Venezuelan Basin. Caribbean Basins. Sedimentology 63 (3), 552–585.
Sedimentary Basins of the World, vol. 4. Elsevier Science B.V., Amsterdam, Jouanneau, J.M., Latouche, C., 1981. The Gironde Estuary: Stuttgart: stuttgart, E.
pp. 419–476. Sehweizerbart’sche verlagsbuchhandlung. Contrib. Sedimentol. 10, 115.
Diez-Canseco, D., Buatois, L.A., Mángano, M.G., Rodríguez, W.J., Solórzano, E.J., 2015. Kelletat, D.H., Scheffers, A.M., May, S.M., 2014. Coastal environments from polar regions
The ichnology of the fluvial–tidal transition: interplay of ecologic and evolutionary to the tropics: a geographer’s zonality perspective. In: Martini, I.P., Wanless, H.R.
controls. In: Ashworth, P.J., Best, J.L., Parsons, D.R. (Eds.), Fluvial-Tidal (Eds.), Sedimentary Coastal Zones from High to Low Latitudes: Similarities and
Sedimentology. Developments in Sedimentology, vol. 68, pp. 283–321. Differences, vol. 388. Geological Society, London, Special Publications, pp. 1–32.
Diez-Canseco, D., Buatois, L.A., Mángano, M.G., Díaz-Molina, M., Benito, M.I., 2016. La Croix, A.D., Dashtgard, S.E., 2014. Of sand and mud: sedimentological criteria for
Ichnofauna from coastal meandering channel systems (Upper Cretaceous Tremp identifying the turbidity maximum zone in a tidally influenced river. Sedimentology
Formation, South-Central Pyrenees, Spain): delineating the fuvial-tidal transition. 61, 1961–1981.
J. Paleontol. 90, 250–268. Lambiase, J.J., 1980. Sediment dynamics in the macrotidal Avon River estuary, Bay of
Ekwenye, O.C., Nichols, G., Nwajide, S.C., Obi, G.C., Onyemesili, O.C., 2017. An insight Fundy. Can. J. Earth Sci. 17, 1628–1641.
into the Eocene tide-dominated estuarine system: implications for Latreille, M., Baumann, P., Audemard, F., Muñoz de, A.N., Aspiritxaga, I., Cassani, F., De
palaeoenvironmental and sequence stratigraphic interpretations. Arabian Journal Menas, J., Isea, A., Taheri, M., Gallagher, M., Canache, M., 1983. Modelo geológico
Geosciences 10 (371), 1–20. integrado de la Faja Petrolífera del Orinoco, 83. Informe técnico PDVSA Intevep INT-
Fan, D., 2012. Open-coast tidal flats. In: Davis, R.A., Dalrymple, R.W. (Eds.), Principles of 00753.
Tidal Sedimentology. Springer, New York, pp. 187–229. Legler, B., Howard, J., Hampson, G.J., Massart, B.Y., Jackson, C., Jackson, M.D.,
Fan, D., Tu, J., Shang, S., Cai, G., 2014. Characteristics of tidal-bore deposits and facies Barkooky, A.E., Ravnas, R., 2013. Facies model of a fine-grained, tide-dominated
associations in the Qiantang Estuary, China. Mar. Geol. 348, 1–14. delta: Lower Dir Abu Lifa Member (Eocene), western desert, Egypt. Sedimentology
Gingras, M.K., Leckie, D.A., 2017. The argument for tidal and brackish water influence in 60, 1313–1356.
the McMurray Formation reservoirs. Reservoir 2, 21–24. Lettley, C.D., Sg Pemberton, S.G., Gingras, M.K., Ranger, M.J., Blakney, B.J., 2009.
Gingras, M.K., Zonneveld, J., 2015. Tubular tidalites: a biogenic sedimentary structure Integrating sedimentology and ichnology to shed light on the system dynamics and
indicative of tidally influenced sedimentation. J. Sediment. Res. 85, 845–854. paleogeography of an ancient riverine estuary. In: MacEachern, J.A., Bann, K.L.,
Gingras, M.K., MacEachern, J.A., Pemberton, G.S., 2006. Latitudinal (Climatic) Controls Gingras, M.K., Pemberton, S.G. (Eds.), Applied Ichnology. Society for Sedimentary
on Neoichnological Assemblages of Modern Marginal Depositional Environments. Geology Short Course Notes, vol. 52, pp. 147–165.
American Association of Petroleum Geologist 2006 Annual Convention, Houston, MacEachern, J.A., Bann, K.L., Bhattacharya, J., Howell, C.D., 2005. Ichnology of deltas:
p. 38. Abstract. organism responses to the dynamic interplay of rivers, waves, storms, and tides. In:
Gingras, M.K., MacEachern, J.A., Dashtgard, S.E., 2012. The potential of trace fossils as Giosan, L., Bhattacharya, J.P. (Eds.), River Deltas-Concepts, Models, and Examples,
tidal indicators in bays and estuaries. Sediment. Geol. 279, 97–106. vol. 83. Society for Sedimentary Geology Special Publications, pp. 49–85.
