Sei sulla pagina 1di 117

Practical Salt Tectonics

A 3-day short course emphasizing the geometry and


evolution of salt structures and their impact on
petroleum systems in salt basins around the world

Mark G. Rowan
Rowan Consulting, Inc., Boulder, CO
Agenda
2:45 PM Diapir initiation (cont.)
- contraction and strike-slip
DAY 1
3:05 PM Diapir reactivation
9:00 AM Introduction
- extension
- contraction
9:20 AM Salt basins
- strike-slip
- deposition of layered evaporite sequences
3:50 PM Break
10:00 AM Break
4:05 PM Exercise 1 – Diapir evolution
10:15 AM Salt basins (cont.)
- tectonic settings of evaporite basins
5:05 PM Passive diapirs
- diapir margins
10:45 AM Mechanics
5:30 PM End of day
11:15 AM Break

11:30 AM Salt welds


DAY 2

11:50 PM Diapir initiation


9:00 AM Passive diapirs
- progradational loading
- diapir margins
- near-diapir deformation
12:30 PM Lunch
10:20 AM Break
1:30 PM Diapir initiation (cont.)
- extension
10:35 AM Allochthonous salt
- emplacement
2:30 PM Break 2
11:30 AM Break 10:00 AM Break

11:45 AM - emplacement (cont.) 10:15 AM Tectonic styles of salt deformation (cont.)


- convergent-margin foldbelts (cont.)
- styles
11:00 AM Break
12:30 PM Lunch
11:15 AM Tectonic styles of salt deformation (cont.)
- passive margins (cont.)
1:30 PM Allochthonous salt (cont.) - processes
- styles (cont.) - proximal extension

2:35 PM Break 12:30 PM Lunch

1:30 PM Tectonic styles of salt deformation (cont.)


2:50 PM Exercise 2 – GoM allochthonous salt - passive margins (cont.)
- distal contraction
3:50 PM Break - strike-slip and role of canopies

2:30 PM Break
4:05 PM Tectonic styles of salt deformation
- thick-skinned extension 2:45 PM Exercise 4 – Espirito Santo foldbelt

5:00 PM Exercise 3 – Central North Sea 3:45 PM Break

4:00 PM Petroleum systems implications


6:00 PM End of day - trap, reservoir, hydrocarbons, seal

5:20 PM Wrap-up
DAY 3
5:30 PM End of course
9:00 AM Tectonic styles of salt deformation
- convergent-margin foldbelts

3
Introduction

There has been an enormous revolution in our Structural restoration. The technique of cross-section
understanding of salt tectonics within the past 25 years (see restoration was first applied to salt structures in the late
Jackson, 1995, for an excellent history of salt tectonics 1980s (e.g., Worrall and Snelson, 1989). In the past 20
research). The beginnings of the revolution occurred within years, numerous people have used restoration to
some of the exploration companies, but their ideas only reconstruct the history of salt movement and associated
became public starting in about 1989. Our increased deformation of surrounding strata.
understanding of the geometry and evolution of salt bodies Field studies. Armed with new ideas, various researchers
and associated strata is due in large part to the fortuitous have reexamined exposed salt basins throughout the world,
convergence of advances in four areas: leading to improved understanding of the processes of salt-
Seismic imaging. There has been a steady improvement in related deformation.
seismic data acquisition and processing over the years. But In this course, we will concentrate on the new ideas about
with the advent of such techniques as pre-stack depth salt tectonics. Many of the illustrations used here are
migration, images of salt bodies became much clearer, with examples of the four areas of advance listed above. The
improved pictures of the bases of salt sheets and overhangs impact of the advances has spread to salt basins worldwide,
and the steep flanks of many diapirs. The newest and in this course we will examine styles and examples
techniques such as reverse-time migration have only added from very different basins.
to the clarity.
Salt tectonics is not just about structural geology. The
Experimental and numerical modeling. Attempts to model history of salt deformation plays a large role in the spatial
salt deformation have been made for many decades, but and temporal distribution of sedimentary facies (reservoirs)
until fairly recently, both salt and its overburden were and in the generation, migration, and entrapment of
modeled as viscous fluids. Starting in the late 1980s, hydrocarbons. These issues will therefore be summarized
however, various researchers started modeling the toward the end of the course. A better understanding of the
overburden as a brittle material, more in keeping with the interactions between salt deformation, sedimentation, and
known mechanical behavior. The results demonstrated fluid flow can only aid in the exploration for, and
salt’s more passive role of reacting to, rather than causing, exploitation of, hydrocarbons in salt basins.
deformation (e.g., Vendeville and Jackson, 1992a, b) and
fundamentally changed the ways most people look at salt
deformation.
4
Salt Basins

Salt basins are found throughout the world (Fig. 1), but a Note that only one evaporite mineral forms throughout the
quick look will show that they occur primarily in rift basins basin at any one time because of hydrological constraints.
and along passive margins. Other settings include Also, deposition is quite rapid, averaging about 1cm/a.
intracratonic basins, foreland basins, and marine basins As an oceanic water mass evaporates, a characteristic suite
that become isolated from the world ocean. of minerals forms (Fig. 5). The first evaporite mineral,
Deposition of layered evaporite sequences after 80% of the water has been evaporated, is gypsum,
which converts to anhydrite (CaSO4) with burial. Halite
Warren (2006) makes the point that modern salt
(NaCl) forms at about 90% evaporation and continues until
depositional environments (sabkha, salinas, etc.) are not
about 98% evaporation, making it the dominant mineral in
good analogues for the giant salt basins of the past. What
salt basins. With extreme amounts of evaporation, rarer
he calls ‘basinwide evaporites’ formed in large,
minerals such as sylvite and carnalite form. Because of
tectonically isolated basins in which there was a drawdown
cyclical evaporation and freshening of the water mass,
of the brine level of up to 1,500 m (Fig. 2) due to arid
most salt basins consist of layered evaporite sequences
conditions (around 30° N and S latitude). Although
(LES), with or without other lithologies. For example, the
shallow basins exist, deep basins are required to generate
Ara evaporites of Oman consist of four main cycles of
large thicknesses of evaporite (Fig. 3). Evaporation of a
halite to anhydrite to carbonate to anhydrite and back to
single mass of ocean water will not produce very much
halite (Fig. 6). In the Pricaspian Basin, there are also
salt; instead, the basin is constantly replenished by seepage
interbedded siliciclastics, especially along the eastern
from the ocean through the intervening barrier or by
margin of the basin where material was shed off the Ural
periodic overflow during major sea-level highstands (Fig.
Mountains (Fig. 7), and volcanic interbeds are common in
3). Water depths can be from essentially zero up to 300 m
some basins.
(1000’), with corresponding topographic relief (Fig. 4).
When the upper water mass is perennially freshened, Once the more mobile halite starts to move, the more
evaporites nucleate only in the shallow water, with the competent beds (e.g., carbonates) are broken up and
deeper basin getting a combination of reworked material incorporated into diapirs. In Oman, these blocks (or
and pelagic rain. With further evaporation, the entire water ‘stringers’) form exploration targets, with the reservoir and
column becomes supersaturated and bottom-nucleating source encased in sealing salt (Fig. 8). Diapirs commonly
prisms can form in the deeper parts of the basin (Fig. 4). have both isolated blocks of non-halite lithologies

