Download as pdf or txt
Download as pdf or txt
You are on page 1of 43

Chapter

Salt Stocks and Salt Walls

5
5.1 Introduction The key feature of a diapir is that it is discordant to its
Salt stocks and salt walls are the most classic of salt structures. overburden: its upper contacts cut across bedding in the over-
They rise as subterranean mountains from the basin floor and burden (Figure 5.1). In maps and depth slices, diapiric salt has
are remarkable for their size: some stocks are more than 10 km an anomalous stratigraphic position: the salt is in discordant
tall, and some walls are more than 100 km long. Their abun- contact with strata younger than the units that stratigraphically
dance is equally impressive: the Precaspian basin alone con- directly overlie autochthonous salt.
tains about 1,800 named stocks and walls. Despite their being The term salt dome (Harris and Veatch 1899) is a loose
prominent in salt-tectonic literature, many aspects of their equivalent of salt diapir but is not synonymous for two
growth were misunderstood for decades. reasons. First, the core of a salt dome can be a diapir or a
These structures are diapirs, which are ductile masses of pillow. Second, a structural dome has a rounded planform,
rock or sediment that have pierced or appear to have pierced whereas a salt diapir can be rounded or highly elongated. Thus
their overburden. The term diapir was first applied to Miocene the term salt dome is both broader and narrower than salt
salt in the Romanian Carpathians (Mrazec 1907), but it has diapir. Although used only rarely in this book, salt dome has
been expanded to include ice, soft sediments (mud, sand, and widespread currency and can be useful. For example, although
peat), soft rocks (gypsum, coal, shale, and limestone), and most salt diapirs are buried, some form low islands or hills,
igneous or metamorphic complexes (granitoids, serpentinite, and this topographic relief is emphasized by dome. The ambi-
gneiss, and migmatite). Because of this broad range of geology, guity of salt dome is useful where the degree of discordance
definitions of diapirs vary widely. This book is concerned only of the salt core is unknown. Because dome rolls easily off a
with evaporite diapirs. layperson’s tongue, it is commonly part of an official name.

Figure 5.1. All diapirs, whether they are salt


a stocks, salt walls, or salt sheets have discordant
contacts against their overburden.
Ridge

Overb
urden
cutoff Discordant
Salt wall contact Overburden
c cutoff

Diapir
b
Dome

Overb
urden Source layer
cutoff
Salt stock

76
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.2 Reactive Diapirism

a Reactive piercement b Active piercement 5.2 Reactive Diapirism


5.2.1 Mechanism of Reactive Diapirism
Regional extension necks the overburden, creating grabens and
Salt half-grabens (Section 10.2). This inhomogeneous thinning
creates a differential load and weakens the overburden by
fracturing and faulting. Salt walls pressurized by overburden
c Thrust piercement d Erosional “piercement” load on adjoining areas begin to rise below the thinned over-
burden. Salt fills the space created by extensional thinning
and separation of overlying fault blocks. In the models in
Figure 5.3, all the layers were initially tabular and horizontal.
Regardless of density, this initial configuration is stable
because the gravitational load is even. During regional exten-
e Ductile “piercement” f Passive “piercement” sion, however, a graben forms. As the graben sinks into the
salt, it is increasingly resisted by two forces: (1) increasing fluid
pressure, and (2) increasing surface area of the graben keel,
which increases the pressure forces supporting the keel. Even-
tually the graben keel sinks to an equilibrium depth, and its
bounding faults cease to slip. Continued extension slices the
Figure 5.2. Six types of diapiric rise. After Hudec and Jackson (2007). graben into smaller pieces as new faults form nearer the graben
axis. Because each new piece sliced from the original graben is
increasingly smaller, each new piece floats with a keel depth
Salt stocks are pluglike bodies (hence often called salt that becomes shallower toward the graben axis. This configur-
plugs) having equant planforms, whereas salt walls are more ation is somewhat analogous to Airy isostasy, in which blocks
elongate. Usage of stocks and walls is gradational. Hudec and of equal density float at different levels depending on their
Jackson (2011) suggested that stocks have axial ratios less than thickness (Bowie 1927). In this way, as faults slice up the
2 and that walls have axial ratios of 2 or more, but any value graben, the underlying salt rises as a triangular diapir. The
proposed is arbitrary and disputable. diapir does not rise by forceful intrusion. Instead, the pressur-
The genesis of salt diapirs is much more complex and ized source layer supports the partial weight of each fault
variable than that of salt pillows. For buried salt to be diapiri- block. The rest of the weight of each fault block is supported
cally emplaced, the overburden occupying the intruded space by its footwall.
must be removed or displaced (Figure 5.2). Three modes of This type of salt upwelling is termed reactive diapirism
diapirism involve true piercement into brittle overburden. because salt rises in reaction to extension (Vendeville and
Regional extension may stretch overburden, which makes Jackson 1992a). In physical models, reactive diapirs stop rising
room for a reactive diapir to rise between fault blocks whenever regional extension stops; they rise again whenever
(Figure 5.2(a)). Flaps of overburden may tilt outward as the extension recommences. Reactive diapirism explains why in
diapir rises by active diapirism (Figure 5.2(b)). Salt may be many basins around the world, salt diapirs began to rise when
emplaced into its overburden in the hanging wall of a thrust the salt layer first extended (Jackson and Vendeville 1994). It
fault (Figure 5.2(c)). may also explain why many cratonic basins never form salt
Three other modes of diapirism have apparent rather than diapirs despite having thick salt and denser overburden. Sedi-
actual piercement of overburden. The roof may be removed by mentary loading was too uniform, and there was no regional
erosion, exposing the salt, which is free to spread discordantly extension to differentially load the salt. Extension controls the
across the unconformity (Figure 5.2(d)). A diapir having a overall shape of reactive diapirs; they are salt walls, not stocks.
weak overburden may thin its roof (Figure 5.2(e)). For Reactive diapirism is complicated by the effects of density
example, diapirs in the Great Kavir, Iran, have an overburden and synkinematic sedimentation. In the simple models of
of mudrocks mixed with halite and gypsum (Jackson et al. Figure 5.3, the overburden was denser than the salt. The same
1990). In some cases this thinned roof is then actively pierced model without a density inversion yielded a similar reactive
(Section 5.3). The third way in which diapirs apparently diapir. The diapir rose reactively without a density inversion
pierce their overburden is passive diapirism, or downbuilding for the same reason that it rose with a density inversion.
(Figure 5.2(f)). Where a passive diapir emerges, it remains at The main differential load is the density contrast between the
or near the surface while surrounding sediments accumulate. overburden and the air-filled (or water-filled) graben. The
The base of the downward-growing diapir sinks – hence the density difference between overburden and salt is much less.
term downbuilding – as the basin subsides. The only effect of an overburden less dense than the salt is to
These diapiric mechanisms are not mutually exclusive, and limit the rise of the dense reactive diapir. Salt having a density
most salt diapirs grow by mixed processes (Section 5.5). greater than the average density of its overburden can emerge

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
77
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

In theory, reactive rise can be prevented altogether if the


a 2 1
Time (h)
0.0 graben is kept completely filled with sediment. In practice, this
case is rare for two reasons. First, most grabens subside fast
enough to maintain some surface relief. Second, graben-fill
sediments are usually less compacted, and thus less dense, than
surrounding sediments. Consequently, a load gradient is pre-
sent to drive salt toward the graben even if the graben is kept
filled with synkinematic sediment.
Salt

b 2 1 2.0
5.2.2 Structural Style of Reactive Diapirs
A reactive diapir formed during synkinematic deposition is
similar to one formed from prekinematic overburden
(Figure 5.4). Both diapirs are triangular, have a sharp apex
overlain by a graben, and are flanked by a fan of inward-
dipping normal faults that end at the salt contact. But reactive
diapirs rising through synkinematic overburdens have add-
c 2 1 3.0
itional features. The inward-dipping normal faults are growth
faults, so that beds thicken across the faults inward toward the
3
salt. In contrast, underlying prekinematic beds maintain con-
stant stratigraphic thickness, although they may thin exten-
sionally, depending on their lithology.
The width of a reactive diapir’s base records the amount of
extension. At the base, extension is accommodated entirely by
flow of salt. Midway up the diapir, extension is partly by flow
of salt and partly by faulting. Above the diapir, extension is
d 2 1 5.0 entirely by faulting. With entirely prekinematic overburden,
3
5 6 4 extension is equal at these three levels, despite taking different
7 8
forms. With synkinematic overburden, extension declines
upward.
All the reactive diapirs discussed so far have symmetric
profiles like an isosceles triangle. However, reactive diapirs
Amount of extension and also include highly asymmetric diapirs. Strongly asymmetric
depth of compensation
reactive diapirs, known as salt rollers (Bally 1981) have
e 2 Rotated flaps from
active stage 7
1 6.5 asymmetric profiles like scalene triangles. Their long sloping
side is concordant to the overburden, and their short
3
6 5 4
sloping side is a normal fault (Figure 5.5). The lower part
of the fault juxtaposes salt against overburden, and its upper
part juxtaposes overburden against overburden. Salt rollers
form whenever one crestal-fault orientation is favored over
another, as might happen if stress axes were rotated from
the vertical. Stress axes rotate above dipping detachments,
0 4 cm so rollers are particularly common above seaward-dipping
salt detachments near the landward ends of divergent
Figure 5.3. A reactive diapir rises because of thin-skinned regional extension. margins.
These cross sections of an evolving reactive wall are from otherwise
identical physical models extended by 0 cm, 2 cm, 3 cm, 5 cm, and 6.5 cm, Figures 5.6 and 5.7 show the key diagnostic features of
respectively. All the overburden is prekinematic. In the final stage (e), active reactive diapiric walls: a triangular shape, pointed crest, and
diapirism had begun. Faults are numbered in order of formation. After Jackson a crestal graben or half-graben, which dominate the map view.
and Vendeville (1994).
An axial trench in the floor of the graben is bounded by fault
terraces that step upward and outward to the upwardly flexed
margins of the graben. Faults are subparallel but may
in the graben floor, but the diapir’s crest can never rise as high anastomose.
as the top of the overburden outside the graben. Most diapirs worldwide began growing during extension
The driving force for reactive diapirism is greatest if the (Jackson and Vendeville 1994), which suggests that regional
reactive graben above the diapir is filled with air or water. Rise extension triggers salt diapirism. If salt in a reactive diapir
rates are less if synkinematic sediments partially fill the graben. eventually reaches the surface to become a passive diapir

78
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.3 Active Diapirism

Faults become Figure 5.4. Synkinematic layers thicken across


younger inward growth faults above a reactive diapir. Extension

Synkinematic
dismembered one marker block so that it was split

layers
each side of the reactive diapir, illustrating the
magnitude of extension. After a physical model by
Bruno Vendeville.

Prekinematic
layers
Segmented marker Salt
0 5 cm Diapir widening offsets Marker block
marker block

(Section 5.4), the faults on the flanks become inactive, as


a further extension is accommodated by diapir widening. The
only evidence of the early reactive phase is a swarm of normal
faults against the deep flanks of the diapir, reflecting its early
extensional history.

5.3 Active Diapirism


5.3.1 Mechanisms of Active Rise
Active rise occurs whenever a salt structure arches, uplifts, or
shoulders aside its overburden. We distinguish between active
rise, in which the roof is uplifted but not necessarily pierced,
and active diapirism, in which the roof is pierced during active
b rise. Active rise can be driven either by halokinetic processes or
by shortening.
Increasing asymmetry

5.3.1.1 Halokinetic Active Rise


Halokinetic active diapirism is driven by overburden load on
the adjoining source layer, which creates an upward pressure
of diapiric salt on its roof. This load can be produced by
buoyancy alone if the average density of the overburden is
greater than that of the salt, or by surface topography
(Figure 5.8). The forces resisting active rise are the weight
and strength of the overburden and the viscosity of the salt.
Resisting forces in the overburden increase as it thickens. The
c roof thickness at which forces are balanced is called the thresh-
old thickness (Figure 5.9). The threshold thickness comes into
play at both the start and the end of halokinetic active rise.
Halokinetic active rise begins owing to increase in driving
force, decrease in resisting force, or both. Driving forces stead-
ily increase as sedimentation increases the pressure on the
source layer. Resisting forces typically decrease owing to
faulting or thinning of the overburden. Thus, reactive diapir-
ism commonly leads to active diapirism (Figure 5.3) because
the roof is thinned and weakened by extensional faulting.
Concordant flank
Overburden can also thin because of small-scale deformation
Discordant flank or erosion. Erosion triggers active rise when erosion bevels the
0 3 cm crest of an anticline or, less commonly, when a canyon cuts
into flat-lying strata.
Figure 5.5. During thin-skinned extension, reactive diapirs can be highly Halokinetic active rise is influenced most strongly by sedi-
asymmetric on a seaward topographic slope. Seaward-dipping faults are
favored at the expense of the conjugate landward-dipping faults, resulting ment density, roof strength and anisotropy, and salt geometry
in a salt roller. Physical model by Lin (1992). (Schultz-Ela et al. 1993). Active rise is promoted by increasing

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
79
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Stairstep faulting Figure 5.6. Diagnostic criteria of reactive


Parallel above buried wall walls in a schematic block diagram. After Jackson
faults et al. (1994).
Regional
extension
pir
dia
rd
owa
gt
Marginal e nin
thick
flexure
allow lts n
Sh l fau n sio
or ma exte
Regional
o fn by
an g
F nin
hin
Pointed
e pt
crest of De
wall

Figure 5.7. Three diapirs show the hallmarks of


reactive diapirism: a triangular salt body overlain by a
fan of inward-dipping normal faults in a
Fan of normal
crestal graben. The diapirs range from symmetric
5.0 (center), to moderately asymmetric (right side),
faults forms
crestal graben to strongly asymmetric (left side). Campos basin,
offshore Brazil. Seismic data courtesy of PGS.
Two-way time (s)

5.5

6.0
Salt weld
Pointed crest of
0 3 km reactive diapir

Figure 5.8. Halokinetic active diapirism can be


a driven by two types of gravitational loading.
(a) Average density of the overburden is greater
ρo < ρs than that of salt. (b) Topographic variation
ρs creating a wedge of water (or air); a diapir can
ρo > ρs rise actively even though the overburden is less
dense than salt.
P2 P2 < P1 P1
b Water wedge ρw Overburden
wedge
ρo ρw < ρo < ρs
ρs
P2 P2 < P1 P1

the average density and thickness of sediments overlying the In contrast, an anisotropic roof (composed of multilayers
source layer, which increases the driving force, and by decreas- separated by slip surfaces) bends easily and forms an anticline
ing thickness and density of sediments above the salt structure, when arched (Figure 5.10(b)). At the outer edge of the anti-
which decreases the resisting force. cline, strata flex but do not fault. Interlayer slip in the roof of
Active rise is also promoted by weak roofs, especially an active diapir has the following effects relative to an isotropic
faulted or unconsolidated sediment. An isotropic roof (com- roof. The diapir rises higher and faster. The roof flexes
posed of massive sediments) does not bend easily. Instead, smoothly as a fold. The diapir crest (top of salt) is rounded
active rise divides the roof into tilted flaps. The hinges of these rather than pointed. Flaps above the shoulders of the diapir dip
flaps are reverse faults that steepen downward (Figure 5.10(a)). more steeply.

