You are on page 1of 26

Kinematics and Dynamics of Salt Tectonics Driven by Progradation1

Hongxing Ge,2 Martin P. A. Jackson,3 and Bruno C. Vendeville3

ABSTRACT Scaled physical models illustrate the importance of progradation as a trigger for salt tectonics and formation of allochthonous sheets. Regional extension and contraction were excluded in the models. In our experiments, prograding wedges above a tabular, buoyant salt layer with a flat base expelled the salt basinward, forming the following structures proximally to distally: (1) sigmoidally distorted initially planar wedges, (2) relict salt pillows and salt welds, (3) basinward-dipping expulsion rollover and crestal graben, (4) rollover syncline, (5) landward-facing salt-cored monocline, and (6) distal inf lated salt layer. This deformation zone amplified and advanced basinward during progradation; however, no diapiric salt structures formed. Over a buoyant salt layer whose basement had steps facing landward, progradation initially formed a broad anticline where salt flow was restricted across each basement step. Distal aggradation pinned the anticline and enhanced differential loading. The anticline actively pierced its crest, which

had been thinned by faulting and erosion. Thereafter, the diapir grew passively, locally sourcing allochthonous salt sheets. This deformation cycle repeated over each basement step so that the age, amplitude, complexity, and maturity of salt-related structures decreased basinward. As each allochthonous salt sheet was buried and evacuated by sediment loading, arcuate peripheral normal faults formed along the sheets trailing edge, detached wrench faults formed along its lateral edges, and active piercement at its leading edge allowed the sheet to break out and climb stratigraphic levels. This process formed a multitiered complex of salt sheets that migrated basinward with time. Restorations of examples from various salt tectonic provinces support our model results. Immense landward-dipping pseudofaults could arise entirely by salt expulsion rather than regional extension. Diapiric families and peripheral sinks are reinterpreted as the result of progradation squeezing salt basinward and laterally. INTRODUCTION

Copyright 1997. The American Association of Petroleum Geologists. All rights reserved. 1Manuscript received October 2, 1995; revised manuscript received May 9, 1996; final acceptance October 16, 1996. 2 Bureau of Economic Geology, University of Texas at Austin, and Department of Geological Sciences, University of Texas at Austin, Austin, Texas 78713. 3Bureau of Economic Geology, University of Texas at Austin, Austin, Texas 78713. All modeling was done at the Applied Geodynamics Laboratory of the Bureau of Economic Geology, with financial support by grant number 3658178 from the Texas Advanced Technology Program and from the following companies: Agip S.p.A, Amoco Production Company, Anadarko Petroleum Corporation, ARCO Exploration and Production Technology, BP Exploration, Chevron Petroleum Technology Company, Conoco and Dupont, Exxon Production Research Company, Louisiana Land and Exploration Company, Marathon Oil Company, Mobil Research and Development Corporation, Petroleo Brasileiro S.A., Phillips Petroleum Company, Socit Nationale Elf Aquitaine Production, Statoil, Texaco, and Total Minatome Corporation. The Department of Geological Sciences and the Geology Foundation at the University of Texas at Austin and Phillips Petroleum Company provided additional financial support for Hongxing Ge. Dan Schultz-Ela helped us depth convert and restore seismic sections. Mark Rowan, Mike Hudec, Lee Fairchild, Sharon Mosher, and Tucker Hentz provided invaluable suggestions for improving the paper. Publication authorized by the Director, Bureau of Economic Geology, University of Texas at Austin.

Salt structures can be triggered by a variety of mechanisms. Several studies have emphasized the role of regional extension in the initiation and growth of salt structures (Duval et al., 1992; Vendeville and Jackson, 1992; Demercian et al., 1993; Jackson and Vendeville, 1994). The triggering effect of sedimentary differential loading has been invoked more often (e.g., Trusheim, 1960; Dailly, 1976; Seni and Jackson, 1983a, b; Jackson and Talbot, 1986, 1991; Worrall and Snelson, 1989; Nelson, 1991) but has received less rigorous attention, especially by workers using scaled physical modeling. Lateral variations of thickness or density of overburden induce sedimentary differential loading, which triggers two types of deformation that occur separately or together. First, the thickest sediments sink deepest into the underlying salt, expelling it laterally, which thins or entirely evacuates the salt, forming a weld. Second, the sediment
AAPG Bulletin, V. 81, No. 3 (March 1997), P. 398423.

398

Ge et al.

399

wedge itself can spread seaward by means of extensional faulting on the shelf and upper continental slope and shortening by thrusting and folding in the lower slope. Progradation of alluvial fans and deltas along the margins of many terrigenous clastic basins is a common and effective way to create differential loading. A prograding system commonly comprises a fairly flat shelf, a shelf break, a middle and upper slope dipping between 0.3 and 5 (e.g., Galloway, 1989; Worrall and Snelson, 1989), and a gently dipping prodeltaic zone of slow, fine sedimentation. Salt diapirs are common in prograding settings, such as in the Paradox Basin (Baars, 1966; Ge, 1996), Gulf of Mexico (Humphris, 1979; Woodbury et al., 1980; Worrall and Snelson, 1989; Wu et al., 1990), west Africa (Duval et al., 1992; Lundin, 1992), and Brazil (Mohriak et al., 1995). However, little is known about the kinematics and dynamics of this relation. Apart from brief descriptions of onlaps rotated into apparent downlaps owing to withdrawal of a laterally migrating mound of salt (Jackson and Cramez, 1989; Wu et al., 1990), we know of no published criteria to distinguish the role of progradation from other mechanisms capable of triggering diapirs. Rettger (1935) was the first to experiment with sedimentary differential loading. He loaded a sand wedge onto a uniformly thick layer of laminated clay under water. Lateral flow of clay thinned it below the thickest part of the wedge. A listric normal growth fault formed below the sand wedge, and minor folds formed in front of the wedge. Synkinematic progradation effects were not examined. Similar structures were experimentally produced by McKee and Goldberg (1969), who used sand wedges dipping at 30 to create serial folds and thrusts in the underlying mud sequences by differential loading. However, the experiments were not scaled to simulate salt tectonics. Centrifuged models by Jackson and Cornelius (1987) and Talbot (1992) simulated progradation on syndepositional salt structures. They used stiff, viscous, or power-law fluids to simulate overburden and viscous fluids to simulate salt. Salt structures were formed by differential loading and gravity spreading, but the fluid behavior of the overburden is unrealistic for the shallow crust where salt tectonics is concentrated (Vendeville and Jackson, 1993; Weijermars et al., 1993). Moreover, for practical reasons, the models used either tabular or unrealistically steeply dipping prograding increments. Nevertheless, these experiments produced some useful insights that were corroborated by our modeling. First, salt was expelled to distal regions from beneath the prograding wedges. In front of the advancing wedges, the source layer was greatly thickened over a broad area, rather than forming a

diapir next to the progradation front. Second, basinal salt pinch-out and basement steps provided effective barriers to lateral salt flow, causing local salt accumulation and diapirism. The modeling results were compared with allochthonous salt sheets of the Gulf of Mexico region. Experiments by Vendeville et al. (1994) examined the effects of sand wedges prograding over a viscous silicone analog for salt. Because the silicone layer was thin or denser than the overburden, deformation was dominated by extension and seaward spreading of the sedimentar y wedge. Extension formed turtle structures separated by small, sagging diapirs in the footwalls of asymmetric grabens. In contrast, our study uses physical models to investigate the effect of prograding wedges advancing over a thick, buoyant salt layer, a setup that typically opposes spreading and extension of the sediment wedge. New concepts illustrate how progradation triggers salt tectonics and how deformation is dominated by sediment subsidence and lateral expulsion of salt where the salt is thick and much less dense than its overburden. We demonstrate how basement steps and distal aggrading sediments control diapir location. The modeling results are used to reinterpret salt tectonics provinces in the South Atlantic, Gulf of Mexico, and Germany. METHODOLOGY Experimental Materials Physical models are simplified laboratory replicas of natural systems. Their usefulness increases with the degree to which they are dynamically scaled to their geologic equivalents. Dynamic scaling theory has matured from the solid foundation laid by Hubbert (1937) and has been successfully adapted to salt tectonics (e.g., Ramberg, 1981; Vendeville et al., 1987; Weijermars et al., 1993). Our models use dry quartz sand to simulate brittle sedimentary rocks. The sand shows near-perfect Mohr-Coulomb behavior and deforms along narrow shear zones, which mimic fault zones (Mandl, 1988; Krantz, 1991). The sand has negligible cohesion, an angle of internal friction of about 30, and a density of about 1700 kg/m3. Sand is a good analog for most sedimentary rocks in the upper continental crust, which obey Mohr-Coulomb behavior (Byerlee, 1978; Weijermars et al., 1993). We used a viscous silicone polymer (RG20, a transparent silicone polydimethysiloxane [PDMS]), and a gray, dense (1390 kg/m3), viscous Silbione gum to simulate damp rock salt that is assumed to be a perfect Newtonian viscous fluid, with a viscosity ranging

400

Progradation and Salt Tectonics

Figure 1Schematic, vertically exaggerated cross section of experimental setup and depositional systems. The base of the tabular salt layer of model 253 is shown as a dashed white line. Aggrading layers are numbered the same as coeval prograding wedges.

