Top Banner
Tectonophysics, 226 (1993) 217-225 Elsevier Science Publishers B.V., Amsterdam 217 Relationship between deeper lithospheric processes and near-surface tectonics of sedimentary basins G. Quinlan a, J. Walsh b, J. Skogseid ‘, W. Sassi d, S. Cloetingh e, L. Lobkovsky f, C. Bois d, H. Stel e and E. Banda g a Memorial University, St. John’s, Canada b University of Liverpool, UK ’ University of Oslo, Norway d Institut Francais du Petrole, France e Vrije Universiteit, The Netherlands f Russian Academy of Sciences, Russia g Institute Jaume Almera, Barcelona, Spain (Received November 20,1992; revised version accepted June 15,1993) Introduction Interactions between upper lithosphere defor- mation, as manifested in the structure and evolu- tion of sedimentary basins, and lower lithospheric deformation presumably reflect large-scale re- gional or global tectonic events. We review the current state of knowledge regarding these inter- actions and suggest areas of research which may lead to an improvement in this knowledge. Understanding the development of sedimen- tary basins requires insight into the processes responsible both for basin subsidence and also for distribution and preservation of sediments within the basins. While some classes of basins are better understood than others, there remain significant gaps both in our ability to account for available data and also in the data available to test current models and develop new ones. Subsidence in some classes of basins can be related to lithospheric scale processes occurring as part of the overall plate tectonic paradigm. For example, it is widely accepted that passive margin basins subside as a result of lithospheric exten- sion (McKenzie, 1978) and that foreland basins are flexural consequences of collision between a continent and outboard terranes (Beaumont, 1981). On the other hand, models for the origin of intracratonic, forearc and trans-tensional basins are essentially still at cartoon level, involving pro- cesses that are difficult to quantify and evaluate. In the following we review a number of funda- mental questions of vital interest for understand- ing of the relationship between deeper litho- spheric processes and near-surface tectonics. Fundamental questions How valid are current kinematic models of basin evolution? Most existing models for basin evolution are only kinematic, i.e., they assume some pattern of lithospheric deformation and predict basin evolu- tion based on the isostatic consequences. Little attention is paid to whether the assumed litho- spheric deformation pattern is mechanically rea- sonable or what conditions are necessary for its realization. The natural variability that appears between superficially similar basins has led to ad hoc introduction of increasing complexity into kinematic models. For example, models of exten- sional basins alternatively assume pure or simple shear extension. Pure-shear models may involve either uniform extension (McKenzie, 1978) or depth-dependent extension (Royden and Keen, 1980). Simple-shear models may involve either whole lithosphere extension (Wernicke, 1985) or 0040-1951/93/$06.00 0 1993 - Elsevier Science Publishers B.V. All rights reserved
9

Relationship between deeper lithospheric processes and near-surface tectonics of basins

Jan 10, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Relationship between deeper lithospheric processes and near-surface tectonics of basins

Tectonophysics, 226 (1993) 217-225

Elsevier Science Publishers B.V., Amsterdam

217

Relationship between deeper lithospheric processes and near-surface tectonics of sedimentary basins

G. Quinlan a, J. Walsh b, J. Skogseid ‘, W. Sassi d, S. Cloetingh e, L. Lobkovsky f, C. Bois d, H. Stel e and E. Banda g

a Memorial University, St. John’s, Canada b University of Liverpool, UK ’ University of Oslo, Norway

d Institut Francais du Petrole, France e Vrije Universiteit, The Netherlands

f Russian Academy of Sciences, Russia g Institute Jaume Almera, Barcelona, Spain

(Received November 20,1992; revised version accepted June 15,1993)

Introduction

Interactions between upper lithosphere defor- mation, as manifested in the structure and evolu- tion of sedimentary basins, and lower lithospheric deformation presumably reflect large-scale re- gional or global tectonic events. We review the current state of knowledge regarding these inter- actions and suggest areas of research which may lead to an improvement in this knowledge.