80
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.3 Active Diapirism

Figure 5.9. Roof thickness controls the ability of


a Halokinetic active rise impossible a diapir to rise actively by halokinesis. If the roof is
too thick (above threshold thickness), the diapir is
Roof thickness
inactive (a). If the roof is thin enough (below
ρo Threshold thickness threshold thickness) (b), the diapir rises actively
Density inversion: ρo > ρs
until its crest rises high enough to pressurize the
Po = pressure below overburden base of the diapir (c).
ρo Salt Pd = pressure below diapir and roof

ρs

b Start of active rise c End of active rise Arched roof

Roof thickness < threshold


Extra thickness
of salt

Po Pd Po Pd

Unstable: Pd < Po Stable: Pd = Po

Wider diapirs are more likely to rise actively than narrow


a Isotropic roof ones, other factors being equal. Cross-sectional shape also
Contour levels are 4 × higher than below 304 000 yr matters: the likelihood of halokinetic active rise is greatest for
rectangular diapirs, and rise becomes progressively less likely
for round-cornered rectangles, semicircles, and triangles.
Halokinetic active rise continues until either the roof
Angular thickens to reach its threshold thickness or the diapir breaks
diapir crest through to the surface and becomes passive. Breakthrough
Salt
typically occurs as a result of roof arching and thinning by
extension or erosion. When the roof is too thin and weak to
resist salt pressure, the remaining roof is uplifted and shoul-
b Layered roof
Steep dips above diapir shoulder 304 000 yr
dered aside to form outward-dipping flaps as the salt emerges
at the surface (Figure 5.11, Section 5.4).
If the overburden is irregular, a diapir can rise actively by
halokinesis even where its roof is less dense. Diapirism can
Flat-topped occur where differential loading is imposed by relief in either
diapir crest the top of the salt (Figure 5.8(a)) or the topography (Figure 5.8
(b)). If the overburden is a flat-topped wedge, its density is
Salt greatest at the thick end of the wedge (Figure 5.8(a)), where
deep sediments have compacted more than at the thin end of
c Full model: isotropic roof the wedge. Even though the diapir is denser than its roof, the
diapir rises actively because the salt is pressurized by the
Passive lines 304 000 yr
distant thick, denser end of the wedge. If the topography
slopes, none of the overburden need be denser than salt for
Higher density an active diapir to rise (Figure 5.8(b)). This is because an
overburden is invariably denser than the same thickness of
Lower density
water or air. This differential loading creates a pressure-head
Salt flow to feed growing diapir
difference that drives salt from below the thick end of the
Fixed
0 5 km side wall overburden wedge into the active diapir.
Figure 5.10. The style of active diapirism is affected by whether its 5.3.1.2 Contractional Active Rise
sedimentary roof is layered, as shown by originally flat-topped diapiric walls.
(a) Crest of diapir arching an isotropic, or massive, roof to form two flaps Regional shortening can initiate contractional active rise of
hinging on downward-steepening thrust faults that propagated up from the salt in two ways. First, shortening adds a convergent displace-
diapir’s shoulders. (b) Crest of diapir arching an anisotropic, or layered, roof after
the same duration; white lines show thin interbeds that allow bedding-parallel ment load to the gravitational load. The sides of the diapir
slip, which promotes folding rather than faulting. (c) Full view of model in (a). converge, and the diapir narrows. Narrowing displaces the
Contours show intensity of plastic strain, where white is the highest strain.
Sediment–salt density ratio is 1.09. Finite-element models by Dan Schultz-Ela.

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
81
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Pierced
a b c crestal graben
Stage 1 Stage 2 Stage 3 Rotated
0 2 cm flap
Crestal graben
Emergent
diapir

Active diapir Active diapir

Figure 5.11. Three stages of active rise of an originally flat-topped salt wall, culminating in complete piercement of the highly extended roof. Traced from
physical models by Bruno Vendeville.

upper diapiric salt upward, where it may arch the roof faster
a Isotropic roof Crestal graben and higher than it would rise by halokinesis alone (Figure 5.12).
Convergence can also displace the lower diapiric salt downward,
304 000 yr
which thickens the source layer and raises the overburden
Flap thrust regionally (Figure 5.12). Salt is more likely to be forced down-
ward when the source layer is still thick because it is difficult to
force salt back into a welded source layer. Second, shortening
Salt can buckle the overburden above salt, causing it to rise up into
anticlines even if not pushed by pressurized salt (Section 4.3).
Contractional active rise is resisted by the same three
factors as is halokinetic active rise: the strength and weight of
the roof and the viscosity of the salt. The difference is that
regional shortening can create much higher stresses than can
b Layered roof Limited extension halokinetic salt rise. Thus, contractional active rise is capable
279 000 yr of lifting much thicker roofs than halokinetic active rise (Ven-
Steep limb
deville and Nilsen 1995). Shortening can thus rejuvenate a
halokinetically stable diapir having a roof too thick to be
Shortening

Flat top pushed up by buoyancy alone, or it can accelerate the rise of


Fold widens a diapir already actively growing by buoyancy.

Salt 5.3.2 Structural Style of Active Diapirs


Active rise forms two types of structures: arched roofs and
upturned collars. Arched roofs result where the salt has not
c Full model: layered roof 279 000 yr pierced to the surface, whereas upturned collars are evidence
that a formerly active diapir once broke through to the surface.
Shortening

Sediments lifted Upward salt flow


Downward salt flow 5.3.2.1 Arched Roofs
The most diagnostic feature of active rise is a roof that has been
Shortening forces salt from diapir into source layer uplifted and arched above regional. The arched roof is a drape
0 5 km fold formed by bending. The roof is typically stretched owing to
bending, which may result in thinning below seismic resolution
Figure 5.12. Regional shortening promotes active rise in ways that depend
on properties of the diapir’s roof. (a) Above an originally flat-topped wall, an
(Davison et al. 2000b) or in crestal normal faults (Figures 5.13
isotropic (massive) roof arched by forming two flaps. These flaps hinged on and 5.14). Arched roofs may also thin by erosion or slumping
downward-steepening thrust faults that propagated up from the diapir’s (Figure 5.15), causing further ambiguities in interpretation.
shoulders. The crestal graben separating the two flaps was minor compared
with an equivalent graben formed without regional shortening. (b) An
Small-scale deformation can be widespread in arched roofs.
anisotropic (layered) roof merely arched into a steep-limbed box fold having Response to deformation depends on strain intensity and
limited extension in its crest. White lines show thin interbeds that allowed lithology. Toward some North Sea diapirs, chalk thins from
bedding-parallel slip, which promoted folding rather than faulting. (c) Full view
of model in (b). Salt in the squeezed diapir was forced upward to the arched
1,500 m on the flanks to 0 to 300 m above the diapir. Here the
crest and downward back into the source layer. Contours show intensity chalk is pervasively fractured in as much as 40 percent of the
of plastic strain, with white being the highest strain; the color range for core interval, forming open and cemented tensile fractures,
plastic strain is different in each example in order to highlight particular
features. Sediment–salt density ratio was 1.09; shortening rate was 1 mm/yr.
small faults, and rarer breccias (Davison et al. 2000b). Frac-
Finite-element models by Dan Schultz-Ela. tures in other brittle rocks, like quartz sandstones, are less
interconnected than in chalk, but breccias are common.

82
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.3 Active Diapirism

Larger-scale normal faults typically cut the arched roof of 1951, 1955; Cloos 1955; Withjack and Scheiner 1982; Sims
an active diapir if sediments have lithified enough. The sim- et al. 2013).
plest crestal fault pattern is a single offset fault slicing obliquely Normal faults above salt walls tend to align parallel to the
through the diapir. More-abundant normal faults form com- long axis of the wall. These faults can be caused by regional
plex patterns, such as a keystone graben. Modeling shows that extension or by local doming. The cross-sectional pattern
a diapir’s planform strongly influences the map pattern of its of the faults varies (Figure 5.16). Most arched roofs above
crestal faults (Link 1930; Cloos 1939; Parker and McDowell natural active diapirs (both halokinetic and contractional)
are anticlines divided into two flaps by a crestal graben
(Figure 5.16(a)). In contrast, most arched roofs in contrac-
Monocline or tional sandbox models have two flaps that hinge upward on
thrust fault
Radial or synthetic reverse faults (Figure 5.16(b)). These flap faults
subparallel steepen downward and root into the sides of the diapir. Rarer
faults in nature are diapirs having two flaps hinging on synthetic
ing
h inn normal faults (Figure 5.16(c)), or asymmetric diapirs having a
T ing single flap and a master normal fault linking the crest to the
Flap ch
Regional Ar
flank of the diapir (Figure 5.16(d)).
Pointed or Normal faults tend to form a broadly radial pattern above a
rounded crest
of stock or wall pluglike stock. Radial patterns are distinctive of diapiric
doming because they do not form by regional extension.
Figure 5.13. Diagnostic features of an active diapir, the most important of Radial faults are most common in the arched roof of an
which is a roof arched above regional. After Jackson et al. (1994). actively rising stock. The throw of radial faults decreases from

Figure 5.14. The arched roof above this diapir


Crestal graben
Strata above regional shows distinctive features of a halokinetic
Topographic bulge active diapir: (1) a bulge in the seafloor in which
2
strata rise above regional, (2) crestal extension, and
(3) thinning and onlap of strata toward the crest.
Gulf of Mexico seismic data courtesy of CGG.

3
Depth (km)

4
Salt

6
V.E. × 2 0 5 km

Headscarp Exposed salt Figure 5.15. A topographically arched roof is


Shallow subject to (a) slumping and (b) erosion. Erosion
Landslide Slump Thrust Eroded
a b erosion
leads to (c) an angular unconformity around
the diapir. This unconformity may be preserved
across the top of the diapir if active rise stops,
Active Former or preserved only on the diapir flanks if active
diapir active rise continues.
diapir

Unconformity
c Deep erosion Truncated flap

Former
active
diapir

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
83
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Figure 5.16. Four styles of faulting above active


a Antithetic crestal graben
Double flaps cSynthetic Double flaps salt walls in schematic cross sections. All these
faults strike parallel to the long axis of the wall.
Arching normal After Schultz-Ela et al. (1993).
fault

Original top
of salt

b
Synthetic reverse fault
Double flaps d Single flap
Normal fault
Arching
Regional

Future
a horst b Horst
Graben
Centerline

Ring
Fu abe
gr

Transport
tu n

Radial normal faults thrust


re
Edge of s

Stock
Lower
hinge
alt

ho ture
Ed

line
rst
g

of
Fu
e

Ring
do
med graben
roof Horst
Upper hinge line along
monocline
Figure 5.17. Radial faults form above an active diapir because, as roof hinge
strata arch and tilt outward, they extend circumferentially (hoop extension).
(a) Before arching (oblique three-dimensional view); (b) after arching
(vertical three-dimensional view, colors delineate individual fault blocks);
after Yin and Groshong (2007).

Figure 5.18. Three types of faults can form in the roof of a plug-shaped
salt stock. (1) Radial normal faults form a graben star by hoop extension.
the crest to zero at the edge of the dome, where a closed hinge (2) A ring graben can form by radial extension in the upper hinge of an
outward-dipping monocline. (3) A ring thrust can form by radial shortening
line separates the dome from its surroundings (Stewart 2006; along the lower hinge of the monocline at the outward limit of doming.
Yin and Groshong 2007). Outward from the diapir’s crest, Oblique view of physical model, courtesy of Tim Dooley.
structural compartments between radial faults widen.
Why do radial faults form in arched roofs? As a salt diapir
rises actively, roof strata rotate upward and radially outward letter Y, the other pattern has three main fault sets that
(Figure 5.17). Strata tilt outward by flexural slip radial to the obliquely slice the dome into seven blocks (three pairs of
dome along the center line of each block (Yin and Groshong alternating horsts and grabens radiating from a triangular
2007). Roof particles move radially outward and diverge. This central graben) (Figure 5.19(b)). As a roof is domed, these
circumferential extension produces radial faults unless the master faults increase their mutual offset, which obscures the
sediments are too soft to break. Radial faulting intensifies as fundamental X or Y pattern. The master faults control the
doming proceeds. Radial faulting is greatest where a diapir’s pattern of younger, smaller faults. The X and Y types of radial
radius of curvature is smallest. This causes radial faults to faulting yield a wide variety of structural styles, depending on
cluster near the ends of elliptical diapirs (Withjack and Schei- where a cross section intersects a domed roof (Yin and
ner 1982; Davison et al. 2000a; Stewart 2006; Sims et al. 2013). Groshong 2007).
Radial faulting above active diapirs may begin with a Radial cross sections through a salt stock commonly inter-
graben star (Cloos 1939; Figure 5.18), or with a simpler pattern sect radial faults tangentially, which confuses interpretation in
(Yin and Groshong 2007), or with a more complex pattern several ways. First, fault traces have apparent dips that are
(Clausen et al. 2014). Like the letter X, one pattern has four much less than the actual dips of the faults. Second, fault traces
radiating faults defining four blocks (a horst, a graben, two can appear more than once in a cross section. Third, fault
half-grabens, and no central graben) (Figure 5.19(a)). Like the traces may dip toward a dome center, although fault surfaces

84
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.3 Active Diapirism

Figure 5.19. Structure


a -92.083° -92.067° -92.050° -92.033° contour maps show that
N radial faults can form
X and Y patterns in the
arched roofs of active

29.467°
-10,60
0
salt diapirs in the U.S.

-10
Gulf Coast. (a) X pattern

-10

,8 0
over Tiger Shoal diapir

,60

29.450°
(Louisiana) drawn on top

0
0
80 of T sand (after Smith
0,
-1
1988). (b) Symmetric
0 0
,6

Y pattern over Clay Neutral Extension Compression


0
-1

29.433°
Creek diapir (Texas),
00 drawn on top of the
0,4
-1 0 Wilcox Formation (after
60
0, McDowell 1951). After Yin
-1
29.417°
and Groshong (2007).
00
0 ,8
-1
-10
-10

,60
,80

0 0 1500 m
0

C.I. = 100 ft

b -96.417° -96.400° -96.383°


30.333°

-3000 Normal fault Strike-slip fault Reverse fault


-3
00
-24 0 Figure 5.20. Fault patterns above circular (upper row) and elliptical
0
-180 00
(lower row) active diapirs are controlled by different regional stress fields
indicated by arrows. After Withjack and Scheiner (1982), based on their physical and
30.317°

analytical models.
0
-240

00
-36

Diapir width strongly affects the style of compressional


-1

active diapirism (Figure 5.22, Nilsen et al. 1995). For identical


80

0
-3

00

30.300°
0
00

-3

N
amounts of shortening, the area of diapiric salt displaced
0

-2400
0 1500 m inward by compression is the same regardless of diapir width,
00
-36 C.I. = 600 ft assuming plane strain. Salt displaced inward in turn displaces
Major horst Major half-graben the same area of salt upward beneath the arching roof. How-
Major graben Central graben ever, salt displaced upward above a wide diapir causes little
Master normal fault Minor normal fault arching because the displaced salt spreads beneath a wide
roof. Conversely, salt displaced upward above a narrow diapir
causes severe arching because the displaced salt is concentrated
actually dip away from the dome center. Fourth, it is difficult beneath a narrow roof.
to differentiate crestal faults from surrounding radial faults in
a cross section. 5.3.2.2 Upturned Collars
Cloos (1968) recognized that far-field (regional) stresses Stratal upturns have long been interpreted from wells on the
can affect faults induced by doming. A fault can be subdued flanks of salt diapirs, even though many of these collars are too
or enhanced, depending on its orientation to far-field stresses. narrow to be seismically resolved (Hanna 1953; Atwater and
Withjack and Scheiner (1982) modeled how regional extension Forman 1959). These upturned units typically comprise brec-
or regional shortening affected the pattern of faults above ciated and sheared deep-marine shale and follow the diapir
gentle active domes (Figure 5.20). The radial pattern of normal margin upward in discordant contact with much younger
faults is suppressed where far-field stresses dominate local strata (Figure 5.23). Commonly known as shale sheath, these
doming stresses. Instead of radiating, normal faults strike collars were for decades attributed to drag-induced shear
perpendicular to the regional extension direction regardless against a diapir rising with respect to surrounding sediments
of the shape of the underlying diapir (Sims et al. 2013). (Bornhauser 1969). However, finite-element models suggest
Arched roofs can occur in both halokinetic and contrac- that slip along the diapir contact affects only extremely weak
tional active diapirs. Each setting has a different combination overburdens having high overpressures and multiple ductile
of forces, which can influence the structural style (Figure 5.21). interlayers (Schultz-Ela 2003).
Some features of arched roofs are diagnostic of contractional Among the first to question the drag dogma were Johnson
active rise, but none clearly indicate halokinetic active rise. and Bredeson (1971) and Worrall and Snelson (1989). They
Most arguments for halokinetic active rise thus focus more concluded that folding, thickness changes, and unconformities
on the absence of contractional features than on the presence adjoining Louisiana diapirs reflected near-surface deformation
of halokinetic features. of rising diapirs. More recent work (Rowan et al. 2003;

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
85
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

a Postkinematic crestal a Before shortening


thinning, no crestal graben Extensional
kink band

Wide Narrow
Arch relief < 500 m diapir diapir
Salt

Tall, narrow
arch
b After shortening Low, wide
arch
Weld
b
Extensional Crestal
bedding-plane Salt displaced
graben Synkinematic upward
slip
Truncation crestal thinning
Roof thinner than
Onlap threshold thickness
Arch relief > 500 m Salt displaced inward
Convex-upward
salt crest Figure 5.22. The same amount of regional shortening affects wide and
Narrow aureole of narrow diapiric walls differently. Diapirs that are initially tall and narrow
upturned strata rise higher and faster than diapirs that are initially wide. The same principle
applies to wide and narrow stocks, but areas of displaced salt do not balance
as they do in walls. Concept by Vendeville and Nilsen (1995).