from 1017 to 1019 Pas, depending mainly on grain size (van Keken et al., 1993). The silicone, RG20, is an almost perfect Newtonian fluid, with a viscosity of about 6 10 4 Pas at a strain rate of 10 4 /s at room temperature and a density of about 950 kg/m3. Its properties are very similar to those of other PDMS products, such as SGM36, that have been used frequently to model salt tectonics (Weijermars, 1986; Weijermars et al., 1993). The gray, dense Silbione gum also has Newtonian viscous behavior. Its viscosity was reduced by a solvent to a value estimated to be similar to that of the other silicone. These two salt analogs were used in different experiments to investigate the influence of density contrast on salt tectonics. The scale ratios between models and geologic equivalents (prototypes) are * 0.50.8 (model materials had densities slightly lower than those of natural rocks, but the density difference in the model was higher than in nature), g* = 1 (models and geologic prototypes both deformed in a natural gravity field), and l* = 105 (1 cm in models represents approximately 1 km in nature). The ratios of time (t *) and associated rates of deformation ( D* ), progradation ( P *) and aggradation ( A*) depended on the viscosity ratio (*). Because natural rock salt has a wide range of viscosity values (van Keken et al., 1993), these time-relevant model ratios also vary widely, so their natural equivalents are not particularly meaningful. Hereafter, the silicone analogs for salt will be referred to as salt for brevity. Experimental Procedures All models were contained in a rigid box 80 cm long and 40 cm wide. Figure 1 is a schematic cross section of undeformed models to show the initial geometry of the salt basin and overburden strata. In the first experiment (model 253), both top and

base of the source layer were f lat (unlike the stepped base shown in Figure 1). The salt source layer was uniformly 1.8 cm thick. In the other experiments (models 263, 270, and 272), the top of the salt was flat, but the base had two steps dipping 60 landward. The steps simulated inactive normal fault scarps in a block-faulted basin filled by salt. Step 1 was 1.0 cm high, and step 2 was 0.5 cm high. The proximal salt was 2.0 cm thick and thinned across each basement step to 0.5 cm distally (Figure 1). The models were then covered by a prekinematic overburden layer uniformly 0.30.6 cm thick. Progradational and aggradational layers were then added episodically to simulate synkinematic sedimentation (Figure 1). The progradational, wedge-shaped sediments had either a clinoformal slope (layers 2, 3, and 4) or a clinoform with a proximal shelf (layers 5, 6, 7, 10, and 11), or a proximal bypass surface (layers 8 and 9). All clinoforms were initially planar and dipped at 5. The wedges prograded from right to left and downlapped basinward (left). We used linear fronts in our models. Although the progradation front of an individual delta or alluvial fan is commonly lobate and curved, single-sourced lobes can shift or coalesce to form a regionally linear or nearly linear progradation front. The Cenozoic deltaic continental margins of the northwestern Gulf of Mexico (Winker, 1982), the Cenozoic Ogallala Formation of the piedmont alluvial deposits in the Great Plains province of the central United States (Frye, 1971), and the present alluvial fans in Death Valley, California (Bloom, 1991), are typical examples of sediment wedges with nearly linear fronts. In our models, the mean rates of progradation ranged from 0.5 to 5.3 cm/day (Figure 2). The distal aggrading layers, however, were tabular except where they thinned and onlapped landward against growing salt structures (Figure 1). They aggraded concurrently with only a few progradation wedges.

Ge et al.

401

Distal
s2 20

Progradation
100 80 s2

Proximal

Model 272

s1 15 s2

60 40 20 0 70 s1

Basement step 1 Basement step 2


s1

Figure 2Plot of distance of progradation front from proximal end wall of models (x axis) vs. time (y axis). The mean rates of progradation ( P) were calculated by excluding the first wedge. A steep slope indicates slow P. The inset (upper right) shows the complete plot for model 272 at reduced scale.

Time (day)

Time (day)

35

Distance (cm)

10

s2

s1 5

Model 253, P = 1.4 cm/day Model 263, P = 1.6 cm/day Model 270, P = 5.3 cm/day Model 272, P = 0.5 cm/day
70 60 50 40 30 20 10 0

Distance from proximal edge (cm)

The mean rates of aggradation ranged from 0 to 0.3 cm/day. Restorations All restorations in this paper used the software, Restore (Schultz-Ela; 1992; Schultz-Ela and Duncan, 1994). Model sections required no depth converting or decompacting. Seismic sections were approximately depth-converted before restoration where possible, using velocity profiles selected to fit the age and lithology of the strata, then adjusted if necessary to conform with other authors depth conversions or well data. Decompaction assumed a shale/sand ratio of 3:1 to 3:2. During backstripping, we unfolded strata using vertical shear because of the paucity of faults and because, in the models at least, no regional extension occurred. We use the term salt expulsion wherever salt is expelled along strike or downdip by overburden differential loading, especially by prograding wedges. This process may or may not involve a change of the cross-sectional area of salt. We use the term salt reduction where salt area has obviously been lost from a cross section by any process. EXPERIMENTAL RESULTS Progradation Over a Tabular Salt Layer Model 253 investigated progradation over a tabular salt layer. Figure 3 shows the initial setup of the

three uniform prekinematic layers (including salt) and the chronostratigraphy for the overlying synkinematic wedges (layers 210). Each thick, horizontal line represents the length of a wedge deposited episodically to prograde from right to left at an average rate of 1.4 cm/day for 20 days. No aggrading layers were deposited. Blank regions represent nondepositional hiatuses. Structural Style and Kinematics The final structure (Figure 4) of the deformed model comprised two domains separated by a landward-facing monocline: (1) a proximal region where prograding wedges overlay a residual low salt pillow and salt weld, and (2) a distal region where the source layer was uniformly thickened. At the proximal end wall, an old reactive diapir (lower right in Figure 4b) flanked by normal faults was buried under the prograding wedges. The synkinematic wedges downlapped onto the prekinematic layer where salt was largely evacuated, whereas the wedge tips onlapped against the monocline. The deformed offlapping wedges were sigmoidal; proximal planar surfaces steepened distally, forming a rollover antiform. Unlike a rollover in the hanging wall of a listric normal fault, the rollover in our model was not induced by extension, but by basinward expulsion of salt beneath the prograding wedges. To sharpen this difference, we refer to this structure as an expulsion rollover. A keystone graben formed at the crest of the expulsion rollover. Distal faults in the graben dipped

402

Progradation and Salt Tectonics

Distal
P = 1.4 cm/day
20 16
Time (day) Starved basin Progradation front
10 9 8

Proximal Progradation
11

A=0

12 8 4 0

7 6 5 4 3 2 1

Basement
Viscous layer (salt analog) 1 Prekinematic layer 2-10 Synkinematic layers

10 cm

11 Postkinematic layer

Figure 3Chronostratigraphy for model 253. The lowermost three layers show the initial setup. Each episodically deposited synkinematic layer is plotted as a numbered, thick, horizontal line. Layer numbers correspond to those in following cross sections (Figures 4, 6). The uppermost layer (11) is the postkinematic layer deposited after the experiment ended. A = mean rate of distal aggradation; P = mean rate of progradation.

more gently than proximal faults because of rotation as the wedges bent. The sigmoidal clinoforms thickened distally and then thinned, forming a rollover syncline above the monocline.

Expulsion Rollover An expulsion rollover (Figure 4) superficially resembles a fault rollover in the hanging wall of a

(a) Distal

Proximal

5 cm

(b)
Salt thickens and lifts overburden Rollover syncline

P = 1.4 cm/day
Expulsion rollover
11 10 0 9

4 3 2 1

Regional datum Early depocenter Proximal faults


10 cm

Basement Rolling monocline Viscous layer (salt analog)

Salt weld

Relict pillow

Prekinematic layer

2-10

Synkinematic layers

11

Future layer

Figure 4(a) Close-up of a cross section through model 253. The anticline in front of the wedges was an artifact induced by salt sagging during section cutting; the actual structure was a monocline. (b) Tracing of a regional cross section of model 253. The dashed horizontal line represents the top of the initially tabular salt layer. The rolling monocline, rollover syncline, and rollover all migrated basinward in tandem with the progradation front. Large black dots show data points plotted in Figure 5a.

Ge et al.

403

(a) Model expulsion rollover


Dip of rollover flank 20 10 0 0 5 10 15 20 25 Distance from progradation front (cm) 30

(b) Model fault rollover


Dip of rollover flank

60

Figure 5Rollover plots of the maximum dips of rollovers. (a) Experimental expulsion rollover (model 253). (b) Experimental fault rollovers associated with simple listric normal faults (profile A, model 188 by B. C. Vendeville and S. T. Lin; profile B, experiment E30 by McClay, 1990). (c) Six geologic fault rollovers. Brazos Ridge, Texas (Christensen, 1983), and Snake Mountain, Nevada (Robison, 1983), are depth-converted sections; fault system 3, Texas (Bruce, 1983); Gulf of Guinea (Gaullier et al., 1993); and Grand Isle, Louisiana (Larberg, 1983), are seismic time sections. (d) Experimental fault rollovers having multiple ramp/flat fault geometry (profile A, experiment E10, McClay, 1989; profile B, experiment E48, McClay, 1990). Open data symbols indicate prekinematic (nondiagnostic) and filled data symbols indicate synkinematic (diagnostic) strata.

