3D Mapping of Intruding Salt Bodies in The Subsurf-5-12

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

A.G. Mattson, M.R. Gani, T. Roesler et al.

Results in Geophysical Sciences 1–4 (2020) 100004

Fig. 3. Major structural and basement-related features in the eastern Gulf of Mexico. The study area is located at the easternmost part of the Central Louann Salt
Basin. The line showing the extent of Louann Salt is taken from Godo (Godo, 2017), and basement features are based on Tew et al. (Tew et al., 1991), Dobson and
Buffler (Dobson and Buffler, 1997), Hudec et al. (Hudec et al., 2013), and Godo (Godo, 2017). Note the location of seismic lines presented in Figs. 4, 5, 6 and 9.

fiable (Fig. 6A). These areas, which we classify as having ‘excellent section under the Horn and Farnella diapirs in Fig. 6C (see Fig. 1 for
seismic quality’, cluster around the southern portion of the study area location). Under parts of the Horn and Farnella diapirs, the signal-
where the basement resides at a shallower depth due to the influence of to-noise ratio decreases so significantly that the seismic section be-
the Middle Ground Arch (Fig. 3) and reduced isostatic effects from the comes almost transparent with only a faint reflector of the Upper Creta-
shelf. ceous/Paleogene boundary. Typically, the Upper Cretaceous/Paleogene
We classify areas where only the deepest stratigraphy in the Meso- boundary is one of the strongest reflections in the subsurface of north-
zoic section becomes more difficult to interpret and correlate as being eastern Gulf of Mexico.
‘acceptable seismic quality’ (Fig. 6B). In these areas, the shallower Meso-
zoic and Neogene intervals are comparatively easy to interpret and cor- 5. Methodology
relate (Fig. 6B). Seismic quality in these areas tends to degrade under
allochthonous salt overhangs, making it difficult to determine the ex- The 3D seismic volume was overlain with the original velocity model
act salt-sediment boundary; however, the general salt body geometry to identify salt bodies (Fig. 5). Petrel’s multi-z tool was used to map
can be deduced by iteratively picking on in-lines and cross-lines, and the Louann salt structures. Salt was first interpreted on the in-lines and
reconciling interpretation differences (i.e., correcting miss ties). then the cross-lines, both in 20 line increments. The initial salt model
‘Poor seismic quality’ is observed under the Horn, Petit Bois, Far- was then quality-controlled by running through the in-lines looking for
nella, and Mitchell salt bodies to the west. The zones of ‘poor seismic discrepancies (miss-ties). If interpretation discrepancies were found, the
quality’ are typically characterized by P-wave to S-wave conversions, model was iteratively reinterpreted until a potential valid solution was
often producing phantom horizons that can be mistaken for base of salt obtained for both the in-line and cross-line directions. The original ve-
(Fig. 6C). Within poor quality areas, the signal-to-noise ratio decreases locity model used halite acoustic velocities for salt bodies that were
dramatically in the Miocene and upper Mesozoic intervals. An example differentiated from the surrounding strata. However, in contrast to the
of P- to S-wave conversions can be seen around the base of Neogene original velocity model, we interpreted some Mesozoic intervals as salt-
A.G. Mattson, M.R. Gani, T. Roesler et al. Results in Geophysical Sciences 1–4 (2020) 100004

Fig. 4. Tying 3D seismic data to the Titan well in the study area. For location, see Fig. 3. Part of the biostratigraphic data of the Titan well are presented in Table 2.

bearing. Therefore, the original velocity model for the seismic volume decreases between the salt and surrounding strata as the sediment com-
cannot be traced directly using the multi-z tool. Nonetheless, our final paction and lithification increase with depth. Where possible, the deeper
interpretation of the salt bodies generally agrees with the original ve- allochthonous and autochthonous salt layers were mapped with the top
locity volume. of salt as a weak peak and the base as a weak trough. The sides of salt
For the shallower section, where a strong impedance contrast ex- bodies were mapped based on where the surrounding sedimentary re-
ists between the salt and surrounding sedimentary rocks, the top of flectors truncate against a relatively transparent (thus, salt) body. 3D
salt was picked as a hard peak, and the base of salt as a hard trough meshed volumes of salt bodies were then created from multi-z interpre-
(Jackson and Hudec, 2017). In the deeper portions, impedance contrast tations to visualize salt-body geometries (Fig. 7).
A.G. Mattson, M.R. Gani, T. Roesler et al. Results in Geophysical Sciences 1–4 (2020) 100004