Weld North South Figure 5.23. Upturned


Not the surface flaps, sometimes called
Contractional
c Fold may
be boxlike Erosional kink band Contractional V
shale sheaths, can
separate some diapirs
truncation bedding-plane slip from discordant younger
Onlap
strata. This Oligocene flap
Postkinematic roof is onlapped by upper
Age of shortening Oligocene and Miocene
Thick prekinematic strata at the contact of
roof W White Castle diapir
(Louisiana). After Johnson
and Bredeson (1971).
Strongly
Sha

V
convex-
Passive rise
le s

upward Z
salt crest
hea

AA
th ?

Weld BB
Wide aureole of Salt OL1
7.2
?
Frio

upturned strata OL
17.
5
Figure 5.21. Diagnostic features of arching produced by (a) differential O O OL1 OL
OL2

OL

L2 L2 9 18
compaction, (b) halokinesis, and (c) active rise. 1 0
22
3

Schultz-Ela 2003) has determined that these salt-flank upturns the thickness of the sediment package involved in the flap.
are a type of sediment flap, formed by active diapirism. A thick wedge of strata produces a wider zone of deformation
When an active diapir breaks through to the surface, its and broader topographic halo than a thin wedge folded to the
broken roof is typically shouldered aside to be eroded or same degree of rotation (Rowan et al. 2003).
preserved as steeply dipping flaps on the flanks of the salt Flap sediments may be intensely deformed by diapirism for
(Figures 5.3(e) and 5.15). These flaps are the evidence of the three reasons. First, the sediments may be weakly consolidated,
final event of a former active phase that ended in an exposed, so rotating them to vertical or overturned dips should involve
passively rising diapir. Flaps typically dip parallel to the salt internal slumping, shearing, and loss of cohesion. Flaps would
face. Flaps can be vertical if breakthrough produced a vertically probably be completely disaggregated if not for being buttressed
rising passive diapir, or even overturned if salt broke out as an against younger sediments deposited while the flap rotated.
extruding salt sheet. The wavelength of flap folds depends on Second, outward movement of flaps shouldered aside by salt

86
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.3 Active Diapirism

2.8 km
Perched flap

Perched flap Salt


37 Myr

Salt diapir
B
me asal
fla ga-
p
Megaflap

Figure 5.24. Wedges of sediment that originally onlapped an emergent 85 Myr


salt diapir become deformed into flaps as the diapir continues to grow.
These flaps form on two scales: (1) megaflaps are kilometers wide and Same volumes, different areas
rest directly on the salt; (2) perched flaps are hundreds of meters wide and Figure 5.25. During active diapirism, a roof of unlithified strata can be
rest on older overburden strata. thinned by extension as the wedge is rotated by rising salt. The cross-sectional
area of the purple polygon shrank over time even though its volume remained
constant. Area was lost because of hoop extension as the diapir expanded
stocks circumferentially stretches the flaps, which thin by plastic outward. The roof gradually rotated to form an upturned basal megaflap.
extension. Third, flaps may be overpressured if composed of Finite-element ELFEN model by Maria Nikolinakou; the overburden was a
mudrocks and connected to deep, overpressured fluids. poro-elastoplastic microcrystalline cellulose (MCC) rheology, and the salt
was viscoplastic.
Brecciated shale around and within flaps has been termed
salt-dome breccia (Kerr and Kopp 1958). Cores show pervasive
fracturing within shale on the flanks of some salt diapirs breakthrough following a brief phase of burial during down-
(Hanna 1953). The brecciated shale is reportedly overpres- building (Section 5.4). Perched flaps formed during downbuild-
sured, but this damage zone can allow water to rise through it. ing are part of larger halokinetic sequences (Section 5.4.2.1).
Three types of upturns form in different settings. The first Finite-element modeling simulates the effects of changing
is the basal megaflap (Giles and Rowan 2012, Graham et al. rates of salt flux and sedimentation. Perched flaps are likely
2012). These thick sequences of sediment are upturned at the where aggradation rates vary greatly and depositional hiatuses
base of a diapir and thin upward along the diapir flank (Fig- allow time for folding before the wedge becomes too thick and
ures 5.24, 5.25, 5.26, and 5.27). The basal megaflap can pinch strong to bend (Schultz-Ela 2003). Fast aggradation or slow
out or be eroded partway up the diapir flank, but in other salt flux causes sediment to encroach and onlap farther across
instances the megaflap can continue all the way up to the crest the salt, which increases the potential for folding and uplift of a
of the diapir, and even overturn beneath the breakout of a salt roof but can eventually bury a diapir and stop its rise. Con-
sheet. Basal megaflaps form as the source layer inflates to form versely, slow aggradation or fast salt flux causes the diapir to
a pillow or anticline, which is then broken through by active bulge outward along vertical to inward-dipping walls, which
diapirism. Sediments in a flap typically thin stratigraphically shortens or inhibits wedges and hence perched flaps.
and by strain as the flap rotates. Third, some flaps form as the limbs of an injection fold,
Basal megaflaps are most likely to originate where a salt which is an active diapir formed in the crests of contractional
contact dips moderately, allowing a wedge of strata to onlap anticlines (Beloussov 1959). As an anticline amplifies, its mobile
across the diapir crest and be folded upward by the rising salt core is tectonically pressurized by displacement loading imposed
(Schultz-Ela 2003) (Figure 5.28). The style of folding depends by the tightening fold limbs. If the overburden is thin enough,
on the depth, rock strength, anisotropy of the overburden, the salt core breaks out actively to the surface. Diapirs formed
diapir shape and tilt, and molding ratio (balance of by fold injection have an ancient pedigree. Such diapirs were
aggradation and salt-rise rates during folding). modeled by Lohest (1921) and Torrey and Fralich (1926), and
The second type of flap is perched flaps that lie partway up injection folding was regarded by Stille (1925) as the main way
a diapir flank (Figure 5.24). Perched flaps are typically much of forming a salt stock. The idea was neglected for many
smaller and thinner than basal megaflaps and form when a decades, but injection folding is being recognized as an
rising diapir breaks through a thinned roof. The perched important type of diapir in fold belts. A diapir formed in this
flap can mark when active diapirism began after lengthy react- way is typically far smaller than the injection fold from which
ive growth (Figure 5.3). Alternatively, the flap represents it emerged (Figure 5.29).

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
87
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

a Seafloor
b
1

2
Diapir
Depth (km)

3
Basal
megaflap
Salt
4

6
0 3 km
Halokinetic unconformity

c V.E. × 1.5

Emirhan Fm.

30.71°N
50
70 Karacaören Fm.

70

30.70°N
30
Karayün Fm.

N
40 30.69°N

0 1 km
L. Oligocene evaporites
37.30°E © 2015 Basarsoft; © 2015 Google; Image © 2015 DigitalGlobe 37.35°E

Figure 5.26. Examples of upturned flaps flanking originally active diapirs. (a) Perched flaps separated by halokinetic unconformities, northern Gulf of Mexico;
seismic image courtesy of CGG, (b) Basal megaflap above allochthonous salt, northern Gulf of Mexico; seismic image courtesy of TGS. (c) Basal megaflaps, especially well
developed on the northeast side of Karayün minibasin (Sivas basin, central Turkey); satellite image from Google Earth shows an oblique erosional cross section
through the minibasin, which is flanked by walls of lower Oligocene evaporites; geologic boundaries after Ringenbach et al. (2013).

88
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.3 Active Diapirism

No
rth a Maximum onlap on
diapir crest Reference point
6

Salt

b Reference datum

n
t tio
es lu
cr sso
12

i
td
la
F
900
9
1500
uth
So 3000 6
5000 Source
t 7000 layer
res
llowc 10,000
of pi Salt
sion pedestal
Ero
12,000
c Stretched and thinned older layers

15,000 18

Contours in feet 15
below sea level
12
Claiborne–Wilcox–Midway Groups
Navarro–Taylor Groups Me 9
ga
fla
Austin Chalk p 6
Upper Eagle Ford Group
Lower Eagle Ford–Woodbine Groups
Washita–Fredericksburg Groups
0 1 km
Paluxy Formation
Figure 5.28. Evolution of a basal megaflap that becomes onlapped by
Massive anhydrite younger strata in the contact zone of a passive diapir. Numbers refer to
synkinematic strata. Finite-element forward model by Dan Schultz-Ela.
Hosston–Cotton Valley Formations
Gilmer–Norphlet Formations
Louann Salt (Figure 5.31(a)). In contrast, the footwall of the diapir is the
subvertical limb of a deep minibasin. The squeezed diapir may
Figure 5.27. Basal megaflaps on the flanks of Hainesville diapir (East Texas)
were uplifted and eroded, producing an angular unconformity on the flanks of
extrude allochthonous salt (Figure 5.31(b)).
the dome. After Jackson and Seni (1984). Where preserved, flaps are diagnostic of active diapirism.
However, they are easily destroyed by erosion or slumping
away as they rise and steepen. Many past phases of active
diapirism may have gone unrecognized because the flaps were
Physical modeling elucidates how injection folds might obliterated.
evolve (Figure 5.30). The crest of a buckling anticline extends
by normal faulting, forming a weak zone that is eventually
actively pierced by a diapir of pressurized salt. If the anticline is
5.3.3 Apparent Active Rise: Arching of a
symmetric, the diapir pierces vertically through its hinge zone. Diapir’s Roof by Compaction
If the anticline is asymmetric, the steeper limb of the fold A distinctive feature of an active diapir is a roof arched above
presses into the source layer as a sagging minibasin. The the surroundings. However, roofs can also be arched purely by
gentler limb of the fold eventually overthrusts the sagging compaction even above an inactive diapir. Two aspects of
minibasin and shears the intervening diapir toward the over- sediment compaction contribute to this arching. Drape com-
ridden minibasin. The result is a shallow hanging wall com- paction arches strata where sediments of varying thickness
posed of deep strata raised by overthrusting the diapir compact by an equal percentage (Billingsley 1982). Thus,

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
89
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Contractional anticline Figure 5.29. The core


of this salt anticline was
pinched shut by regional
a Seaward
Slight
Landward
Inclined slab
shortening, during which injection of older strata
the salt core broke out as Asymmetric
an injection fold, forming growth syncline Diapiric salt
at crest
a small diapir at its crest. 3
Campos basin, Brazil.
Seismic data courtesy of PGS.

Small diapir

Two-way time (s)


Welded neck of anticline
4
Concordant
limbs of
injection
fold
Deep
Anticline limb minibasin
5

Base of salt
Pedestal of
salt anticline
0 5 km
Salt weld

b Seaward Landward
Inclined older strata in hanging wall
Salt
extrusion Eroded

Prekinematic Synkinematic
Figure 5.30.
a Prekinematic overburden Asymmetric injection
folds form as diapiric salt
thrusts over a deepening Concordant
Salt limbs of
asymmetric growth
syncline in its footwall. injection
b From a physical model fold
by Brun and Fort (2004,
figure 12). Welded Asymmetric
injection fold growth syncline 0 3 cm

Figure 5.31. Diapirs interpreted to have formed as injection folds during


c moderate regional shortening. (a) Geoseismic section of a weak injection fold
in the Kwanza basin (offshore Angola); the crest of an otherwise concordant
salt anticline is locally diapiric. (b) A strong injection fold expelled a salt
extrusion as the diapir was overthrust by its hanging wall in a physical model.
After Brun and Fort (2004, figure 11).

d Injection
fold
Growth syncline

Shortening
does not compact during deep burial and even expands slightly
Salt
overthrust
as it experiences heating. In contrast, the overburden next to
the diapir subsides by compaction, once again arching sedi-
ments above the diapir (Figure 5.32(d), (e)). The edge of the
arched roof overlies the edge of the diapir.
Compactional arching is widespread and marked. Because
during compaction, thinner sediments in the diapir roof sub- of the additive effects of drape compaction and differential
side less than adjacent thicker sediments, thereby arching the compaction, most diapirs have arched roofs, irrespective of
roof (Figure 5.32(a), (b)). Folding by drape compaction whether or not they actively rose while the roof accumulated.
decreases upward, resulting in an apparent growth anticline The amount of arching increases with the height of the
even in postkinematic strata above a deep static anticline diapir and decreases with the degree of undercompaction
(Figure 5.32(c)). In contrast, differential compaction results (Figure 5.33(a)). These calculations assume that every incre-
from the juxtaposition of incompressible crystalline salt in a ment of sedimentation is followed by an increment of com-
diapir against compacting sediments next to the diapir. Salt paction. The amount of arching also increases with the

90
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.3 Active Diapirism

Drape compaction Differential compaction Figure 5.32. Drape compaction and differential
compaction overlap, but the terms make a useful
a Before compaction
Datum
d Before compaction
Datum
distinction between (a)–(c) compacting layers
varying laterally in thickness (drape compaction)
and (d), (e) adjacent rocks compacting by different
h1 h1 amounts (differential compaction). (a)–(c) After
Billingsley (1982).
h2 h2
Salt
Deep static anticline
Φ = 0% h2
Onlapping postkinematic interval

b After 30% compaction


Compactional drape
Φuc = 45%
Datum

0.7h1
e After 30% compaction
Datum
0.7h2 0.7h2
Deep static anticline 0.7h1 Edge of arch

c Apparent growth anticline


Weakly compacted
Salt
increases downward
Drape compaction

Φ = 0%
0.7h2

Φc = 21%
Deep static anticline
Weld
Strongly compacted Edge of salt pedestal

Figure 5.33. Differential compaction around a


a 450 b1400 diapir arches its roof to produce topographic relief
that depends on roof thickness, diapir height, and
ted the depth/porosity ratio. (a) For a fixed roof
pac
om ed thickness of 1 km, compaction generates
400
rma
lc 1200 act
No o mp increasing topographic relief for increasing diapir
rmal c
Topographic relief (m)

Topographic relief (m)

350 d 1000 No heights. (b) For a fixed diapir height of 5 km,


cte compaction generates increasing topographic
mpa ted
erco pac
300 Und 800 om relief for increasing roof thickness. Normal
rc
Unde compaction creates more arching than
undercompaction. The analysis here uses the
250 600
depth/solidity functions for shale of Baldwin and
Butler (1985). Analytical models by Maria Nikolinakou.
200 Roof thickness = 1 km 400 Diapir height = 5 km

150 200

100 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10
Diapir height (km) Roof thickness (km)

thickness of the diapir’s roof (Figure 5.33(b)). This analysis decrease in structural relief of an arched roof is com-
suggests that compaction alone can produce arched roofs patible not only with compaction but also with a dying
having relief of as much as about 1 km where thick roofs occur active diapir in which the pressure drive declines to an
above tall diapirs. In most geologic settings, relief of a few equilibrium level.
hundred meters is more typical. By contrast, an active origin should be seriously considered
How can we distinguish between arching caused by if an arched roof has structural relief of more than a few
compaction and arching caused by active diapirism? The hundred meters above regional (1 km in special cases,
surest approach is to assume that all gentle arching above Figure 5.33), which is the maximum relief estimated for differ-
a diapir is caused by compaction because its effects are ential compaction around a diapir. Similarly, any signs of
inescapable (Figure 5.32). An apparent growth anticline shortening in the roof (Figure 5.21) point to an active contri-
above a diapir is not a reliable criterion. A gradual upward bution to arching.