Understanding the development of sedimen- tary basins requires insight into the processes responsible both for basin subsidence and also for distribution and preservation of sediments within the basins. While some classes of basins are better understood than others, there remain significant gaps both in our ability to account for available data and also in the data available to test current models and develop new ones.

Subsidence in some classes of basins can be related to lithospheric scale processes occurring as part of the overall plate tectonic paradigm. For example, it is widely accepted that passive margin basins subside as a result of lithospheric exten- sion (McKenzie, 1978) and that foreland basins are flexural consequences of collision between a continent and outboard terranes (Beaumont, 1981). On the other hand, models for the origin of intracratonic, forearc and trans-tensional basins

are essentially still at cartoon level, involving pro- cesses that are difficult to quantify and evaluate. In the following we review a number of funda- mental questions of vital interest for understand- ing of the relationship between deeper litho- spheric processes and near-surface tectonics.

Fundamental questions

How valid are current kinematic models of basin evolution?

Most existing models for basin evolution are only kinematic, i.e., they assume some pattern of lithospheric deformation and predict basin evolu- tion based on the isostatic consequences. Little attention is paid to whether the assumed litho- spheric deformation pattern is mechanically rea- sonable or what conditions are necessary for its realization. The natural variability that appears between superficially similar basins has led to ad hoc introduction of increasing complexity into kinematic models. For example, models of exten- sional basins alternatively assume pure or simple shear extension. Pure-shear models may involve either uniform extension (McKenzie, 1978) or depth-dependent extension (Royden and Keen, 1980). Simple-shear models may involve either whole lithosphere extension (Wernicke, 1985) or

0040-1951/93/$06.00 0 1993 - Elsevier Science Publishers B.V. All rights reserved

Page 2: Relationship between deeper lithospheric processes and near-surface tectonics of basins

ramp-flat-ramp style extension (Gibbs, 1989). The apparent need to introduce such variety sug- gests that the lithosphere probably does deform in a variety of different ways (Buck, 1991). Why is

this so? Are the proposed models reasonable kinematic equivalents of the dynamic processes by which the lithosphere actually deforms‘?

Some progress in evaluating the dynamics of extensional basin formation has been made using finite element techniques (Braun and Beaumont, 1989; Dunbar and Sawyer, 1989). In this ap- proach, a simplified rheological structure for the lithosphere is represented by a finite element grid, specific stresses are applied to this grid and deformation is followed as it evolves. Explicit in these models is the presence of crustal and man- tle strong and weak layers (Carter and Tsenn, 1987); stress is partitioned into the strong layers and strain into the weak layers during deforma- tion. The virtue of such models is not so much that they do a superior job of explaining basin architecture as that they allow examination of how basic properties of the lithosphere influence the style of deformation. The most productive approach with such models has been to test the relative importance of factors such as crustal and mantle weaknesses and geothermal gradients in controlling the style of lithospheric deformation. It becomes possible to ask in what tectonic envi- ronments lithospheric deformation could approxi- mate that assumed as the starting point of the various kinematic models. In general, variations in geothermal gradient, strain rate and the pres- ence or absence of imperfections within the litho- spheric strong layers can produce extension pat- terns resembling one or another of the various kinematic deformation patterns. It must be re- membered, however, that construction of dy- namic models is severely limited by our knowl- edge of lithospheric rheology. It is probably pre- mature to say that dynamic models have put kinematic ones on a firm theoretical footing but the footing is certainly firmer.

Very little work has been done on dynamic models of compressional tectonics. Jamieson and Beaumont (1989) have used such techniques to investigate large-scale deformation within com- pressional orogens and to relate this deformation

to metamorphism and uplift history 01’ near- surface rocks. In this work an orogen is treated as an elasto-plastic material the yield strength ot which varyies with depth according to Byerlee’s Law. Brittle failure is thus approximated as a ductile process and the deformation that occurs

in reality on systems of dicrete faults is averaged over a large area of ductile deformation. This model does not concern itself with the details of deformation on a lithospheric scale but rather only on an upper-crustal scale.

