You are on page 1of 16

Complex Petroleum Systems Developed

by Subduction Process, Offshore Talara Basin,


Northwest Peru

Fernando Zúñiga-Rivero
Hugh Hay-Roe
BPZ & Associates, Inc.
11999 Katy Freeway, Suite 560
Houston, Texas 77079
(e-mail: bpz@bpzenergy.com)
Linda Jenkins
10 McCarty Island Road
Picayune, Mississippi 39466
(e-mail: libertybelle55@hotmail.com)
Allen Lowrie
238 FZ Goss Road
Picayune, Mississippi 39466-9458
(e-mail: alowrie@webtv.com)

Abstract

Study of 2-D marine seismic data of 1993 vintage along the Peruvian coast has led to the idea that
subduction of marine sediments beneath a continental plate may be a potentially important mechanism
for petroleum generation. The concept can be embodied in a new geologic model involving five essen-
tial factors:
1. Subduction of both continental and oceanic sediments along the Peru Trench.
2. Incorporation into the subducted material of rich organic material from two different sources:
3. From the west: subducted deep-ocean sediments.
4. From above: organic matter precipitated from the surficial Humboldt Current.
5. Probable duration of this previously unexamined mechanism during the past 100 million years.
6. Upward migration of the thermally mature, buoyant hydrocarbon fluids (first as liquid petroleum,
later as natural gas) through the shattered sediments and along normal faults cutting the slope mar-
gin.
7. Entrapment of rising hydrocarbons in anticlinal closures, normal fault-blocks and fluviodeltaic
stratigraphic traps.
Hydrocarbon formation must have benefited from unique oceanographic-geologic conditions asso-
ciated with the “El Niño” phenomenon, and probably effective for the past 20-25 Ma. El Niño
periodically generates a massive erosion of the coastal plain, which provides reservoir-grade sediments
to the continental slope. Normal organic output from the nutrient-rich Humboldt Current is a primary
factor in the creation of source rocks; the other source is subducted deep-sea sediments. Shallow cores
taken during Leg 112 of DSDP (Yeats et al, 1976) indicate total organic carbon (TOC) ranging from
0.5% to 2.0%. Hydrocarbon generation could take place after incorporation into the subduction zone.
Subsidence and burial of Tertiary, Mesozoic, and Paleozoic sediments would also provide hydrocarbon
sources, whether derived from the paleo-shelf or the paleo-slope. The presence of two separate hydro-
GCSSEPM Foundation 21st Annual Research Conference 755
Petroleum Systems of Deep-Water Basins, December 2–5, 2001
Complex Petroleum Systems Developed by Subduction Process, Offshore Talara Basin, Northwest Peru

carbon-generating systems would enhance the prospectiveness of the Peruvian continental margin. The
logical place to commence exploration is westward from the prolific and long-developed onshore and
shallow offshore sectors of the Talara Basin.

Introduction

The “petroleum system” concept (Magoon and Dow, 1994a and 1994b; Dow 1974) is here applied
to a regional study of the offshore Talara Basin, northwest Peru (Fig. 1). The on-shore Talara basin has
been a petroleum producer since the 1860s and has yielded some 1.6 billion barrels of oil.1 This petro-
leum is produced within an extensive forearc basin that became shallow in the early Tertiary and
emergent in late Miocene and Quaternary times. Those tectonic effects are related to the growth of the
accretionary wedge associated with the subduction of the Nazca plate, part of the Pacific Basin since at
least the Cretaceous, and possibly as far back as the Late Paleozoic.
A petroleum system may be defined as “a natural system that encompasses . . . active source rock
and all related oil and gas and which includes all the geologic elements and processes that are essential
if a hydrocarbon accumulation is to exist.” Petroleum, as used here, includes concentrations of the fol-
lowing (Magoon and Dow, 1994b, p. 10):
1. Thermal or biogenic gas found in conventional reservoirs or in gas hydrates, tight reservoirs, frac-
tured shale and coal.
2. Condensates, crude oils, and asphalts found in nature.
The term petroleum system “describes the interdependent elements and processes that together form
the functional unit that creates hydrocarbon accumulations. The essential elements include petroleum
source, reservoir, seal, and overburden (rocks and/or unconsolidated sediments), while the geologic pro-
cesses are trap formation and generation–migration–accumulation of petroleum. These elements and
processes must occur in a range of time and space such that organic matter in a source rock can be con-
verted to a petroleum accumulation” (Magoon and Dow, 1994b, p. 11).
This paper proposes that by acting with increased temperature and pressure on the marine and conti-
nental sediments being subducted under the South American plate, the plate-collision process has
facilitated the generation, migration, and entrapment of hydrocarbons.

Source Rocks

The Peruvian continental margin appears to have two main sources of organic matter. One source is
the sediments brought into the subduction zone. The Talara basin extends from around 3°30’S to
6°30’S, a north-south distance of some 330 km. Given a subduction rate of about 10 cm/year (Lowrie
and Hey, 1981; Vogt et al., 1976) since the Cretaceous—65 million years—6500 km of oceanic crust
and sediments from the Nazca plate could have been subducted beneath Peru along the western South
American continental margin. A realistic thickness for sediment cover over oceanic crust of 0.1 km
seems reasonable (Yeats et al., 1976).
Thus, for each 1 km of trench length, a total of only 650 km3 of Tertiary deposits could have been
added to the Peruvian accretionary wedge or subducted. The material subducted offshore from the
Talara basin could have had a volume equivalent to 215,000 km3. With a TOC content of 0.5% to 2.0%
in these sediments, TOC volume brought into subduction could range from around 1,070 to 4,300 km3.
If these volumes of carbon were entirely converted to hydrocarbons, they would yield from 6.7 trillion
barrels to 27 trillion barrels of oil2 (without considering migration, trapping, and preservation). The 1.6

1. After the initial drilling success in 1859 in Pennsylvania, one of Colonel Drake’s partners went to NW
Peru and applied the nascent technology of petroleum production to an area near the oil seeps known there
since the pre-colonial era.
2. Conversion: 1 m3=6.2898 (≈ 6.3) barrels of oil.