30

A B

0 0 5 10 15 20 Distance from master fault (cm) 25

Dip of rollover flank

(c) Geologic fault rollover 90 Brazos Ridge 1, TX Brazos Ridge 2, TX Grand Isle, LA 60

Gulf of Guinea Fault system 3, TX Snake Mountain, NV

30

0 0 5 10 Distance from master fault (km) 15

Ramp

Ramp

(d) Model fault rollover 90


Flat Flat A

60

A B

30

0 -20 Ramp 0 5 Flat Ramp Flat B 30

10 15 20 25 Distance from master fault (cm)

listric growth normal fault. However, dip analysis of the rollover (rollover plot) provides an objective and seemingly robust technique to differentiate them. Figure 5a plots the maximum dip of each wedge in the expulsion rollover vs. distance from the

present progradation front in Figure 4b. In the rollover syncline, the dip was measured on the basinward-dipping flank. The dip of the youngest wedge (layer 11 in Figure 4b) retained the initial depositional slope of 5. The rollover plot is shaped like an asymmetric hill (Figure 5a). Dips rapidly increase away from the front to a maximum and then decrease less rapidly. The maximum dip corresponds to where the thickest salt was expelled. Because the salt plateau thickened over time, the thickest salt was below the youngest rollover syncline. This maximum migrated basinward with time as each new syncline formed. The oldest wedges preserved the initial 5 depositional slope and planar shape even through they were folded and unfolded by the monocline that rolled through them. Subtle residual salt pillows affected the dips. For example, the dip of layer 2 in Figure 4b is about 3slightly less then the original 5because a pillow was left beneath it. The plot of a fault rollover in the hanging wall of a simple listric normal growth fault is quite different: there is no decrease of dip away from the present progradation front. Instead, the dips of hangingwall cutoffs abruptly then gradually increase to reach a maximum value away from the surface trace of the listric normal fault both in physical models (Figure 5b) and in nature (Figure 5c). The constant maximum dip far from the fault corresponds to flattening of the underlying fault. The maximum dip depends on the initial sedimentation slope, initial fault dip, and dip of the sole fault (Crans et al., 1980; Jackson and McKenzie, 1983; Vendeville, 1991). This type of analysis should be carried out only on synkinematic strata, which are distinguished from prekinematic strata by having local lateral variation in thickness. Listric faults having multiple ramps and f lats yield more complex rollover plots (Figure 5d). Both plots include a tail of data points from

Dip of rollover flank

404

Progradation and Salt Tectonics

Distal

Proximal
Prograding wedges dip 5 Downlap

(a)

Starved basin

A=0

P = 1.4 cm/day
2+3 1

Figure 6Restored, nonexaggerated sections showing the evolution of model 253 as wedges prograded above a tabular salt layer.

Basement

(b)
Salt thickens and lifts overburden

Progradation continues

Depocenter
4 2+3 1

Proximal faults

(c)

Rolling monocline Downlap

Incipient rollover syncline


1

5+6+7 4 2+3

Relict pillow

Reactive diapir

(d)

Monocline advances Rollover syncline

Expulsion rollover

8+9+10 5+6+7 4 2+3

Salt weld

(e)
Inflated plateau

Rotated downlap

Rollover advances

8+9+10 5+6+7 4 2+3 1

10 cm

Rotated downlaps
1

Source layer

Prekinematic layer

2-10

Synkinematic wedges

prekinematic strata (open squares and circles). These prekinematic data points are not diagnostic, but they hint that dual ramps merely double the plot profiles of single ramps. However, we have not yet been able to test this prediction on a natural example containing clearly defined dual ramps and substantial synkinematic strata. Dynamics of Salt Tectonics Driven by Progradation Over Tabular Salt The structural evolution (Figure 6) incorporates relevant three-dimensional aspects derived from overhead views photographed during the experiment. Rollover keystone grabens were omitted in structural reconstruction for simplicity. The oldest sand wedge (layer 2+3) initially had a planar upper surface dipping 5 basinward (Figure 6a). Differential loading by the wedge squeezed initially tabular salt distally to where the load was less.

Salt therefore thinned beneath the wedge and uniformly inflated in the distal starved basin. Both the wedge and the underlying prekinematic layer were rotated clockwise, causing the prekinematic layer to tilt landward and the initially basinward-dipping wedge to become horizontal then tilt landward as well (Figure 6b). Normal faults and a reactive diapir formed at the proximal edge where the contact between the model and the end wall acted as a zone of weakness. The overburden subsided deepest where it was thickest, forming a depocenter at the proximal end (layer 4 in Figures 4b, 6b). A dynamic equilibrium was reached when distal salt thickening balanced proximal overburden subsidence. Salt evacuation by layer 4 caused the prekinematic layer (layer 1) to ground onto the basement, forming a proximal pr imar y salt weld (Jackson and Cramez, 1989). Subsequent partial evacuation left a subtle residual salt pillow (Figure 6c). At this stage, the landward-tilted wedge (layer 2+3) rotated back to its initial depositional slope angle, whereas

Ge et al.

405

Distal
Progradation front
25

Proximal
P = 0.5 cm/day
Progradation

90 80

A = 0.02 cm/day
70 60

20

50 40
15

30
12

20
7

9 6

10

10 0

B y p a s s
5

Time (day)

Basement

Step 2
1

Step 1 Prekinematic layer


2-24

10 cm

Viscous layer (salt analog)

Synkinematic layers

25

Postkinematic layer

Figure 7Chronostratigraphy of model 272. The lowermost three layers show the initial setup of the experiment. Each episodically deposited synkinematic layer is plotted as a numbered, thick, horizontal line; proximal wedges extend from the right, whereas distal aggrading layers extend from the left. The distal bypass region represents a nondepositional hiatus. Layer numbers correspond to those in the following cross section. See Figure 3 for additional explanation.

wedge 4 tilted basinward more steeply than its initial slope of 5. This steepening set the pattern for all younger wedges that also accumulated over the monocline between the inflated plateau and the subsiding wedge front. The depocenter shifted to the frontal region above the monocline (Figure 6c) as a rollover syncline (layer 5+6+7). Lack of salt supply hindered further rise of the proximal reactive diapir. Lack of proximal salt meant that subsequent wedges thickened and deformed mostly at their distal parts. Basinward expulsion of salt below the thickened wedges caused the monocline to roll forward by folding, then unfolding, the overburden (Figure 6ce). Evacuation of salt landward of the monocline resulted in maximum subsidence of the overburden at the tip of the younger wedges (Figure 6d, e). Originally straight wedges

became increasingly bent into sigmoidal clinoforms basinward. A keystone graben formed by local stretching of a migrating expulsion rollover anticline. The primary salt weld also propagated basinward in the wake of the monocline. Although the salt layer in the distal region inflated over time, no diapirs or even salt-cored anticlines formed there. This supports Nelsons (1991) proposal that progradation over tabular salt is an ineffective initiator of diapirism where the salt is buried by a constant-thickness overburden in the distal part of the basin. Progradation Over Stepped Salt Basins A combination of basinward salt thinning and landward-dipping, subsalt normal faults could

406

Progradation and Salt Tectonics

Distal

Proximal
P = 0.5 cm/day A = 0.02 cm/day
Progradation

(a)

Early crest VE=1


0 10 cm

Salt weld

(b)
24 23 22 21 20 19 18 17 16 15 14 13 12 11

12 9 7 6 1

Downlap surface

Ea
rly

10 9 8 7 6 5 4 3 1 2

cr e
st

Step 2

V E = 3X
0 10 cm

Basement Viscous layer (salt analog)


1

Step 1 Prekinematic layer


2-24

Synkinematic layers

Figure 8Regional cross sections of model 272 showing depositional patterns and structural style. (a) True scale, (b) vertically exaggerated by 3.

provide an effective buttress against lateral salt flow. The following three models examine how such basement steps can influence salt tectonics driven by progradation. Structural Style and Kinematics Slow Progradation Over Moderately Buoyant Source Layer In model 272, the density of the salt analog was 1390 kg/m3, compared with 1700 kg/m3 for sand. The prekinematic layer was 0.5 cm thick. The model deformed over 97 days with 24 synkinematic wedges at an average progradation rate of 0.5 cm/day, the lowest of all models (Figure 2). Only four aggrading layers (6, 7, 9, and 12) were added distally at an average rate of 0.02 cm/day (Figures 2, 7). A vertical section of the deformed model is illustrated in Figure 8. As in model 253, the oldest

prograding wedges (215) were deformed into an expulsion rollover above a salt weld. However, a salt anticline (rather than a monocline) formed above basement step 1. Thinning of the prekinematic layer in the proximal flank of the salt anticline was caused by erosion (sand was scraped off) of the former crest that subsequently migrated basinward. The salt never extruded laterally, although parts of the salt crest were often exposed. The distal region was little deformed. Synkinematic aggrading layers (6, 7, 9, and 12) thinned toward and onlapped against the distal f lank of the salt anticline. Young sediment wedges 1624 prograded across the anticline and buried it. Overburden above basement step 2 tilted proximally. Despite continued differential loading, only a subtle salt pillow formed above basement step 2 because the overburden was too thick to bend much (Figure 8b). Wedges that downlapped onto aggrading layer 12 preserved their oblique clinoformal geometry.

Ge et al.

407

Distal
Progradation front
8

Proximal
P = 5.3 cm/day
14

Progradation

A = 0.3 cm/day
6
7 8

10

Time (day)

4
5 5

2
3

0
1 1

Basement Viscous layer (salt analog)

Step 2
1

Step 1 Prekinematic layer


2-13

0
14

10 cm

Synkinematic layers

Postkinematic layer

Figure 9Chronostratigraphy of model 270. See Figures 3 and 7 for additional explanation.