In order to evaluate such models considerable information is needed on the variation in meta- morphic grade and cooling ages as a function of distance from an orogen. Field studies such as described by Kamp et al. (1989) from the South- ern Alps of New Zealand are important. As a general field area for studying the dynamics of compressional orogeny, the Southern Alps area is among the world’s best because of the rapid uplift and good exposure. Deep seismic reflection data across the area would be particularly wel- come as providing a glimpse of the internal struc- ture of an actively deforming orogen.

What is the influence of regional thermo-mechani- cal history?

Most models assume a homogeneous starting configuration for the lithosphere although there is fairly clear evidence that compressional oro- gens form on passive margins and that continen- tal rifting tends to be localized near the site of earlier compressional orogens. Insufficient atten- tion has been paid to how the thermo-mechanical history of an area affects basin localization and development (Stockmal et al., 1986). These ef- fects can be studied in the areas of southern Spain (e.g., Van Wees et al., 1992) and France where basins of extensional and compressional style have alternated on time-scales that are short compared with thermal relaxation time of the lithosphere. Another important topic, which can be studied excellently in the western Mediter- ranean basins is the formation of extensional basins in a tectonic regime of overall conver- gence.

Page 3: Relationship between deeper lithospheric processes and near-surface tectonics of basins

RELATIONSHIP BETWEEN DEEPER LITHOSPHERIC PROCESSES AND NEAR-SURFACE TECTONICS 219

How do sedimentological factors inj7uence basin

evolution?

Although the success of any individual basin model is often gauged by its ability to reproduce the observed sedimentary record, few models deal realistically with sediment transport and preser- vation. For example, patterns of unconformities within foreland basin stratigraphy that have been attributed to rheological behaviour of the litho- sphere (Quinlan and Beaumont, 1984) may in fact be related to neglected aspects of sediment trans- port (Flemming and Jordan, 1989). Sediment transport and post-depositional alteration within the basin also have significant influence on the evolution of large-scale basin architecture through time (Watts, 1989). This is true both because the sediment load itself modifies basin subsidence and because the post-depositional compaction and diagenesis of sediment affects the room available for additional sediment. Since erosion and sediment transport can be modelled to first order as diffusional processes dependent on local topography (Turcotte and Kenyon, 19851, the way is clear to incorporate such effects into quantita- tive basin models.

What is the influence of magmatism at rifted mar- gins?

Although the relation between rifting, magma- tism and subsidence can be modelled quantita- tively on a large scale (McKenzie and Bickle, 1988), there is a poor understanding of the de- tailed geometric and mechanical processes occur- ring during the formation of rift-related basins. What are the effects of non-conservative pro- cesses, such as partial melt diffusion and man- tle-crust interaction, on crustal geometry, me- chanical behaviour and thermal regime? Dynamic models suggest that these are important consider- ations for the ultimate behaviour of an extending region. Detailed structural, geochemical and geochronological investigations in fossil rift zones should be compared with surface magmatism, geophysical data and drilling results from recent rifts such as the northeastern Atlantic volcanic rifted margins and the Rhine graben. The Oslo

rift zone is a suitable area for studies of fossil

rift-magmatism relations. The Jotnian relict basins of Sweden and Finland are suggested for

evaluation of the relation between anorogenic granites and surface subsidence, a relation sug- gested to control the subsidence of cratonic basins (Klein and Hsui, 1987). Both these areas are located on the Baltic Shield where an excellent set of refraction seismic data is available (Luosto et al., 1990). Significant studies of gravity, geo- chronology and geochemistry have been carried out (Nurmi and Haapala, 1986) and the tectonic framework is closely related to the well docu- mented evolution of the North Sea basin (Zie- gler, 1990). The northeastern Atlantic region, in particular the early Tertiary volcanic rifted mar- gins, are suitable for studying a succession of rifting episodes since Caledonian collision, of which the last was associated with significant and widespread magmatism (Morton and Parson, 1990). Formation of large igneous provinces, in- cluding volcanic margins, is at present considered to reflect major global events, whose genesis and evolution are directly linked to mantle dynamics (White and McKenzie, 1989; Coffin and Eldholm, 1991). This opens the possibility to study the interplay between mantle processes, lithospheric extension and basin formation (Pedersen and Ro, 19921, how igneous activity modifies patterns of basin subsidence (Skogseid et al., 1992a,b) and, more indirectly, how intrabasinal stress regime influence, or guide the emplacement of intrusives (Sundvoll et al., 1990).