Fig. 5. (A) A representative seismic section of the study area. For location, see Fig. 3. (B) The same seismic section overlain with the RTM (reverse-time migration)
velocity model. Intra-Neogene salt bodies show a strong impedance contrast between the salt and surrounding strata, while deeper Mesozoic salt bodies show little
to no impedance contrast to surrounding strata. (C) The same section with interpreted stratigraphic intervals and salt bodies based on biostratigraphic data from
nearby wells (e.g., Fig. 4 and Table 2).
A.G. Mattson, M.R. Gani, T. Roesler et al. Results in Geophysical Sciences 1–4 (2020) 100004

Fig. 6. A comparison of seismic data quality across the study area. For locations of the seismic lines, see Fig. 3. (A) Seismic quality is ‘excellent’ in the southern
portions of the study area where there are fewer allochthonous salt bodies and the basement is at a shallower depth. Pre-salt basement is readily identifiable in these
areas. (B) An example of ‘acceptable’ seismic quality. Note that the lower Mesozoic reflectors are less clear. (C) An example of ‘poor’ seismic quality, showing P- to
S-wave conversions and seismically transparent zone under the Horn and Farnella salt bodies. Typically, the areas with ‘poor’ seismic quality cluster around multiple
salt stocks and directly under allochthonous canopies near the original salt feeder.
A.G. Mattson, M.R. Gani, T. Roesler et al. Results in Geophysical Sciences 1–4 (2020) 100004

Fig. 7. 3D salt-body visualization across the study area on top of the pre-salt basement.

Fig. 8. 3D view of the salt bodies in the study area showing three different levels of salt: lowermost autochthonous layer (Middle Jurassic) forming salt anticlines and
pillows; an Upper Cretaceous/Paleogene layer forming localized sheets, canopies, and walls; and an intra-Neogene layer comprised of diapirs, sheets and canopies.

In addition to mapping all the salt bodies in 3D in the study area, comprised of salt pillows and anticlines; an Upper Cretaceous to Pa-
all of the observed normal faults and 14 chronostratigraphic horizons leogene interval with salt sheets, canopies, diapirs, and walls; and an
were mapped. Geological ages of these horizons were determined based intra-Neogene level containing diapirs, sheets and canopies. These three
on well biostratigraphic data obtained from the Bureau of Ocean En- salt-levels are connected together by a vertical to slightly inclined salt
ergy Management (e.g., Fig. 4, Table 2). These horizons were used to feeders, most of which are interpreted to be un-welded. Salt anticlines
determine the ages of the allochthonous salt features. are elongated salt mounds with concordant overburden (Jackson and
Hudec, 2017; DeGolyer, 1925), whereas salt walls have discordant over-
6. Application and discussion burden contacts (Jackson and Hudec, 2017; Trusheim, 1960). Generally,
salt walls transition to salt anticlines near the margins of the salt struc-
Using a novel approach described above, we generated the 3D salt ture in the study area. Full 3D salt-body visualization can significantly
geobodies in the Gulf of Mexico (Fig. 7). We were able to identify improve our understanding of salt geometries at various scales, which
three prominent salt levels: autochthonous Middle Jurassic salt-level we describe next.
A.G. Mattson, M.R. Gani, T. Roesler et al. Results in Geophysical Sciences 1–4 (2020) 100004

Fig. 9. Tertiary welding between two separate Upper Cretaceous/Paleogene canopies. (A) 3D salt geobody model showing two allochthonous salt sheets that appear
to parallel each other without any connection. (B) A cross section through these two salt sheets showing a relatively-uniform minibasin forming above the two
allochthonous bodies, suggesting that these two salt bodies were once together. Locations of salt welding are denoted by a pair of black dots. For location of the
seismic line, see Fig. 3.