Progradation Over Highly Buoyant Source Layer Models 270 (Figures 9, 10) and 263 (Figures 11, 12) had setups and depositional histories similar to those of model 272, but the density of the salt was only 950990 kg/m 3 . The mean rates of progradation were 5.3 cm/day for 8 days in model 270 and 1.6 cm/day for 24 days in model 263. Although the deformed models had similar overall structural styles, the detailed salt structures varied greatly, even between different sections in the same model. Model 270 was tilted 1 landward before deformation to simulate the slightly dipping basement typical of foreland basins and half grabens. Tilting caused gravity-driven extension at the distal end where a salt roller formed (left end of Figure 10). This extension was balanced by buckling the prekinematic layers in front of the prograding wedges. These minute buckle folds were masked by younger deformation and became invisible. Proximal gravity gliding stopped early because the overburden thickened rapidly there. The effect of tilting was negligible compared with that of progradation. Salt structures were well developed above basement step 1, but were developed only sporadically next to step 2 (Figure 10). Above step 1, salt structures varied from a salt anticline (Figure 10a), to a salt wall (Figure 10bd), to a salt tongue and salt sheet (Figure 10eh). These changes along strike resulted partly from the

initial model setup, but were mainly controlled by synkinematic deposition. The prekinematic layer was 0.51.0 mm thicker and the source layer 0.51.0 mm thinner in section (a) than in section (h). This differential loading caused salt to flow along strike. As in model 272, prograding wedges and distal aggradation localized a salt anticline above step 1. The anticline plunged beneath the thickest overburden (Figure 10a), and younger wedges prograded over and buried the anticline where it was lowest. Salt later flowed from this area, and the anticline gradually subsided. Eventually, only a relict, salt-cored anticline was preserved above step 1 (Figure 10a). However, the sigmoidal expulsion rollover unambiguously recorded the thickening, then thinning, of the salt. A salt wall with large overhangs dominated the structure above basement step 1 in Figure 10bd. The stem was pinched off locally (Figure 10d) when the basinward, inf lated diapir ic f lank sagged as the supporting salt escaped by extrusion. A salt tongue and sheet formed above basement step 1 where salt was initially thickest (Figure 10eh). The basal contact of the advancing salt sheet between steps 1 and 2 has a flat/ramp geometry, where the base of salt climbs basinward overall but locally cuts back landward wherever a pulse of sediment accumulated. A younger salt anticline and diapir formed next to basement step 2 (Figure 10fh). Its proximal overhang coalesced with the salt sheet from the older

Distal

Proximal
P = 5.3 cm/day
Downlap surface
8 11 12

(a)
A = 0.3 cm/day
1 3 5 7

Progradation

Rollover and crestal faults


13 10 9 8 7 6 5 4

Basement Salt roller Allochthonous salt Rotated downlaps Salt wall Relict salt anticline

(b)

Salt weld

(c)

Salt diapir with overhangs

(d)

Detached salt diapir

Relict pillow

(e)

Salt sheet Allochthonous salt

(f)

Salt ramp

Salt flat

Salt anticline

(g)

Salt canopy

Foundered roof

(h)

Step 2
70 60 50 40 30

Step 1
20 10 0

Distance from proximal edge (cm) Viscous layer (salt analog)


1

Prekinematic layer

2-13

Synkinematic layers

Figure 10Serial nonexaggerated cross sections of model 270 at the end of the experiment. Salt structures increase in maturity, complexity, and area from (a) to (h). Each section was 3 cm apart.

Ge et al.

409

Distal
Progradation front
17 20 16 14 15 15 10

Proximal
P = 1.6 cm/day
Progradation
Bypass

A = 0.2 cm/day
Time (day)

12 6 8 5 5

0 1 1

Basement

Step 2 1 Prekinematic

Step 1 2-16 Synkinematic layers

10 cm

Viscous layer (salt analog)

17 Postkinematic layer

Figure 11Chronostratigraphy of model 263. See Figures 3 and 7 for additional explanation.

diapir above step 1 to form a salt canopy (Jackson and Talbot, 1989). A slab of overburden foundered and was engulfed in the salt (Figure 10g, h). In the overburden, an expulsion rollover, crestal graben, and salt weld dominated proximally (Figure 10). Along the distal flank of the diapir above step 1, the aggrading layers tilted landward, rotating the downlap surface on the top of layer 8 (Figure 10a) to become an apparent onlap surface (Figure 10b). The wedge sediments in the distal flank of the diapir above step 1 apparently onlapped against and thinned toward the anticline next to step 2. Because of this geometry, these strata could be misinterpreted as a secondary peripheral sink (Trusheim, 1960). However, the process is distinctively different. The domeward thickening here was primarily due to progradation rather than diapirically induced salt withdrawal. Furthermore, rollover synclines migrated away from the diapir in contrast to secondary or tertiary peripheral sinks that generally migrate toward the rising diapir. We return to this significant difference in the section dealing with geologic analogs. Little deformation occurred distally of step 2. In the slowly prograding model 263 (Figures 11, 12), which had similar a setup to models 270 and 272, a subsided salt anticline was preserved above basement step 1. The tiny notch in layer 1 in the proximal f lank of the relict anticline

(Figure 12bh) marked early reactive piercement below crestal faults, although the salt did not emerge. The expulsion rollover and the onlaps of aggradational sedimentation in the distal basin record the early growth and subsidence of the anticline. In Figure 12ik, where the source layer was initially thickest, a diapir above step 1 emerged and extruded a salt sheet laterally as far as section (c). This salt sheet evolved into a multitiered sheet complex (Figures 12ik, 13). Each tier was bounded by normal faults at the trailing edge of the evacuated sheet (TFS1TFS3 in Figure 13) and by buckle folds (Figure 12j), frontal monocline (Figure 12k), active diapirs, and breakouts to higher stratigraphic levels at the leading edge of the sheet. A second salt structure formed above step 2 (Figure 12). The anticline initiated early before deposition of the first aggrading layer 5. This layer thinned widely over the anticline above step 2 and toward the anticline above step 1 (Figure 12ad, i, j) because these two folds were initially a single broad anticline, which was then divided by prograding sediments. Here, there was time for salt to escape basinward of step 1 to form large distal structures because the progradation rate was much slower than in model 270, where salt was trapped against step 1. The crest of the old anticline above step 2 was eroded, which enhanced

410

Progradation and Salt Tectonics

Distal

Progradation A = 0.2 cm/day Rollover and crestal faults Downlap surface Onlap

P = 1.6 cm/d

Proximal

(a)
6 1 5

14

15

16

15 14 13

1211 10

9 8 76

5 4

3 1

Basement Salt anticline

Allochtho

Salt weld Allochthonous salt

Relict salt anticline

Figure 12Serial nonexaggerated cross sections of model 263 at the end of the experiment. Each section was 2 cm apart except between (j) and (k), which were 4 cm apart. As in Figure 11, salt structures increase in maturity, complexity, and area from (a) to (k).

(b)
Rotated downlap Rotated downlap

(c)
Evacuated allochthonous sheet

(d)
Truncation Passive wall TFS

(e)

(f)

(g)

(h)
Onlap Salt anticline Salt-sheet complex

(i)

Buckle fold Salt ramp

Salt flat

(j)
Salt canopy

(k)
Step 2
70 60 50 40 30

Step 1
20 10 0

Distance from proximal edge (cm) Viscous layer (salt analog)


1

Prekinematic layer

2-16

Synkinematic layers

active, then passive, diapirism during deposition of layers 615. The youngest aggrading layer 16 unconformably rested on the truncated flanks of

t h e d i a pi r a bove s t e p 2 ( Fi g u re 1 2 e h ) . An expulsion rollover also formed in the proximal flank of this diapir.

Ge et al.

411

Figure 13Tracing of the structure above step 1 in Figure 12i showing the salt-sheet complex formed by sediment progradation. TFS1TFS3 denote peripheral faults at trailing edges of salt sheets. LEB1LEB3 denote leading edge breakouts of salt sheets where these climbed stratigraphic section.

Dynamics of Salt Tectonics Driven by Progradation Over Stepped Salt A conceptual model based on models 263, 270, and 272 shows the evolution of salt structures as sediment wedges prograded over a distally thinning salt basin with basement steps dipping landward (Figure 14). Initially, a uniform prekinematic layer 1 covered the stepped source layer. Synkinematic wedge 2 initially dipped 5 (Figure 14a). Distal aggradation was slow. Early on, differential loading by the wedge expelled underlying salt basinward, as in the model without basement steps (Figure 6). However, the lateral flow of salt was restricted by the basement step. Thus, salt backed up there, forming a broad, asymmetric salt-cored anticline (Figure 14b). As the salt anticline rolled forward across step 1, the source layer thickened in its distal flank. Aggrading layer 4 thinned and onlapped against the anticline. These distal aggrading sediments had two major influences: (1) they resisted basinward rolling of the anticline by buttressing its leading flank, so that the influx of more salt caused the anticline to amplify, and (2) the onlapping layers created differential loading that expelled distal salt to feed the growing anticline. In nature, aggradation is less episodic than in our experiments, but, even so, the raised bathymetry above the distal inflated salt could prevent appreciable distal aggradation until late in the history, as in our following examples from the Gulf of Mexico and Santos Basin. Arching of the anticline stretched its outer arc to form crestal faults, which initiated reactive diapirism above step 1. Erosion of the crest caused further thinning. After emerging, salt remained at the surface during sedimentation of layer 5, and the diapir grew passively into a tall salt wall by downbuilding (Figure 14c) (Nelson, 1989; Jackson and Talbot, 1991). The passive diapir could extrude to

form a salt sheet (Figure 12d) and even an evacuated salt-sheet complex. However, rapid progradation could bur y and depress the old anticline (Figures 10a, 12ah). Progradation of wedge 5 across the diapir overlying step 1 triggered a younger anticline above the next constriction, basement step 2 (Figure 14c). Again, repeated crestal faulting, erosion, and synkinematic sedimentation promoted passive diapirism above step 2, while the diapir above step 1 continued downbuilding or extrusion (Figure 14df). Complete evacuation of the source layer caused the overburden to ground onto the basement, forming a basinward-younging salt weld. Expulsion rollovers formed against the proximal flanks of each diapir. The amplitude of the rollover adjoining step 2 was smaller than that adjoining step 1 (Figures 12, 14f) because the source layer initially thinned distally. However, the initially planar wedges still deformed to sigmoidal clinoforms, as was discussed in model 253 (Figure 6a). In this way, progradation created a series of saltrelated structures subparallel to the progradation fronts and basement steps. The ages of progradation fronts and salt-related structures decreased basinward. Accordingly, the amplitude, complexity, and maturity of salt structures decreased basinward. However, as the overburden thickened and the source layer thinned distally, diapir initiation became more difficult. Evacuation of Salt Sheets During Progradation As each extruded sheet was rapidly buried, it deformed under the influence of basinward gravity spreading and differential loading by new wedges. Normal faults formed along the salt sheets trailing edge, whereas active piercement at its leading edge