5
embedded in salt and discrete bedded intervals that
commonly form isoclinal folds (Figs. 9, 10).
The more mobile lithologies (halite, carnalite, etc.) are
intensely folded and sheared (Fig. 11). This is rarely seen
on seismic data, although a notable exception is found in
the deepwater Santos Basin of Brazil (Fig. 12), where the
LES consists of 86% halite, 7% anhydrite, and 7%
carnalite and yet there are excellent images of contorted
beds. The more transparent parts of the sequence are not
diapirs or salt tongues, but simply areas where the beds
have been completely broken up during ductile flow.
Figure 13 shows undeformed and deformed examples of
the Zechstein Evaporite in the Central North Sea. Note that
in the deformed case, the upper part of the evaporite
sequence forms minibasins between more mobile diapiric
boides. In other words, the older salt strata were already
moving while younger evaporites were still being
deposited. Seeing a thick between diapirs tells you
absolurely nothing about lithology – it just means that
some salt was evacuating and the associated minibasin was
being filled with whatever lithologies were being deposited
in that area at that time.

6
Figure 1. Locations of salt basins around the world (compiled by M. Jackson).
Figure 2. Ancient salt basins (Warren 2006).
Figure 3. Basinwide evaporites (Warren 2006).
Figure 4. Lateral and depth variations (Warren 2006).
Figure 5. Progressive evaporation leads to different minerals (Rouchy & Caruso 2006).
Figure 6. Ara evaporites, Oman (Warren 2006).
Figure 7. Precaspian Basin (Gralla & Marsky 2000).
broken-up carbonate “stringers”

Figure 8. ‘Stringers’ within Oman diapirs


(Peters et al. 2003).
1 km

“Older salt” with


blocks of mafic
igneous rocks

“Younger salt” with


layered evaporites
and mudstones

Figure 9. Aerial photograph of diapir in Great Kavir Desert, Iran (Jackson et al. 1990).
Scattered blocks of
carbonate, quartzite,
siltstone, shale, &
volcanics

diapir

Isoclinally folded
strata

Figure 10. Hyperspec image of distorted cross section through Witchelina diapir, South Australia.
Figure 11. Convoluted folding of evaporites, Germany (Richter-Bernburg 1980).
Figure 12. Internal deformation in layered evaporites of the
deepwater Santos Basin, Brazil (Fiduk & Rowan 2011).
Figure 13. Undeformed and deformed examples of the Zechstein Evaporites in the North
Sea showing interbedded facies and variations in mobility (Stewart and Clark, 1999).
Tectonic settings of evaporite basins the basement architecture controls the spatial and thickness
distribution of evaporites (Fig. 18). Alternatively, salt may
There are two broad settings for basinwide evaporites. In
be part of the post-rift thermal-subsidence sequence,
the first, an older basin already exists and is connected to
resulting in more widespread salt layers and less basement
the world ocean. Isolation of the basin, through some
control (Fig. 18).
combination of tectonics or glacio-eustatic sea-level
fluctuations, creates drawdown and evaporite deposition Salt commonly occurs in paired basins on either side of
(Fig. 14). The classic example is the Messinian salinity oceanic crust, such as across the Gulf of Mexico, the South
crisis in the Mediterranean near the end of the Miocene Atlantic, and the North Atlantic. Thus, it has generally
(Fig. 15), when tectonics in northern Morocco effectively been thought that evaporite deposition occurs only on
closed the connection to the Atlantic and up to 2 km of salt continental crust, with subsequent oceanic spreading
formed in less than 500,000 years. Because of the complex separating a once contiguous basin into two parts (Fig. 19).
tectonics and consequent basin morphology, there are Some passive margins are interpreted to have seaward-
chronological and facies variations between the basins dipping reflections, or SDRs – subaerial basalts considered
even though they formed due to the same process (Fig. 15). to be the initial expression of oceanic spreading – beneath
the salt, leading to a model in which salt deposition
The other broad setting for basinwide evaporites includes
postdates the onset of oceanic spreading (Fig. 20). Salt on
various tectonically created basins such as foreland basins
the Angolan margin, for example, is interpreted to occur
and rift basins, in which subsidence in ongoing while
above the breakup unconformity and thus is underlain by a
evaporites are deposited (Fig. 16). Foreland basins form in
combination of proximal continental crust and distal proto-
convergent-margin settings. A classic example is the Ebro
oceanic crust (Fig. 21).
Basin south of the main Pyrenees mountain belt in northern
Spain (Fig. 17). Loading of the crust created a basin in There are some relatively new models for highly attenuated
which evaporites formed, but there are actually three passive margins (e.g., Karner et al. 2007, Péron-Pinvidic &
evaporite layers of different ages that partly overlap Manatschal 2008) in which extension is depth-dependent,
spatially (Fig. 17). We will examine how this affects occurring at different levels of the crust at different
shortening geometries later in the course. locations (Fig. 22). These models explain, for example, the
great thickness of thermal-sag fill relative to the amount of
Rift basins have a distinct basement architecture made up
upper-crustal extension seen in such basins as the Angolan
of grabens and half-grabens segmented by transverse
margin and therefore have important implications for the
structures such as accommodation zones or transfer faults
distribution of salt (Karner et al. 2003, Karner & Gamboa
(Fig. 18). Salt may form part of the syn-rift fill, such that
2007). Depending on just when the
20
basin is isolated and evaporation occurs, salt may be
deposited entirely above normal rifted continental crust
(Fig. 22a), it may be deposited over highly attenuated crust
(Fig. 22b), or it may be locally deposited over proto-
oceanic crust or even exhumed mantle (Fig. 22c).
If oceanic spreading post-dates salt deposition, the salt is
pulled apart, creating a hypothetical salt chasm (Fig. 23).
However, salt is not strong enough to hold up any kind of
topographic relief and will flow out over the oceanic crust,
something happening today in the Red Sea (Fig. 24). New
models for the northern Gulf of Mexico (e.g., Imbert &
Philippe 2005, Pindell & Kennan 2007, Hudec & Peel
2010) suggest that the most distal deep salt is not actually
the original depositional salt, but rather an early
allochthonous salt nappe that extruded out over oceanic
crust and then oceanic sediment for a distance of possibly
as much as 200 km.
In all tectonic settings, the basin morphology controls the
spatial and thickness distribution of evaporites, and this in
turn will impact subsequent salt-related deformation such
as the locations of major faults or the presence or absence
of diapirs and canopies. But stratigraphic relationships are
generally different in the various settings: (1) an ocean
basin that gets isolated will generally have deepwater
sediment both below and above the salt; (2) rift basins and
passive margins will typically transition upward from
subaerial or lacustrine facies through the evaporites to
marine facies (Fig. 21); and (3) foreland basins will often
display the opposite, with marine facies below and
continental facies above (Fig. 17).
21
Figure 14. Evaporite deposition in an open marine basin that gets
isolated from the world ocean (Warren 2006).
Figure 15. The Mediterranean became isolated during the Messinian
(latest Miocene) due to a combination of sea-level fall and contractional
tectonics (from Ryan 2009 and Rouchy & Caruso 2006).
Figure 16. Different tectonic settings for evaporite basins (Warren 2006).
Figure 17. The Ebro Basin is a foreland basin
south of the Pyrenees in which three different,
partially overlapping evaporite levels formed
(from Vergés et al. 1992 and Sans et al. 2003).
Figure 18. Crustal-scale extension creates
a complex rift-basin architecture that
controls evaporite deposition (Stonely
1981). Salt may be part of the syn-rift fill
or part of a larger post-rift sag basin (Tari
et al. 2003).
Figure 19. Model for paired salt basins in which one originally contiguous basin
is separated by oceanic spreading (Fonck et al., 1998).
Figure 20. Alternative model for paired salt basins in which salt is deposited after
oceanic spreading begins, so that the basins were never contiguous (Fonck et al., 1998).
Figure 21. Regional cross section across Angolan margin showing salt
deposition above both continental and oceanic crust (Jackson et al., 2000).
Figure 22. Four-stage, depth-dependent stretching model for development of highly
extended margins (Péron-Pinvidic & Manatschal 2008). Salt will have different
relationships to underlying crustal type depending on when it is deposited in this history.
Figure 23. Salt basin separated by oceanic spreading forms hypothetical salt chasm
(Pindell & Kennan 2007), but in reality salt flows over oceanic crust as it is formed
so that it appears to have been deposited over oceanic crust.
Figure 24. Salt flowing over newly formed oceanic crust in the
Red Sea (Mitchell et al. 2010).
Mechanics