756
Zúñiga-Rivero et al.

Figure 1. Peruvian Coastal Basins: shallow-water continental shelf basins, Salavery, Huacho, and Pisco, separated from the
continental slope basins—Pimentel, Trujillo, Lima, and Paracas—with the Talara, Tumbes-Progreso, and Moquegua-
Mollendo basins straddling on-shore and shelf. Modified from Zuniga-Rivero et al. (1998a, Fig. 2).

757
Complex Petroleum Systems Developed by Subduction Process, Offshore Talara Basin, Northwest Peru

Figure 2. Northward migration of Australia: impacting the evolution of the Indonesian volcanic arc establish a continuous
Austral-Asian landmass and terminating the equatorial Tethys Sea, conditions appropriate for initiation of the El Niño
phenomenon. Ages before present given in Ma, modified from Daly et al. (1991), their Figure 2.

758
Zúñiga-Rivero et al.

Figure 3. The bathymetric outline of the Nazca Ridge compared to a similarly shaped and sized portion of the Mascarene
Plateau, western-central Indian Ocean: comparison with assumed pre-subduction Nazca Ridge extended. Note that the
northern Mascarene is continental crust and therefore harder to subduct than oceanic crust. If this model is correct for the
extended Nazca Ridge, then subduction would have been complex. [Adapted from Nur and Ben-Avraham (1981), their
Figure 6.].

billion barrels extracted so far could thus be as little as one ten-thousandth of the total hydrocarbons that
could hypothetically be generated.
The second carbon source is associated with the Humboldt Current. The west- to east-flowing Cir-
cum-Antarctic Current is hindered by the South American continent, and the resultant northward-
flowing Humboldt Current produces abundant upwelling. This upwelling off Peru and Chile brings
nutrients near the sea surface, resulting in a major surface-water fishery. The resulting organic matter
deposited onto the continental slope contains up to 2% to 3% TOC. During the periodic El Niño events,
massive die-offs contribute to the organic total, although at the same time, upwelling ceases and there is
no deposition of organic matter from that source.

Possible Geologic Duration of the El Niño Phenomenon

A question arises concerning how long “El Niños” have occurred. Published reports speculate that
they originated during historic times (Quinn et al., 1987; Rodbell et al., 1990; Sandweiss et al., 1996).
By employing first principles, it is possible to provide reasonable estimates of when the El Niño phe-
nomenon began. First, a review of what El Niño is, and the processes that create it, is in order.
The equatorial east- to west-flowing trade winds push the equatorial waters of the Pacific Basin
westward until the waters hit the Asian landmass. These warm waters remain in the western Pacific as
long as the trade winds flow westward. Occasionally the trade winds cease, resulting in an expansion of
the equatorial convergence zone associated with light winds. The trade winds can even reverse, flowing
from west to east and bathing both the continental margin of Peru in particular and the western margins
of the Americas in general. These warm waters terminate the upwelling of the Humboldt Current’s cold
waters, leading to a massive kill-off of marine organisms and precipitation over the coastal desert.
What, then, are the inputs into creating the El Niño phenomenon? The two major inputs should be
the global wind patterns, especially the equatorial, and a continental-insular rampart against which these
wind-driven waters can accumulate. The first is the result of the main surficial heat source, which is the
sun impacting a spherical planet with heat migrating from the equator to the two poles, and then being

759
Complex Petroleum Systems Developed by Subduction Process, Offshore Talara Basin, Northwest Peru

modified by the fragmentation of Pangea and the imposition of a north-south global oceanic pattern
(Scotese, 1997; Smith et al., 1981, 1994).
The second input would be the configuration of Austral-Asia to serve as a catchment for the west-
ward-driven equatorial waters. A review of plate tectonic reconstructions (Smith et al., 1981, 1994;
Scotese, 1997) provides an answer. Reconstructions show the same general pattern as that of today back
to the Early Miocene (Burdigalian-Aquilan-20 Ma) (Smith et al., 1981; 1994). By then, the South China
Sea had come into being and the Indonesian archipelago was beginning (Daly et al., 1991). Here, the
subduction process had created volcanic arcs that, with time, coalesced into ever-larger emergent areas.
The forward wedge of the northward-migrating complex of New Guinea was approaching the
emerging Indonesian Island arc in mid-Tertiary times (Smith et al., 1981, 1994; Daly et al., 1991—see
Fig. 2). When the closure became complete is not certain, but it probably occurred between 30 to 20 Ma
(Smith et al., 1994, Figs. 4 and 5). Thus, as interpreted here, by 20 Ma a relatively continuous rampart
of the Asian landmass, the Indonesian archipelago/New Guinea islands complex, and the Australian
landmass marks the western limit of the Pacific Basin. Consequently, by the Early Miocene there was a
wall against which westward-migrating equatorial currents could be lodged and the El Niño phenome-
non as it is known at present could have begun. This does not mean that an evolving and possibly
weaker El Niño did not become active in the Paleogene.
The working hypothesis employed here is that an El Niño mechanism has been active for the past 20
to 25 Ma (Casey et al., 1988) starting as a weak but ever-strengthening El Niño in the Paleocene. The
major impact of the phenomenon is in providing reservoir-grade sediments to the continental margin to
be transported downslope by mass wasting. Evidence strong erosive activity from vigorous mass wast-
ing, including turbidity currents, consists of discrete concave reflectors (interpreted as channel bases)
underlying regionally continuous seismic horizons. Numerous individual concave reflectors are up to 10
km across. These sizable rivers, whether subaerial or submarine, would represent massive outflows
caused by the El Niño process.