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
91
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Box 5.1– Bryant Canyon Turned Inside-Out

a
27.0°

Inverted Salt
canyon
N
b
Minibasins

Canyon
erosion
Saddle

26.5°
Canyon
c inversion
Upturned flap

Salt
upwells

Figure 5.35. How salt upwelling can invert a canyon. (a) Flat-lying
overburden above salt. (b) Erosion of a submarine canyon replaces the
sedimentary load with less-dense water. (c) As a result, salt rises below
26.0° the canyon floor and inverts it as younger sediments accumulate.

of the Sigsbee Escarpment as an unconfined fan and associated


Bryant
Fan Bryant
channel–levee complex on the continental rise (Figure 5.34). The
Sigsbee Canyon fan contains more than 1,100 m of Pleistocene strata, which
Escarpment
suggests a large supply canyon. The vestigial canyon incising
-92.5° 0 20 km -92.0° the fan is as deep as 400 m just inboard of the Sigsbee Escarp-
Map data: http://www.geomapapp.org/, Ryan et al, 2009 ment. Remarkably, though, Bryant Canyon has been largely des-
Figure 5.34. Salt arching the floor of a canyon can invert and destroy it. troyed upstream of the first 15 km of incised salt canopy. Some
This bathymetric map of the submarine Bryant Canyon (northern Gulf of segments of the former canyon are preserved as subtle depres-
Mexico) shows the dashed trace of its former course down the lower sions or chains of minibasins. Other segments are flat or even
continental slope to breach the Sigsbee Escarpment. Where it was have positive relief. Minibasins guided the path of the incising
shallowly buried, the salt canopy bowed up the former canyon floor and
inverted the topography, leaving only a vestigial canyon. Where the canyon, and that erosion breached the saddles between the
canopy was loaded by deep minibasins, the canyon floor did not invert. minibasins. After the canyon was cut, the thick minibasins
Bathymetry from GeoMapApp, canyon course from Lee et al. (1996). remained as depressions. In contrast, the saddles were underlain
by culminations in the salt canopy, where salt was only thinly
Erosion can trigger diapirism by thinning the overburden and buried. Because of this differential loading, salt welled up in
removing its weight from the salt. This process can occur when a these saddles, arching the floor of the canyon and everting it
canyon cuts into flat-lying strata or when erosion bevels the crest inside-out (Figure 5.35). Eventually rise of the canyon floor des-
of an anticline. Bryant Canyon in the northern Gulf of Mexico is troyed or subdued the canyon in patches along its length. Salt-
an advanced stage of the former process, although this is not tectonic inversion thus healed the effects of canyon erosion and
evident from the present-day bathymetry. The Pleistocene Bry- subdued the relief of the seafloor. The now-discontinuous path of
ant Canyon was fed by a late Wisconsinan delta of the ancestral Bryant Canyon can still be traced by the seismic character of its
Mississippi River at the modern shelf margin. The canyon incised chaotic, discontinuous, or onlapping fill.
across the Sigsbee salt canopies and emerged in an embayment Key references: Lee et al. (1996); Suter and Berryhill (1985); Winker (1993).

5.4 Passive Diapirism diapirism began (Figure 5.36). This process, a form of active
diapirism, is known to occur if the diapir roof is relatively thin
5.4.1 Mechanism of Passive Diapirism (Section 5.3.1.1). However, rising through a thick, strong over-
5.4.1.1 Historical Review burden by gravity alone is mechanically implausible for two
Before 1933, salt diapirs were universally thought to rise reasons. First, the diapir must force its way through a much
through thick sedimentary rocks that accumulated before stronger prekinematic overburden, a process as difficult as

92
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.4 Passive Diapirism

Figure 5.36. Schematic


a Deposition of a salt layer rise of a diapir through its a Thickness
Crest of diapir
Surface
thick overburden by the
process of upbuilding. Height
In this portrayal, the salt
was thought to be too Base of diapir
strong to rise until it had b
b Burial of the salt layer by
additional sedimentation been buried and heated Subsidence Down-
until it could flow by building
creep. This notion is
rheologically unrealistic
because salt flows readily
c
at surface temperatures.
The depiction would
entail a massive room
problem as the diapir is

c Growth of the instability of the salt layer


emplaced, as well as
insufficient force to
d
penetrate any lithified
overburden. After Turcotte
and Schubert (2002).

Figure 5.37. A downbuilding (passive) diapir grows in height by maintaining


its crest at or near the sedimentary surface while its base sinks along with
the surrounding sediments. Forward kinematic model after Hudec and Jackson (2011).
d Formation of salt domes
accumulates above the salt. Most of the world’s tall salt domes
and walls spent most of their histories as passive diapirs.

5.4.1.2 Initiation of Passive Diapirs


Passive diapirism can occur in any tectonic setting. However,
for passive diapirism to begin, salt must be at the surface.
Passive diapirism can initiate very shortly after salt deposition,
sourced from salt that remained exposed as surrounding areas
began to be buried. However, generally salt must break
driving a putty nail through a wooden board (Wade 1931). through its overburden to reach the surface and emerge as a
Second, the volume of overburden rock displaced by the passive diapir.
diapir’s volume is largely unaccounted for – an anomaly The most common way for salt to reach the surface is by
equivalent to the room problem in pluton emplacement. active diapirism (Section 5.3.1). This is especially likely after a
Barton (1933, p. 1029) termed this process upthrusting, long phase of reactive diapirism, where extension has established
whereby “the base of the salt core is regarded as stationary a load gradient and weakened the roof (Figure 5.3, Section 5.2).
and the crest is regarded as uplifted.” Aware of the mechanical However, the active phase may also be triggered by regional
implausibility of upthrusting (now more commonly known as shortening or by sedimentary topography (Figure 5.25).
upbuilding), he turned diapirism on its head, so to speak, by Passive diapirs can also be initiated by deep erosion of an
discerning that “the crest of the salt core is regarded as station- anticline crest, which exposes much of its salt core. An early
ary and the base is regarded as moving down.” This elegant stage of this process is shown by Cardona diapir in the Ebro
insight envisaged a diapir growing syndepositionally, main- foreland basin of the Spanish Pyrenees. The diapir is less than
taining its crest near the sediment surface, while its base sank, 300 m high and wide but rests on a salt anticline 1,900 m high
along with the surrounding sediments accumulating in the and 7,000 m wide (Sans and Koyi 2001) (Figure 5.38). This
basin. Barton (1933) termed this process downbuilding, which mismatch in size between a small diapir and a large anticline is
is synonymous with passive diapirism (Nelson 1991; Vende- typical of diapirs formed from injection folds. The late Eocene
ville and Jackson 1992a). Cardona salt was deposited in the Ebro foreland basin after
Downbuilding solves both major mechanical problems of shortening in the Pyrenees had already started in the Late
upbuilding. Because the diapir’s crest is continually emergent Cretaceous (Puigdefabregas et al. 1992). However, the thick
(except for ephemeral veneers of sediment), little country rock prekinematic sequence in the anticline (Figure 5.38) shows that
is displaced or lifted (Figure 5.37). There is no room problem. shortening reached this part of the Ebro foreland only during
As a result, in cross section a downbuilding diapir cuts across or after deposition of the Oligocene Solsona Formation. Other
many kilometers of thickness of strong overburden, but the diapirs are scarce here, and other major anticlines in the Ebro
piercement is only apparent. Sedimentary strata may onlap the basin have salt cores close to piercing, but their overburdens
flank of the exposed diapir, but negligible permanent roof are too thick to be pierced by halokinetic active diapirs.

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
93
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Pinós-Cardona anticline Figure 5.38. Erosion of an anticline crest can


trigger the rise of a small diapir. Cardona diapir
South-southeast North-northwest (Spain) breached the unroofed crest of the
Cardona diapir much larger Pinós–Cardona anticline. After Sans
Foreland and Koyi (2001).
Marba
Marba Alberto
Llardella 1 Cardona
Topography

Solsona Formation
Oligocene
0 1 km
Castelltallat Formation
V.E. × 1
Barbastro and Surla Formations
Eocene
Cardona evaporites

Future erosion rate backward indicates that the Cardona diapir


a erosion
level
pierced only 2 Myr ago.
Later stages of erosional initiation of passive diapirs are
Salt common in the Zechstein basin of the southern North Sea.
Salt-cored anticlines began to grow by gravity-induced buck-
b ling during the Late Jurassic to Early Cretaceous. The folds
amplified further in the Late Cretaceous and Paleogene during
inversion and thermally induced regional tilting, which
Shortening
restored the Lower Permian strata to horizontal (Coward and
Stewart 1995; Stewart and Coward 1995, 1996). The North
German basin was also shortened and deeply eroded in the
Unroofing Early Cretaceous and less so in the early Paleogene (Mohr et al.
c unconformity
2005). Mesozoic erosion cut deeply into the anticlines, and salt
tectonics transformed them to diapiric walls. Two processes
Shortening thinned the anticline crests (Coward and Stewart 1995; Stewart
and Coward 1995). Some crests were thinned by extension as
their limbs steepened and failed along weak beds such as the
Röt Halite, which allowed the salt core to intrude the crest as a
Extension
d Rotational
subsidence
Dissolution reactive diapir (Figure 5.39(a), (b)). The most potent trigger
for diapirism was erosional breaching, which beveled the dip-
ping fold limbs (Figure 5.39(c)). Once the salt had been
breached, it was expelled as a diapir, which drained the anti-
Unroofing unconformity tilts inward cline’s salt core. Deprived of salt support, the fold limbs sub-
Figure 5.39. Erosional unroofing of a buckling salt anticline triggers a
sided by rotating inward and unfolded back to horizontal
diapiric wall. (a) An immature salt anticline. (b) The mature salt anticline’s (Figure 5.39(d)). The originally horizontal unroofing uncon-
crest thins by extension, with the future erosion level marked by the wavy line. formity also rotated inward and cut deeper toward the salt wall
(c) The mature salt anticline‘s crest thins further by erosion, breaching the
underlying salt. (d) Escape and passive rise of a diapir deflates the unsupported
(Figure 5.40).
salt anticline, tilting inward the erosional surface (unroofing unconformity).
Subsidence of the anticline’s limbs creates local extension, which must 5.4.1.3 Growth of Passive Diapirs
be balanced by shortening elsewhere. After Coward and Stewart (1995). In order for a passive diapir to grow, salt within it must be
pressurized enough to flow upward at the sediment surface.
Sans and Koyi (2001) estimated that the Cardona diapir This pressure may be generated in at least three ways
pierced its overburden when erosion had thinned its roof (Figure 5.41). First, if the average density of overburden above
to a threshold thickness of about 300 m, at which point the the source layer is greater than that of salt, salt is then forced to
diapir actively broke through it. Extrapolating the present the surface. Second, the salt diapir may be laterally shortened,

94
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.4 Passive Diapirism

causing salt to rise at the surface by displacement loading. to this global trend is the diapirs in the Zagros fold belt, which
Third, salt may be loaded by sedimentary topography, for build unusually high freeboards because of regional shortening
example, near the toe of a continental slope or within a graben induced by collision of the Arabian and Iranian plates and the
having topographic relief. preponderance of dense carbonates in the overburden (Talbot
Where a passive salt stock extrudes faster than it dissolves, 2005). Another exception is the unusually high freeboards
emergent salt swells to form a bun-shaped topographic dome of 400 m or more of some diapirs on Axel Heiberg Island
(Talbot 1998; Dooley et al. 2015). The dynamic pressure (Arctic Canada). Depending on the proportion of anhydrite–
exerted by rising salt builds a dynamic bulge in these summit carbonate to halite at depth within the diapirs, which is
domes up to about 600 m in relief in the Zagros fold belt. unknown, these high-standing diapirs could be elevated by
A passive diapir rises as sediments accumulate around its mild post-Eurekan compression or purely by halokinesis
margin. In most environments sedimentation is episodic, so (Harrison and Jackson 2014a, b). Theoretical diapiric free-
the diapir is periodically buried beneath a sedimentary veneer boards can be calculated assuming static, thick salt in a hori-
during rapid sedimentation. Each burial triggers a brief period zontal source layer and particular densities and thicknesses
of active diapirism, during which the diapir breaks through for salt and overburden. At static equilibrium, an emergent
the ephemeral sedimentary veneer and continues to grow salt diapir is supported by the weight of the surrounding
by downbuilding. Passive diapirs thus have three growth overburden on the source layer (Talbot 2005), so that
modes: (1) true passive diapirism, when the salt is emergent;
(2) temporary diapir burial after sediments aggrade faster ρs ghs ¼ ρo gho
than the diapir crest rises; and (3) active piercement as aggra- where ρs and hs are the density and height of a salt diapir above
dation is outpaced by diapiric rise (Jackson et al. 1988, 1994). its base, ρo and ho are the density and thickness of its overbur-
These three processes are repeated as cycles that keep the den, and g is the acceleration due to gravity.
crest of the diapir at or near the sediment surface during
growth, so that the overall process is one of downbuilding.
This complex history turns out to have important implica-
5.4.2 Structural Style of Passive Diapirs
Because there is no room problem, passive or downbuilding
tions for sedimentary structures adjacent to passive diapirs
diapirs are noted for the mild deformation of their country
(Section 5.4.2.1).
rocks. The diapir resembles a cutout stamped out of encasing
5.4.1.4 Diapiric Freeboard layered strata, whether in vertical section (Figures 5.42 and
A diapir’s freeboard is its relief above the surrounding plain 5.43) or in horizontal view (Figure 5.44). Deformation in the
(Talbot 1998). Most emergent diapirs around the world have overburden is typically limited to subtle thickness changes
freeboards less than 200 m (Davison et al. 1996a). An exception related to flow of underlying salt and upturned or sheared
strata near the diapir contact.
0
Cenozoic
5.4.2.1 Halokinetic Sequences
The complex rise history of most passive diapirs means that
Two-way time (s)

Unr
oof
ing they undergo multiple phases of active rise, as ephemeral roofs
1 unc
onf
orm Cretaceous are successively pierced. These repeated phases of uplift and
ity Zechstein salt
breakout each produce tilted beds, which can (if preserved)
Triassic–Jurassic stack to form halokinetic sequences adjacent to the diapir.
2 Each halokinetic sequence records one episode of burial,
Pre-Zechstein breakout, and downbuilding.
0 5 km The type area for halokinetic sequences is La Popa basin in
Mexico (Lawton and Giles 1997; Giles and Lawton 2002). Here
Figure 5.40. A diapir triggered in the manner of that in Figure 5.39 has
a diagnostic inward tilt of the unroofing unconformity. Traced from a seismic halokinetic sequences contain a systematic stratigraphy. Each
section in Stewart (2006). sequence begins and ends with an angular unconformity that

Displaced Figure 5.41. Passive diapirs can grow by three


a b salt c mechanisms: (a) density inversion in which the
average density of the overburden is greater
than that of the salt; (b) displacement loading,
ρo ρs in which lateral contraction squeezes diapiric salt
upward; and (c) topographic relief, in which the
thickness of diapiric salt is less than the thickness
of the overburden.
ρo > ρs Driven by displacement loading Sediment wedge, ρo
Driven by density inversion
Water wedge, ρw
ρo > ρw

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
95
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Figure 5.42. Diagnostic features of a passive


diapir, the most important of which are the
Residual soil stratal cutoffs and paucity of deformation in
or salt exposed the country rock. After Jackson et al. (1994).

ge
e flan
siv
Upturned tru
Ex
flap ss
ne
Regional
h i ck
e t es
btl ng
Su cha
Flat crest of
stock or wall

Figure 5.43. Seismic examples of the cutout


a appearance of (a) a leaning passive diapir and
(b) an hourglass-shaped diapir. Both diapirs exhibit
1 flaps of various scales along their flanks, indicating
phases of active diapirism. Northern Gulf of Mexico
seismic data courtesy of CGG.
2
Depth (km)

3
Salt

0 5 km
V.E. × 1
Crestal graben
b2

3 Salt
Depth (km)

Base allochthonous salt 0 3 km


V.E. × 1

grades outward into correlative conformities (Figures 5.26, The unconformities bracket two end-member types of
5.45, and 5.46). Halokinetic sequences like these tend to be halokinetic sequences, known as hook and wedge, as distin-
only a few hundred meters wide, but in some parts of the guished in Table 5.1 and Figure 5.47(a), (b).
world they are several kilometers wide, possibly as a result of These stratigraphic variations depend on changes in
lateral shortening. topography above an emergent or shallowly buried diapir.