Rock salt (halite) is a very different material from other, Gulf of Mexico needs to be approximately 2,300 m (7,500
more typical sedimentary rock types. There are several key ft) thick before the average density is greater than that of
factors that play roles in dictating its behavior. First, salt is the salt (Hudec et al. 2009b).
much weaker than other lithologies under both tension and The relatively low density of buried salt was historically
compression (Fig. 25). Even overpressured shale almost used as a rationale for models in which salt punches its
always has more strength than salt. In fact, the curve for way through more dense overburden until it reached the
wet salt in Figure 30 falls essentially on the axis of zero surface, creating a series of peripheral sinks (or rim
strength. The reason is that salt deforms as a viscous synclines) (Fig. 29). Buoyancy was also invoked to cause
material that effectively flows, with flow rates up to 15 m/a the diapir to keep rising as a teardrop, with the surrounding
in exposed diapirs in Iran (Talbot et al., 2000). Flow is by a strata collapsing back into the necking-off feeder (Fig. 30).
combination of Poiseuille flow due to overburden loading These models envision the overburden as another, more
and Couette flow due to overburden translation (Fig. 26). dense viscous fluid, but in fact, the overburden is a brittle
The viscous nature of salt means that it forms a constant- material with real strength (Figs. 25 and 27). The under-
strength, albeit very weak, layer between normal consolidated nature of shallow sediments in places like the
sedimentary layers whose strength increases with depth northern Gulf of Mexico may suggest a viscous nature, but
(Fig. 27). Thus, salt serves as an excellent detachment this is incorrect. In fact, there are plenty of fault scarps at
surface into which faults sole. the sea floor that attest to the brittle style of deformation of
Second, salt has a relatively constant density of about 2200 even very young sediments. Moreover, if density is the
kg/m3, irrespective of burial depth (it actually gets slightly drive, how does a minibasin in the Gulf of Mexico get
less dense with depth due to temperature effects). This is in started if it doesn’t become more dense than salt until it is
contrast to other sedimentary strata, such as sands and quite mature?
shales, that become compacted during burial and thus Vendeville and Jackson (1992a) argued that the strength
become more dense (Fig. 28). Thus, salt is more dense than and brittle nature of the overburden means that density
its immediately surrounding strata when it is near the contrast plays only a limited role in salt tectonics. Instead,
surface, but is less dense than siliciclastics at burial depths they showed that salt should be viewed as a pressurized
of 500-900 m. However, if you integrate the density fluid and that it is the differential fluid pressure that drives
profile of overburden from the surface or seafloor down to
salt flow (Fig. 31). There are four key messages in this
the top of salt, a minibasin in a passive margin such as the figure. First, density is only a second-order factor. Yes, it
33
plays a role in determining the pressure on the salt, but it or just below the surface, the dip of the salt-sediment
can be misleading. For example, a less dense block sinks interface is a function of the interplay between the
more (Fig. 31b) and a denser block rises higher (Fig. 31d) sediment accumulation rate and the salt-supply rate (Fig.
because it is the differential pressure that is important. 33), and the salt-supply rate increases as the load increases.
Second, if there is a differential load on the fluid (salt) and For the same reason, many diapirs have hourglass shapes,
there is a place for it to go, salt will flow. Third, and with a deep pedestal, a narrow waist, and a flaring top (Fig.
conversely, the salt cannot just push its way into the 34). If we keep sediment accumulation rate constant, we
overburden because of the overburdens strength. In other move up a vertical line on Fig. 35 as the load increases,
words, for salt evacuation and diapirism to occur: (1) there going from the tapering field (lower right) into the flaring
must be an open path to a near-surface salt body; (2) the field (upper left).
overburden must be thin and weak enough for the
Eventually, as a salt layer thins, viscous drag forces inhibit
differential fluid pressure to overcome the overburden
lateral flow of salt, and minibasin subsidence rates slow
strength, in which case it will deform the thin overburden;
even though there can be a large differential load.
or (3) some other process (e.g., tectonic) must create space.
Minibasin subsidence ceases once the underlying salt is
The fourth message, or bottom line, in Figure 31 is that salt
depleted and a salt weld forms (indicated by pairs of dots,
does not drive salt tectonics; instead, salt merely facilitates
Fig. 32).
and reacts passively to external forces. This is a key point
that underlies almost all subsequent discussions.
Let’s look at a (more dense) minibasin subsiding into salt
(Fig. 32). As the minibasin grows, the load increases
compared to the thin overburden above the flanking highs
(not shown). Thus, all other things being equal, subsidence
and diapirism rates gradually increase through time. Note
that subsidence is not driven by the sediment added during
any short time interval, but by the net differential load that
exists at any time. Thus, it is the existing basin that drives
salt flow – active sedimentation just adds to the differential
load and is thus a second-order effect.
You can see that the edge of the minibasin gets steeper
over time (Fig. 32). This is because, for a diapir growing at
34
Figure 25. Strength of various rock types in both tension and compression
(Jackson and Vendeville, 1994). Note that wet salt falls effectively on the
axis of zero strength because it is a viscous material.
Poiseuille flow