6.1. Spatial distribution of salt Middle Ground Arch on the Louann Salt deposition, where less salt was
initially deposited in the southwest resulting in smaller salt structures.
By mapping the salt in 3D, we can understand the true spatial-
distribution of the salt bodies. Since most of these salt bodies have ver- 6.2. Allochthonous canopy geometry
tical to near-vertical feeders, we approximate current-day salt-structure
distribution as a proxy for initial salt-distribution. Most of the salt bod- One of the most prominent features of our 3D salt model is the
ies are located in the west-central portion of the study area. This agrees extensive layer of allochthonous salt at the Upper Cretaceous to Pa-
with the basement-geometry models proposed by Tew et al. (1991) and leogene interval (Fig. 8). While K-T (Cretaceous-Paleogene boundary)
Dobson and Buffler (1997) where the Apalachicola Basin begins transi- salt sheets and canopies were previously documented (Bouroullec et al.,
tioning into the Central Louann Salt Basin. Areas to the southeast show 2017; Bouroullec et al., 2017), there is no publication on the extent or
only smaller Mesozoic salt structures that failed to significantly pierce geometry of this layer. A possible reason for the lack of proper investiga-
the Neogene strata. We attribute this to the continued influence of the tion of this layer is the discreteness of the Upper Cretaceous salt bodies.
A.G. Mattson, M.R. Gani, T. Roesler et al. Results in Geophysical Sciences 1–4 (2020) 100004

Fig. 10. Geometries of Neogene salt feeders. While all the tops of the diapirs are tilted basinward, not all the salt feeders are flared in the same direction. Feeders
can tilt basinward (Mobile and Whiting Domes), landward (Horn Dome), or stay vertical (Kings Peak, VK909, and Pascagoula).

Moreover, because of the salt-body’s small size and frequent variation Fig. 8 shows the most likely area containing a tertiary welding where
in geometry, they are difficult to visualize properly using standard 2D two allochthonous salt sheets appear to parallel each other. It seems
cross-sections. unlikely that two independent salt sheets would flow parallel to each
By mapping the salt structures in 3D, most of these challenges were other without suturing (connecting). This interpretation is further sup-
overcome. Our 3D salt model (Fig. 8) shows a regionally discrete, but ported by the fact that a single, relatively-uniform minibasin formed
locally continuous, Upper Cretaceous salt layer with possible tertiary above the two separate salt sheets, indicating a uniform salt-withdrawal
welding occurring in certain areas (see next section). Additionally, the from the two salt sheets. These observations suggest that the two salt
number and size of salt structures within the Upper Cretaceous to Pale- sheets were initially one salt canopy where a minibasin formed on top
ogene layer increases eastward, possibly in relation to the proximity to of the allochthonous salt body and subsided within the allochthonous
the Cretaceous shelf-edge (Fig. 3). salt canopy before welding onto the lower Cretaceous strata.

6.3. Allochthonous salt-weld applications


6.4. Salt feeder geometry
Jackson and Hudec (2017) defined salt weld as “a surface or thin
zone marking a vanished salt”, with tertiary weld specifically referring Traditionally, horizon-mapping tools in seismic interpretation soft-
to welds occurring in gently-dipping allochthonous bodies. Understand- ware are not useful in mapping steeply-dipping (>70°) beds since these
ing salt welds is import since they represent paleo-salt locations and beds approach the limit of a single depth-value per spatial point. As salt
can form hydrocarbon traps and affect migration pathways. Incomplete feeders typically have dips around 90°, we need multi-z tool like the one
salt-welds with a residual salt layer can act as an impermeable barrier used in this study to map salt feeders. With the 3D mapping of salt bod-
forming a hydrocarbon trap (Pilcher et al., 2011), such as the Yucatan ies, salt feeder geometries became distinct in the study area (Fig. 10).
field in the western Gulf of Mexico (Weimer et al., 2016). In contrast, In our 3D salt model, all of the tops of the diapirs, sheets, and
complete salt-welds can be porous and permeable allowing for hydro- canopies show a basinward inclination (Fig. 10). However, not all the
carbons to migrate across the salt weld and into supra-weld reservoirs, salt feeders tilt in the same direction. Salt feeders for the Mobile and
as is the case for the Marco Polo field in the central Gulf of Mexico Whiting Domes are inclined basinward, while the salt feeders for the
(Mount et al., 2006). Pascagoula, Kings Peak, and VK909 Domes are nearly vertical. The Horn
Salt welds on seismic data are typically characterized by high- Dome feeder actually tilts landward, likely due to a nearby minibasin
amplitude events often associated with what appears to be angular un- deforming the feeder (Fig. 10).
conformities occurring at the weld surface, demarking the strata above Salt feeder geometry has important implications for three-way clo-
and below the now absent salt. Identifying potential salt welds within sures against salt feeders and base-of-salt truncation-traps for hydro-
Upper Cretaceous to Paleogene interval in the northeast Gulf of Mexico carbons (Pilcher et al., 2011). Three-way closures against salt feed-
is difficult, since the interval tends to contain strong seismic reflections ers are a common hydrocarbon trapping-mechanism in the central and
(Fig. 4). These strong reflections are likely caused by the reduced sed- western Gulf of Mexico, like in the Anchor, Triton, Heidelberg, K2/K2-
imentation rates during Upper Cretaceous through Paleogene, forming North and Genghis Khan fields (Mount et al., 2010), and in the Shenzi
highly-condensed sections. Furthermore, the Chicxulub impact caused and Tahiti fields (Pilcher et al., 2011). Furthermore, petroleum sys-
localized erosion within Upper Cretaceous to Paleogene sedimentary tem modeling by Mount et al. (2006) shows the potential for hydro-
hiatus, forming localized angular unconformities and thus making it dif- carbons to migrate along the salt-sediment interface of salt feeders and
ficult to distinguish between welded and non-welded areas. to charge salt-truncation traps where the salt intrudes and creates a ver-
By mapping the salt system in 3D, we were able to identify prob- tical conduit from the source-rock below to a potential reservoir-rock
able areas of salt welding within the Cretaceous-Paleogene canopy. above.
A.G. Mattson, M.R. Gani, T. Roesler et al. Results in Geophysical Sciences 1–4 (2020) 100004