How can brittle deformation be modelled?

An important question is how strain rate af- fects the formation of new faults as opposed to the reactivation of old ones (Knipe, 1985) or influences stress regimes that are not parallel to those in which the original faults formed. It is not obvious how discontinuities such as faults can be treated quantitatively (see Vilotte et al., 1993-this volume). The rotating domino model of faulting has been shown to be generally consistent with observations of faulting in both active and inac- tive (ancient) extensional sedimentary basins (Jackson and White, 1989). Large normal faults

Page 4: Relationship between deeper lithospheric processes and near-surface tectonics of basins

are approximately planar and appear to rotate as they move with tilting of intervening fault blocks (Jackson, 1987). Observed vertical movements of fault blocks are in general agreement with this model, although it is recognised that the motion of rigid block is an over-simplification (more real- istic variants of this geometric model are avail- able). Two features of the model are that it allows estimates of regional extension to be made and that the spacing of large faults may be regu- lar, possible with the maximum size of fault blocks (spacing of faults) controlled by the thickness of the seismogenic layer. Recent work has, however, shown that the scaling properties of fault popula- tions (expressed in terms of displacements or dimensions) are fractal (Kakimi, 1980; Villemin and Sunwoo, 1987; Childs et al., 1990; Sassi et al., 19921, a conclusion which casts some doubt on the reliability of extension estimates. Small faults, with displacement below the limit of seismic reso- lution, may account for apparent discrepancies between estimates of extension from the upper- crustal faulting and estimates derived from obser- vations of crustal thickness and subsidence (Mar- ret and Allmendinger, 1990; Walsh et al., 1991). Of prime interest are:

(1) the segmentation of faults, which may con- trol sediment distibution and fluid migration;

(2) the relation between fault size and spacing, which controls the strain distibution within the basin;

(3) the spatial evolution and growth of fault systems, which is central to the geological history of the basin.

Further research should be directed towards a better understanding of the nature and origin of fault “size” and spacing populations. A theoreti- cal basis for a fractal nature has yet to be estab- lished and this will require integration of earth- quake studies and the quantitative analysis of ancient faults.

A better understanding of the scaling proper- ties of fault populations requires the quantitative analysis of data from reflection seismics, on both regional and reservoir scale, and from outcrop and core studies. Particular emphasis should be placed on datasets which allow the spatial varia- tions in fault populations to be analysed and

provide data covering a broad range of scales. The North Sea and Gulf Coast are identified as key areas for investigation because faulting in these areas is the result of two contrasting tec- tonic regimes and because of the abundance of data and the availability of a relatively well de- fined geological context.

How valid are section balancing techniques?

The past decade has seen the emergence of section balancing as a widely accepted technique for the analysis of extensional basins. The main principles of section balancing were first devel- oped for regions of contractional tectonics and mainly derive from a requirement to produce cross-sections for which there is no change in bed lengths and areas when it is restored to the original undeformed state. There are several rea- sons why section balancing, despite its recogni- tion as a valuble tool, is not easily applied to extensional basins. Compaction and the synsedi- mentary nature of many faults, involving sedi- mentation and erosion, means that section bal- ancing of the syn-rift sequence cannot be per- formed routinely. Additional complications are that the analysis generally requires the depth migration of seismic sections and that field expo- sures are usually not available to provide con- straints on the balancing procedure. These prob- lems have led to the recognition that section balancing of extensional basins is best performed by backstripping techniques in which the section is restored by sequential decompaction, sediment removal and restoration of fault movements. Al- though this procedure will generally provide valu- able insights into the evolution of a basin and will allow gross errors in interpretation to be identi- fied, the quality of available data dictates that balanced sections, even without the complications of sedimentation and compaction, are not neces- sarily correct. The accuracy of a restored section depends not only on the quality of data but also on the validity of various fundamental assump- tions incorporated in the balancing/ restoring process. Since fault restoration models are cen- tral to the construction of balanced cross-sections it is this aspect with requires particular attention.