412

Progradation and Salt Tectonics

Distal

Proximal
Progradation
1 2

(a)

Semi-starved basin

(b)
Slow aggradation
4 1 2

Salt anticline
3

Basement

(c)
Salt anticline Starved
4 5

Expulsion rollover Passive wall


5 4 3 2 1

Salt weld

(d)
Slow aggradation
6 4

Salt sheet Truncated crest


6 5 5 4 3 2 1

(e)
Slow aggradation
7 6 4

Trailing fault system Passive wall


6 5 5 4 3 2 1

(f)
Slow aggradation
7 6 4

Diapir with overhang


7 6 5 5 4 3 2 1

20 cm

Step 2

Step 1

Source layer

Prekinematic layer

2-7

Synkinematic layers

Figure 14Cross sections showing the schematic evolution of salt structures during progradation above a salt basin underlain by stepped basement.

allowed salt to climb to a higher stratigraphic section over the basal cutoffs and spread glacially during each nondepositional or erosional hiatus. Repetition of the sometimes coeval processes of burial, sheet evacuation, sheet climb, and sheet

spreading formed a multitiered complex of salt sheets (Figure 13). This complex of salt sheets migrated basinward along with the progradation front (Figures 12ik, 13) (Ge et al., 1994; Ge, 1996). Extrusions flowed

Ge et al.

413

GEOLOGIC ANALOGS Salt structures form during progradation in many salt basins, but have been overprinted by either extension on divergent margins or shortening in foreland basins. The closest analogs of our models should be salt structures in progradational environments that have been minimally affected by other processes. The following examples from various basins have salt structures strikingly similar to our model structures. We therefore suggest that salt tectonics in those basins was driven by progradation. Santos Basin Figure 16 presents two seismic sections of the salt-related structures in the northern Santos Basin, offshore Brazil. The salt is Aptian (Demercian et al., 1993; Mohriak et al., 1995; Szatmari et al., 1996) and the subsalt basement is fairly f lat. A giant rollover downlaps onto a flat salt weld. The distal part of the rollover overlies a diapiric wall. Demercian et al. (1993) inferred that the rollover in Figure 16a formed during Late Cretaceous extension as a fault rollover against the Cabo Frio fault, a landward-dipping listric growth fault (left center of Figure 16a). The fault zone created a lateral gap in the lowermost Albian sequence of more than 25 km (Demercian et al., 1993; Szatmari et al., 1996). In contrast, Szatmari et al. (1996) attributed the rollover to the bending of the Upper Cretaceous prograding wedges during evacuation of the underlying salt. However, they too interpreted an extensional listric fault bounding the rollover that hindered seaward progradation. In contrast, Mohriak (1995) and Mohriak et al. (1995) proposed several hypotheses and favored a combination of progradation and extension. The structures in Figure 16 closely resemble our model results (Figures 4, 6). In this section, we test the hypothesis that the Cabo Frio fault zone and related structures in the northern Santos Basin resulted primarily from progradation and salt expulsion rather than from regional extension. Rollover plots (Figure 16c) of two adjoining lines (Figure 16a, b) support an origin by progradation and salt withdrawal rather than by regional extension. Each plot has the shape of an asymmetric hill characteristic of an expulsion rollover (compare Figure 5a), unlike the progressive dip increase typical of fault rollover plots (Figure 5b, c). The small hollows in the broad crest of plot B coincide with the relict pillows, where incomplete evacuation of salt resulted in less bending of the overburden. We tried restoring these lines incorporating regional extension, but were unable to produce plausible solutions. Hypothetical extension either

Figure 15Map of model 263 after 19 days showing linked systems with basinward-spreading salt sheets. Ticks and hachures indicate downthrown side of normal faults. LEM = leading edge monocline; TFS = trailing fault system; KFS = keystone fault system; WFS = wrench fault system. See Figure 13 for cross-sectional structure.

in any direction, depending on the local slope, which was commonly landward into the adjoining rollover syncline. However, the breakout sites, where salt climbed stratigraphic section, systematically shifted basinward. That may be a useful rule of thumb for interpreting similar structures on seismic profiles. In map view (Figure 15), the traces of the trailing faults (TFS) were arcuate in contrast to the straight keystone faults (KFS) because they soled out on the rounded trailing edges of salt sheets. This extension at the trailing edge of the sheet was balanced at the leading edge by mild buckling (Figure 12j), active diapiric piercement, contraction of the diapir neck, and transfer of allochthonous salt to higher stratigraphic levels. Subtle wrench structures formed along the sides of the roof over a sheet because of its basinward movement relative to the adjoining overburden (Figure 15). All these structures were linked during sheet evolution.

414

Progradation and Salt Tectonics

(a)
Southeast
0 0 1 5 km

Northwest

TWT (s)

2 3 4 5

(b)
0 1 0 5 km

TWT (s)

2 3 4 5

(c)
60

Dip of rollover flank

Distal Relict pillow

Proximal Relict pillow

40

20

0 0 5 10 15 20 25 30 35

Data from Figure 16a Data from Figure 16b


40 45 50 55

Distance from progradation front (km)

Figure 16(a, b) Time-migrated seismic sections through the Cabo Frio fault zone, Santos Basin, offshore Brazil (from Demercian et al., 1993, and Mohriak et al., 1995, respectively). (c) Rollover plot of dips of rollover flanks vs. distance from the frontal salt structures, showing the hill profile typical of expulsion rollovers. See Figure 17 for their structural evolution. Compare with the model structures in Figure 4 and the model rollover plot in Figure 5a.

violates basin mechanics (for example, requiring upslope gravity gliding) or generates unrealistic strains. For example, 25 km of downdip shortening are required to balance just the Albian gap. Some contraction has been reported in deep water (Cobbold et al., 1995), but the magnitude seems too low. Moreover, an extensional origin for the Albian gap requires the initial salt layer there to be an absurd 36 km thick, assuming conserved salt area.

Conversely, by excluding regional extension, our decompacted reconstruction of the depth-converted seismic section in Figure 16b plausibly explains the origin of the Albian gap (Figure 17). Sediment wedges prograded southeastward (Mohriak et al., 1995) over initially tabular Aptian salt, progressively expelling it seaward from the Albian to the Paleocene. The salt layer uniformly thickened distally up to 4 km by the end of the Campanian (Figure 17d). Condensed abyssal sediments were probably

Ge et al.

415

Figure 17(ah) Evolution of the Cabo Frio fault zone, Santos Basin, offshore Brazil. (i) Speculative future development of a landward-dipping normal fault produced by removal of underlying salt on the proximal flank of the diapir; such a structure is the analog of that in Figure 16a. The nonexaggerated present depth section (h) was converted from the seismic section in Figure 16b using a velocity profile from the Gulf of Mexico shelf (Schultz-Ela and Duncan, 1994) slightly adjusted to correspond with the depth section in Mohriak et al. (1995). Wedges were assumed to initially dip at 1. Thin abyssal sediments above the starved distal salt plateau were omitted in (ad).

416

Progradation and Salt Tectonics

South-Southwest (a) Thickened Louann Salt


Rollover syncline

North-Northeast
Progradation

(b)

(c)

Future fault

Expulsion rollover

(d) Cenomanian
Rollover syncline

Rotated onlap

(e) End Miocene

Relict pillow

(f) Present
0

Inverted rollover syncline

Depth (km)

TP TUM

5
TMO TLC

TLM

10

TUM

Top Pliocene (1.6 Ma) Top Miocene (5.5 Ma) TLM Top lower Miocene (15.5 Ma)
TP

TM TLC

Top middle Oligocene (30 Ma) Top Lower Cretaceous (94 Ma)

10 km

Figure 18Reconstruction of a depth section from the northeastern Gulf of Mexico (Wu et al., 1990, their foldout 3c). Prograding wedges between the Upper Jurassic and the Lower Cretaceous were slightly reinterpreted based on the original seismic line (Wu et al., 1990, their foldout 3a, b). Wedges were restored to 1 initial dip; other horizons were restored to 0.5 initial dip. Section (c) was slightly shortened compared with section (d) to compensate for extensional faulting. Minor faults were omitted.

deposited on the distal plateau above the inflated salt. This condensed section represents the presentday Albian gap. Evacuation from the thickened salt

layer formed an expulsion rollover and rollover syncline (Figure 17c), both of which increased in amplitude and migrated seaward (Figure 17ah).

Ge et al.

417

Southwest

Progradation

Northeast

Salt flow Zechstein Salt Anhydrite and younger salt Triassic

Relict pillow

Figure 19Relation between Triassic prograding wedges and underlying salt structures in the Mansfeld Basin, Germany (after Trusheim, 1960). Compare with the similar model 253 in Figure 4.