Gingras, M.K., MacEachern, J.A., Dashtgard, S.E., Ranger, M.J., Pemberton, G.S., 2016. Mackay, D.A., Dalrymple, R.W., 2011. Dynamic mud deposition in a tidal environment:
The significance of trace fossils in the McMurray Formation, Alberta, Canada. Bull. the record of fluid-mud deposition in the Cretaceous Bluesky Formation, Alberta,
Can. Petrol. Geol. 64, 233–250. Canada. J. Sediment. Res. 81, 901–920.
Goldring, R., Cadee, G.C., et al., 2004. Climatic control of trace fossil distribution in the Mángano, M.G., Buatois, L.A., 2004. Ichnology of Carboniferous tide-influenced
marine realm. In: McIlroy, D. (Ed.), The Application of Ichnology to environments and tidal flat in the North American Midcontinent. In: McIRoy, D.
Palaeoenvironmental and Stratigraphic Analysis, vol. 228. Geological Society, (Ed.), The Application of Ichnology to Palaeoenvironmental and Stratigraphic
London, Special Publications, pp. 77–92. Analysis, vol. 228. Geological Society of London Special Publications, pp. 157–178.
Gugliotta, M., Flint, S., Hodgson, D., Veiga, G., 2016a. Recognition criteria, Martini, I.P., 2014. General considerations and highlights of low-lying coastal zones:
characteristics and implications of the fluvial to marine transition zone in ancient passive continental margins from the poles to the tropic. In: Martini, I.P., Wanless, H.
deltaic deposits (Lajas Formation, Argentina). Sedimentology 63, 1971–2001. R. (Eds.), Sedimentary Coastal Zones from High to Low Latitudes: Similarities and
Gugliotta, M., Kurcinka, C.E., Dalrymple, R.W., Flint, S.S., Hodgson, D.M., 2016b. Differences, vol. 388. Geological Society, London, Special Publications, pp. 1–32.
Decoupling seasonal fluctuations in fluvial discharge from the tidal signature in Martinius, A., Hegner, J., Kaas, I., Bejarano, C., Mathieu, X., Mjøs, R., 2012.
ancient deltaic deposits: an example from the Neuquén Basin, Argentina. J. Geol. Sedimentology and depositional model for the early Miocene Oficina Formation in
Soc. 173, 94–107. the Petrocedeño Field (Orinoco heavy-oil belt, Venezuela). Mar. Petrol. Geol. 35,
Hamilton, D., 1979. The high-energy, sand and mud regime of the Severn estuary, S.W. 354–380.
Britain. In: Severn, R.T., Dineley, D., Hawker, L.E. (Eds.), Tidal Power and Estuary Martinius, A., Hegner, J., Kaas, I., Bejarano, C., Mathieu, X., Mjøs, R., 2013. Geologic
reservoir characterization and evaluation of the Petrocedeño Field, early Miocene

24
E.J. Solórzano et al. Journal of South American Earth Sciences 112 (2021) 103515

Oficina Formation, Orinoco heavy oil belt, Venezuela. In: Hein, F.J., Leckie, D., Rodríguez, W.J., Buatois, L.A., Mángano, M.G., Solórzano, E.J., 2018. Sedimentology,
Larter, S., Suter, J.R. (Eds.), Heavy Oil and Oil-Sand Petroleum Systems in Alberta ichnology, and sequence stratigraphy of the Miocene Oficina Formation, Junín and
and beyond, American Association of Petroleum Geologists Studies in Geology, vol. Boyacá areas, Orinoco Oil Belt, Eastern Venezuela Basin. Mar. Petrol. Geol. 92,
64, pp. 103–131. 213–233.
Martinius, A.W., Fraticelli, C.M., Markwick, P., Suter, J.R., 2014. Latitudinal Controls on Rodríguez-Tovar, F.J., Mayoral, E., Santos, A., Dorador, J., Wetzel, A., 2019. Crowded
Stratigraphic Models and Sedimentary Concepts. American Association of Petroleum tubular tidalites in Miocene shelf sandstones of southern Iberia. Palaeogeogr.
Geologists. Search and Discovery Article #41424. Palaeoclimatol. Palaeoecol. 521, 1–9.
Martinius, A.W., Jablonski, B.V.J., Fustic, M., Strobl, R., Van den Berg, J.H., 2015. Sarma, J.N., 2014. The Brahmaputra River in Assam: The Outsized Braided Himalayan
Fluvial to tidal transition zone facies in the McMurray Formation (Christina River, River. Landscapes and Landforms of India. Springer, Dordrecht, pp. 165–172.