Page 5: Relationship between deeper lithospheric processes and near-surface tectonics of basins

RELATIONSHIP BETWEEN DEEPER LITHOSPHERIC PROCESSES AND NEAR-SURFACE TECTONICS 221

Several topics are identified below, although it is recognised that valuable insights will be provided simply from the analysis of well constrained, depth-converted sections.

(1) Rigid domino models are increasingly find- ing applications in the analysis of extensional basins. It is generally accepted, however, that the concept of “rigid” dominos is unrealistic and work should be directed towards the develop- ment of models which incorporate internal defor- mation of the fault blocks, incorporating elastic dislocation theory and flexural isostasy (e.g., King et al., 1988).

(2) Methods which use hangin~all fold ge- ometries to predict the shapes of faults and calcu- late depths to detachment are often used in sec- tion balancing (Gibbs, 1983, 1984; Williams and Vann, 1987). Available constructions often pre- dict listric fault shapes with sub-horizontal de- tachment surfaces, which are geometries inconsis- tent with those of large faults in many areas (Jackson, 1987). Should the application of these techniques be restricted to certain types of faults? The assumed mechanisms of hangin~all defor- mation are many and varied but a mechanical basis is entirely lacking for most of them. These shortcomings need to be addressed either through detailed analysis of faults from outcrop and from well constrained seismic datasets, or by analogue modelling.

(3) The significance of small-scale faulting, with displacements below the limits of seismic resolu- tion, needs to be analysed and, if necessary, their effects should be incorporated in balancing pro- cedures.

(4) Displacement analysis is a new technique which may provide constraints on the restoration process (Beach and Traynor, 1991; Chapman and Meneilly, 1990; Walsh and Watterson, 1991). Em- phasis is placed on the requirement for consider- ation of structure in 3-D, an approach which, when combined with sequential restoration, will help to decrease the number of acceptable resto- rations.

How do eva~~te layers i~~~e~ce basin evolution?

A related issue concerns the role of evaporite/ clay layers in controlling extensional basin geom-

etry. Syn-extensional activation of evaporites in sedimentary basins is common but difficult to evaluate. Little is known of how ductile layers transmit fault-controlled extension of basement to the cover sequences or how such layers react to differential loading by uneven sedimentation (Jenyon, 1986). Both questions are basic to un- derstanding the relation between large-scale tec- tonics and detailed basin geometry and evolution. In fact, more precise knowledge of the geometry and structural evolution of areas containing evap- orite/clay layers may shed light on the behaviour of brittle-ductile-brittle layered systems as a whole and so become scaled-down models for the lithosphere. Field areas where such observations can be made are the Sverdrup basin of Arctic Canada and the Prebetic zone of Spain. Evapor- ite structures in the Sverdrup basin are both well exposed and particularly well preserved by the dry climate (Schwerdtner and Osadetz, 1983).

Construction and evaluation of kinematic and dynamic models is presently limited as much by our knowledge of lithospheric rheology, composi- tion and structure as by difficulties of incorporat- ing such info~ation into the models.

What is really known about the lower lithosphere?

Information about the lower lithosphere has come from three basic sources: (1) direct observa- tion in tectonically exposed sections, (2) labora- tory studies of supposedly representative rock types, and (3) geophysical investigations, primar- ily seismic and electro-magnetic.