Subduction Processes Along the Peruvian Margin

A review of plate tectonic reconstructions (Smith et al., 1981, 1994; Scotese, 1997) reveals that dur-
ing Early Carboniferous time, 356 Ma, if not earlier, the western margin of South America was the
eastern edge of the Panthalassic Ocean, the direct ancestor of the Pacific Basin. Subduction may have
been active by the Early Carboniferous, certainly by Late Carboniferous (306 Ma) and Late Permian
(255 Ma) times (Scotese, 1997). Thus, subduction at some intensity has characterized the major geo-
logic process of the Peruvian continental margin since Late Paleozoic times.
The contemporary Peruvian coastline shows a long and narrow seaward concavity (this slim embay-
ment is approximately 50 km wide) from circa 6°S. to 14°S.). Cretaceous-age granites and diorites
along the coast around 15°S do not reappear until north of 6°S (Mapa Geológico del Perú, 1975). These
upper Mesozoic granitic rocks characterize a low-lying coastal range. They have formed in the magma
chambers and/or throats of the then-volcanic arc (Pichler and Zeil, 1969); the Sechura-Salaverry, Hua-
cho, and Pisco basins along the continental shelf are separated from the Pimentel, Trujillo, Lima, and
Paracas basins, which underlie the present-day continental slope—see Fig. 1).
The emergent basins north of 6°S (the northern—totally onshore—portion of the Sechura-Salavery
basin and the partially onshore and offshore Talara basin, and the Tumbes-Progreso basin extending into
Ecuador to the north) are separated by uplifted Paleozoic metamorphics and Mesozoic granites (Shep-
herd and Moberly, 1981, their Fig. 3). South of 15°S, the Moguega basin lies totally onshore, landward
of the coastal range (“coastal cordillera” in Shepherd and Moberly, 1981), which is composed of Creta-
ceous granites and diorites (Mapa Geológico del Perú, 1975). Seaward of the coast range, the Mollendo
basin extends across the narrow coastal plain and onto the continental shelf and upper slope. Subsequent
interpretation of offshore seismic data (Zúñiga-Rivero et al., 1998a, 1998b, 1998c, 1998d, 1999) shows
that the coastal range lies under the outer shelf-upper slope and separates the continental shelf and slope
basins between 6°S and 15°S (Fig. 1).

760
Zúñiga-Rivero et al.

Figure 4. Present subduction of the Nazca Ridge and its proposed southward migration from the upper Miocene to the
present, the last 9 Ma: present ridge position indicated by dark patterns and former positions farther back in time, shown
with successively lighter patterns, and with a mirror image of the Tuamotu Ridge, French Polynesia, circa 20°S, 140°W.
The Tuamotu Ridge was created at the same site along the East Pacific Rise, as the Nazca Ridge. Inserts show the
paleobathymetric records, derived from paleontology, from Ocean Drilling Project (ODP) cores along the Peruvian
continental margin plotted against time. Adapted from von Huene et al. (1996), their Figure 2.

At present, the aseismic Nazca Ridge enters the Peru Trench between 16°S and 15°S (Mapa
Geológico del Perú, 1975; Vogt et al., 1978). The Nazca Ridge subduction apparently has caused a
reduction in seismicity over a broader area than the actual zone of present-day ridge subduction. Spe-
cific earthquake patterns that change in this same region include the southern limit of deep (between
650 and 500 km) Peruvian earthquakes and the northern limit of the deeper intermediate Chilean earth-
quakes (those earthquakes between 200 and km deep; Ocola, 1966).
Following the tectonic segmentation of the Andes (Sillitoe, 1974), the present Nazca Ridge intersec-
tion with the Peru Trench marks the northern limit of the Quaternary volcanic characteristics of northern

761
Complex Petroleum Systems Developed by Subduction Process, Offshore Talara Basin, Northwest Peru
Figure 5. Geotectonic features across a plate convergence boundary: stipples indicate undeformed sediments; subduction complex is synonymous with accretionary prism.
Adapted from Dickenson (1995), his Figure 6.1.
762
Zúñiga-Rivero et al.

Chile–southernmost Peru (Lowrie and Hey, 1981). Farther northward, Quaternary volcanics recom-
mence at 3°S in southern Ecuador (Shepherd and Moberly, 1981). The northern limit of the longitudi-
nal valley west of the volcanic arc and west of the flat-floored, elevated Altiplano also occurs in the
same region. A trans-Andean discontinuity and deflection, marking geologic changes, the Pisco discon-
tinuity, also exists in this region (Vogt et al., 1978; Shepherd and Moberly, 1981). Vogt et al. (1978, p.
49) suggest that the “above coincidences may be fortuitous, but it is hard to believe they all are. The
Nazca Ridge subduction may be responsible for some of these effects.”
Nur and Ben-Avraham (1981) and von Huene et al. (1996) have employed an exciting idea about the
subduction of the Nazca Ridge; both assume that much of the Nazca Ridge has already been underthrust
beneath the South American continental plate. How large the Nazca Ridge was prior to subduction is
conjecture. Nur and Ben-Avraham (1981) suggest that what remains of the Nazca Ridge could have
originally been a feature analogous to the Mascarene Plateau, underlying the Seychelles Islands, west-
ern equatorial Indian Ocean (Fig. 3). The plateau consists, at least in part, of a continental fragment. The
density thus would range from 3.3–3.4 (that of oceanic basalt) to 2.6–2.8 for continental crust. The
facility for subduction would vary greatly, ranging up to 20% (2.7/3.3).
In 1996, von Huene et al. suggest that an analog for the Nazca Ridge would be the Tuamotu Ridge
of French Polynesia, near 20°S, 140°W. The hypothesis is based on the ridges common origin at a melt-
ing anomaly on the [East Pacific] spreading mid-ocean ridge (p. 20; see Pilger and Handschumacher,
1981, their Fig. 2). The Tuamotu Ridge is now topped by less dense coral atolls, making the overall
ridge less dense (more buoyant), hence less easy to subduct.
As they portray their suggestion, the paleo-Nazca Ridge originally extended as far north as 6°N, the
area of Punta Negra and the islets of Lobos de Tierra and Lobos de Afuera. A backtracking calculation
was conducted using present plate convergence rates from De Mets et al. (1990) “and the anomaly 3
pole from Conde, 1985, to revise the ridge crest subduction history along Peru” (p. 20—see Fig. 4).
The seismic data used by von Huene et al. (1996) include two industry lines collected by Shell Oil
Company (The Hague) in 1972: #1017 and #1018, as well as academic-quality seismic data associated
with the Ocean Drilling Project (ODP Log 112). Those data do not incorporate the industry seismic grid
from 1993, used here. The investigation of Von Huene et al suggests that the paleo-Nazca Ridge sub-
duction began south of 6°S, and caused an erosion of prior accretionary wedge sediments as the ridge
was underthrust beneath the Peruvian continental margin. Once part of the ridge had been absorbed by
subduction, a new phase of accumulating accretionary wedge sediments began and continues to the
present. Evidence of vertical movement along the margin includes studies of benthic foraminifera “as
the subducting ridge migrated south relative to any individual site causing water depths to increase
across benthic foraminiferal depth zones” (von Huene et al., 1996, p. 21, data from Resig, 1990).
A review shows that Resig’s (1990) data are also compatible with the notion that shallow sediments
were mass-wasted downslope to their present position. On the 9°S transect, at ODP Site 683, there has
been a 1500-m subsidence for middle Eocene sediments, which marks the commencement of a 26 Ma
hiatus. If the Nazca Ridge began subducting some 15 Ma ago, then much older sediments have been tec-
tonically eroded. There are raised terraces along the coastal plain in southern Peru near 15°-16°S.,
opposite the approaching and compressing Nazca Ridge, indicating “current uplift from this” subduc-
tion process (von Huene et al., 1996, p. 21; data from Hsü, 1988).
Hydrocarbon generation could take place when the subducted material has been buried to a depth of
some 3.5-4 km, assuming (1) a thermal gradient of 2.4°C. per 100m; (2) that the continental slope rises
above the subduction zone; and (3) that there has been no heat generated by friction—surely an unreal-
istic assumption.