Couette flow

Figure 26. Types of flow of an undeformed viscous material (a): (b) Poiseuille flow, in which
the overburden pressure drives lateral flow in an unconfined layer (velocities are slower at the
boundaries because of viscous drag); and (c) Couette flow, in which the viscous material is
sheared due to lateral translation of the overburden. Figure from B. Vendeville.
Figure 27. Simple 3-layer model of the crust with a weak, but constant-
strength, salt layer between two brittle layers whose strength increases
with depth (Vendeville and Jackson, 1993).
Figure 28. Curves of density vs. depth for various lithologies (Jackson and Talbot, 1986).
Salt has an approximately constant density, whereas sands and shales get more dense
with depth, so that buried salt is less dense than its overburden.
Figure 29. Classic model of salt diapirism in which density inversion causes buoyant rise
of the salt, first forming a salt pillow and then punching through to the surface (Trusheim
1960). Different stages in the evolution are recorded by primary (I), secondary (II), and
tertiary (III) peripheral sinks (or rim synclines).
Figure 30. Numerical model of buoyant rise of a less dense viscous material
through a more dense viscous material (Podladchikov et al. 1993).
Figure 31. Model of salt as a pressurized fluid (Vendeville and Jackson, 1992a),
with the basic set-up in (a) and three scenarios with different density contrasts
between the pressurized fluid and the overburden in (b)-(d).
Figure 32. Evolution of a denser minibasin subsiding into salt: (a) initiation; (b)
subsidence and growth of basin even in the absence of sediment input because of the
differential load on the salt; and (c) cessation of subsidence once salt weld forms.
Figure 33. Dip of salt-sediment interface as a function of the interplay between
sedimentation rate and salt-supply rate (McGuinness & Hossack 1993).
Figure 34. Hourglass-shaped diapir (prestack depth migration from
cover of Veritas Newsletter, Dec. 2004).
Salt Welds

Salt welds represent completely or mostly evacuated salt


bodies and were initially classified as primary
(autochthonous level), secondary (vertical diapir), or
tertiary (allochthonous level) (Fig. 35, Jackson & Cramez
1989). Rowan et al. (1999) later divided welds in the
northern GoM into primary, roho, counterregional, bowl,
thrust, and wrench welds depending on their geometry and
origin.
Primary welds are generally subhorizontal because the
depositional salt is generally subhorizontal (Fig. 36).
Tertiary welds represent the base of salt canopies and can
thus have highly variable geometries (Fig. 37). Secondary
welds often form by shortening-induced squeezing of salt
walls or diapirs, leading to vertical welds (Fig. 38) or
inclined thrust welds (Fig. 39). Steep secondary welds can
also form without shortening, as illustrated in Figure 40.
Welds are typically defined on seismic data, so that it is
generally unknown whether salt is totally evacuated. A
weld may be complete (no salt), incomplete (up to 50 m of
salt; see Wagner 2010), discontinuous (areas of more than
50 m of salt), or apparent (greater than 50 m of salt but
interpreted as a weld) (Rowan et al. 2011). Their impact on
hydrocarbon migration/sealing will be examined in a later
portion of the course.

45
Figure 35. Formation of primary, secondary, and tertiary salt welds
(Jackson & Cramez 1989).
Figure 36. Primary salt weld from Block 33, Lower Congo Basin
(data courtesy of C. Fiduk and CGGVeritas).
Figure 37. Tertiary salt weld from the northern GoM
(Hudec et al. 2009).
Figure 38. Vertical secondary weld formed by
squeezing of salt wall in La Popa Basin, Mexico
(Rowan et al. 2011).
Figure 39. Inclined secondary weld
(thrust weld) formed by squeezing
and thrusting of a salt diapir in the
Lower Congo Basin, Gabon (Jackson
et al. 2008).
Figure 40. Inclined secondary diapir in the
northern GoM (Dyson & Rowan 2004) and
the genesis of counterregional welds by
progressive expulsion rather than shortening
(Rowan & Inman 2005).
counterregional weld
Diapir Initiation