Since the 1970’s deep seismic reflection profil- ing of the continental lithosphere has produced many records in the range of 15-20 s two-way- time, adequate to see a few tens of kilometers into the upper mantle. Unfortunately, acquisition of such data is much farther advanced than the secure interpretation of what the data mean. This is due in large part to the failure of most deep reflectors to connect with surface features where their age and significance can be inferred from traditional geological approaches. In the absence of direct correlation with the surface, one can construct self-consistent but rarely unique models of lithospheric tectonics. Poorly constrained in-

Page 6: Relationship between deeper lithospheric processes and near-surface tectonics of basins

222

ferences are often exacerbated by a failure to complement seismic reflection surveys with re- lated studies such as seismic refraction, gravity and magnetic surveying and electromagnetic sounding which can be useful in further con- straining reflection interpretations,

The more deep seismic reflection data are gathered, the more diversity appears in the re- flection patterns obtained. However, some pat- terns appear frequently enough to demand expla- nation by even first-order modeis of lithospheric deformation. Frequently seen features include: dipping reflectors in the upper crust that can often be traced into mapped faults or shear zones and that are therefore considered to represent faults or shear zones themselves; sub-horizontal bands of reflectivity in the middle or, more com- monly, in the lower crust; dipping reflectors that appear to cross-cut the base of the crust and extend a short distance into the upper mantle. Occasionally these latter reflectors can be traced to greater depths as with the BIRPS data north and west of Britain (Warner and McGeary. 1987).

If dipping reflectors in the upper crust do represent faults, then their termination at depth has certain implications. Since motion on a fault cannot terminate abruptly, this motion must be taken up either by other faults of different style and orientation or by distributed ductile deforma- tion. In the former case one might expect the orientation of reflectors to change but the reflec- tors still to be visible. In the latter case ductile distributed strain may not necessarily produce seismically reflecting horizons and so the dipping reflectors could simply terminate in a seismically transparent zone. In any case it is valid to ask what prevents the dipping reflectors from contin-

uing unchanged to greater depth. Rheological models of the lithosphere based

on laboratory deformation of appropriate rock types suggest possible explanations. According to these studies, the upper crust and upper mantle are rheologically brittle layers separated by layers that deform by ductile processes (Carter and Tsenn, 1987). These are the alternating strong and weak zones incorporated in dynamic models of lithospheric deformation. Seismic profiles showing reflectors (faults) that terminate at

depths roughly corresponding with these ex- pected weak zones are interpreted as defining lithospheric detachments across which there are major changes in the orientation and possibly style of deformation (e.g., Keen et al,, 19871. These detachment surfaces are essential compo- nents of simple shear extension models in which they are expected to link offset, dipping faults or shear zones (Gibbs, 19871.

It should be remembered, however, rhat de- tachment surfaces are only a plausible and setf- consistent interpretation of the available seismic and rheological data. Results of laboratory defor- mation experiments must be extrapolated over 8-10 orders of magnitude in strain-rate when applied to the earth. Even accepting this extrapo- lation, the expected depth of these detachments depends strongly on mineralogy and on the ther- mal regime at the time of deformation, With current knowledge, it is not possible, for example, to predict whether dipping faults should detach at specific levels on specific seismic profiles and use the sections as a test of such predictions. Instead, the observed termination of dipping re- flectors at middle and/or lower crustal depths is taken as supporting the existence of detachments that are beyond our present capacity to predict in any detailed manner.

Lower-crustal reflectivity is generally attrib- uted to one or more of three causes: (11 fluid-filled pore space; (2) lenticular mafic intrusions; and (3) shearing within the lower crust. High-pressure rocks exhumed from lower-crustal depths are characteristically granulite-facies anhydrous mafic silicates in which primary lithoiogic layering has often undergone ductile shearing. Interestingly, such reflectivity patterns are not as obvious in near-surface granulite terranes exhumed from lower-crustal depths Nlemperer and Matthews, 1987). This suggests that the reflectivity of the lower crust is at least partially controlled by other, in situ, factors that do not survive transport to the surface.

The presence of fluid layers in the lower crust is often invoked to explain both the seismic re- flectivity pattern and the high electrical conduc- tivity inferred for some regions of lower crust (Hyndman, 1988). The existence of fluids in the

Page 7: Relationship between deeper lithospheric processes and near-surface tectonics of basins

RELATIONSHIP BETWEEN DEEPER LI~OSPHERIC PROCESSES AND NEAR-SURFACE TECTONICS 223

deep crust is a controversial topic however since they are difficult to reconcile with the dry miner- alogy of granulite facies rocks.