Complete salt evacuation created a salt weld that propagated seaward as the expulsion rollover advanced. Incomplete salt evacuation, possibly because of rapid progradation, left two residual pillows (Figure 17a, e). Wedges that initially onlapped against the proximal flank of the rolling monocline later rotated to form apparent downlaps onto the salt weld (Figure 17c). A passive diapir was pushed up by expelled salt ahead of the prograding wedges during Maastrichtian time (Figure 17e), possibly because of shifting depocenters. The diapir was probably prevented from migrating farther basinward by the additional differential loading from the extreme distal sediments. The diapir grew passively by maintaining its crest near the surface during sedimentation. The contact between the diapir and the proximal rollover is diapiric rather than faulted (Figures 16b; 17g, h). Forward modeling of the present structure (Figure 17h) indicates how this diapiric flank could evolve in the future (Figure 17i). If salt were expelled seaward or along strike, the diapir and its roof would subside. Being brittle, the roof can subside only by faulting. We carried out this forward modeling (Figure 17i) because an almost identical landward-dipping fault along strike is imaged in a neighboring seismic profile (Figure 16a) (Demercian et al., 1993). No regional extension is necessary to form the fault, although we cannot eliminate the possibility of some extension. Similar arguments for an origin by salt reduction rather than by regional extension have been made after restoring similar landward-dipping faults in the Gulf of Mexico (Diegel et al., 1993, 1995; Schuster, 1993, 1995). Gulf of Mexico Basin The Gulf of Mexico Basin contains some of the most numerous and complex salt structures in the world (e.g., Worrall and Snelson, 1989). Sediment progradation has been proposed to provide differential loading that deformed both autochthonous and allochthonous salt (e.g., Humphris, 1979; Wu et al., 1990). This proposed effect is generally obscured by the enormous thickness of the overburden, by extensional and contractional tectonics, and by emplacement of allochthonous salt sheets.

In most parts of the Gulf of Mexico, progradational effects can only be displayed by accurate restorations of regional sections (Worrall and Snelson, 1989; Diegel et al., 1995; Peel et al., 1995). However, an exceptionally clear example of the relation between progradation and salt tectonics in the northeastern Gulf of Mexico is illustrated by a seismic line published by Wu et al. (1990, their foldout 3), which we have slightly reinterpreted, depth-converted, and restored (Figure 18). Qualitative reconstruction by Wu et al. (1990, their figures 7, 10) suggested that Upper Jurassic and Lower Cretaceous pseudo-clinoforms were initially subhorizontal slope sediments that rotated and grounded onto basement after basinward withdrawal of Louann Salt. Our reconstruction (Figure 18a) shows that initially oblique (dipping 12) Upper Jurassic prograding wedges expelled most of the underlying salt and mounded it into a distal salt plateau, where salt was apparently thickened to about 4 km. As in model 253 (Figures 4, 6), oblique wedges were deformed into sigmoidal clinoforms as underlying salt was expelled seaward (Figure 18bd). Originally onlapping wedges were rotated to apparent downlaps onto largely evacuated salt. A rollover syncline formed above the landward-facing rolling monocline. However, unlike in the Santos Basin (Figure 17) and our models, salt subsequently thinned rather than thickened distally (Figure 18bd). Salt must have migrated beyond or out of this plane of section. The salt diapir appears to have been initiated in the middle Cretaceous (Figure 18c, d). What triggered this diapiric wall? Uniform thicknesses of Jurassic and Lower Cretaceous overburden adjoining the diapir eliminate differential loading as a plausible explanation. The absence of evidence for any underlying basement fault also eliminates a stepped base of salt as a likely cause. Instead, we propose that the diapir was initiated reactively by regional extension. The hypothetical landwarddipping fault depicted in Figure 18c is not recognizable in the present-day section. However, along strike (Wu, 1993, his lines D2-D6, plate 1), this diapiric wall passes into a major normal fault, whose growth is recorded by thickened middle Cretaceous sediments in the proximal flank (Figure 18d). The diapir grew passively during the early

418

Progradation and Salt Tectonics

Figure 20Cross sections of a salt structure in Lower Saxonian Basin, Germany. Restored to the (a) pre-Albian and (b) present-day structure (both after Trusheim, 1960). The finely stippled strata from the Dogger to the Neocomian were interpreted as a primary peripheral sink by Trusheim (1960). In contrast, we interpret them as wedges prograding north-northeastward. Double arrows denote rollover synclines that also advanced basinward.

Miocene. Rapid seaward spreading of salt during the late Miocene formed an allochthonous salt sheet (Figure 18e). The experimental arcuate trailing fault system detaching on the landward periphery of evacuated salt sheets (Figures 13, 15) may have a counterpart in some of the seaward-dipping, arcuate growth faults of offshore Louisiana (Worrall and Snelson, 1989), where a series of arcuate faults define the landward edge of some sheets. North German Basins The famous salt structures in northern Germany have traditionally been attributed to halokinesis without significant lateral tectonic forces (Trusheim, 1960). Trusheim envisaged that during sedimentary differential loading, primary peripheral sinks developed during the pillow stage, whereas secondary and tertiary sinks formed during diapirism. Sannemann (1968) coined the term saltstock family to describe how daughter and granddaughter diapirs could develop in succession from a central mother diapir as peripheral sinks migrated outward. This outward migration was envisaged by Trusheim (1960, p. 1523) as due to a large-scale rhythmical phenomenon like ripples from a stone thrown into water. In contrast, based on structural geometry similar to those in our experiments, we think that the German salt structures and diapiric families were initiated by prograding sediment wedges. Trusheims (1960) turtle structures were ascribed to inversion of the primary peripheral sink as an adjoining pillow evolved into a diapir. We reinterpret this sink as an

expulsion rollover. Moreover, the sequential formation of daughter and granddaughter diapirs is probably a response to the progradational shift of depocenters between the Keuper and the early Tertiary (Trusheim, 1960, his figure 20). The Mansfeld Basin is a fine example of Triassic prograding wedges expelling underlying salt southwestward (Figure 19) in a form closely resembling our models (Figures 4, 6). Laterally expelled salt accumulated as a broad pillow near the salt pinch-out at the southwest margin. A salt pillow was overridden by progradation and trapped in the northeast margin. Trusheim (1960) saw this geometry as evidence for earlier mobilization of deeper salt in the center of the saucer-shaped basin. In the Lower Saxonian Basin (Figure 20) (Trusheim, 1960), the oldest Triassic overburden (Figure 20, layer 2) is uniformly thick and thus prekinematic. The overlying Lias (Figure 20, layer 3) varies in thickness, but the thinnest section is on the present limb of the monocline (Figure 19a). This anomaly probably records the former crest of the monocline, as in model 272 (Figure 8); the crest later became its limb as the fold rolled beneath it. Northnortheastward progradation of the Middle Jurassic Dogger (Figure 20, layer 4) to Lower Cretaceous Neocomian (Figure 20, layer 7) wedges uniformly thickened Zechstein salt. This formed a south-southwestwardfacing monocline that separated a salt plateau in front of the wedges. On the distal salt plateau, sediments coeval with the prograding wedges were sparse. In the proximal flank of the monocline, rollover synclines (double-ended arrows in Figure 19a) created by salt expulsion migrated basinward. As in model 250 (Figures 4, 6), the thickened salt did

Ge et al.

419

Figure 21Restored evolution of Gorleben salt wall, Germany. (ae) Initiation by the Middle Triassic Muschelkalk to Upper Jurassic Malm wedges (fine stipple) prograding northwestward, and (fi) passive and glacial growth. The present structure (j, after Trusheim, 1960) was not depth converted or decompacted because the traveltime scale is unknown. The wedges were restored to 5 initial dip in (cf), which is smaller than the present-day 810 dips (in time) of the deformed Keuper reflections.

not pierce the uniform thickness of the Albian sediments. In contrast to Trusheims (1960) interpretation, the section shows no record of significant diapirism in the form of a salt-filled gap in

the restoration or diagnostic thickness changes above the incipient diapir. The Albian (Figure 20, layer 8) is fairly uniform in thickness, which suggests that progradation ended

420

Progradation and Salt Tectonics

or migrated further basinward at this time. In contrast, the overlying Upper Cretaceous is also greatly thickened on each side of the incipient diapir and on the proximal side of the turtle structure to the south-southwest. Accordingly, we interpret the Upper Cretaceous on the proximal flank of the diapir not as a rollover syncline, but as a peripheral sink formed in response to salt reduction. The incipient diapirs crestal graben is bounded by reverse faults. Because of the great thickness of the diapirs roof compared with its height, we ascribe these reverse faults not to active diapirism (SchultzEla et al., 1993) but to basin inversion during the Late Cretaceous. Salt probably migrated laterally out of section, deforming the progradation wedges into a turtle structure anticline when the overburden was welded onto the basement (Figure 19b). In contrast, Middle Triassic Muschelkalk to Upper Jurassic Malm wedges prograding northwestward over a subtle basement ramp created the Gorleben salt wall (Figure 21). An incipient salt anticline with an eroding crest was inflated by salt expelled from beneath the Muschelkalk sediments (Figure 21b). Lateral salt flow was restricted over the basement ramp, which, together with distal aggradation, probably stabilized the position of the anticline. As the Gorleben anticline amplified, salt inflated the distal plateau (Figure 21cf ), which was sediment-starved because of its elevation and because the anticline dammed the prograding wedges. Crestal erosion and active diapirism (Figure 21e) created the Gorleben passive diapir during the Late Triassic (Figure 21f), which spread as a salt glacier (Figure 21g, h). Salt extruded landward (southeast) because the proximal anticlinal flank sank faster than the distal flank owing to sedimentar y loading; similar landward extrusion occurred in our models (diapir above step 2 in Figure 12eh). The distal salt partly reversed flow to feed the extrusion because of distal aggradation. Evacuation of salt caused the proximal overburden to rotate counterclockwise. The prekinematic Buntsandstein was partly unfolded and welded onto the basement; the wedges distorted into sigmoidal clinoforms and formed an expulsion rollover during salt reduction (Figure 21gj). The proximal weld below the wedges propagated basinward from the Early Jurassic to the Tertiar y (Figure 20dj), whereas the weld below the distal aggrading sediments formed during the Tertiary (Figure 21j). CONCLUSIONS Physical models indicate that progradation alone can initiate salt anticlines that can mature into salt walls, stocks, glaciers, or linked, multitiered