7. Conclusions Gee, M.J.R., Gawthorpe, R.L., 2006. Submarine channels control by salt tectonics: exam-
ples from 3D seismic data offshore Angola. Mar. Pet. Geol. 23, 443–458.
Godo, T., 2017. The Appomattox field: Norphlet Aeolian sand dune reservoir in the deep-
As better seismic acquisition and processing techniques are being -water Gulf of Mexico. In: Merrill, R.K., Sternbach, C.A. (Eds.). In: Giant Fields of the
developed for allochthonous salt basins, better sub-salt 3D seismic data Decade 2000-2010, 115. AAPG Memoir, pp. 29–54.
will become available. These improved 3D seismic volumes will play a Hudec, M.R., Jackson, M.P.A., 2006. Advance of allochthonous salt sheets in passive mar-
gins and orogens. AAPG Bull. 90, 1535–1564.
critical role in future investigations of salt bodies. In this study, we show Hudec, M.R., Jackson, M.P.A., 2007. Terra infirma: Understanding salt tectonics. Earth
how true 3D mapping and modeling of subsurface salt bodies can be per- Sci. Rev. 82, 1–28.
formed in a basin. 3D salt models have applications in studies of salt vol- Hudec, M.R., Norton, I.O., 2019. Upper Jurassic structure and evolution of the Yucatan
and Campeche subbasins, southern Gulf of Mexico. AAPG Bull. 103, 113–1151.
ume estimation, allochthonous canopy geometries, allochthonous weld
Hudec, M.R., Norton, I.O., Jackson, M.P.A., Peel, F.J., 2013. Jurassic evolution of the Gulf
identification, and salt feeder geometries. Using 3D seismic dataset in of Mexico salt basin. AAPG Bull. 97, 1683–1710.
the northeastern Gulf of Mexico, we were able to define the 3D spa- Jackson, M.P.A., Hudec, M.R., 2017. Salt tectonics: Principles and Practice. Cambridge
University Press, Cambridge, United Kingdom.
tial distribution of salt structures, the geometry of K-T salt canopy, and
Jones, I., Davison, I., 2014. Seismic imaging in and around salt bodies. Interpretation 2-4,
show potential salt-feeder conduits for the Upper Cretaceous and Neo- SL1–SL20.
gene salt bodies. These findings have implications in hydrocarbon ex- Kramer, K.V., Shedd, W.W., 2017. A 1.4-billion-pixel map of the Gulf of Mexico seafloor.
ploration and salt tectonic interpretation of the basin. Similar studies of Eos 98. https://doi.org/10.1029/2017EO073557.
Mancini, E.A., Mink, R.M., Bearden, B.L., Wilkerson, R.P, 1985. Norphlet formation (Up-
3D salt modeling can be conducted in other parts of the Gulf of Mexico per Jurassic) of southwestern and offshore Alabama: environments of deposition and
and/or in other basins in the world. petroleum geology. AAPG Bull. 69, 884–898.
Mount, V.S., Rodriguez, A., Chaouche, A., Crews, S.G., Gamwell, P., Montoya, P.,
2006. Petroleum systems observations and interpretations in the vicinity of the
Declaration of Competing Interest K2/K2-North, Genghis Khan, and Marco Polo fields, Green Canyon, Gulf of Mexico.
Gulf Coast Assoc. Geol. Soc. Trans. 56, 613–625.
Mount, V.S., Mahon, K.I., Mentemeier, S.H., 2010. Structural restoration and basin mod-
None. eling in the north-central Gulf of Mexico deepwater subsalt plays. Gulf Coast Assoc.
Geol. Soc. Trans. 60, 503–510.
Pilcher, R.S., Kilsdonk, B., Trude, J., 2011. Primary basins and their boundaries in the