Oceanographic Conditions Conducive to Carbon Accumulation and Preservation

The principal source of organic material along the Peruvian continental margin is the massive
upwelling associated with the cold northward-flowing Humboldt Current. As Antarctica became sta-
tionary over the South Pole while its connected crustal appendages migrated northward, the “northward
flight of continents” (R. W. Fairbridge3, personal communication, 3/2000) and there was regional cool-
ing, the clockwise circum-Antarctic current commenced and then strengthened. The southernmost part

763
Complex Petroleum Systems Developed by Subduction Process, Offshore Talara Basin, Northwest Peru

of South America deflects a significant portion of the current’s waters, sending them northward along
the Pacific margin of Chile and Peru. The upwelling ensures that the waters of the Humboldt Current
are biologically very productive and accounts for the abundant TOC that settles through the water mass
and is deposited along the Peruvian continental margin.
The Humboldt Current began to flow during the Eocene and has been active throughout the Neo-
gene. The relatively steady supply of organically rich deposits is hindered during El Niño periods,
when the supply of reservoir-grade sediments increases, while the amount of organic material decreases
because the upwelling is interrupted.
The 13C isotopic record shows that between 17.5 and 13.5 Ma, more carbon was preserved in the
sediments; a “temporary shift of about 1% to more positive values” (Fulthorpe and Schlanger, 1989, cit-
ing data in Vincent and Berger, 1985; Berger and Vincent, 1986; Vincent et al., 1985; Shackleton,
1987). There were similar 13C events in the early Miocene (Shackleton, 1987).
That these Miocene changes are measured in both planktonic and benthic foraminiferal records
“means that the entire oceanic carbon reservoir was affected” (Fulthorpe and Schlanger, 1989, p. 733)
and that the entire world oceanographic volume was included in this change. Slower oceanographic cur-
rents led to reduced aeration of intermediate waters, thus enlarging the oxygen-minimum zone and
accelerating organic carbon burial and preservation worldwide (Berger and Vincent, 1986). The global
increase in Miocene phosphoric deposits, from the silled, semi-enclosed basins around the north Pacific
and deposition off the southeastern United States does require a worldwide mechanism (Vincent and
Berger, 1985).

Forearc Basin, Accretionary Wedges and Critical Cohesive Coulomb Wedges

A forearc basin lies between the axes of submarine trenches, marking subduction zones, and the par-
allel magmatic-volcanic arcs, the loci of igneous activity initiated by descent of oceanic lithosphere into
the underlying mantle (Dickinson, 1995, p. 221—see Fig. 5). The term “forearc” reflects the convention
that arc-trench systems, where inherent polarity is defined by the asymmetry of subduction, are said to
“ ‘face’ from arc toward trench” (p. 221).
The term arc-trench system describes the entire plate convergence (Dewey, 1980); sometimes it is
mislabeled a “subduction zone,” whereas that term applies only to the actual underthrusting of oceanic
crust. Dickinson (1973) notes that as subduction proceeds along a converging margin, the spacing
between the trench and “igneous” arc grows in width and igneous activity has a periodicity of some
10 Ma. of alternating magmatic activity and quiescence. Northern Peru is not characterized by contem-
porary volcanism, yet is marked by uplifted mountains (the Andes) composed of Tertiary igneous and
metamorphic rocks (Mapa Geológico del Perú, 1975). These are granitoid batholiths and their wall
rocks, exposed as erosion removed surficial volcanic cover such as ash, andesites, and ignimbrites
(andesites are chemically analogous to the underlying diorites and granites: Hamilton, 1969; Pichler and
Zeil, 1969).
In Dickinson’s terminology, the subduction complex is composed primarily of those deep ocean sed-
iments scraped off from the underthrusting oceanic crust by the overlying active continental margin
with the forearc basin landward and distinct from the subduction complex. A synonym is “accretionary
wedge.”
A mechanism to describe the evolution of a subduction complex/accretionary wedge has been pro-
posed by Davis et al. (1983) and Dahlen et al. (1984). It has been amplified by many, including Lalle-
mand et al. (1994), and based on observations by Chapple (1978), which emphasized commonality of
characteristics of much varying accretionary wedges and fold-and-thrust belts. These commonalities are
listed:
1. “A basal surface of detachment or decollement, below which there is little deformation and dipping
toward the interior of the [adjacent] mountain belt;
2. A large horizontal compression in the material above the decollement;