complexes of evacuated sheets and salt welds. The prograding wedges expelled the salt basinward by differential loading regardless of density relations in ways that reveal the most subtle interplay between sedimentation and lateral salt flow. Over a f lat base of salt, expelled salt evenly inflated the salt layer in front of the wedges, forming a sediment-starved distal salt plateau. Deformation was concentrated in the frontal region of the wedges in three adjoining and migrating domains, comprising in a seaward direction: (1) initially straight, oblique wedges were folded into sigmoidal clinoforms, forming an expulsion rollover containing crestal grabens, and initial onlaps were rotated into apparent downlaps by salt evacuation; (2) a rollover syncline; and (3) a progressively steepening landward-facing limb of a salt-cored, rolling monocline. These deformation zones amplified and advanced basinward together with the progradation front. Complete evacuation of salt left basinward-propagating salt welds below the wedges; incomplete evacuation left salt pillows as residual highs; however, no diapirs formed. Over a salt basin with basement steps representing landward-facing fault scarps, the prograding wedge initially formed structures similar to those in the model with a flat basement. However, further lateral flow of salt was restricted by each basement step, causing salt to accumulate above the steps as a broad, salt-cored anticline, especially where progradation was too fast for salt to escape seaward. Aggrading distal sediments pinned the salt anticline by hindering its basinward migration and enhanced salt upwelling by providing additional differential loading. The salt anticline evolved into an active, and then a passive, diapir. An expulsion rollover formed in the proximal flank of the diapir. This defor mation cycle was repeated as the wedges prograded across each basement step, creating a series of salt structures next to major basement steps. Again, the defor mation zones advanced basinward so that the age, amplitude, complexity, and maturity of salt-related structures decreased basinward. Rollover plots, introduced here, provide an objective way of distinguishing expulsion rollover anticlines formed by salt expulsion during progradation from conventional rollover anticlines in the hanging walls of listric growth faults formed by regional extension. Expulsion rollovers in synkinematic strata have a plot shaped like an asymmetric hill, whereas fault rollovers above simple listric faults have plots that abruptly, then gradually, increase away from the surface fault trace. Rollover analysis can be applied to model results and to seismic sections in both time and depth. Glacial extrusions from emergent diapirs flowed in any direction, depending on the local slope,

Ge et al.

421

including landward into the adjoining rollover syncline. However, the breakout sites, where salt climbed stratigraphic section, systematically shifted basinward because of progradation, as in the Gulf of Mexico. During differential loading by new wedges, a salt sheet carried its overburden basinward by gravity spreading as it became evacuated of salt, creating arcuate peripheral normal faults that soled out on the rounded trailing edge of salt sheets. Subtle wrench faults in the roof overlay the sides of the sheet. Extension at the sheets trailing edge was contractionally balanced at its leading edge by subtle buckling, active diapirism, contraction of diapiric necks, and upward transfer of allochthonous salt. Repetition of burial, sheet evacuation, sheet climb, and sheet spreading created a multitiered complex of salt sheets that migrated seaward. Consistency of restored geologic sections from around the world with model geometries allows us to propose a different explanation for some salt structures. In the northern Santos Basin (Brazil), the 25-km-wide missing section, known as the Cabo Frio fault zone, is reinterpreted to result almost entirely from progradation rather than from regional extension. The Gulf of Mexico Basin, too, contains an excellent example of a rolling deformation front pushed by progradation; a diapir was triggered by regional extension of uniformly thin strata over the distal salt plateau. Some of the Louisiana-style seaward-dipping, listric growth fault systems are interpreted to sole out on the landward margins of arcuate, buried salt extrusions that became largely evacuated by progradation. In the Zechstein Basin (northern Germany), classical features such as diapiric families and peripheral sinks are reinterpreted as the result of progradation squeezing salt laterally. REFERENCES CITED
Baars, D. L., 1966, Pre-Pennsylvanian paleotectonicskey to basin evolution and petroleum occurrences in Paradox basin, Utah and Colorado: AAPG Bulletin, v. 50, p. 20822111. Bloom, A. L., 1991, Geomorphology: a systematic analysis of late Cenozoic landforms (2d ed.): Englewood Cliffs, New Jersey, Prentice-Hall, 532 p. Bruce, C. H., 1983, Shale tectonics, Texas coastal area growth faults, in A. W. Bally, ed., Seismic expression of structural styles: AAPG Studies in Geology Series 15, v. 2, p. 2.3.1/1 2.3.1/6. Byerlee, J. D., 1978, Friction of rocks: Pure and Applied Geophysics, v. 116, p. 615626. Christensen, A. F., 1983, An example of a major syndepositional listric fault, in A. W. Bally, ed., Seismic expression of structural styles: AAPG Studies in Geology Series 15, v. 2, p. 2.3.1/36 2.3.1/40. Cobbold, P. R., P. Szatmari, L. S. Demercian, D. Coelho, and E. A. Rossello, 1995, Seismic experimental evidence for thin-skinned horizontal shortening by convergent radial gliding on evaporites, deep-water Santos Basin, in M. P. A. Jackson, R. G.

Roberts, and S. Snelson, eds., Salt tectonics: a global perspective: AAPG Memoir 65, p. 305321. Crans, W., G. Mandl, and J. Haremboure, 1980, On the theory of growth faulting: a geomechanical delta model based on gravity sliding: Journal of Petroleum Geology, v. 2, p. 265307. Dailly, G. C., 1976, A possible mechanism relating progradation, growth faulting, clay diapirism and overthrusting in a regressive sequence of sediments: Canadian Petroleum Geology Bulletin, v. 24, p. 92116. Demercian, S., P. Szatmari, and P. R. Cobbold, 1993, Style and pattern of salt diapirs due to thin-skinned gravitational gliding, Campos and Santos basins, offshore Brazil: Tectonophysics, v. 228, p. 393433. Diegel, F. A., J. F. Karlo, D. C. Schuster, R. C. Shoup, and P. R. Tauvers, 1993, Cenozoic structural evolution and tectonostratigraphic framework of the northern Gulf Coast continental margin (abs.): AAPG Annual Convention, Program with Abstracts, v. 2, p. 91. Diegel, F. A., J. F. Karlo, D. C. Schuster, R. C. Shoup, and P. R. Tauvers, 1995, Cenozoic structural evolution and tectonostratigraphic framework of the northern Gulf Coast continental margin, in M. P. A. Jackson, R. G. Roberts, and S. Snelson, eds., Salt tectonics: a global perspective: AAPG Memoir 65, p. 109151. Duval, B., C. Cramez, and M. P. A. Jackson, 1992, Raft tectonics in the Kwanza basin, Angola: Marine and Petroleum Geology, v. 9, p. 389404. Frye, J. C., 1971, The Ogallala Formationa review, in Ogallala aquifer symposium: Texas Technology University International Center for Arid and Semi-Arid Land Studies Special Report 39, p. 514a. Galloway, W. E., 1989, Genetic stratigraphic sequences in basin analysis II: application to northwest Gulf of Mexico Cenozoic basin: AAPG Bulletin, v. 73, p. 143154. Gaullier, V., J. P. Brun, G. Gurin, and H. Lecanu, 1993, Raft tectonics: the effects of residual topography below a salt dcollement: Tectonophysics, v. 228, p. 363381. Ge, H., 1996, Kinematics and dynamics of salt tectonics in the Paradox basin, Utah and Colorado: field observations and scaled modeling: Ph.D. dissertation, University of Texas at Austin, Austin, Texas, 317 p. Ge, H., M. P. A. Jackson, and B. C. Vendeville, 1994, Experimental initiation of salt structures and salt sheet emplacement by prograding sediment wedges (abs.): Geological Society of London Salt Tectonics Meeting, September 1994, Programme and Abstracts, p. 5152. Hubbert, M. K., 1937, Theory of scale models as applied to the study of geologic structures: Geological Society of America Bulletin, v. 48, p. 14591520. Humphris, C. C., Jr., 1979, Salt movement on continental slope, northern Gulf of Mexico: AAPG Bulletin, v. 63, p. 782798. Jackson, J., and D. McKenzie, 1983, The geometric evolution of normal fault systems: Journal of Structural Geology, v. 5, p. 471482. Jackson, M. P. A., and R. R. Cornelius, 1987, Stepwise centrifuge modeling of the effects of differential sediment loading on the deformation of salt structures, in I. Lerche and J. J. OBrien, eds., Dynamical geology of salt and related structures: Orlando, Florida, Academic Press, p. 163259. Jackson, M. P. A., and C. Cramez, 1989, Seismic recognition of salt welds in salt tectonics regimes, in Gulf of Mexico salt tectonics, associated processes and exploration potential: Gulf Coast SEPM Foundation Tenth Annual Research Conference Program and Abstracts, p. 6671. Jackson, M. P. A., and C. J. Talbot, 1986, External shapes, strain rates, and dynamics of salt structures: Geological Society of America Bulletin, v. 97, p. 305323. Jackson, M. P. A., and C. J. Talbot, 1989, Salt canopies, in Gulf of Mexico salt tectonics, associated processes and exploration potential: Gulf Coast Section of SEPM Foundation Tenth Annual Research Conference Program and Abstracts, p. 7278.