We have introduced what drives salt movement, but how will sink until the salt is welded, at which point the
does it get initiated? Global tectonics can trigger salt depocenter shifts basinward and the process is repeated
movement, whether under extension, contraction, or strike- (Fig. 43). The result is an expulsion rollover, with salt
slip deformation. But even in the absence of a tectonic expelled basinward to areas of inflation. An example from
drive, salt can begin moving if there is hydraulic head. the Pricaspian Basin shows the progressively shifting
Hydraulic head consists of two components: elevation head depocenters that developed as salt was expelled up and
and pressure head (Fig. 41). Either one, or a combination basinward into an allochthonous salt body (Fig. 44). Note
of the two, will trigger salt flow as long as the head is great that this process must initiate early for salt flow to occur;
enough to overcome the strength of the overburden and otherwise, the pressure head is insufficiently large to
viscous drag effects in the salt. Note that elevation head overcome the resistance of a thicker, stronger overburden
can drive salt movement without any overburden, a process (Fig. 45).
termed salt drainage (Quirk et al. 2008).
The “Albian Gap” in the Santos Basin, Brazil, is a
Salt flow driven by hydraulic head typically involves updip basinward-dipping growth monocline (Fig. 46) that has
or proximal extension and downdip or distal contraction. been interpreted as an expulsion rollover by Ge et al.
This is certainly true in the case of elevation head, when it (1997). Although it can be restored accordingly (Fig. 46), it
is called gravity gliding, but is also common in the case of can also be restored as a large counterregional normal fault
pressure head, when it is called gravity spreading (see later (Fig. 47). Only the salt moves basinward in the evacuation
section on passive margins). Of course, the overburden interpretation, but both salt and overburden move
never moves as a solid mass, so there will also be zones of basinward in the extensional interpretation. Significant
strike-slip movement as the deformation is shortening of equivalent age in more distal positions
compartmentalized (Fig. 42). In the following sections, we supports the extensional origin (Fig. 47).
examine three different generators of salt flow, be they due
to hydraulic head or due to regional tectonics. If the salt layer thins over basement steps, instead of being
constant-thickness, inflation is localize over the basement
steps, thereby creating salt highs that ultimately evolve into
diapirs (Fig. 48). An example of this is found in the
Progradational loading Paradox Basin of the western USA, where there is a series
If sediment progrades such that there is a surface slope, of salt walls located over basement faults (Fig. 49). The
pressure head is developed. Even if the overburden is less evaporites were loaded immediately after deposition by
dense than the salt, it is more dense than the laterally prograding coarse clastics, derived from the adjacent
adjacent water or air. The area beneath the sediment wedge Uncompahgre Uplift, which thinned toward the SW (Fig.
50). The shifting depocenters and the growth of one salt
wall are nicely illustrated in Figure 51. 52
Figure 41. Hydraulic head, comprising (a) pressure head and/or
(b) elevation head (Hudec & Jackson 2007).
basinward

landward
Shortening Translation Extension
(salt evacuation and diapirism)

Figure 42. Linked systems of deformation on a passive margin; diapirs


may be triggered by extension, contraction, or strike-slip deformation.
Figure 43. Progradational trigger for salt movement (pressure head),
producing basinward-shifting depocenters in an expulsion rollover, with
salt moving basinward and inflating the overburden (Ge et al. 1997).
Allochthonous
Suprasalt salt

Thickening
onto weld/salt
Thinning Subsalt
onto weld

Weld

Presalt

Figure 44. Expulsion rollover in the Pricaspian Basin, Kazakhstan.


Note terminology of suprasalt, subsalt, and presalt.
salt will flow because
hydraulic head large
enough to overcome
strength of overburden

salt will not flow


because overburden
too thick and strong –
get continued
progradation

Figure 45. Progradational loading must be early in order to drive


salt movement.
Depth section
VE = 1 10 km

End Cretaceous

End Albian

Figure 46. “Albian Gap” in Santos Basin, Brazil (Mohriak & Szatmari 2008), interpreted
as an expulsion rollover by Ge et al. (2007). Restoration from Rowan & Ratliff (2010).
Depth section
VE = 1 10 km

End Cretaceous

End Albian

Albian Gap

Figure 47. Alternative restoration as a counterregional normal fault (Rowan & Ratliff 2010).
Distal shortening of the same age (Modica & Brush 2004) supports this interpretation.
Figure 48. If salt thins over basement steps, progradational loading
causes local inflation and diapir initiation over the steps (Ge et al. 1997).
Figure 49. Gravity gradient map of Paradox Basin with hot colors
denoting edges of salt walls (Trudgill 2011). Note that salt walls
are located over interpreted presalt basement faults.
Figure 50. Cross sections showing progradational geometry extending from
Uncompahgre Front and diapirs located over basement faults (Trudgill 2011).
Figure 51. Progressive growth of salt wall driven by sediment
prograding into the basin from uplifted basement (Trudgill 2011).
Extension many extensional provinces are asymmetric, dominated by
basinward-dipping faults with triangular salt rollers in their
Analyses of salt basins combined with experimental
footwalls (Fig. 56). The dominant control is the geometry
models suggest that many diapirs are initiated by, and grow
of the detachment, with dipping detachments having faults
during, regional extension (Vendeville & Jackson 1992a,
that dip in the same direction (Fig. 57). Down-to-the-basin
Jackson & Vendeville 1994). The models show that diapirs
faults typically have displacement that is a maximum near
may go through three evolutionary stages: reactive, active,
the center of the fault and decreases to zero at the fault tips
and passive diapirism (Fig. 52).
(Fig. 58). A cross section where displacement is small and
When a horizontal salt layer is buried by constant- late shows a small salt roller, whereas a section where the
thickness strata, nothing will happen (even if the displacement is greatest shows a larger salt roller and a
overburden is more dense) until there is sufficient thinner overburden (Fig. 58). This is where a diapir will
hydraulic head or some external force is applied. In the break through if there is enough salt and enough
case of extension, the overburden is lengthened and differential load established by the faulting and
thinned, which is accommodated by graben formation at synkinematic deposition. An example is shown in Figs.
the surface and an “inverse graben” at the salt-sediment 59-61, which are parallel seismic profiles that cross the
interface (large-scale boudinage). Salt reacts to the same counterregional (landward-dipping) fault.
extension by overall thinning and by passively filling the Displacement is small in Fig. 59 but much greater in Fig.
space in the inverse graben (Figs. 52, 53). The result is a 60. Where it was even larger, a diapir broke through and
triangular diapir (elongate in the strike direction) that has continued rising over time (Fig. 61). Note that the
flanking growth faults that get younger toward the diapir extensional fault(s) are abandoned and rotated on one side
crest. The diapir grows in size as extension progresses. An of the diapir.
example from the northern Gulf of Mexico is shown in Fig.
54 – again, note the triangular shape, the overlying seafloor Figs. 60 and 61 show turtle structures on either side of the
graben, and the increase in fault age down the diapir counterregional fault or diapir. Turtles form when an initial
flanks. Restoration shows how this diapir formed as the minibasin touches down and can no longer subside, but the
overburden moved basinward above the salt, pulling away minibasin flanks collapse, thereby inverting the early basin
from the deeper minibasin to the north (Fig. 55). The width (Fig. 62). But what drives the formation of the initial
of the salt at any given stratigraphic level shows how much minibasin and ultimate turtle formation? The most likely
extension is hidden in the salt. candidate is extension: experimental models with no lateral
movement resulted in simple minibasins, whereas
Classic reactive diapirism is symmetrical, in that conjugate extension helped flank collapse and the formation of turtle
sets of normal faults develop over the growing diapir. But structures (Fig. 63). Turtle structures from the Pricaspian
64 margin,
Basin (Fig. 64) are elongate, parallel to the basin
and located along the basin margin (Fig. 65), a province
typically characterized by extension as salt and overburden
slides into the basin. Moreover, Fig. 66 shows a
prekinematic layer (between salt and the blue horizon) that,
when rotated back to horizontal, does not fit back together
as it should. The implication is that early extension pulled
the thin overburden apart, allowing salt to break through
and rise and the minibasin to subside, with continued
extension likely leading to flank collapse and turtle
formation. A turtle structure from the northern Gulf of
Mexico (Fig. 67) shows that initial subsidence didn’t even
start until ~100 Ma after salt deposition, despite the more
dense overlying carbonates. The trigger was probably
gravitational failure of the passive margin in the early
Cenozoic.
Turtle structures may or may not be charged with
hydrocarbons. In the Pricaspian Basin (Fig. 66), the source
rocks are beneath the salt. Hydrocarbon discoveries around
the tops of the diapirs demonstrate that hydrocarbons
moved across the welds, but the turtle structures are not
charged because the section immediately above the weld
comprises shale, anhydrite, and halite and is thus relatively
impermeable. In the northern GoM, the source rocks are
just above the salt. In some turtles, they are well developed
(Fig. 67) overlying sands are charged (e.g., Thunder
Horse). In others, source rocks are thin to absent (Fig. 68)
and, moreover, overlying sands have a synclinal geometry
so that there is no charge into higher sands in the turtle
(e.g., Mensa Deep).