96
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.4 Passive Diapirism

Figure 5.44. Exposed passive diapirs in the


a Great Kavir (Iran) have discordant cutouts of strata
against their flanks. (a) Aerial photograph of a
smooth, circular stock, Dome 14; after Jackson et al.
(1990, figure 1.42). (b) Satellite image of a ragged,
irregular wall, Dome 51; from Google Earth.
34.96°

34.92°

0 2 km
53.72° 53.79°
Center 34.941°N, 53.754°E
b 34.83°

Rising
salt N

Wasting
salt

Sharp, discordant contact

34.81°
0 1 km
53.20° 53.22°
Center 34.818°N, 53.216°E

Topographic relief increases and decreases as sedimenta- tapered composite sequence in the Zagros fold belt are
tion rate fluctuates rapidly against a background of slowly shown in Figure 5.48.
varying rates of salt flux (Rowan et al. 2003). Halokinetic The width of a deformation halo around a diapir is
sequences can be stacked into two types of composite expected to correlate loosely with lithology and aggradation
sequences: tabular and tapered (Figure 5.47(c), (d)). Hook rate; the latter affects the thickness of the halokinetic wedge
halokinetic sequences stack to form tabular composite (Rowan et al. 2003). Slow, mud-dominated sedimentation is
sequences during slow aggradation. Wedge halokinetic likely to result in halokinetic aureoles about 200 to 300 m wide,
sequences stack to form tapered composite sequences where sands thin from regional thickness only near the diapir.
during fast aggradation. Well-exposed examples of a Fast sedimentation should be associated with halokinetic

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
97
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

a
100.80°W 100.75°W Viento Formation

Adjuntas Formation (Eocene)

Upper Sandstone Member


110 99
El

Maastrichtian–Paleocene
G 114 Upper Mudstone Member

Potrerillos Formation
or El Upper Gordo lentil
do 102 Papalote 155
an 115 Delgado Sandstone
tic 90 180
lin
e Middle Siltstone Member
Middle Gordo lentil
26.05°N Lower Mudstone Member
Lower Gordo lentil
Lower Siltstone Member

Muerto Sandstone (Maastrichtian)


78 Parras Shale (Campanian)
110
84 N
El Minas Viejas salt
100 Gordo
90 Axial trace of folds
124 100
Limit of diapiric deformation
Dip
114 Rotation of flank strata
0 3 km
26.00°N

b
West-northwest East-southeast
Hook
0 500 m halokinetic
sequences
Cusp at contact
Eroded crest

El Papalote 4 5
diapir
2
Composite 3
halokinetic
sequence

Wedge
halokinetic
sequences
Figure 5.45. Halokinetic sequences in the 500- to 1,000-m-wide contact zone around El Papalote and El Gordo diapirs (La Popa basin, Mexico). Strata are
locally overturned in hook halokinetic sequences; angular unconformities bounding the sequences are subvertical and grade into correlative conformities within
about 250 m of the diapir’s contact. (a) Map view, (b) cross section. After Rowan et al. (2003).

aureoles about 600 to 800 m wide, where sands onlap and eventu- erosion and is in contact with sediments accumulating
ally cover the diapir. Diapirs growing along contractional anti- around the diapir. The competition between rising salt and
clines could have deformation aureoles as much as 3 km wide. aggrading sediments is the main control of the cross-
sectional shape of a passive diapir. This competition can be
5.4.2.2 Controls on Cross-Sectional Shapes of Passive Diapirs described in three mathematical ways, each of which has
After a diapir emerges at the surface it changes growth advantages and disadvantages (Figure 5.49). None of the
dynamics. The salt becomes exposed to dissolution or methods is rigorously applicable to initiation of diapirs, to

98
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.4 Passive Diapirism

Figure 5.46. Hook A salt sheet forms when R/ _ A_  1. The approach was applied
a halokinetic sequences to both upbuilding and downbuilding diapirs. These ideas
Extension and fracturing form on the flanks
of thin roof of a salt diapir being were expanded by McGuinness and Hossack (1993), Jackson
Future failure surface
and halokinetic
repeatedly buried et al. (1994), and Koyi (1998), and especially by Talbot (1995),
sequence boundary by a thin sedimentary who introduced the term molding ratio for R/ _ A_ (Figure 5.50).
Diapir roof, which is then
inflation diapirically breached Although these studies were all merely two-dimensional ana-
Hook 1 _ A_ ratio have remained the
and shouldered aside lyses, the implications of the R/
to form drape folds. preeminent explanation for the cross-sectional shapes of pas-
Debris flows mark
b Onlap and overlap Debris flow from failed the base of each sive diapirs for the past three decades.
roof and diapir halokinetic sequence. Despite this preeminence, the R/ _ A_ concept is fatally flawed
Flexural slip below each as an explanation for cross-sectional shape of passive diapirs. It
Hook 2 unconformity increases
Diapir the original erosional has long been known that R/ _ A_ = 1 for all passive diapirs,
Hook 1 angle of truncation and regardless of cross-sectional shape (Jackson et al. 1988).
creates a salt cusp. After Thus, workers writing about molding ratio must have had
Giles and Rowan (2012).
c Overlap strata some other concept in mind, at odds with the original defin-
_
_ A.
ition of R/
Future failure surface We infer that this other concept was volumetric flux, q
Diapir
inflation
Rotated beds
(Figure 5.49). This is the rate of volume flow across a unit
Cusp Hook 2 area at the top of a diapir. It has units of meters per second (the
same units as R). _ Volumetric flux for a rising diapir can be
Hook 1
thought of as the average vertical velocity of salt through
the top of the diapir, or as the hypothetical rise rate of the
d Debris flow from failed
diapir if it maintained vertical sides in defiance of gravity.
roof and diapir As with R/ _ net volumetric flux would account for salt loss
_ A,
Hook 3
due to erosion or dissolution. The advantage of volumetric
flux is that q/A_ obeys all the rules originally assigned to R/ _
_ A.
Diapir Hook 2 That is, a diapir narrows upward if q/A_ < 1, has vertical sides if
q/A_ = 1, and widens upward if q/A_ > 1. A salt sheet forms
Hook 1 when q/A_  1. We therefore propose that molding ratio be
redefined as q/A_ and that R/ _ A_ be restricted to discussions of
e Failure scarp
evolving topography.
A third way of measuring diapiric rise rate is by volumetric
Diapir Hook 3 rate of salt supply (Weijermars et al. 2015). This is the volume
inflation of salt entering the base of the diapir per unit time, Q. The
Hook 2 advantage of using Q is that it varies systematically through
time as a function of regional parameters such as changes in
Hook 1
basin topography, salt thickness, or gravitational or
displacement loading on the source layer. Despite their simpli-
city, analytical models can simulate in three dimensions the
diapirs buried by a thin roof or minibasin, or to laterally growth of walls having rectangular planforms and stocks
squeezed or stretched diapirs. having circular planforms. The effects of salt spreading over
The oldest, best known, and most misused measure is the a dipping seafloor were explored in the more sophisticated
ratio of rise rate to aggradation rate, R/ _ Rise rate (R)
_ A. _ was analytical models of Weijermars et al. (2015). However, the
defined by Seni and Jackson (1983a) as the change in relief (R) shape of a diapir can be calculated only if Q/A_ is known for the
of a diapir over time, although the concept had been implicit entire history of the diapir. These models resulted in diapirs
since Barton (1933) introduced the concept of downbuilding. typically having curved sides in vertical section rather than the
If the base of salt is set as a fixed reference level, R_ is the rate at straight sides depicted in R/ _ A_ analyses.
which the top of the diapir moves up (or down) (Figure 5.49). If the molding ratio varies over time, more-complex shapes
As a refinement, R_ was defined as the net rate of diapir rise, result. A common shape is an hourglass diapir (Figure 5.51),
which is the total rise rate minus dissolution and erosion of formed when a basal salt pedestal rises to a slimmer stem
salt. Jackson and Cornelius (1987), Jackson et al. (1988), and surmounted by a bulb forming an overhang. However, hour-
Jackson and Talbot (1991) proposed that the interplay between glass diapirs can also form in active diapirism with megaflaps
sedimentation and upwelling salt controlled the cross-sectional (Section 5.3.2.2). For hourglass diapirs formed by passive rise,
shape of a passive diapir. A diapir narrows upward if R/ _ A_ < 1, initially the thin overburden exerts little weight on the source
has vertical sides if R/_ A_ = 1, and widens upward if R/ _ A_ > 1. layer, so salt rises slowly and is onlapped and partly buried by

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
99
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Table 5.1. Types of halokinetic sequences

Feature Hook sequence Wedge sequence


Facies in a single halokinetic sequence Outer-shelf black shale (top) Middle-shelf shale (top)
Middle-lower shoreface Tidal or lagoonal sandstone
Outer-shelf black shale Upper-shoreface sandstone
Subaqueous debris flow (base) Lower-shoreface sandstone/shale
Middle-shelf shale (base)
Internal folding Folding common Folding rare
J-hook unconformities, salt cusps Truncation angles <15°
Truncation angles <90°
Slip on sequence-bounding unconformity Significant slip Negligible slip
Abundant brittle shear
Relative aggradation rate Very low, during transgression Moderately high, during regression
Termination of halokinetic sequence Terminates directly against diapir Terminates against older, folded halokinetic sequences;
diapir averages 250 m from truncation
Width of upturn Up to 200 m Up to 800 m
Note: Adapted from Giles et al. 2004, Giles and Rowan 2012.

Figure 5.47. Characteristic features of (a) a hook


a b halokinetic sequence, which can be stacked into
Axial trace (c) a tabular composite halokinetic sequence;
Axial trace (b) a wedge halokinetic sequence, which can
be stacked into (d) a tapered composite halokinetic
Wedge sequence. After Giles and Rowan (2012).
Diapir

Hook Diapir

300–1,000 m
50–200 m
Drape folding in zone 50–200 m wide Drape folding in zone 300–1,000 m wide
Up to 90° angular unconformity Up to 30° angular unconformity
Facies change abruptly near diapir Broad zone of gradational facies changes
Results from relatively slow aggradation Results from relatively fast aggradation
Zone of stacked Zone of stacked
c monoclinal
axial traces
d monoclinal
axial traces

Hook 4 Wedge 4

Hook 3 Wedge 3
Diapir

Hook 2 Wedge 2
Diapir
Hook 1 Wedge 1
50–200-m width of folding
300–1,000-m width of folding
Subparallel base and top boundaries Convergent base and top boundaries
Narrow zone of thinning and upturn near diapir Broad zone of thinning and upturn near diapir
Axial traces are near diapir and offset in Axial trace is continuous and climbs away
zone parallel to diapir margin from diapir
Relatively slow aggradation Relatively fast aggradation

encroaching strata (Figure 5.51(a)). As surrounding sediments increases the driving force of diapirism, but depletion of the
thicken, their extra weight on the source layer causes the diapir source layer offsets this increase. As the source layer thins,
to rise faster, so its contact steepens to vertical (Figure 5.51(c)) boundary drag within it increases exponentially. Consequently,
and then widens upward (Figure 5.51(d)). During this vigor- the thinning source layer starves the diapir ever more effect-
ous stage, gentle upturning of surrounding strata commonly ively. The diapir rises more sluggishly, and its bulb abruptly
results in angular unconformities centered on the diapir narrows upward, allowing the crest of the diapir to be buried
(Figure 5.27). The extra weight of the thickening overburden when passive growth ends (Figure 5.51(e)).

100
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.5 Evolution of Diapirs

During passive rise other changes occur. Sediments onlap


a and encroach across the future salt pedestal, and depocenters
N
of each unit migrate inward, pushing salt toward the diapir
axis. Strata originally onlapping the pedestal (Figure 5.51(a))
rotate as the underlying salt flows toward the diapir axis. After
salt has been entirely expelled locally, a salt weld forms; ori-
Diapir
conta
ct ginal onlaps rotate to form apparent downlaps (Figure 5.51(d)).
The cutoffs of these apparent downlaps become younger in the
Diapir direction in which the expelled salt flowed (toward the diapir).
center Welding cuts off supply of salt from the outer part of the
Road source layer. From initial welding onward, the diapir can only
downbuild by redistributing salt already trapped in its pedestal
(Figure 5.51(b)–(d)).
90
m 5.4.2.3 Salt Wings
The term salt wing was used by Lowrie et al. (1991) to refer
to small protuberances of salt from the flanks of a diapir.
They did not specify how a salt wing forms but appear to have
had intrusion in mind, which was driven by thermal changes
in the salt or density changes in the overburden. Such out-
growths had been previously recognized (for example, Jenyon
b 1986a, figure 7.7) and attributed to lateral injection of salt.
Where overlying strata onlap the salt wing, this geometry
N

Diapir
center suggests that the wings formed by periodic brief extrusion
of salt on the seafloor rather than by intrusion, as in the
Dip slope “Christmas-tree” structures of buried but periodically extrud-
ing mud volcanoes (Figure 5.52). These effects of pulsed
t changes in aggradation rate in forming repeated salt flanges
tac or wings around passive diapirs are seen in physical and
on
rc mathematical models using all three approaches described
pi
ia in the previous section.
D
However, other types of salt wings appear to have formed
by lateral intrusion of salt during diapir shortening and liftoff
folding, especially in the Zechstein basin (Hudec 2004). During
SB the Late Cretaceous regional shortening, Upper Permian salt
intruded laterally along thin Triassic Röt evaporite layers,
which acted as weak surfaces that were “unzipped” and wedged
19 apart by intruding salt (Section 11.2.1.4).
0
up -m-
tu wi
rn de
SB 5.5 Evolution of Diapirs
Salt can rise by reactive, active, or passive diapirism. A diapir
may continue growing in one of these modes for its entire
history. However, any growth mode can switch to any other
mode, which yields a composite growth history. Evidence of
changing growth modes is preserved in sediments surround-
SB ing the diapir, as first recognized by Ferdinand Trusheim
Road SB
(Box 5.2). Examples of changing modes of growth in
Figure 5.53 are arranged according to the initial mode of
growth. Four main agents described in the following four
sections can change the growth mode.
Figure 5.48. Hook halokinetic sequences along the southern margin of
Herang diapir (Zagros fold belt, Iran; SB is sequence boundary. (a) Overview of 5.5.1 Changes in Sedimentation Rate
diapiric margin; the diapir ruin has negative relief. (b) Four hook halokinetic
sequences forming a tabular composite halokinetic sequence. Outside the Diapiric growth is sensitive to changes in aggradation rate.
~200-m-wide zone of upturn (purple line), the halokinetic sequences have a Local and far-field changes in aggradation rate typically
subparallel top and base; axial traces (yellow lines) are offset but stacked parallel
to the diapir contact. Photographs courtesy of Jean Letouzey.

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
101
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Definition Relatively slow Relatively fast Figure 5.49. The cross-sectional shape of passive
sedimentation sedimentation diapirs depends on the interplay of rising salt
and aggrading sediments. The flow of salt can be
R /A = 1 R /A = 1
measured in three ways: by R_ (the one-dimensional
R A rise rate of the diapir), by q (the volumetric
R R A R A
flux of salt at the diapir crest), and by Q (the
volume of salt added to the base of a diapir per
unit time). L and T are the dimensions of length
and time.