3. Dept. of Geology, Columbia University (ret.)

764
Zúñiga-Rivero et al.

3. A characteristic wedge shape of the deformed material, tapering toward the margin of the mountain
belt (Davis et al., 1983, p. 1153).”
A simple mechanical model, applicable to these deformation belts, is considered to be analogous to
that of a wedge of soil or snow in front of an advancing bulldozer (Fig. 6). The material within the
wedge deforms until a critical taper is attained, after which it slides stably, continuing to grow at con-
stant taper as additional material is encountered at the toe. The critical taper is the shape for which the
wedge is on the verge of failure under horizontal compression everywhere, including the basal decolle-
ment. A wedge of less than critical taper will not slide when pushed but will deform internally,
steepening its surface slope until the critical taper is attained” (Davis et al., 1983, p. 1153). A critically
tapered wedge that is not accepting fresh material is the thinnest wedge that can be thrust over its basal
decollement without internal deformation. A critically tapered wedge that is accreting fresh material
deforms internally while sliding in order to accommodate the influx and to maintain its critical taper.
At present, the east-facing flank of the East Pacific Rise is covered by approximately 100m of
pelagic sediment that gradual thickens to 150–200 m on approaching the Peru Trench. An exception is
the thinner cover over the Neogene (now extinct) Galapagos Rise. A question to be asked concerns the
possible variations in sediment-cover thickness and type of oceanic basement during the Tertiary and a
possible maximum subduction of as much as 6,500 km. A reasonable assumption about these sediments
is that they have been fine-grained silts (largely from biological tests) and clays (marine “snow”) and
overall are structurally weak, with large local variations in structural strength (R. Faas, personal com-
munication, spring 1999).
The cohesive critical Mohr-Coulomb wedge model (e.g., Davis et al., 1983; Dahlen et al., 1984) is
pressure-dependent, time-independent behavior by brittle fracture (Paterson, 1978) or frictional sliding
(Byerlee, 1978). These deformation mechanisms are similar save for cohesion. Strength of cohesion
within siliciclastic sediments (5 to 20 mPa; Hoshino, 1972) is up to three orders of magnitude weaker
than that of granites and metamorphic rocks. These published cohesive values for sediments probably
may be regarded as overestimates because the measurements are commonly done on minute and pristine
samples, and rock strength decreases as the samples become larger and more heterogeneous (Davis et
al., 1983).

Figure 6. Schematic diagram of a cohesive critical Mohr-Coulomb wedge: sedimentary material is compressed and on the
verge of Coulomb failure. Simplified from Davis et al. (1983), their Figure 5.

765
Complex Petroleum Systems Developed by Subduction Process, Offshore Talara Basin, Northwest Peru

The maintenance of a critical wedge along a subduction zone depends on the amount and type of
sediments brought into the region, both by accretion along the trench and gravity-driven, mass wasting
downslope; the rate of plate convergence and the roughness of the subducted oceanic crust; and the
amount of fluid within the accreted sediments. Basal friction along the decollement is influenced by
pore pressure, the strength of the sediments subducted, and vertical relief along the underthrusting plate.
As overpressured pore fluid is incorporated by underthrusting into the subducted material, the
charged fluids will migrate into the accretionary wedge and cause hydrofracturing in the overriding
plate. Given rapid rates of underthrusting (around 8 cm/yr along the Peruvian margin), sediments are
buried—underthrust faster than they can dewater, and thus they cause overpressuring in pore fluids
(Lallemand et al., 1994). The potential for overpressuring will be highest in regions characterized by
thin trench sediment fill and fast convergence rates (Lallemand et al., 1994), both characteristic of Peru.
Variations in grain size due to different sediment sources (fine-grained from pelagic sources and
coarser grained from continent-derived mass-wasted sediments—most of the latter probably deposited
during sea-level lowstands) can affect the dynamics within the accretionary wedge and the basal decol-
lement. Should there be a decrease in grain size, pore pressure decreases, leading to an overall drop in
basal friction; the critical taper of the Coulomb wedge will decrease and therefore promote internal
deformation throughout the wedge. If there is an increase in grain size, a pore pressure increase, and an
overall increase in basal friction, the wedge’s critical taper will increase, and that will also promote
internal deformation throughout the wedge (Davis et al., 1983, p. 1169).
In the former instance with wedge thinning, overall compressive forces may be operating, promot-
ing intra-wedge thickening and erosion along the wedge base. The converse should occur in the latter
instance of a wedge thickening; overall extensional forces may be operating with increased normal
faulting and uplift caused by underthrusting along the wedge base.
An important question is, what periodicity of sedimentary grain-size variations could affect intra-
wedge dynamics? A classic third-order sea level oscillation lasts approximately a million years (Haq et
al., 1987). At an underthrusting rate of 10 cm/yr, 100 km of subduction/underthrusting zone could be
affected by a sediment change in the subducted material. If this proposition is true, then the entire
wedge of the offshore Talara basin could experience appreciable change in vertical thickness.
A remaining caveat to be discussed concerns the possible implications of massive down-slope
slumping and the concomitant reducing and increasing of overburden stress and their relationships to
the maintenance of a sedimentary wedge. Interpretation of both strike and dip lines along the continen-
tal margin of the Talara basin suggests that discrete slump blocks, as deep as 2–2.5 sec (2–2.5 km thick)
and as broad as 75–150 km on seismic strike line 93-20, may have moved onto the wedge over a short
geologic time span of perhaps some millennia. With such massive blocks of material potentially migrat-
ing downslope there must be dynamic implications for the stability of the sedimentary wedge.