Alberta, Canada), with emphasis on the reflection of flow intensity in bottomset Shaw, J., Courtney, R.C., 2002. Postglacial coastlines of Atlantic Canada: digital images.
architecture. In: Ashworth, P.J., Best, J.L., Parsons, D.R. (Eds.), Fluvial-Tidal Geological Survey of Canada. Open-File Report 4302.
Sedimentology. Developments in Sedimentology, vol. 68. Elsevier, pp. 445–480. Sisulak, C.F., Dashtgard, S.E., 2012. Seasonal controls on the development and character
Melnyk, S., Gingras, M.K., 2020. Using ichnological relationships to interpret heterolithic of inclined heterolithic stratification in a tide-influenced, fluvially dominated
fabrics in fluvio-tidal settings. Sedimentology 67, 1069–1083. channel: Fraser River, Canada. J. Sediment. Res. 82, 244–257.
Méndez, J.B., 2000. Delta del Orinoco (Geología), 13. Instituto Nacional de Geología y Solomon, S., Qin, D., et al. (Eds.), 2007. Climate Change 2007: the Physical Science Basis.
Minería, República Bolivariana de Venezuela, Boletín Geologico Publicación Contribution of Working Group I to the Fourth Assessment Report of the
Especial, p. 127. Intergovernmental Panel on Climate Change. Cambridge University Press,
Miall, A., 2010. The Geology of Stratigraphic Sequences. Springer, London, p. 522. Cambridge.
Moreton, D.J., Carter, B.J., 2015. Characterizing alluvial architecture of point bars Solórzano, E., Farías, A., 2017. Modelo Cronoestratigráfico para las áreas de Carabobo y
within the McMurray Formation, Alberta, Canada, for improved bitumen resource Ayacucho en la Faja Petrolífera del Orinoco. Informe Técnico PDVSA Intevep INT-
prediction and recovery. In: Ashworth, P.J., Best, J.L., Parsons, D.R. (Eds.), Fluvial- 2270, 2017.
Tidal Sedimentology. Developments in Sedimentology, vol. 68. Elsevier, Solórzano, E., Farías, A., Cabrera, D., Buatois, L.A., 2015. Chronostratigraphic
pp. 529–559. Framework for the Ayacucho and Carabobo Areas in the Orinoco Oil Belt. Society of
Musial, G., Reynaud, J., Gingras, M.K., Féniès, H., Labourdette, R., Parize, O., 2012. Petroleum Engineers International, western Venezuela Petroleum Section, 3er South
Subsurface and outcrop characterization of large tidally influenced point bars of the American Oil and Gas Congress. SPE WVS-614.
Cretaceous McMurray Formation (Alberta, Canada). Sediment. Geol. 279, 156–172. Solórzano, E.J., Buatois, L.A., Rodríguez, W.J., Mángano, M.G., 2017. From freshwater to
Obi, G.C., Okogbue, C.O., 2004. Sedimentary response to tectonism in the Campanian- fully marine: exploring animal-substrate interactions along a salinity (Miocene
Maastrichtian succession, Anambra Basin, southeastern Nigeria. J. Afr. Earth Sci. 38, Oficina Formation of Venezuela). Palaeogeogr. Palaeoclimatol. Palaeoecol. 482,
99–100. 30–47.
Olariu, C., Steel, R.J., Olariu, M.I., Choi, K., 2015. Facies and architecture of unusual Suarez, L., Solórzano, E., 2014. Descripción sedimentológica de los núcleos del pozo
fluvial–tidal channels with inclined heterolithic strata: Campanian Neslen MFK- 6. Informe Técnico PDVSA Intevep INT-14815, 2014.
Formation, Utah, USA. In: Ashworth, P.J., Best, J.L., Parsons, D.R. (Eds.), Fluvial- Taylor, A.M., Goldring, R., 1993. Description and analysis of bioturbation and
Tidal Sedimentology. Developments in Sedimentology, vol. 68. Elsevier, ichnofabric. J. Geol. Soc. 150, 141–148.
pp. 353–394. Thomas, R.G., Smith, D.G., Wood, J.M., Visser, J., Caverley-Range, E.A., Koster, E.H.,
Parnaud, F., Gou, Y., Pascual, J.C., Capello, M.A., Truskowski, I., Passalacqua, H., 1995. 1987. Inclined heterolithic stratification: terminology, description, interpretation
Stratigraphic synthesis of western Venezuela. In: Tankard, A.J., Suarez, S.R., and significance. Sediment. Geol. 53, 123–179.