R = LT-1
Horizontal area q / A = 1.2 q / A = 0.5
Volume Height neglecting
q added A
gravity Volume added A

q = L3L-2T-1 = LT-1

High Q Low Q
Q A
A

Q = L3T-1 Horizontal area Volume added


Q = volume added

Time Increasing net diapir rise rate Upward Upward Base-salt flat during
stages narrowing widening relatively rapid diapir rise
6 E F G H
5
D
4
C
3
Slow aggradation
2 B
A
1

0
Salt flare during
relatively rapid diapir rise Steep ramps = relatively slow diapir rise, fast aggradation
Gentle ramps = relatively fast diapir rise, slow aggradation
Figure 5.50. Many workers have used R/ _ A_ to explain cross-sectional shape of passive diapirs. In these explanations, changing the ratios of diapiric rise and
aggradation produces diapirs of widely different shapes (A through H). The problem with this technique is that, between each pair of horizontal time lines, the diapir
and the sediment surface rise by the same amount. Thus, R/ _ A_ = 1 for all diapirs shown. After Talbot (1995).

have opposite effects. Faster far-field aggradation accelerates rapid effect when the overburden is still thin, but only a
overburden thickening and increases pressure of the source gradual effect if the overburden is already thick.
layer, which favors diapirism. In Figure 5.53(c) pressurizing Faster local aggradation slows the rise of a passive diapir. In
the source layer accelerates the rise of salt, so the passive diapir Figure 5.53(a) faster aggradation buried the diapir (aided by
widens upward and extrudes salt. However, doubling the far- depletion of the source layer). Salt continued to rise as an
field aggradation rate does not immediately double the salt active diapir by arching its roof. The local effect of slower
pressure in the source layer. The salt pressure is proportional aggradation is to accelerate the rise of a passive diapir, which
to the total overburden thickness, not to the rate of thickening. widens upward, as long as salt can be imported from the
Thus, doubling the rate of far-field aggradation will have a source layer rapidly enough (Figure 5.53(c)).

102
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.5 Evolution of Diapirs

Figure 5.51. Evolution of an hourglass-shaped


a Outward-
dipping flanks
Onlapping
strata Thin basin depocenter
wall that grew entirely as a passive diapir.
As depocenters migrate inward, adjoining
strata onlap then uplap the salt before being
Salt overstepped by the diapir. q/A_ increases through
most of the diapir’s history because the source
layer is increasingly loaded, and then wanes
near the end as the source layer depletes.
b Primary withdrawal basins touch down

c Withdrawal basins
shift toward diapir
Vertical diapir
flanks
Uplapping
strata Large total thickness

d Source layer welded, no


additional salt available to diapir Inward-shifting depocenters

Pseudo-downlap is rotated onlap


e Buried wide crest

Pedestal

Younger in direction of basal cutoffs

5.5.2 Changes in Salt Supply picture can be complicated. For example, if crustal uplift
caused the far-field erosion, increased basinward tilt could
If salt flows more slowly into a diapir, its rise is slowed, which
increase salt pressure and promote diapirism downdip from
is, in effect, equivalent to increasing the local aggradation rate.
the exhumed uplift.
The most common cause of decreased salt supply is exhaustion
Local erosion thins a diapir’s roof and lessens its resistance
of the source layer (Figure 5.53(a)), but salt supply can also be
to diapirism. An originally passive diapir can be buried by
decreased by far-field erosion.
faster aggradation, then rise actively by arching its roof
Conversely, increasing salt supply favors diapirism. The
(Figure 5.53(b)). Erosion of the arched roof could thin it to
most common cause of increased salt supply is sediment
less than threshold thickness, causing the diapir to pierce fully
aggradation above the source layer, but salt supply can also
and then resume passive growth (Figure 5.53(d)). A diapir
be increased by regional tilting, causing salt in the source layer
initiated as an active injection fold (Figure 5.53(h)) or a dela-
to flow more rapidly downdip.
mination intrusion (Figure 5.53(i)) can switch to passive
growth after erosion has stripped off its roof strata.
5.5.3 Erosion
Here again, the effects of far-field and local erosion are oppos- 5.5.4 Regional Extension
ite. Far-field erosion thins the overburden and reduces its load. Extension is a potent trigger to initiate diapirs. Local structural
This decreases the pressure of the salt source layer, which thinning causes reactive diapirism, regardless of density, as
retards diapirism fed by the source layer. An outcome of far- long as aggradation is slow enough not to completely fill the
field erosion could thus be burial of the diapir, causing a switch graben above the diapir. A reactive diapir continues rising as
from passive to active growth (Figure 5.53(a)). However, the long as salt can be imported fast enough from the source layer.

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
103
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Seafloor Growth Agent


Mode

a Active
Burial

Incipient Passive
salt wing

b Active Burial then


shortening
Passive
Salt wing
Extrusive

Salt
c passive Faster rise or
slower local
Emergent aggradation
passive

Aggradation
d Passive
Active
and rise rates
balanced
Burial then
Passive faster rise
or erosion
Figure 5.52. Salt wings typically record periodic extrusion of a wedge of
salt that subsequently became buried, as in this example from the northern
Gulf of Mexico. However, in the Zechstein salt basin, most wings represent salt
intrusions along weak layers during contractional delamination. Seismic image
courtesy of TGS.
e Falling
Extension
accelerates
or source layer
depletes
Passive
The supply of salt may not keep pace with the growing diapir
if extension is too fast or if the source layer becomes thin
enough that boundary drag slows salt flow. If salt cannot be Extension stops,
supplied fast enough and extension continues, a reactive diapir f Active
aggradation
continues
starts to fall (Figure 5.53(g)) (Vendeville and Jackson 1992b).
Reactive
An originally passive diapir also falls if extension is too fast
(Figure 5.53(e)). If the roof becomes too thin to resist salt
pressure, a reactive diapir evolves into an active diapir Extension
(Figure 5.53(f)). g Falling accelerates
or source layer
depletes
5.5.5 Regional Shortening Reactive

On a geologic time scale, shortening has immediate effects.


The source layer is further pressurized as the overburden
thickens by folding and thrusting. The source layer itself h Passive
Erosion then
thickens by layer-parallel shortening. Increases in pressure aggradation
and thickness of the source layer promote diapirism. More- Injection
over, buckling of a thick roof above a diapir and squeezing of fold
its flanks force salt to rise as an active diapir. Thus, shortening
can rejuvenate a buried, dormant diapir (Figure 5.53(b)). i Passive
Erosion then
5.6 Diapir-Flank Faults Delamination
aggradation

intrusion
5.6.1 Radial Faults
Figure 5.53. The effects of changing modes of diapiric growth.
Faults radiating outward from diapirs are a fundamental
aspect of salt tectonics and are seen in nearly all diapiric salt
basins. Radial faults abut the diapir contact or curve inward

104
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.6 Diapir-Flank Faults

Box 5.2 – Ferdinand Trusheim – The Father of Modern Salt Tectonics


Ferdinand Trusheim (Figure 5.54) was born on April 24, 1906, in strata thin by erosion and deposition toward a salt pillow, which
Frankfurt am Main and died on June 28, 1997, in Hannover. He forms a topographic high. This unit is overlain during the diapir
was a pioneer of German petroleum geology and can be stage by a secondary peripheral sink in which strata thicken
regarded as the founder of modern salt tectonics. Trusheim toward a salt diapir (the former salt pillow) except near the salt
began studying geology at Frankfurt, then received a doctorate contact. This thickening records subsidence due to inward flow
from the University of Würzburg for research on the Mittenwald of underlying salt, which evacuates the outer parts of the
Karwendel trough. After working at Senckenberg am Meer former salt pillow. Subsidence deforms the primary peripheral
research institute, he returned to the University of Würzburg as sink from a convex-downward base and a flat top into a flat
an assistant in the Mineralogical-Geological Institute, then base and a convex-upward top. This inversion forms what
became an associate professor in geology and paleontology Trusheim called a turtle-shape structure (nowadays called a
there in 1942. In 1943 Trusheim became a military geologist in turtle structure or turtle-structure anticline) (Figure 5.56).
various countries occupied by the Third Reich before being The secondary peripheral sink is in turn overlain during the
captured and imprisoned by the Soviet army until 1950. He then postdiapir stage by a tertiary peripheral sink. The outer strata
shifted his career to petroleum geology, joining Brigitta (a fore- in this unit are isopachous or thicken slightly toward the salt in
runner of BEB Erdgas und Erdöl). He became their head of response to inward salt flow in the source layer (as in a second-
exploration in Hannover in 1957, the year he published his ary peripheral sink). However, the innermost strata thin across
seminal study of halokinesis. the crest of the diapir.
In these papers, Trusheim (1957, 1960) laid the tectonostrati- Trusheim’s use of the geometry of flanking sediments, rather
graphic foundation to analyze growth of salt pillows and diapirs. than the salt itself, to infer the history of salt flow was a seminal
His landmark papers show how lateral thickness variations in innovation. His approach is now used by virtually everyone
flanking strata record salt movement through time. Using geo- interpreting the growth history of salt diapirs. But his idea took
physical data from more than 200 Zechstein salt walls and stocks another 23 years to disseminate beyond the Zechstein salt basin
in northern Germany, Trusheim coined the term peripheral sink (Seni and Jackson 1983a; Lobao and Pilger 1985). Since then,
for synkinematic depocenters flanking diapirs; such depocenters additional knowledge and insights have suggested modifica-
are also known as withdrawal basins. The geometry of these tions to Trusheim‘s ideas (Vendeville 2002). However, his work
sinks suggested a three-stage history: the pillow stage, the diapir has remained the basis for all modern analyses of salt tectonics.
stage, and the postdiapir stage (Figure 5.55). Trusheim‘s
sequence begins with a prekinematic, isopachous unit deposited
on the Zechstein Salt. During the pillow stage, synkinematic
sedimentation begins in a primary peripheral sink in which
End Muschelkalk a

b Lower Cretaceous (Dogger)

c End Lower Cretaceous

d End Cenozoic

Salt weld

Tertiary peripheral sink Prekinematic unit


Secondary peripheral sink Salt
Primary peripheral sink Presalt
Figure 5.55. Trusheim‘s (1957, 1960) concept of evolving peripheral sinks
during the growth of a salt pillow and diapir, based on examples from North
Germany. His original restoration has been mirrored about the center of the
Figure 5.54. Ferdinand Trusheim. Photo courtesy of Senckenberg Forschung, line to give equal emphasis to the growth of diapirs and the intervening
Wilhelmshaven, Germany. turtle-structure anticline.

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
105
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Box 5.2 – (cont.)


0 Figure 5.56. Seismic example of a
turtle-structure anticline produced by
1 halokinesis. Northern Gulf of Mexico seismic
data courtesy of TGS.
2
3 Axial trace of
turtle structure
4
5
Depth (km)

6
7
8
Salt
9
10
11
12
V.E. × 2 0 10 km

to intersect it tangentially (Figure 5.57). The salt contact is Carruthers et al. 2013) (Figure 5.59). These fault patterns result
cuspate where intersected by tangential faults (Van Beukel from stress perturbation induced by diapiric doming. The
et al. 2000). Radial faults form in at least four ways. inward transition from polygonal to aligned faults records
First, some radial faults may have originated as crestal the outer limit of the former stress field around an actively
radial faults in a domed roof that has subsequently been rising stock at any one time (Carruthers et al. 2013).
pierced. An arched roof is typically somewhat wider than the Third, contracting diapirs may develop radial thrust faults,
underlying diapir, so the outermost parts of the crestal fault like the iris leaves of a camera when the f-stop is changed. For
system may be preserved in a salt-flank position even if most of example, Upheaval Dome in Utah, which some interpret as a
the roof is destroyed during breakthrough. If so, these remnant pinched-off diapir and some others as an impact crater, is
faults would have only small displacements and should not ringed by outer normal faults and lies at the hub of inner
extend far from the salt face. converging thrusts (Box 5.3) (Figure 5.60).
Second, some radial faults may be caused by hoop (circum- Finally, preexisting salt structures are mechanically weak,
ferential) strain around an expanding diapir (Figure 5.58(b)). so they are the first to deform in regional extension or
This expansion may occur as a flap rotates during active dia- shortening. Faults thus initiate at diapirs and propagate out-
pirism, or it may simply result from the outward force of salt ward, often connecting via relays with faults propagating out
pushing on its country rocks (Nikolinakou et al. 2014). Next to from nearby diapirs (Sections 11.2.1.3 and 12.5). Faults propa-
a radially expanding diapir, country rock typically undergoes gating out from isolated diapirs tend to form at high angles to
radial compaction, which diminishes outward; a compres- regional stress, but this tendency often gives way to a stronger
sional rim syncline might also form (Stewart 2006). While preference to follow the trend of buried salt ridges. As with
country rock is radially shortening, it stretches parallel to the faults formed by circumferential strain, normal or reverse
expanding diapir contact. This hoop extension is proportional faults formed as a result of regional deformation of preexisting
to the strike curvature of the salt contact (Stewart 2006), which diapirs have their largest displacements adjacent to salt.
is why radial faults cluster at the end of elliptical diapirs. Faults
formed by this process should have largest displacement next
to the diapir and should become smaller outward. 5.6.2 Ring Faults
The limit of hoop extension around a stock can be gauged Circumferentially trending ring faults are not common on salt
by the limit of polygonal faults created by compaction of flanks. Those few that exist form by inward collapse around
the overburden. Polygonal faults are confined to a mud-rich the constricting stem of a diapir (Branney 1995; Malthe-
interval, in which dewatering causes sediments to shrink verti- Sørenssen et al. 1999; Stewart 2006) (Figure 5.58(c)). Still other
cally and horizontally. The faults overlie and flank diapirs. concentric normal faults lie many kilometers distant from a
These polygonal fault patterns record an isotropic far-field diapir contact on the outer edges of large withdrawal basins
stress. Within about 2 km (or one to two diapir radii) of (Maione 2001). These distant faults form mainly by bending
a diapir contact, the polygonal pattern changes to a radial into the withdrawal basin rather than by diapiric constriction.
pattern centered on the diapir, or a parallel array connecting Localized on this hinge, the fault swarm dies out both inward
two adjoining stocks (Davison et al. 2000a; Stewart 2006; and outward from the diapir.