65
Figure 52. Model of diapir initiation and growth during extension (Vendeville and Jackson,
1992a): (1) necking down of the overburden creates a reactive diapir (large-scale boudinage);
(2) once the diapir is tall enough and the overburden is thin and weak enough, the diapir will
actively punch through to the surface; and (3) once at the surface, the diapir will grow
passively as surrounding minibasins subside into and displace the salt in the source layer.
Figure 53. Steps in the evolution of a reactive diapir with no synkinematic sedimentation
(Jackson and Vendeville, 1994). The diapir widens and gets taller with time, and the faults
get younger toward the crest of the diapir (note that it is a passive diapir by step (e)).
Figure 54. Reactive diapir from the northern GoM, with triangular shape,
overlying graben, and older faults down the flanks (Rowan 1995).
Figure 55. Restoration of the reactive diapir in Fig. 54, showing its growth
as the overburden moves basinward relative to the minibasin (Rowan 1995).
0

2
1 km

Figure 56. Down-to-the-basin faults from offshore Angola soling into a basinward-
dipping salt detachment with salt rollers in fault footwalls (Brun & Mauduit 2009).
Figure 57. Extensional style as a function of the dip of the detachment
(modified from Stewart 1999).
Figure 58. Listric normal fault with displacement at a maximum near the center of the fault and
decreasing toward the fault tips. The salt roller is largest and the overburden is thinnest at the
point of maximum displacement, so that salt break-through and diapir initiation will occur here.
W E
counterregional fault

Figure 59. 3-D, time-migrated seismic profile from the Espirito Santo Basin in Brazil
(data courtesy of C. Fiduk and CGGVeritas). This line crosses a counterregional fault
near its termination, where displacement is small.
W E
counterregional fault

Figure 60. 3-D, time-migrated seismic profile from the Espirito Santo Basin in Brazil (data
courtesy of C. Fiduk and CGGVeritas), crossing the same counterregional where displacement is
larger. Note the larger salt roller and the thinner overburden at the time of the yellow horizon.
W E

Figure 61. 3-D, time-migrated seismic profile from the Espirito Santo Basin in Brazil (data courtesy of C.
Fiduk and CGGVeritas), <2 km from the line in Fig. 60. Note the asymmetric thickness of the red-yellow
interval, similar to that across the counterregional fault, indicating diapir initiation along the fault.
Figure 62. Formation of a turtle structure by flank collapse and inversion
of a minibasin whose center has touched down and formed a weld.
(a) (b)

Figure 63. Model results of minibasin subsidence showing no turtle formation in the absence
of extension (a) and turtle formation triggered by extension (b) (courtesy of B. Vendeville).
Crestal
faulting Flank collapse
Salt Flank collapse
Salt
Initial depocenter

Weld

Figure 64. Turtle structure from the Pricaspian Basin located between adjacent
diapirs. Note the crestal faulting and erosion. Also note that subsidence and
minibasin formation began immediately after salt deposition.
basin margin
basin center

20 km

Turtle structure

Figure 65. Bouguer gravity map of eastern Pricaspian Basin (Barde el al. 2002) modified to
show linear trends of turtle structures (orange lines) along and parallel to the basin margin,
probably triggered by thin-skinned extension due to gravity gliding into the basin center.
Figure 66. Diapir from the Pricaspian Basin with a prekinematic layer between the salt and the
blue horizon. If this layer is restored to horizontal (in depth), there is a gap that suggests
diapirism was triggered by extension, snapping the thin overburden and allowing salt to move.
SW NE
20 km

meters
Figure 67. Turtle structure from the northern GoM (data courtesy of ION/GXT).
Note that initial minibasin formation did not occur until the time of the red horizon,
approximately 100 my after salt deposition.
source rocks

Figure 68. Turtle structure from the northern GoM (data courtesy of ION/GXT). The
source rocks are thin or even absent and the deepest sands have a synclinal geometry.
Contraction
Contraction (shortening) can also initiate diapirism. When
shortening occurs above a salt décollement, simple
detachment folds typically form. If the overburden is
eroded and/or weakened by faulting, salt may break
through to the surface and subsequently grow as a diapir
(Figs. 69, 70). A striking example of a breached fold from
offshore Angola is illustrated in Fig. 71. Any further
upward movement of salt will result in a diapir (or an
allochthonous salt sheet). It is important to stress that this
mechanism of diapir initiation, like progradation and
usually extension, can only happen when there is a thin
overburden. Otherwise, the differential load is insufficient
and the thickness and strength of the overburden is too
great and only a fold will form (Fig. 72).
Strike-slip
A variation of reactive diapirism occurs in wrench-fault
settings. Pull-apart basins form where there is a releasing
bend or stepover in strike-slip faults. These are the sites of
very rapid thinning of the overburden and are ideal
locations for the generation of reactive diapirs (Fig. 73).
An example of diapirs at releasing bends is seen in the
Zagros Mountains of Iran (Fig. 74). Of course, a
restraining bend will create local contraction, which can
also trigger diapirism if early in the history.