Acknowledgments deep-water northern Gulf of Mexico: Origin, trap types, and petroleum system impli-
cations. AAPG Bull. 95, 219–240.
Pilcher, R.S., Murphy, R.T., Ciosek, J.M., 2014. Jurassic raft tectonics in the northeast Gulf
We like to thank anonymous reviewers and the Editor C. Ozgen of Mexico. Interpretation 2, SM39–SM55.
Karacan for constructive reviews. This research was partially sup- Pindell, J., Kennan, L., 2001. Kinematic evolution of the Gulf of Mexico and Caribbean.
ported by the American Chemical Society – Petroleum Research Fund In: Fillon, R.H., Rosen, N.C., Weimer, P., Lowrie, A., Pettingill, H., Phair, R.L.,
Roberts, H.H., van Hoom, H.H. (Eds.). In: Petroleum Systems of Deep-Water Basins—
(PRF#545000-UN18) and Western Kentucky University (WKU) – RCAP- Global and Gulf of Mexico Experience, 21. Gulf Coast Section SEPM, pp. 193–220.
I grant (#18–8014) to N. Gani, and by the WKU Graduate School Re- Rojo, L.A., Escalona, A., 2018. Controls on minibasin infill in the Nordkapp basin: evidence
search Grant to A. Mattson. We thank IHS for Kingdom and Schlum- of complex Triassic synsedimentary deposition influenced by salt tectonics. AAPG
Bull. 102, 1239–1272.
berger for Petrel software donations. We also thank National Archive of Rowan, M.G., Jackson, M.P.A., Trudgill, B.D., 1999. Salt-related fault families and fault
Marine Seismic Surveys for providing the 3D seismic data, and Bureau of welds in the northern Gulf of Mexico. AAPG Bull. 83, 1454–1484.
Ocean Energy Management for well and biostratigraphic data. Michael Salvador, A., 1987. Late Jurassic paleogeography and origin of the Gulf of Mexico. AAPG
Bull. 71, 419–451.
Hudec helped us in interpreting some of the salt welds. Salvador, A., 1991. Triassic-Jurassic. In: Salvador, A. (Ed.), The Gulf of Mexico Basin: the
Geology of North America. Geological Society of America, Denver, pp. 131–180.
Sanford, J.C., Snedden, J.W., Gulick, S.P.S., 2016. The Cretaceous-Paleogene boundary
References
deposit in the Gulf of Mexico: Large-scale oceanic basin response to the Chicxulub
impact. J. Geophys. Res. Solid Earth 121, 1240–1261.
Bouroullec, R., Weimer, P., 2017. Geometry and kinematics of Neogene allochthonous salt Schlager, W., Camber, O., 1986. Submarine slope angles, drowning unconformities, and
systems in the Mississippi Canyon, Atwater Valley, western Lloyd Ridge, and western self-erosion of limestone escarpments. Geology 14, 762–765.
DeSoto Canyon protraction areas, northern deep-water Gulf of Mexico. AAPG Bull. Snedden, J.W., Galloway, W.E., Ganey-Curry, P., Blum, M.D., 2015. The geologic history
101, 1003–1034. of submarine fans in the deepwater Gulf of Mexico: Mesozoic to modern. Gulf Coast
Bouroullec, R., Weimer, P., Serrano, O. , 2017a. Regional structural setting and evolution Assoc. Geol. Soc. Trans. 65, 521–527.
of the Mississippi Canyon, Atwater Valley, western Lloyd Ridge, and western DeS- Steier, A., Mann, P., 2019. Late Mesozoic gravity sliding and Oxfordian hydrocarbon reser-
oto Canyon protraction areas, northern deep-water Gulf of Mexico. AAPG Bull. 101, voir potential of the northern Yucatan margin. Mar. Pet. Geol. 103, 681–701.