106
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.7 Diapir Pinch-Off

a Shallow structural level through roof a Coordinate


Axial

system
Radial

Hoop

Crestal
faults undiffer-
entiated
b
Radial
fault

ta
Tip of

s tr a
radial
fault

ed
om
fd

o
in
M arg Thrust
or
pin line

b Deep structural level through flanks

Radial fault

Rim syncline or
compactional halo

c
Tangential
Salt fault
Central
graben Radial
ta

thrust
s tr a
ed
om

Central
horst
fd

o
in
rg
Ma
Figure 5.57. Radial faults on a diapir flank commonly have a different
pattern and origin from radial faults in an arched diapir roof. (a) Shallow radial
faults are caused by hoop extension of the domed roof. (b) The curvature
of deep tangential faults indicates a changing balance between radial and
hoop stresses.
Area of diapir expansion Strain ellipse
or contraction
5.7 Diapir Pinch-Off Figure 5.58. Schematic maps showing types of radial faults and ring faults
There is strong evidence that many salt diapirs are pinched around expanding and contracting diapiric stocks. (a) Coordinate system for
axial, radial, and hoop directions centered on a diapir. (b) Faulting induced by
off from their source layers (Figure 5.61, Section 11.2.1.3). lateral expansion of a stock. (c) Faulting induced by lateral contraction of
A diapir can pinch off in two main ways, depending on a stock. Inspired by Stewart (2007).
whether the stem is tilted or vertical.
First, a diapir can be pinched off by halokinesis if its stem strata subside into the stem and pedestal of the diapir, they
tilts strongly (Figure 5.62). Strata in the hanging wall above the fold into an expulsion rollover dipping toward the diapir. Salt
inclined stem subside vertically under gravity. As hanging-wall expulsion creates space for more sediment in the hanging wall,

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
107
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

producing an isopach thick above the shrinking stem. Salt is


a displaced chiefly to the diapir’s crestal bulb, which expands
N upward and over the footwall of the diapir, which is little
affected by salt expulsion. As salt is expelled in this way, the
sides of the inclined stem converge and can eventually weld,
resulting in a pinched-off diapir.
The most robust way to pinch off a vertical diapir stem is
by shortening (Figure 5.63). During shortening, opposing
flanks of the diapir move toward each other until they touch,
North Pierce pinching off the stem. In cross section, this closure seems
straightforward. However, pinch-off requires a major change
in the shape of the diapir and adjoining country rocks in plan
view. The overburden along strike from the diapir must
shorten as much as the diapir does, so the strength of these
country rocks must be overcome in order for the diapir to
pinch off. Thus, one test for contractional pinch-off of a diapir
is to look for contractional structures in the nearby wall rocks
South Pierce (Section 11.2.1.3). Salt walls appear to be much easier to pinch
off than circular stocks, because the long central section of the
wall can pinch off without deforming the wall rocks, leaving
only the high-curvature ends of the walls to deform. Several
physical models and natural examples show walls that pinch
off in the middle, leaving pods of salt at the ends of the wall
(Figure 5.63).
If a stock were to pinch off, deformation in the encasing
0 2 km country rock would be complex and increase inward to the
shrinking stem. During pinch-off, the country rock undergoes
b bulk constriction: a unit sphere transforms into a prolate
N ellipsoid, like a vertical cigar. In the outer rim of deformation,
concentric extensional faults form where strata converge on
the closing stock. Some convergence could be by lateral com-
paction and vertical thickening, but the convergence is mostly
by radial thrusting. Like a camera iris, thrust plates override
each other as they converge, so that thrust heaves increase
inward to the diapir. A rare exposed example of inward-
converging thrust plates is Upheaval Dome in the Paradox
Fram salt basin (Box 5.3).
Can a vertical stem of a diapir pinch off during halokinesis?
For this to occur, the static pressure generated by sediments
pushing inward on a steep salt contact would have to exceed
the pressure of salt pushing out, and the pressure difference
would have to be enough to overcome the strength of the wall
rocks so that they would fail and move inward. This scenario is
unlikely, for three reasons (Heidari et al. 2016; Figure 5.64).
First, the pressure produced by the weight of the overlying
0 2 km units is likely to be greater in the salt than in the wall rocks.
Salt behaves like a viscous fluid, so stresses within it must be
Figure 5.59. Radial faulting above or around stocks is a response either isotropic. Thus, the horizontal stress pushing out on a vertical
to arching in the diapir’s roof or to hoop extension caused by outward salt contact must be equal to the vertical stress produced by the
pressure of diapiric salt against flanking sediments. (a) Radial faults above
North Pierce diapir and radial faults next to South Pierce diapir; dip map at weight of the overlying salt, which is ρsalt gz in an emergent
1,500-m-depth in East Central graben, North Sea; seismic data courtesy of Shell diapir. In the wall rocks, however, only a fraction of the
UK and Dan Carruthers. (b) Radial faults above Fram diapir, grading outward vertical stress is transferred to horizontal stress. This fraction,
into polygonal faults; two-way-time map at 1,600 ms in West Central graben,
North Sea; seismic data courtesy of Shell UK and Dan Carruthers. K, has a value of 0.7 to 0.9 for most conditions in sedimentary
basins. The horizontal stress of wall rocks pushing against a
avg
vertical salt contact is thus Kρsed gz. For the inward pressure of

108
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.7 Diapir Pinch-Off

109.933°W 109.925°W Figure 5.60. The constricted core of Upheaval


Dome (Utah) showing a radial thrust system in
which throw increases toward a highly deformed
center dominated by sandstone dikes injected
from below. After Jackson et al. (1998), who interpreted
38.440°N

this as the stem of a pinched-off stock.

N
38.435°N

Center 38.437°N, 109.929°W


Upper Chinle Fm. Upper Moenkopi Fm.
Lower Chinle Fm. Middle Moenkopi Fm. Contact
Fault 0 100 m
Moss Back Member Lower Moenkopi Fm.
of Chinle Fm. Turret Rock
White Rim Sandstone dikes

1 Seabed Figure 5.61. A pinched-off salt diapir overlain


by a salt canopy. The salt pedestal below the
2 pinched-off stem passes laterally into a primary
3 salt weld. Gulf of Mexico seismic data courtesy of CGG.
4
5 Top
Allochthonous salt
Depth (km)

6 Base
7
8
9
10
11
Salt
12 pedestal
13 Base autochthonous salt
14
V.E. × 1 0 5 km

wall rocks to exceed the outward pressure of salt, given a salt Second, salt pressure in a vertical diapir is further elevated
density of 2,200 kg/m3, the average density of sediments above by upward transmission of salt overpressures generated by
the pinch-off must exceed 2,450 kg/m3. This is above the upper loading the source layer. If the average density of sediments
limit for siliciclastic rocks in sedimentary basins. Thus, salt loading the source layer exceeds the salt density, then the
stress typically exceeds horizontal sediment stress, so, if any- source layer becomes overpressured with respect to salt. This
thing, vertical diapirs have a tendency to expand outward until overpressure is transmitted into any diapirs connected to the
checked by elastic resistance in the country rock. source layer, adding to the salt pressure imposed by the vertical

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
109
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Figure 5.62. An inclined stock can pinch off


a b as its hanging-wall overburden settles by gravity
and expels salt into an expanding bulb. Salt
Bulb displacement accommodates an isopachous thick
Bulb and an expulsion anticline in the hanging wall.
Stem

Pedestal

cKeystone graben d Expanded


bulb
Expanded

Keystone graben
bulb

Thick
Pinched-off
Ex stem
p
an ulsio
ticl
ine n
Sidewall fault
Expanding weld

Hinterland Foreland pinch off, it is not enough for horizontal sediment stress to
Strongly shortened Weakly shortened overcome the strength of the wall rocks; it must also overcome
Brick the outward pressure of the salt diapir.

Depth Section Brick 5.8 End of Diapir Growth


line
As discussed in Section 3.8, salt flow is driven by differential
Narrowest loading and resisted by salt supply and roof weight and
in center
strength. Diapirs initiate when driving forces exceed resisting
forces, and stop growing when the reverse occurs. For clarity,
what we mean here by “end of diapir growth” is that the
system achieves static equilibrium, so that salt is no longer
Regional shortening

Ellipse flowing. It is possible for diapirs not to grow taller even though
salt continues to rise, as, for example, when the top of the
Ellipse
diapir is vigorously eroded or dissolved.
Depth Section
line Most diapiric salt basins continue to evolve until the source
layer is largely evacuated, so it’s likely that depletion of salt is
Not welded the most common reason why diapirs stagnate. The ability of
at tips salt to flow out of a source layer into a diapir is limited by
viscous shear stress (boundary drag) near the top and bottom
Hourglass contacts of the salt layer. For a salt layer of given thickness, this
Bulbous end boundary shear zone is wide if the salt flow is Newtonian
Depth Section viscous (having dynamic viscosity independent of strain
Hourglass rate). The boundary shear zone is narrower and more intensely
Welded line
in center strained if the flow is power-law viscous, where viscosity
0 2 cm decreases by shear thinning as the rate of shear increases
toward the boundary. The resistance supplied by viscous shear
Figure 5.63. How a diapiric wall pinches off during regional shortening forces is less important for thick salt layers, where only a small
depends on its planform shape. A brick-shaped wall shortens more in the center percentage of salt thickness is intensely sheared along the
than at the ends, which are braced by enclosing sediments. An elliptical wall
is the most difficult to pinch off. An hourglass-shaped wall is most easily welded boundary layers. However, for thin salt layers, viscous resist-
in its center. Left column shows vertical sections through the center of each ance can immobilize the salt. Assuming Newtonian behavior,
wall; map on the right is a horizontal section. Physical model by Tim Dooley. the volumetric flux of laminar flow is proportional to the third
power of the layer thickness (Section 3.8.4.2). Thus, halving the
load of the salt in the diapir. The magnitude of the over- layer thickness retards flow by a factor of 8. For power-law
pressure decreases linearly upward. flow, drag resistance is smaller – although certainly present –
Third, the strength of the wall rocks resists any change in because of shear thinning. Once salt expulsion has reduced the
shape that would accompany pinch-off. Thus, for a diapir to thickness of a layer below a threshold thickness of, say, a few

110
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.8 End of Diapir Growth

Figure 5.64. The balance of static pressures


Sediment surface acting on the vertical sides of a salt wall precludes
the possibility that a vertical diapir can pinch off
z merely by static loading. Outward pressure from
Vertical
z Salt the diapir exceeds inward pressure from the
pressure Vertical country rock. After Heidari et al. (2016).
stress = stress =
ρavg gz due to
sed vertical ρsalt gz
load = Overburden
kρavg
sed
gz ρsalt gz load generates
source-layer
Horizontal stress overpressure if
is a fraction Raised salt pressure ρavg > ρsalt
of vertical stress in diapir due to sed
overpressured
source layer

Box 5.3 – The Upheaval Dome Controversy


Upheaval Dome is located in Canyonlands National Park, south- Impact crater hypothesis
eastern Utah. Spectacular canyon relief and superb exposures Almost all publications on Upheaval Dome agree that any resem-
make Upheaval Dome a popular attraction for tourists and geolo- blance of the present topography to an impact crater is merely
gists alike (Figure 5.65). Its present exposed diameter is about 5 fortuitous. Any crater was removed by Tertiary uplift and exhum-
km. The principal structures in Upheaval Dome are three circular ation of the Colorado Plateau. However, there is much disagree-
folds, which are concentric about the core of the dome: an ment as to the timing of impact and the size of the original
extensional outer rim monocline, an intermediate rim syncline, crater. The impact is variously estimated over a time range from
and a constrictional central uplift. as early as Jurassic to as late as the last few million years. The
Upheaval Dome is one of the most controversial geologic diameter of the modified crater is inferred to be as small as 5 km
structures in North America. Hypotheses for its origin include to as large as 9 km. Estimates of peak confining pressures range
subsurface salt diapirism, salt dissolution, cryptovolcanic explo- from <1 GPa to >10 GPa.
sion, meteoritic impact, and pinch-off of a salt diapir. A robust Various features at Upheaval Dome that were proposed as
genetic case can be made only for meteoritic impact and salt evidence of impact have since been discounted. Kriens et al.
diapir pinch-off. Table 5.2 summarizes the evidence for both (1997) described a few localized fan-shaped arrays of fractures,
hypotheses (see Jackson et al. 2001 for a detailed assessment which they called “shatter surfaces,” but they do not resemble
of the evidence). Many features, present or absent, support both true shatter cones produced by shock (Kenkmann 2003). Kriens
origins, and both the pinch-off and impact hypotheses have their et al. (1999) and Kenkmann and Scherler (2002) inferred planar
flaws, which has helped to keep the controversy alive. microstructures in quartz in the White Rim Sandstone, which

1 km Figure 5.65. This oblique view of Upheaval


Dome (Utah) shows the trace of its rim
Rim syncline surrounding its constricted core
sync N
line (white and dark-brown strata). Satellite image
from Google Earth.

1 km
© 2015 Google; Image Landsat

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
111
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Box 5.3 – (cont.)


Table 5.2. Evidence for pinch-off and impact hypotheses for Upheaval Dome

Observation Pinch-off favored Impact favored

Incompatible Compatible Compatible Incompatible


with impact with pinch-off with impact with pinch-off
POSITIVE EVIDENCE
Circularity X X
Central uplift X X
Clastic dikes X X
Crushed quartz grains X X
Inner constrictional zone X X
Outer extensional zone X X
Fracturing X X
Radial flaps (“dog tongues”) X X
Presence of underlying salt X X
Magnetic anomaly off-center X X
Contiguous anticline X X
Nearby salt structures X X
Rim syncline X X
Inward increase of deformation X X
Asymmetry of tectonic transport X X
Rim monocline X X
Growth folds X X
Growth faults X X
Shifting rim synclines X X
Truncations and channeling X X
Onlap X X
Multiple fracturing and cementation X X
Steep zones X X
Outward-verging extension X X
Volume imbalance X X
Gravity high X X
Pressures much higher than lithostatic X X
Planar deformation features X X
No seismic evidence of deep root X X
Difference from nearby salt structures X X
Statistical probability of impact X X
NEGATIVE EVIDENCE
Lack of salt at the surface X X
Lack of nearby piercement diapirs X X
Lack of meteoritic material X X
Lack of melt and pseudotachylite X X

112
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.8 End of Diapir Growth

Box 5.3 – (cont.)


Table 5.2. (cont.)

Observation Pinch-off favored Impact favored

Incompatible Compatible Compatible Incompatible


with impact with pinch-off with impact with pinch-off
Lack of widespread in situ breccia X X
Lack of ejecta X X
Lack of shock-metamorphic minerals X X
Lack of siderophile-element anomalies X X
Lack of outer fault terracing X X
Lack of overturned peripheral flap X X
Lack of shatter cones X X
Lack of diaplectic glass X X

forms clastic dikes in the central uplift. However, these micro- structure (Table 5.2). They include stratigraphic thickness
structures are actually thin deformation bands (Kenkmann 2003), changes, angular truncations, onlap surfaces, channeling, growth
which are common tectonic features on the Colorado Plateau. faults, growth folds, shale diapirs, and shifting rim synclines. This
Kriens et al. (1997, 1999) interpreted cobbles of vesicular quartz stratigraphic evidence indicates protracted growth of Upheaval
in the rim syncline as formerly molten impactite ejecta. However, Dome over at least 20 Myr, which is incompatible with geologic-
a study of the same samples by Koeberl et al. (1999) concludes ally instantaneous deformation during hypervelocity impact.
that these were merely hydrothermal quartz nodules unrelated The hypothesis for pinch-off envisages an emergent passive
to impact. Koeberl et al. (1999) also refuted an origin by impact stock less than 1 km in diameter that was surrounded by a gentle
for the clastic dikes, originally thought to be fluidized by transi- rim syncline in Pennsylvanian time. The inferred diapir remained
ent shock. An additional problem for the impact hypothesis is near the surface throughout the Permian, Triassic, and Early
the paucity of allochthonous breccia or in-place breccia, which Jurassic as sediments accumulated around it. Increasing sedi-
should be widespread at the structural level exposed today. mentary load on the Paradox source layer below increased the
Although they have been exhaustively examined for evidence flux of salt up the diapir. Abortive salt glaciers are inferred to
of impact, most of the quartz grains in the quartz-rich sandstones have spread from a passive salt stock during the Late Triassic and
at Upheaval Dome do not have shock features (Buchner and Early Jurassic, based on speculative evidence of small salt welds
Kenkmann 2008). In their study of 120 thin sections of Kayenta around the central uplift. The diapir widened over time, as
sandstones on the eastern flank of the central uplift, these authors recorded by concentric steep zones marking the limits of the
documented only three quartz grains containing shock-induced diapir, which migrated outward over time.
planar deformation features. The proportion of quartz grains The central uplift is inferred to be the toe of a convergent
showing shocked features is so minuscule that they are likely to gravity spreading system (as in the impact hypothesis). Radial
have been weathered from an impact crater in a distant source growth folds in the lower Wingate sandstones indicate that the
area, and even at a distant time, before being transported by walls of the diapir began to converge inward during the Early
Jurassic rivers to the area of Upheaval Dome. This possibility could Jurassic. The country rocks collapsed inward as the dome periph-
be tested by examining Kayenta sandstones well away from the ery extended radially. This movement produced intense
dome for similarly minute traces of planar deformation features. constrictional strain and structural thickening in the center of
Seismic reflection data should weigh heavily as evidence of the dome. Sediments in front of the extruding salt steepened to
the origin of Upheaval Dome. Interpreting a seismic line across near vertical. Salt extrusion accompanied pinch-off of the diapir
part of the rim syncline and the outer ring monocline, Kanbur et al. during the Middle Jurassic. By analogy with emplacement of
(2000) detected flat reflectors above the Paradox Salt and dis- allochthonous salt sheets on passive margins, the constriction
missed any possibility of a pinched-off salt pedestal. This may well of diapiric pinch-off would have forced salt from the shrinking
be so. However, greatly superior 3D imaging of pinched-off salt in diapir stem to extrude at the surface. During the Middle Jurassic,
the Gulf of Mexico shows that the base of a welded diapiric stem the allochthonous salt spread to the farthest extent. This formed a
need not have a pedestal large enough to be imaged. This would pancake-shaped extrusion about 3 km in diameter, whose limits
be especially true for a single 2D seismic line across rugged are recorded in the youngest and widest concentric steep zone.
topography and only covering a part of Upheaval Dome.
Conclusion
Diapir pinch-off hypothesis Given the excellent three-dimensional exposure of Upheaval
The strongest argument for diapirism is the wide range of Dome, it is remarkable that there is evidence of two such differ-
synsedimentary structures spatially restricted to this circular ent proposed origins. As is often so in a controversy, each faction