Depositional and Tectonic Summary

In this review and interpretation, six novel observations are offered:


1. The El Niño process apparently has been active since the early Miocene and may have began during
the Oligocene.
2. The heavy rains that characterize the El Niño process have been a source of reservoir-grade sedi-
ments for the Neogene and perhaps the Early Paleogene.
3. Upwelling associated with the Humboldt Current throughout the Neogene—and probably with
increasing strength ever since the Eocene—has been a principal source of organic carbon deposited
along the Peruvian continental margin. A second source of organic material has been the pelagic
sediment carried into the Peruvian subduction zone by the Nazca oceanic plate.
4. Global oceanographic conditions associated with increased oxygen minimums have increased the
potential for carbon preservation in Miocene marine sediments and thus improve hydrocarbon
sources.
5. A cohesive critical Mohr-Coulomb wedge may describe as a single unit the mechanics of sediments
accumulated along the Peruvian margin.

766
Zúñiga-Rivero et al.

6. Two geological caveats are suggested for application of the wedge model to the Peruvian margin:
• Changes in deposition patterns from finer to coarser sediments during the sea-level oscillations
as quantified by the Vail (Exxon) curve may change internal wedge and wedge-base tectonics,
modifying the entire wedge.
• As interpreted from commercial seismic data, the continental margin off the Talara coast have
undergone massive slumping during the Neogene. Individual slump blocks are as thick as 2.5
km and as wide as 150 km. Such huge slumps must have changed the dynamics of the entire sed-
imentary wedge.
Thus, the entire active margin of northwest Peru and offshore in the prolific Talara basin appears to be
more active than previously envisioned. Such increased activity affects the petroleum systems of the region.

Petroleum Systems of the Offshore Talara Basin, NW Peru

The unique oceanographic-geologic conditions created by the El Niño phenomenon and the
upwelling of the Humboldt Current have been present since at least the Early Miocene. The El Niño at
3–7 year intervals causes massive erosion along the coastal plain, yielding reservoir-grade sediments to
be incorporated into the sediments of the continental margin and trench floor. During inter-El Niño
periods, the near-surface, northward-flowing Humboldt Current causes upwelling; the newly raised
deeper waters provide nutrients for a zone of biological productivity.
These organic materials are incorporated into sediments from three sources: subducted deep-ocean
sediments; continent-derived mass-wasted debris and turbidity currents descending across the shelf and
slope; and fine-grained autogenic detritus from the water mass itself, especially the Humboldt Current.
Both the organic matter and the sediments from their various sources must be incorporated into the
continental margin and transported to depth through the two main mechanisms: normal deposition as
the area subsides, and subduction.
Active plate margins are characterized by low geothermal gradients, such as 2.4°C/100 m or 96°C at a
depth of 4 km. At circa 100°C, hydrocarbon maturation commences (Tissot and Welte, 1984). Normal
subsidence and burial incorporate the organic matter and sediments deposited along the shelf. Review of
seismic dip line 93-16 (Fig. 7) reveals that relatively undeformed shelf sediments extend to depths of at
least to 6–8 km beneath the upper slope (Waples, 1980). Hydrocarbons generated in this environment
would ascend along migration routes until they reached reservoirs with an effective trap and seal.

Figure 7. Line drawing of dip seismic reflection record: extending from center of Sechura Bay (5°37'S, 81°20'W) across the
Peru Trench (5°35'S, 82°06'W): offshore Talara Basin sequences, extending from Quaternary through Paleozoic near the
eastern edge of the record to the accretionary prism to the west, reaching to the landward trench wall with east-to-west
thrusting within the middle slope. Modified from von Huene et al. (1989), their Figure 2.

767
Complex Petroleum Systems Developed by Subduction Process, Offshore Talara Basin, Northwest Peru

Hydrocarbon Traps, Offshore Talara Basin

Within the outer shelf and slope provinces of the offshore Talara basin, the Neogene and Upper
Paleogene seismic sequence is characterized by distinct, continuous reflections. This good reflector
quality is appropriate for detailed interpretation. Such interpretations reveal hydrocarbon traps
described here from offshore. Delta-type sequences descend eastward, and western onlap terminations
of Talara basin-fill sediments onlap onto the deltas. Abundant reservoir-grade sediments are known to
exist in the Talara basin and logically could extend into this yet undrilled sector.
Farther west, some 10%-15% of all reflectors show evidence of extensive fluvial erosion, as dis-
cussed in the earlier section Possible Geologic Duration of the El Niño Phenomenon. The fluvial
sediments and intersecting eastward-dipping faults created reservoirs and traps
Along the outer shelf, and separate from the east-facing limit of the Talara basin, are a series of shelf
deltas—some mid shelf and others outer shelf, of probable Miocene age. Sequence stratigraphy sug-
gests that these deltas are caused by sea-level oscillations; thus, fine-grained transgressive units cap
each coarse-grained delta. Many such deltas are cut by normal faults that dip seaward. Thus, these shelf
deltas, especially those along the outer shelf, are petroleum traps that are partly stratigraphic and partly
structural.
The present shelf break and various paleo-shelf breaks are characterized by numerous and pro-
nounced growth faults that dip toward the ocean basin. These growth faults often have dramatic offsets
and also cut paleo-shelf break deltas and separate blocks of heterogeneous sediments.
The presumably Mesozoic- and Paleozoic-age sediments provide reflector images at depths of 4-6
seconds that are suggestive of deltas dipping eastward toward the onshore part of the Talara basin, and
the Mesozoic continental margin that faces to the west appears cut by paleo-shelf break growth faults.
These extensive faults can serve as migration routes and as creators of fault traps. Data quality is an
important issue here.