422

Progradation and Salt Tectonics

Jackson, M. P. A., and C. J. Talbot, 1991, A glossary of salt tectonics: University of Texas at Austin, Bureau of Economic Geology Geologic Circular 91-4, 44 p. Jackson, M. P. A., and B. C. Vendeville, 1994, Regional extension as a geologic trigger for diapirism: Geological Society of America Bulletin, v. 106, p. 5773. Krantz, R. W., 1991, Measurements of friction coefficients and cohesion for faulting and fault reactivation in laboratory models using sand and sand mixtures: Tectonophysics, v. 188, p. 203207. Larberg, G. M. B., 1983, Contra-regional faulting: salt withdrawal compensation, offshore Louisiana, Gulf of Mexico, in A. W. Bally, ed., Seismic expression of structural styles: AAPG Studies in Geology Series 15, v. 2, p. 2.3.2/422.3.2/44. Lundin, E. R., 1992, Thin-skinned extensional tectonics on a salt detachment, northern Kwanza basin, Angola: Marine and Petroleum Geology, v. 9, p. 405411. Mandl, G., 1988, Mechanics of tectonic faulting: models and basic concepts: Amsterdam, Elsevier, 407 p. McClay, K. R., 1989, Physical models of structural styles during extension, in A. J. Tankard and H. R. Balkwill, eds., Extensional tectonics and stratigraphy of the north Atlantic margins: AAPG Memoir 46, p. 95110. McClay, K. R., 1990, Extensional fault systems in sedimentary basins: a review of analogue studies: Marine and Petroleum Geology, p. 206233. McKee, E. D., and M. Goldberg, 1969, Experiments on formation of contorted structures in mud: Geological Society of America Bulletin, v. 80, p. 231244. Mohriak, W. U., 1995, Salt tectonics structural styles: contrasts and similarities between the south Atlantic and the Gulf of Mexico, in C. J. Travis, B. C. Vendeville, H. Harrison, F. J. Peel, M. R. Hudec, and B. F. Perkins, eds., Salt, sediment and hydrocarbons: Gulf Coast Section of SEPM Foundation Sixteenth Annual Research Conference, p. 177191. Mohriak, W. U., et al., 1995, Salt tectonics and structural styles in the deep-water province of the Cabo Frio region, Rio de Janeiro, Brazil, in M. P. A. Jackson, R. G. Roberts, and S. Snelson, eds., Salt tectonics: a global perspective: AAPG Memoir 65, p. 273304. Nelson, T. H., 1989, Style of salt diapirs as a function of the stage of evolution and nature of the encasing sediments, in Gulf of Mexico salt tectonics, associated processes and exploration potential: Gulf Coast Section of SEPM Foundation Tenth Annual Research Conference, p. 109110. Nelson, T. H., 1991, Salt tectonics and listric-normal faulting, in A. Salvador, ed., The Gulf of Mexico Basin: Geological Society of America, The Geology of North America, v. J, p. 7389. Peel, F. J., C. J. Travis, and J. R. Hossack, 1995, Genetic structural provinces and salt tectonics of the Cenozoic offshore U.S. Gulf of Mexico: a preliminary analysis, in M. P. A. Jackson, R. G. Roberts, and S. Snelson, eds., Salt tectonics: a global perspective: AAPG Memoir 65, p. 153175. Ramberg, H., 1981, Gravity, deformation and Earths crust in theory, experiments and geological application (2d ed.): London, Academic Press, 452 p. Rettger, R. E., 1935, Experiments on soft-rock deformation: AAPG Bulletin, v. 19, p. 271292. Robison, B. A., 1983, Low-angle normal faulting, Marysriver Valley, Nevada, in A. W. Bally, ed., Seismic expression of structural styles: AAPG Studies in Geology Series 15, v. 2, p. 2.2.2/122.2.2/16. Sannemann, D., 1968, Salt-stock families in northwestern Germany, in J. Braunstein and G. D. OBrien, eds., Diapirism and diapirs: AAPG Memoir 8, p. 261270. Schultz-Ela, D. D., 1992, Restoration of cross-sections to constrain deformation processes of extensional terranes: Marine and Petroleum Geology, v. 9, p. 372388. Schultz-Ela, D. D., and K. Duncan, 1994, Users manual and software for Restore (version 3.04): Bureau of Economic Geology, University of Texas at Austin, Austin, Texas, 97 p.

Schultz-Ela, D. D., M. P. A. Jackson, and B. C. Vendeville, 1993, Mechanics of active salt diapirism: Tectonophysics, v. 228, p. 275312. Schuster, D. C., 1993, Deformation of allochthonous salt and evolution of related structural systems, eastern Louisiana Gulf Coast (abs.): AAPG Annual Convention, Official Program, v. 2, p. 179. Schuster, D. C., 1995, Deformation of allochthonous salt and evolution of related salt-structural systems, eastern Louisiana Gulf Coast, in M. P. A. Jackson, R. G. Roberts, and S. Snelson, eds., Salt tectonics: a global perspective: AAPG Memoir 65, p. 177198. Seni, S. J., and M. P. A. Jackson, 1983a, Evolution of salt structures, East Texas diapir province, part 1: sedimentary record of halokinesis: AAPG Bulletin, v. 67, p. 12191244. Seni, S. J., and M. P. A. Jackson, 1983b, Evolution of salt structures, East Texas diapir province, part 2: patterns and rates of halokinesis: AAPG Bulletin, v. 67, p. 12451274. Szatmari, P., M. C. M. Guerra, and M. A. Pequeno, 1996, Genesis of large counter-regional normal fault by flow of Cretaceous salt in the South Atlantic Santos Basin, Brazil, in G. I. Alsop, D. J. Blundell, and I. Davison, eds., Salt tectonics: Geological Society of London Special Publication 100, p. 259264. Talbot, C. J., 1992, Centrifuged models of Gulf of Mexico profiles: Marine and Petroleum Geology, v. 9, p. 412432. Trusheim, F., 1960, Mechanism of salt migration in northern Germany: AAPG Bulletin, v. 44, p. 10681093. van Keken, P. E., C. J. Spiers, A. P. Van Den Berg, and E. J. Muyzert, 1993, The effective viscosity of rocksalt: implementation of steady state creep laws in numerical models of salt diapirism: Tectonophysics, v. 225, p. 457475. Vendeville, B. C., 1991, Mechanisms generating normal fault curvature: a review illustrated by physical models, in A. M. Yielding and B. Freeman, eds., The geometry of normal faults: Geological Society of London Special Publication 56, p. 241249. Vendeville, B. C., and M. P. A. Jackson, 1992, The rise of diapirs during thin-skinned extension: Marine and Petroleum Geology, v. 9, p. 331353. Vendeville, B. C., and M. P. A. Jackson, 1993, Some dogmas in salt tectonics challenged by modeling (abs.): AAPG Hedberg International Research Conference on Salt Tectonics, Bath, England, September 1993, p. 263265. Vendeville, B. C., P. R. Cobbold, P. Davy, J. P. Brun, and P. Choukroune, 1987, Physical models of extensional tectonics at various scales, in M. P. Coward, J. F. Dewey, and P. L. Hancock, eds., Continental extensional tectonics: Geological Society of London Special Publication 28, p. 95107. Vendeville, B. C., H. Ge, and M. P. A. Jackson, 1994, Experimental deformation of prograding sedimentary wedges above a viscous source layer (abs.): AAPG Annual Convention, Official Program, v. 3, p. 276. Weijermars, R., 1986, Flow behaviour and physical chemistry of bouncing putties and related polymers in view of tectonic laboratory applications: Tectonophysics, v. 124, p. 325358. Weijermars, R., M. P. A. Jackson, and B. C. Vendeville, 1993, Rheological and tectonic modeling of salt provinces: Tectonophysics, v. 217, p. 143174. Winker, C. D., 1982, Cenozoic shelf margins, northwestern Gulf of Mexico basin: Gulf Coast Association of Geological Societies Transactions, v. 32, p. 427448. Woodbury, H. O., I. B. Murray, Jr., and R. E. Osborne, 1980, Diapirs and their relation to hydrocarbon accumulation, in A. D. Miall, ed., Facts and principles of world petroleum occurrence: Canadian Society of Petroleum Geologists, p. 119142. Worrall, D. M., and S. Snelson, 1989, Evolution of the northern Gulf of Mexico, with emphasis on Cenozoic growth faulting and the role of salt, in A. W. Bally and A. R. Palmer, eds., The Geology of North Americaan overview: Boulder, Colorado, Geological Society of America, v. A, p. 97138. Wu, S., 1993, Salt and slope tectonics, offshore Louisiana: Ph.D.

Ge et al.

423

thesis, Rice University, Houston, Texas, 251 p. Wu, S., A. W. Bally, and C. Cramez, 1990, Allochthonous salt struc-

ture and stratigraphy of the northeastern Gulf of Mexico, part II: structure: Marine and Petroleum Geology, v. 7, p. 343370.

ABOUT THE AUTHORS


Hongxing Ge Hongxing Ge received his B.S. degree from Nanjing University, Peoples Republic of China, in 1985, and his M. S. degree from Colorado State University in 1990. He completed his Ph.D. at the University of Texas at Austin in 1996. His dissertation on Paradox basin salt tectonics was nominated for the Outstanding Dissertation Award. He is currently a postdoctoral fellow at the Bureau of Economic Geology. His interests include salt tectonics, tectonic modeling, strain analysis, and petroleum geology. Martin Jackson Martin Jacksons early career included lunar structures, mineral exploration, and Precambrian geology. He received his Ph.D. from the University of Cape Town in 1976, and joined the Bureau of Economic Geology in 1980, where he directs the Applied Geodynamics Laboratory funded by a consortium of oil companies. A recipient of AAPGs Sproule Award (with S. J. Seni) and Matson Award, he lectured in AAPGs Structural Geology School, was an AAPG Distinguished Lecturer, and served 6 years as associate editor for the AAPG Bulletin and GSA Bulletin. Bruno C. Vendeville Bruno C. Vendeville received his Ph.D. from the Universit de Rennes, France, in 1987. He specializes in experimental modeling of tectonic processes with emphasis on gravity tectonics, salt tectonics, and extensional tectonics. He was the co-recipient of two honorable mentions from the SEPM/AAPG (1993) and AAPG (1990) with M. P. A. Jackson. He is currently a research scientist at the Bureau of Economic Geology, University of Texas at Austin.

You might also like