83
Figure 69. Formation of a diapir by erosion and salt breakthrough in a
contractional detachment fold (Coward and Stewart, 1995).
Figure 70. Physical model showing contractional folding, deep
erosion, and salt break-through (Sans & Koyi 2001).
Figure 71. Breached anticline in Lower Congo Basin (data courtesy of C. Fiduk and CGG-
Veritas). Further squeezing or subsidence into salt will initiate diapirism or lateral salt flow.
salt will break through
and initiate diapirism
because overburden is
thin and weak

salt will not break through


because overburden too
thick and strong

Figure 72. Shortening must be early in order to trigger break-through and diapirism,
unless there is significant erosion of thick prekinematic overburden as in Fig. 70.
Figure 73. Origin of a diapir in a releasing bend or stepover in a strike-slip fault
system (Letouzey and Sherkati, 2004). This deformation effectively creates a hole
in the overburden, allowing salt a pathway to the surface.
Figure 74. Location of diapirs at pull-apart steps in right-lateral tear faults
in the Zagros Mts. fold-and-thrust belt, Iran (Talbot and Alavi, 1996).
Diapir Reactivation

Existing diapirs commonly get reactivated during Shortening of diapirs


extension, shortening, or strike-slip movement. The diapirs
If contraction continues or initiates after diapirs are
are the weakest parts of the rock volume and thus localize
formed, the diapirs again will localize the contractional
any deformation in the area.
strain (cryptic shortening). In the case of a buried diapir,
the salt body narrows and the overburden folds, with salt
displaced from the diapir filling the core of the overlying
Extensional collapse
fold (Fig. 78). This is termed diapir rejuvenation, and
If extension continues or takes place after diapirs have although it appears as though the salt is actively raising the
already formed, the diapirs will widen with time because overburden (active diapirism), the deformation is actually a
this is easier than faulting the stronger minibasins. As long response to imposed lateral forces. An example of a
as there is still salt coming from the source layer, the diapir rejuvenated diapir with a folded thick roof is shown in
can grow and widen at the same time (‘cryptic’ extension). Figure 79; it can be shown to be the result of shortening
However, if the extension continues after the source layer because thrust faults and contractional folds extend
is depleted, the diapir will collapse as it widens, thereby laterally away from the diapirs. Other models have created
forming a young depocenter over what was the diapiric more asymmetric reactivation, with the development of
high (Fig. 75). Taken to the extreme, this produces what is thrust faults that effectively use the squeezed diapirs (Fig.
called a “mock-turtle” anticline. Serial sections through a 80). An example from Gabon is shown in Figure 39. Yet
model show the sequential widening and collapse of a salt other models have shown that salt is not just forced upward
diapir (Fig. 76) – note that the narrow salt ‘horns’ in the during shortening, but that there are plumes of salt that are
lower panel (Fig. 76) are not intrusions of salt up the driven downward and back into the source layer if the
normal faults, but are simply remnants of the original diapir roof is thick (Fig. 81). However, the models are
diapir after downdropping of the hanging walls during somewhat unrealistic in that they start with a significant
diapir collapse. An example of a slightly fallen diapir from thickness of deep salt and a thick roof; typically, a diapir
the extensional province of the Central North Sea is shown gets buried only when it runs out of salt. But if flanking
in Fig. 77. strata are folded/thrust upward, delaminating at the salt
detachment, then salt will indeed fill the space created.
Note that diapir rejuvenation refers strictly to contraction
of diapirs that are already buried. However, diapirs can
90
also accommodate shortening while salt is still flowing Summary
from the source layer. In this case, both differential loading
The weakness of diapirs compared to the stronger
and shortening contribute to diapir growth, so that salt-
minibasins means that diapirs often have complex
flow rates are increased significantly. Shortening can also
histories, as we will see in the next exercise. In the same
drive diapir rise even after the deep salt layer is depleted
vein, Fig. 88 shows a reinterpretation of the classic
by squeezing the diapir until if forms a secondary weld
geometries identified by Trusheim (1960). The diapir is
(Fig. 82). Modeling by B. Vendeville has shown that very
envisioned by B. Vendeville as having been triggered by
minor shortening can drive salt rise, even when there is no
extension, with the formation of a turtle structure with its
geometric evidence for shortening in immediately
central depocenter (primary peripheral sink) and flank
surrounding strata, as in the example of Fig. 83. This is
collapse (secondary peripheral sink), and later shortened to
similar to early doming of a diapir roof due to shortening
create the tertiary peripheral sink. The drive is not buoyant
even when there are no lateral shortening structures (Fig.
rise of the salt, but is external to the salt.
84).
Can a vertical diapir get welded in the absence of
shortening? This has been proposed for Upheaval Dome in
the Paradox Basin (Fig. 85), and requires a system of
surrounding normal faults that accommodate movement
back into the evacuating feeder (Fig. 86). However, a
system of radial thrust faults is required in 3-D, something
not observed in nature, and most people reject the squeezed
diapir interpretation for Upheaval Dome and favor instead
an impact origin.

Reactivation in strike-slip
Salt diapirs can also, of course, be reactivated in strike-slip
movement at bends or stepovers in strike-slip faults (Fig.
87). Restraining bends or stepovers result in squeezing of
diapirs, and releasing bends or stepovers lead to
extensional collapse of diapirs.
91
Figure 75. Diapir fall driven by continued extension after the source layer is depleted; as the diapir
widens and collapses, it is replaced by a young depocenter (Vendeville and Jackson, 1992b).
Figure 76. Serial cuts through a model showing different stages in the collapse
of a diapir (Vendeville and Jackson, 1992b). Note the salt horn in (b) and (c)
that is a remnant of the original diapir, not an intrusion up the fault.
Figure 77. Partially collapsed diapir from the Central North Sea
(seismic from Jones et al. 2005).
Figure 78. Rejuvenation of a buried diapir during shortening; the diapir is squeezed as the
overburden is folded, driving salt higher into the fold core (Vendeville and Nilsen, 1995).
W E