Conclusions

This paper relates the geologic processes that strongly affect hydrocarbon generation and preserva-
tion—i.e., the maintenance of a petroleum system—along the continental margin offshore within the
Talara basin of northwest Peru. Two sources of organic carbon have been identified.
Probably since the Paleozoic and definitely since the Mesozoic, underthrusting beneath the South
American plate has been the dominant tectonic process, modified by the subduction of aseismic ridges
in the Late Neogene. A key factor is the time of origin of the El Niño phenomenon in geologic history,
which is not yet known. Using first principles, a crude duration estimate has been derived in which El
Niño has affected both source and reservoir generation at least throughout the Neogene. Organic preser-
vation is evidently enhanced through oceanographic changes during the Miocene.
Sediments accreted along a subduction zone may be mechanically modeled as a single unit,
although variations in internal and external deformation may markedly complicate the details of the
petroleum system involved. A preliminary conclusion is that such preliminary modeling, even with only
a single unit, suggests possibilities of wedge-deforming tectonics and increasing opportunities for
hydrocarbon generation, fluid migration, and oil and gas entrapment.
Two petroleum systems span the continental margin: one is associated with regional subsidence and
compaction and geothermal temperatures rising with depth under the shelf, the other with the subduc-
tion process and sediments both accreted against the slope and underthrust beneath the margin.
Hydrocarbon generation commences at approximately 100°C. Forearc basins within an arc-trench sys-
tem are geothermally cool, and gradients are around 2.4°C/100 m or 96°C at a depth of 4 km. The shelf
and paleo-shelves of the Talara basin offshore region reach depths of 6-7 sec two-way travel time
(depths of 6-7 km). Thus, the available industry seismic data may be imaging 2-3 km into the petroleum
“kitchen.”
Within the subduction process itself, the matter of geothermal gradients is much more complex. On
being subducted, sediments move almost laterally, being dewatered, increasing fluid pressure and tem-
peratures and gaining heat from two sources: (1) the vertical geothermal gradient through the

768
Zúñiga-Rivero et al.

accretionary wedge and through the oceanic lithosphere, and (2) the lateral gradient along the subduc-
tion path itself, with fluids both rising and descending.
Accretionary wedge deformation provides local compression and extension and migration routes,
components essential to any petroleum system. Thus, the intricate and complicated active continental
margin geology and petroleum systems of the outer offshore Talara basin appear to be promising for
petroleum exploration.

References

Berger, W. H., and E. Vincent, 1986, Deep-sea carbonates: reading the carbon-isotopic signal: Geologische Rund-
schen, v. 75, p. 249-269.
Casey, R. E. et al., 1988, El Niño-like events during Miocene: AAPG Bull, v. 72, no. 3, p. 377.
Conde, S. C., 1985, Nazca-South American plate interactions since 50 my B.P., in D. M. Hussong et al., eds., Atlas
of the Ocean Margin Program, Peru Continental Margin Region VI: Marine Science International, Woods Hole,
Mass., 14 p.
Dahlen, D., J. Suppe, and D. Davis, 1984, Mechanics of fold-and-thrust belts and accretionary wedges: cohesive
Columb theory: Jour. of Geophysical Research, v. 89, no. B12, p. 10,089-10,101.
Daly, M. C., M. A. Cooper, I. Wilson, D. G. Smith, and B. G. D. Hooper, 1991, Cenozoic plate tectonics and basin
evolution in Indonesia: Marine and Petroleum Geology, v. 8, no. 1, p. 2-21.
Davis, D., J. Suppe, and F. A. Dahlen, 1983, Mechanics of fold-and-thrust belts and accretionary wedges: Jour. of
Geophysical Research, v. 88, no. B2, p. 1153-1172.
De Mets, C., R. G. Gordon, D. F. Argus, and S. Steins, 1990, Current plate motions: Geophysical Jour. Int., v. 101,
p. 425-478.
Dewey, J. F., 1980, Episodicity, sequence, and style at convergent plate boundaries: Geological Association of Can-
ada, Special Paper 20, p. 553-573.
Dickinson, W. R., 1973, Widths of modern arc-trench gaps proportional to past duration of igneous activity in asso-
ciated magmatic arcs: Jour. of Geophysical Research, v. 78, p. 3376-3389.
Dickinson, W. R., 1995, Forearc basins, in C. J. Busby and R. V. Ingersoll, eds., Tectonics in sedimentary basins:
Blackwell Science, Inc., p. 221-261.
Dow, W. G., 1974, Application of oil correlation and source rock data to exploration in Williston basin, AAPG
Bull., v. 58, no. 7, p. 1253-1262.
Fulthorpe, C. S., and S. O. Schlanger, 1989, Paleo-oceanographic and tectonic settings of Early Miocene reefs and
associated carbonates of offshore Southeast Asia: AAPG Bull., v. 73, no. 6, p. 729-756.
Hamilton, W., 1969, The volcanic central Andes—a modern model for the Cretaceous batholiths and tectonics of
western North America, in A. R. McBirney, ed., Proceedings of the Andesite Conference, Oregon, July 1968,
International Upper Mantle Project, Science Report Number 16: Oregon, Department of Mineral Industries,
Bull. No. 65, p. 175-184.
Haq, B. U., J. Hardenbal, and P. R. Vail, 1987, Chronology of fluctuating sea levels since the Triassic: Science, v.
235, no. 4794, p. 1156-1167.
Hsu, J. T., 1988, Emerged quaternary marine terraces in southern Peru: Sea level changes and continental margin
tectonics over the subducting Nazca Ridge: Ph.D. dissertation, 309 p., Cornell University, Ithaca, New York.
Lallemand, S. E., P. Schnurle, and J. Malavieille, 1994, Coulomb theory applied to accretionary and nonaccretion-
ary wedges: possible causes for tectonic erosion and/or frontal accretion: Jour. of Geophysical Research., v. 99,
no. B6, p. 12,033-12,055.
Lowrie, A., and R. Hey, 1981, Geological and geophysical variations along the western margin of Chile near lat 33°
to 36°S and their relation to Nazca plate subduction, in L. D. Kulm, J. Dymond, E. J. Dasch, and D. M. Hus-
song, eds., Nazca Plate: Crustal Formation and Andean Convergence: Geological Society of America Memoir
154, p. 741-754.
Magoon, L. B., and W. G. Dow, eds., 1994a, The Petroleum System—From Source to Trap: AAPG Memoir 60, 655 p.
Magoon, L. B., and W. G. Dow, 1994b, The Petroleum System, in L. B. Magoon and W. G. Dow, eds., The Petro-
leum System—From Source to Trap: AAPG Memoir 60, p. 3-24.
Mapa Geologico del Peru, 1975, Instituto de Geologia y Mineria, Ministerio de Energia y Minos, Republica del
Peru, scale 1:1,000,000, Mercator Projection, 4 sheets.
Nur, A., and Z. Ben-Avraham, 1981, Volcanic gaps and the consumption of aseismic ridges in South America, in L.
D. Kulm, J. Dymond, E. J. Dasch, and D. M. Hussong, eds., Nazca Plate: Crustal Formation and Andean Con-
vergence: Geological Society of America Memoir 154, p. 729-740.
Ocola, L., 1966, Earthquake activity of Peru, in J. S. Sleinhart and T. J. Smith, eds., Geophysical Monograph #10:
American Geophysical Union, p. 509-528.