Welsink, H.J. (Eds.), Petroleum Basins of South America, vol. 62. American Toro, M., Casanova, M., Marquina, M., Chang, R., Brito, F., Quintero, M., Canache, M.,
Association of Petroleum Geologists Memoir, pp. 741–756. González, O., Asencio, E., Benzaquen, I., Velásquez, A., Rodulfo, F., Aguado, B.,
Parra, K., Santiago, N., Escorcia, L., Hernandez, Z., Luna, H., Marcano, E., Marcano, J., Crespo, J., 2001. Modelo estático integrado de Cerro Negro, Area Bitor (Faja
Moya, M., Oliveros, R., Peña, Y., Rivas, Y., 2018. Proyecto evaluación del sistema Petrolífera del Orinoco). Informe Técnico PDVSA Intevep INT-08380, 2001.
petrolífero, Subcuenca de Guarico. Informe Técnico PDVSA Gerencia General de Törő, B., Pratt, B., 2015. Eocene paleoseismic record of the Green River Formation, fossil
Exploración. Basin, Wyoming, U.S.A.: Implications of synsedimentary deformation structures in
Pemberton, S.G., Wightman, D.M., 1992. Ichnological characteristics of brackish water lacustrine carbonate mudstone. J. Sediment. Res. 85, 855–884.
deposits. In: Pemberton, S.G. (Ed.), Applications of Ichnology to Petroleum Weimer, R.J., Howard, J.D., Lindsay, D.R., 1981. Tidal flats and associated tidal
Exploration: A Core Workshop, vol. 17. Society for Sedimentary Geology Core channels. In: Scholle, P.A., Spearing, D. (Eds.), Sandstone Depositional
Workshop, pp. 141–167. Environments, vol. 31. American Association of Petroleum Geologist Memoir,
Pemberton, S.G., Spila, M., Pulham, A.J., Sauders, T., MacEachern, J.A., Robbins, D., pp. 191–246.
Sinclair, I.K., 2001. Ichnology and Sedimentology of Shallow to Marginal Marine Weslawski, J.M., Szymelfenig, M., 1999. Community composition of tidal flats of
Systems. Ben Nevis and Avalon Reservoirs. Jeanne d’ Arc Basin. Geological Spitsbergen: consequence of disturbance? In: Gray, J.D., Amborse, W.,
Association of Canada Short Course Notes 15, St. John’s. Szaniawska Jr., A. (Eds.), Biochemical Cycling and Sediment Ecology. Kluwer
Plink-Bjorklund, P., 2005. Stacked fluvial and tide-dominated estuarine deposits in high- Academic, Dordrecht, pp. 185–193.
frequency (fourth-order) sequences of the Eocene Central Basin, Spitsbergen. Wetzel, A., Carmona, N., Ponce, J., 2014. Tidal signature recorded in burrow fill.
Sedimentology 52, 391–428. Sedimentology 61, 1198–1210.
Plummer, P.S., Gostin, V.A., 1981. Shrinkage cracks: desiccation or syneresis. Yang, B.C., Dalrymple, R.W., Chun, S.S., 2005. Sedimentation on a wave-dominated,
J. Sediment. Res. 51, 1147–1156. open-coast tidal flat, south-western Korea: A summer tidal flat–winter shoreface.
Pritchard, D.W., 1967. What is an estuary? Physical viewpoint. In: Lauff, G.H. (Ed.), Sedimentology 52, 235–252.
Estuaries. American Association for the Advancement of Science, vol. 83, pp. 3–5. Yang, B.C., Dalrymple, R.W., Chun, S.S., Lee, H.J., 2006. Transgressive sedimentation
Proyecto Orinoco Magna Reserva, 2012. Atlas de Integración Regional de la Faja and stratigraphic evolution of a wave-dominated coast, western Korea. Mar. Geol.
Petrolífera del Orinoco. Informe Técnico PDVSA CVP. 235, 35–48.
Quiroz, L.I., Buatois, L.A., Mangano, M.G., Jaramillo, C., 2010. Ichnology and Yang, B.C., Gingras, M.K., Pemberton, S.G., Dalrymple, R.W., 2008. Wave-generated
sedimentology of the late Miocene Urumaco Formation, Falcón Basin, northwestern tidal bundles as an indicator of wave-dominated tidal flats. Geology 36, 39–42.
Venezuela: sediment-organism interactions in tropical deltas. In: Abstracts, 18th Zaitlin, B.A., Dalrymple, R.W., Boyd, R., 1994. The stratigraphic organization of incised-
International Sedimentological Congress, Mendoza, Argentina. valley systems associated with relative sea-level changes. In: Boyd, R., Zaitlin, B.A.,
Reineck, H., Wunderlich, F., 1968. Classification and origin of flaser and lenticular Dalrymple, R. (Eds.), Incised Valley Systems: Origin and Sedimentary Sequences:
bedding. Sedimentology 11, 99–104. Society of Economic Paleontologists and Mineralogists. Special Publication No. 51,
pp. 45–60.

25

You might also like