Figure 79. 3-D, time-migrated seismic profile showing rejuvenated diapir in the
Espirito Santo Basin, Brazil (data courtesy of C. Fiduk and CGGVeritas).
Figure 80. Rejuvenation models with thrust development (Callot et al. 2007).
If the cover is thin, salt can break through to the surface.
Figure 81. Summary cartoon of models of diapir rejuvenation in which
salt flows back down into the source layer (Dooley et al. 2009a).
Figure 82. Shortening can drive further passive diapirism even after the source layer is depleted
(Hudec & Jackson 2007). Growth ceases and the diapir is buried once a secondary weld forms.
Figure 83. Diapir from the northern Gulf of Mexico growing with no adjacent evacuation
basins or shortening structures (data courtesy of C. Fiduk and CGGVeritas).
Figure 84. Surface view of modeled shortening-induced squeezing of a buried diapir
showing that laterally adjacent folds/thrusts are absent in early stages (Dooley et al. 2009b).
Figure 85. Proposed pinch-off of a vertical diapir at Upheaval Dome
in the Paradox Basin (Jackson et al., 1998).
Figure 86. Pinch-off by inward collapse of surrounding strata requires
concentric normal faults and radial reverse faults (Jackson et al., 1998).
Figure 87. Reactivation of diapirs during strike-slip deformation at either
a restraining stepover or a releasing stepover (Hudec & Jackson 2007).
Figure 88. Reinterpretation of the classic Trusheim (1960) model in
which early extension and late contraction are invoked (Vendeville 2002).
Exercise 1

In this exercise, you will interpret the three seismic profiles Exercise 1.2
in the following pages. They are from different tectonic
This time-migrated, 2-D profile from offshore Brazil
regimes from different parts of the world. The goal is to do
shows a diapir rising from the autochthonous (Aptian) salt
quick interpretations of the profiles and use the observed
layer. Basinward is to the right. Again, sketch an
geometries to deduce the history of diapir growth. Don’t
interpretation of the salt (including welds), any faults, and
forget to use the concept of “regional:, i.e., the level of a
the indicated horizons. When and how did the diapir
horizon where it has not gone down, due to salt evacuation
initiate and how did it evolve through time? Sketch the
or extension, or up due to salt inflation or shortening.
geometry at the times of the different horizons. Seismic
profile from Demercian et al. (1993).
Exercise 1.1
This is a time-migrated inline from a 3-D survey over an Exercise 1.3
allochthonous salt sheet in the northern Gulf of Mexico.
This time-migrated, 2-D seismic profile from the
Basinward is to the left. Sketch an interpretation of the top
Pricaspian Basin in Kazakhstan shows a diapir between
salt and overburden geometry using the indicated horizon.
two turtle structures. Basinward is to the left. Interpret the
Don’t worry about the base salt or deeper geometries.
salt (and weld) geometry. What was the timing and
What is the history of the prominent diapir on the sheet?
mechanism of diapir initiation? Was salt movement caused
What style(s) of diapirism were active? Assume: first, that
purely by vertical loading or was there any lateral
there are young extensional faults off-section to the north
(tectonic) trigger?
(right) that sole into the salt layer; and second, that no such
faults exist. Seismic data courtesy of D. Ratcliff and
Diamond Geophysical.

106
S N

Exercise 1.1 courtesy of Diamond Geophysical


W E
W E

Exercise 1.3
Passive Diapirs

We have seen that salt flow and diapirism may be triggered upper 3000-5000 ft. of the earth’s surface (Fig. 93). This is
by progradational loading, extension, contraction, or strike- because dissolution is a consequence of the circulation of
slip deformation. In all cases, areas of inflated salt with meteoric water, which dissolves and carries the salt away
thin overburden are created. At some point, the differential and is constantly replenished. Deeper diapirs below the
pressure on the salt is great enough to overcome the zone of meteoric-water circulation. Likewise, caprock is
strength of the overburden, and the salt punches through to rare to absent in deepwater environments. One explanation
the surface (e.g., Figs. 48, 52, 69). It then grows as a is that the surrounding water is already very saline; another
passive diapir at or near the ground surface or sea floor is that condensed muds covering diapirs inhibit the
(Figs. 89, 90). It is generally very difficult to determine transport away of any dissolved salt (Fletcher et al., 1995).
when a salt body breaks through its conformable Having said that, there are brine pools known in the
overburden and becomes a true diapir (Fig. 91). In any deepwater Gulf of Mexico, so at least some dissolution is
case, most tall diapirs in every salt basin spend the bulk of occurring, but this is where the salt is exposed at the sea
their history growing passively, with or without later floor (Pilcher & Blumstein, 2007).
reactivation during extension, contraction, or strike-slip Another known feature of diapir margins is the so-called
movement. In the following sections, we examine passive shale sheath that is found as a flanking skirt around the
diapirs more closely. deeper portions of some diapirs (Fig. 94). This typically
consists of a thin zone of overpressured shale that is older
than onlapping, more shallow-water facies. It was
Diapir margins
traditionally interpreted as fault gouge formed as the diapir
We have already discussed the internal deformation of punched through the overburden, but we now recognize it
diapirs, but what do we see right at the edges? The tops of as a remnant of the condensed section found on top of
many diapirs have a zone of caprock (Fig. 92), typically many salt bodies. This overburden is condensed and mud-
consisting of anhydrite, limestone, igneous rocks, clastics, rich because it forms on the bathymetric highs above
and minerals such as pyrite and barite. This is usually diapirs, and ends up as shale sheath as the diapir flank
interpreted as the insoluble residue after the halite has been collapses during minibasin growth (Fig. 95).
removed by dissolution. A plot of caprock thickness (which
gets up to 300 m in South Louisiana) versus diapir burial
depth shows that caprock is best developed in the
Figure 89. Passive diapir from the northern GoM (Rowan 1995).
Figure 90. Restoration of a passive diapir showing how it stays at the sea floor as
flanking minibasins subside into the source salt layer (Worrall and Snelson, 1989).
Vertical simple shear Flexural slip

  Early diapir   Conformable overburden

Figure 91. GoM diapir (Giles & Rowan 2011) can be restored in different ways,
leading to different interpretations of when the diapir formed (Rowan & Ratliff 2010).
Figure 92. Sampling of onshore Louisiana salt diapirs with caprock development (Fails et al., 1995).
Figure 93. Thickness of caprock as a function of the present burial depth of diapirs; diapirs deeper than
about 5,000’ have effectively no caprock because of a lack of circulating meteoric water (Murray, 1966).
Figure 94. Geometry of shale sheaths around some Louisiana onshore and offshore diapirs
(Fails et al., 1995). The sheaths typically consist of overpressured, mud-dominated section
that is older than onlapping strata but is stratigraphically in place.
Figure 95. Explanation of shale sheath as simply representing preserved portions of the
original condensed section that was deposited on top of the salt body when it was at the
sea floor (Worrall and Snelson, 1989).

Potrebbero piacerti anche