769
Complex Petroleum Systems Developed by Subduction Process, Offshore Talara Basin, Northwest Peru

Pichler, G., and W. Zeil, 1969, Andesites of the Chilean Andes, in A. R. McBirney, ed., Proceedings of the Andesite
Conference, Oregon, 1968, International Upper Mantle Project, Science Report Number 16: Oregon, Depart-
ment of Mineral Industries, Bulletin No. 65, p. 165-174.
Pilger, R. H., and D. W. Handschumacher, 1981, The fixed hotspot hypothesis and the origin of the Easter-Sala y
Gomez-Nazca trace: Geological Society of America Bull, v. 92, no. 7, p. 437-446.
Quinn, W. H., V. T. Neal, and S. E. Antunez de Mayolo, 1987, El Niño occurrence over the past four and a half cen-
turies: Jour. of Geophysical Research, v. 92, no. C13, p. 14,449-14,462.
Resig, J., 1990, Benthic forminiferal stratigraphy and paleoenvironments off Peru, Leg 112: Proceedings, Ocean
Drilling Program Scientific Research, v. 112, p. 263–282.
Rodbell, D. T. et al., 1990, An approximately 15,000-year record of El Niño-driven alluviation in southwestern
Ecuador: Science, v. 283, no. 5401, p. 516-520.
Sandweiss, D. H. et al., 1996, Geoarchaeological evidence from Peru for a 5000 B.P. onset of El Niño: Science, v.
273, no. 5281, p. 1531-1533.
Scotese, C. R., 1997, Paleogeographic Atlas, PALEOMAP Project, Progress Report 90-0497: Department of Geol-
ogy, University of Texas at Arlington, Arlington, Texas, 45 p.
Shackleton, N. J., 1987, The carbon isotope record of the Cenozoic; history of organic carbon burial and of oxygen
in the ocean and atmosphere, marine petroleum source rocks: Geological Society Special Publication 26, p.
423-434.
Shepherd, G. L., and R. Moberly, 1981, Coastal structure of the continental margin, northwest Peru and southwest
Ecuador, in L. D. Kulm, J. Dymond, E. J. Dasch, and D. M. Hussong, eds., Nazca Plate: Crustal Formation and
Andean Convergence: Geological Society of America Memoir 154, p. 351-391.
Sillitoe, R. H., 1974, Tectonic segmentation of the Andes: Implications for magmatism and metallogeny: Nature, v.
250, p. 542-545.
Smith, A. G., A. M. Hurley, and J. C. Briden, 1981, Phanerozoic paleo-continental world maps: Cambridge Univer-
sity Press, Cambridge Earth Science Series, 102 p.
Smith, A. G., D. G. Smith, and B. M. Funnell, 1994, Atlas of Mesozoic and Cenozoic Coastlines: Cambridge Uni-
versity Press, 99 p.
Tissot, B.P., and D. H. Welte, 1984, Petroleum formation and occurrence, Springer-Verlag, 699 p.
Vincent, E., and W. H. Berger, 1985, Carbon dioxide and polar cooling in the Miocene—the Monterey hypothesis,
in E. T. Sunquist and W. S. Broeker, eds., The carbon cycle and atmosphere CO2-natural variations Archean to
present: American Geophysical Union Monograph 32, p. 455-468.
Vincent, E., J. S. Killinglely, and W. H. Burger, 1985, Miocene oxygen and carbon isotope stratigraphy of the trop-
ical Indian Ocean, in The Miocene ocean-paleoceanography and biogeography: Geological Society of America
Memoir 163, p. 103-130.
Von Huene, R., I. A. Pecher, and M.-A. Gutscher, 1996, Development of the accretionary prism along Peru and
material flux after subduction of Nazca Ridge: Tectonics, v. 16, no. 1, p. 19-33.
Waples, D. W., 1980 Time and temperature in petroleum formation: application of Lopatin’s method to petroleum
exploration: AAPG Bull., v. 64. p. 916-926.
Yeats, R. S., S. R. Hart, and others, 1976, Initial Reports of the Deep Sea Drilling Project: v. 34: Washington, D.C.,
U.S. Government Printing Office.
Zúñiga-Rivero, F., H. Hay-Roe, and T. Vargas, 1999, Potential untested under 50 million acres in Peru: Oil and Gas
Jour., 15 March, p. 67-72.
Zúñiga-Rivero, F., J. Keeling, and H. Hay-Roe, 1998a, Attractive potential seen in 10 sub-basins off Peru: Oil and
Gas Jour., 7 Sept., p. 117-122.
Zúñiga-Rvero, F., J. A. Keeling, and H. Hay-Roe, 1998b, Peru onshore-deepwater basins should have large poten-
tial: Oil and Gas Jour., 19 Oct., p. 88-95.
Zúñiga-Rivero, F., J. A. Keeling, and H. Hay-Roe, 1998c, Oil, gas potential in shallow water: Peru’s continental
shelf basins: Oil and Gas Jour., 16 Nov., p. 92-96.
Zúñiga-Rivero, F., J. A. Keeling, and H. Hay-Roe, 1998d, New hunting grounds studied on Peru’s continental
slope: Oil and Gas Jour., 7 Dec., p. 87-92.

770

You might also like