1035–1071. Tew, B.H., Mink, R.M., Mann, S.D., Bearden, B.L., Mancini, E.A., 1991. Geologic frame-
Bouroullec, R., Weimer, P., Serrano, O. , 2017b. Petroleum geology of the Mississippi work of Norphlet and pre-Norphlet strata of the onshore and offshore eastern Gulf of
Canyon, Atwater Valley, western Lloyd Ridge, and western DeSoto Canyon protraction Mexico. Gulf Coast Assoc. Geol. Soc. Trans. 41, 590–600.
areas, northern deep-water Gulf of Mexico: traps, reservoirs, and tectono-stratigraphic Triezenberg, P.J., Hart, P.E., Childs, J.R., National Archive of Marine Seismic Surveys
evolution. AAPG Bull. 101, 1073–1108. (NAMSS): A USGS Data Website of Marine Seismic Reflection Data Within the
Carter, R.C., Gani, M.R., Roesler, T., Sarwar, A.K.M., 2016. Submarine channel evolution U.S. Exclusive Economic Zone (EEZ), U.S. Geological Survey Data Release, 2016.
linked to rising salt domes, Gulf of Mexico, USA. Sediment. Geol. 342, 237–253. 10.5066/F7930R7P.
DeGolyer, E., 1925. Origin of North American salt domes. AAPG Bull. 9, 831–874. Trusheim, F., 1960. Mechanism of salt migration in northern Germany. AAPG Bull. 40,
Diegel, F.A., Karlo, J.F., Schuster, D.C., Shoup, R.C., Tauvers, P.R., 1995. In: Jack- 1519–1540.
son, M.P.A., Roberts, D.G., Snelson, S. (Eds.). In: Cenozoic Structural Evolution and Volozh, Y., Talbot, C., Ismail-Zadeh, A., 2003. Salt structures and hydrocarbons in the
Tectono-Stratigraphic Framework of the Northern Gulf coast Continental Margin, 65. Pricaspian Basin. AAPG Bull. 87, 313–334.
Salt tectonics: a global perspective, AAPG Memoir, pp. 109–151. Weimer, P., Zimmermann, E., Bouroullec, R., Cumella, S., Adson, J., Gutterman, W.,
Dobson, L.M., Buffler, R.T., 1997. Seismic stratigraphy and geologic history of Jurassic Payeur, T., Snyder, B., Hirsh, H., Bettinger, D., 2016. Temporal and spatial evolu-
rocks, northeastern Gulf of Mexico. AAPG Bull. 81, 100–120. tion of reservoirs, northern deepwater Gulf of Mexico. Gulf Coast Assoc. Geol. Soc.
Galloway, W.E., Whiteaker, T.L., Ganey-Curry, P., 2011. History of Cenozoic North Ameri- Trans. 66, 539–555.
can drainage basin evolution, sediment yield, and accumulation in the Gulf of Mexico Weimer, P., Bouroullec, R., Adson, J., Cossey, S.P.J., 2017. An overview of the petroleum
basin. Geosphere 7, 938–973. systems of the northern deep-water Gulf of Mexico. AAPG Bull. 101, 941–993 (2017).
Galloway, W.E., 2008. In: Miall, A.D. (Ed.). In: Depositional Evolution of the Gulf of Mex- Wu, S., Bally, A.W., Cramez, C., 1990. Allochthonous salt structure and stratigraphy of
ico sedimentary Basin, 5. Sedimentary Basins of the World, pp. 505–549. the north-eastern Gulf of Mexico. Part II: Structure. Mar. Pet. Geol. 7, 334–370.
Garcia, S.F.M., Letouzey, J., Rudkiewicz, J.L., Filho, A.D., Lamotte, D.F., 2012. Structural Zhang, Y., 2015. Promises and superiority of reverse time migration.
modeling based on sequential restoration of gravitational salt deformation in the San- https://www.cgg.com/technicalDocuments/cggv_0000008633.pdf, (accessed 17
tos Baasin (Brazil). Mar. Pet. Geol. 35, 337–353. October 2018).

You might also like