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
113
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

Box 5.3 – (cont.)

focuses on evidence favorable to its position and discounts pinched-off salt diapir. Either way, the complex exposures pro-
adverse evidence, typically from a simplistic perspective of pro- vide a stark warning to petroleum geoscientists that impact
cesses with which it is unfamiliar. The significance of Upheaval craters or pinched-off diapirs are fraught with exploration risks.
Dome is that it is likely to display one of the world’s best Key references: Kriens et al. (1997, 1999), Jackson et al. (1998, 2001),
three-dimensionally exposed roots of either an impact crater or a Kenkmann et al. (2005).

tens of meters, the salt flows little, even if a large differential


load is applied. In addition, circulating groundwater can
a Widening
b
passive Extension
Extension
dissolve the remaining salt to form a complete salt weld diapir Sagging crest
(Section 9.5.2).
The other resisting factor that can terminate diapirism is
the strength and weight of the diapir roof. If the roof is too Widening
thick or strong, then the diapir stops rising even if there is still diapir
salt left in the source layer. The roof typically thickens and Abundant salt supply Restricted salt supply
strengthens by aggradation, although, in contractional settings, Figure 5.66. The influence of extension on diapirism depends on the diapir
thrusting can place a competent roof over the top of a diapir. If geometry and salt supply. (a) After a diapir has broken through to the surface to
a diapir is trapped in the subsurface by a strong roof, then rise passively, regional extension widens the diapir. The salt can still keep rising
as long as salt supply remains abundant. (b) Once salt supply becomes
anything that disrupts this roof can restart diapir rise, some- restricted, continued extension widens the diapir and leads to diapir fall.
times very rapidly if the salt is overpressured.
Diapirs can also stop rising if driving forces decrease,
although this is less common than increasing resistance forces. be circulated against a salt face. Dissolution only causes a
The most common example of this is in reactive diapirism, diapir crest to subside when dissolved salt is not being replaced
where diapir rise is driven by differential loading caused by by the rising diapir.
extensional thinning of the diapir roof. At each stage in the
reactive process, the diapir rises until equilibrium is reached,
and then it stops. Further extensional thinning is required to 5.10 Dissolution of Salt Diapirs
perturb this equilibrium and create additional driving force for All evaporites are soluble, so salt structures are easily dissolved
the diapir to continue rising. by undersaturated surface water or groundwater. Diapirs near
the surface are particularly prone to salt dissolution. The
5.9 Diapir Fall following section deals with the effects of dissolution on land.
The next section covers dissolution under water. A third
Because diapirs by definition rise through younger rocks, it
section describes how dissolution forms cap rock.
seems counterintuitive for a diapir crest to subside. However,
there are at least three situations in which a diapir can fall.
First, although extension can cause diapirs to rise reactively 5.10.1 Dissolution on Land
if salt supply is plentiful (Section 5.2.1), extension can also Only the broad facts were known about subaerial dissolution
cause diapirs to fall once the source layer is thin or depleted of salt diapirs until work on the geomorphology of Zagros
(Section 10.2.4.2). Stretching a diapir faster than the source diapirs by geologists at Charles University (Prague) and their
layer can supply new salt causes a diapir crest to subside – akin Iranian colleagues (Bosák et al. 1998; Bruthans et al. 2000,
to a drop in water level if the walls of a container are moved 2006, 2009, 2010) and by Talbot (in a series of papers, synthe-
apart (Vendeville and Jackson 1992b; Figure 5.66). sized by Talbot 1998).
Second, part of a diapir can fall if salt flows from it to The geomorphology of salt diapirs is affected by climate
another part of the same diapir. For example, parts of a salt and by the vigor of diapiric rise. An actively rising salt plug
wall may subside to allow some culminations within the wall to has positive relief, a periclinal stream network, and well-
keep growing (Figure 5.67). Another example is the subsidence exposed evaporites. A salt plug that recently became inactive
of minibasins into a salt sheet (which is a type of diapir, has only relict exposures of salt, and a substantial gypsum
Section 7.3), allowing other parts of the salt sheet or canopy mantle. A diapir that has long stopped rising degrades to a
to inflate or extrude. salt ruin, which is a flat or slightly sunken area dotted with
Finally, salt diapirs can fall by dissolution collapse (Section rounded hills of exotic blocks and having negligible rock salt
5.10). Such structures can be impressive at the surface but are at the surface.
usually small on the scale of the entire diapir. Furthermore, In contact with fresh water or undersaturated brine, rock
large-scale dissolution is usually restricted to the near-surface, salt dissolves to form striking geomorphic features (Figure 5.68).
in the zone where a large volume of undersaturated water can Surface-lowering rates in salt outcrops are commonly several

114
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.10 Dissolution of Salt Diapirs

Rising
Graben above culmination
sinking part of wall of salt wall
Rising
culmination
of salt wall

n
rb urde
Ove flow
lel salt
e - paral Extension
Salt k
Stri
Dip-parallel salt flow Fallen diapir
Regional extension
Figure 5.67. Commonly different parts of a salt wall behave differently during regional extension. Depressions in the wall tend to fall because they are overlain by the
thickest overburden, which drives salt along strike to culminations in the wall. These culminations therefore rise more vigorously. After Vendeville and Jackson (1992b).

a b

Bedding d
c

Figure 5.68. Exokarst and endokarst structures are common in the crest of dissolving salt diapirs in the Zagros (Iran). (a) Centimeter-scale rillenkarren
(solution grooves) separated by sharp ridges and spires in mylonitic salt of Kuh-e-Namak (Bushehr Province). (b) Meter-scale sinkholes revealed by removal of
residual gypsum soil before salt quarrying; Garmsar salt nappe, Great Kavir (Iran). (c) Entrance to a cave into Namakdan diapir, Qeshm Island (Iran). (d) 3N Cave,
Namakdan diapir (Iran). Photographs (a)–(c) by Martin Jackson, (d) courtesy of Jiří Bruthans.

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
115
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

centimeters per year and may reach more than 10 cm/yr. faults should be carefully evaluated in inverted basins, where
As near-surface salt dissolves, its sediment mantle is breached they might have regional causes.
by exokarst and endokarst landforms that are similar to Most diapirs are 95 percent covered by surficial sediments a
those in limestones but evolve much more rapidly. Sinkholes few decimeters to several tens of meters thick. Most commonly
merge to form badlands, which form closed depressions these sediments are a weathering residue formed in the unsat-
having sediment-filled bottoms as much as several kilometers urated zone. The residue is typically a gypsum–anhydrite mix-
wide. Some sinkholes lead into caves or act as swallow holes for ture, and a wide range of sedimentary, igneous, and
streams draining underground at the end of blind valleys and metamorphic rocks. The next most common sediments on
in large depressions. Caves vary widely from small embryonic subaerial diapir crests are fluvial. The weathering residue is
cavities to large multilevel systems. The world’s longest salt reworked to form stratified and sorted fluvial deposits in
cave is the 6,580-m-long 3N Cave in Namakdan diapir. This valleys, flat surfaces, and sinkholes on diapirs.
cave contains at least 9 km of explored passages created in the
last three to six thousand years by episodic streams after heavy 5.10.2 Dissolution under Water
rainfall on the normally arid Qeshm Island (Iran).
The Mediterranean Sea and the Red Sea contain many depres-
Sinkholes that form beneath diapir roofs are commonly
sions attributed to salt dissolution (Ross and Uchupi 1973;
rimmed by extensional faults. Physical modeling suggests that
Schoell et al. 1974; Bertoni and Cartwright 2005, 2015). The
this is only part of the kinematic picture and that the most
superheated “black smokers” venting mineral-laden brine
diagnostic structures are contractional thrusts near the middle
from the abyssal Red Sea have been attributed to hydrothermal
of the collapse (Ge and Jackson 1998). Conical, concentric
circulation of seawater leaching metals from new oceanic crust.
rings of inner contraction and outer extension form above a
Some of these salts and dissolved metals may have been
subsiding salt stock, regardless of the shape of the subsiding
derived from dissolution of evaporites.
diapir’s crest (Figure 5.69). The zones of extension and con-
Brine dissolved by salt diapirs mostly dissipates as it
traction balance each other. The contraction differentiates
mixes with seawater. However, on reaching a closed topo-
collapsed roofs from those stretched by regional extension.
graphic basin, the dense brine settles and forms a long-lived
However, the diagnostic inner contractional zone is likely to
submarine brine lake if undisturbed by bottom currents.
be obscured by colluvium and alluvium. The origin of reverse
A prime example of a brine lake is the Orca basin, a mini-
basin on the continental slope of the northern Gulf of
Mexico (Shokes et al. 1977; Trabant and Presley 1978;
a Undeformed rectangular wall bExtensional
Subsided rectangular wall
Extensional Pilcher and Blumstein 2007). A striking seismic reflector
zone zone marks the top of the dense brine lake and the base of normal
Contractional
zone seawater (Figure 5.70).

Salt wall
5.10.3 Cap Rock
Cap rock has several geologic meanings that refer to a layer
c Undeformed semicircular wall dExtensional
Subsided semicircular wall that is either more resistant to weathering or less permeable
zone
Extensional
zone
than the rock it caps. In salt tectonics, though, cap rock has a
Contractional specific designation as rock formed at the crest of many diapirs
zone as halite dissolves and the residue is chemically altered. Cap
rock has been extensively studied because it can contain eco-
nomic minerals, especially sulfur and base-metal sulfides.
Salt wall
Moreover, vugs and fractures in cap rock also provide petrol-
eum reservoirs, as in the epochal Spindletop oil discovery in
e Undeformed triangular wall fExtensional
Subsided triangular wall
Extensional
1901, which was the first over a diapir, and the world’s most
zone zone productive at the time (Box 1.3). Most research on cap rocks
Contractional
zone has been on U.S. Gulf Coast diapirs, where cap rock is espe-
cially widespread. This section draws heavily from Feely and
Kulp (1957) and Posey and Kyle (1988).
Salt wall Cap rock forms as diapiric evaporites are altered by warm,
saline, formation fluids from deep-basin sources mixing with
0 5 cm Salt weld
cool, dilute, meteoric waters. As halite dissolves, the crest of
Figure 5.69. As salt diapirs dissolve, their roofs collapse. Cross sections show the diapir and all its internal structures are truncated to form
(a, c, e) initial geometry and (b, d, f) effects of salt subsidence on the roofs a near-horizontal solution table. Salt dissolves inward from the
of diapiric walls of three basic profile shapes in physical models. In each shape,
a central contractional zone is bounded by a balancing extensional zone. top and upper sides of the diapirs. Cap rock is thickest on
After physical models by Ge and Jackson (1998). the crest of diapirs and thins down the diapir’s shoulders as

116
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008
5.10 Dissolution of Salt Diapirs

West East

Orca minibasin
2.5 Rotated Brine Exposed salt
extensional runoff
faults
Brine lake
3.0
Two-way time (s)

Thin roof
Chaotic, slumped interval
3.5
Salt diapir

4.0
Salt diapir
Minibasin
4.5

5.0
0 5 km

Figure 5.70. Where brine leaks from dissolving salt diapirs into a closed depression, the brine ponds, forming salt flats in deserts or brine lakes under water.
The seismic profile shows a brine lake in Orca minibasin, northern Gulf of Mexico. The diapir on the eastern flank of the minibasin has a concave summit where salt
is partly exposed by extensional slumping and dissolution, which supplies brine for the lake. After Pilcher and Blumstein (2007).

meteoric waters grade with depth into saltier compactional and


thermobaric waters. Only salt domes shallower than 2 km have
anhydrite cap rock in the U.S. Gulf Coast. Negative relief is
absent over any Texas dome that either is deeper than 600 m or
has cap rock thicker than 200 m.
Cap rock typically has two main zones: an anhydrite cap
and a calcite cap (Figure 5.71). The microlaminated anhydrite
cap is the residue of halite dissolution. The cap can be thicker
than 300 m. It rests directly on salt or overlies cavities of
dissolved salt in the crest of the diapir. Anhydrite layers
become younger downward because anhydrite forms by
underplating as halite dissolves at the base of the cap rock.
As salt dissolves, the salt diapir can compensate by
continued rise.
Calcite replaces the upper or outer parts of the anhydrite
cap, probably from the top downward. A thin transitional
gypsum zone forms by hydration of anhydrite. Above the
gypsum the calcite cap generally comprises two rock types
(Figure 5.71): an upper variegated limestone and a lower
banded limestone. The upper variegated limestone formed
early and commonly contains siliciclastic and variegated car-
bonate clasts in a massive and fine-grained carbonate matrix.
Only about half of Gulf Coast cap rocks contain the lower
banded calcite or “zebra-textured” zone. The lower banded False calcite cap Gypsum (transitional) cap
zone is lithologically similar to the variegated limestone, except
Marine calcite cap Anhydrite cap
that the calcite phases form subhorizontal bands crosscut by
younger calcite veins and vugs. Limestones termed false calcite Variegated calcite cap Halite dissolution (anhydrite
sand and cavities)
and marine false calcite can form above the variegated and
banded calcite zones (Figure 5.71). False calcite is composed of Banded calcite cap Salt stock
calcite-cemented clasts from wall rocks. Marine calcite forms Figure 5.71. Different stages of alteration produce a zoned cap rock, which is
on topographic bulges above salt diapirs. typical of U.S. Gulf Coast diapirs. After Posey and Kyle (1988).

Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
117
https://doi.org/10.1017/9781139003988.008
Salt Stocks and Salt Walls

As the alteration front moves downward, the cap rock cap and native sulfur are commonly on the basinward side of
reacts in complex ways. In the transitional gypsum zone, salt domes. Free sulfur and hydrogen sulfide are formed by
anhydrite is hydrated to gypsum and dissolved in water. The numerous types of anaerobic sulfate-reducing bacteria, chiefly
freed sulfate ions mix with organic compounds and hydrocar- Desulfovibrio desulfuricans. Sulfate-reducing bacteria flourish
bons, including biogenic or thermogenic methane, and are at depths generally less than 750 m and temperatures below
altered by sulfate-reducing bacteria to hydrogen sulfide, cal- 60 °C, given an energy source. Some twenty-four onshore
cium carbonate, and water to form the calcite cap. The textural domes in the Gulf Coast have produced sulfur commercially
variety of the calcite cap indicates that preexisting sediments, and contain the biggest deposits of native sulfur in the world.
topography, and timing of calcite formation affect the oxida- Even more sulfur is recovered from petroleum refining and
tion to calcite, which can preserve original anhydrite textures. from sour natural gas.
As anhydrite dissolves, the calcite–anhydrite interface moves Iron sulfides, sphalerite, galena, barite, celestite, and stron-
downward. As anhydrite dissolves, voids open and collapse to tianite are common in the calcite and anhydrite caps of Gulf
form breccias cemented by younger phases of calcite. Coast diapirs; rare acanthite, realgar, and uranium minerals
The hydrogen sulfide is ultimately oxidized to native are also known. This mineralization may be related to that
sulfur by several intermediate reactions. Native sulfur forms in salt diapirs in northwest Africa (Rouvier et al. 1985) and
throughout a cap rock but is generally concentrated in the to Mississippi Valley-type and sedimentary exhalative metal
lower part of the calcite cap. Thickest accumulations of calcite deposits (Kyle and Price 1986).

118
Downloaded from https://www.cambridge.org/core. University of Leeds, on 17 Jun 2019 at 15:53:27, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
https://doi.org/10.1017/9781139003988.008

You might also like