The true significance of lower-crustal reflectiv- ity is important for an understanding of litho- spheric deformation. If, for example, there is connected ~uid-filled porosity in the lower crust, this pore fluid must be unable to escape to the surface in geologically relevant timescales. Pore space could not be supported by the rock matrix at the ductile conditions thought to prevail in the lower crust. Any fluid present would be at litho- static pressure and may play a poorly understood, but potentially major, role in modifying the strength of the lithosphere. ~ternatively, mafic lenses could result from magmatic intrusion dur- ing extension and so be as much an effect of, as a contributor to, deformation.

Knowledge of Poisson’s ratio in the lower crust would be useful in assessing the physical signifi- cance of the reflectors. In principle, this can be obtained from P- and S-wave velocities and depth; however, S-wave studies are still in their infancy. Changes in Poisson’s ratio influence pre-critical reflection amplitude as a function of offset and so amplitude vs. offset studies of high-quality reflec- tion data may shed light on this parameter.

There are legitimate questions as to whether reflectors appearing to originate in the upper- most mantle actually originate in the lower crust. There is often poor control on lower-crustal seis- mic velocity coincident with deep reflection data. This translates into uncertainty in where these dipping reflectors migrate. It is rare for them to remain unambiguously within the upper mantle after migration. In addition, reflections from shal- lower features that are out of the plane of the seismic section may appear to come from mantle depths although some theoretical studies suggest that these out-of-plane reflections should be of low amplitude. This type of problem is difficult to compensate for in the absence of laterally con- necting seismic data.

Mantle reflectors may be a genuine rarity or only apparent caused by the limitations of avail- able data. The few undisputed examples of man- tle reflectors come from depths at or beyond the 20 s two-way-time commonly reached in deep

seismic reflection surveys. Reflection sections

recorded to greater times using appropriately modified energy sources and recording parame- ters are needed. Refraction data suggest that there is significant velocity layering in the upper 100 km of the mantle and that the velocity con- trasts can be as great as across the Moho (Fuchs et al., 1987). This does not guarantee the exis- tence of reflectivity contrasts but does suggest that a search for mantle reflectors could well be fruitful (Lie and Husebye, 1991).

Dipping reflectors in the upper mantle are key elements of various tectonic models for extension and compression. They may, for example, repre- sent the sub-crustal portion of shear zones in simple-shear models of lithospheric extension (Wernicke, 1985). Alternatively they may record shortening of the upper lithosphere above a de- lamination surface in the late stages of conti- nent-continent collision (Stockmal et al., 1986). Sub-horizontal reflectors in the mantle (Lie and Husebye, 1991) are even more ambiguous. Do they have a structural meaning? Are they indica- tors of strain distribution in the mantle? What is their age? What is the role of melting and/or fluids with regard to these reflectors? Answers to these questions require more data on the fine structure of the subcrustal lithosphere.

References

Beach, A. and Traynor, P., 1991. The geometry of normal

faults in a sector of the offshore Nile Delta, Egypt. In: A.M. Roberts et al. (Editors), The Geometry of Normal Faults. Geol. Sot. London, Spec. Publ., 56: 173-182.

Beaumont, C., 1981. Foreland basins. Geophys. J.R. Astron. Sot., 65: 291-329.

Braun, J. and Beaumont, C., 1989. A physical explanation of the relation between the Bank uplifts and the breakup unconformity at rifted continental margins. Geology, 17: 760-764.

Buck, W.R., 1991 Modes of ~ntinental lithospheric exten- sion. J. Geophys. Res., 96: 20,161~20,178.

Carter, N.L. and Tsenn, M.C., 1987. Flow properties of conti- nental lithosphere. Te~onoph~i~s, 136: 27-63.

Chapman, T.J. and Meneilly, A.W., 1990. Fault displacement analysis in seismic exploration. First Break, 8: 11-22.

Childs, C., Walsh, J.J. and Watterson, J., 1990. A method for estimation of the density of fault displacements below the limit of seismic resolution in reservoir formations. In:

Page 8: Relationship between deeper lithospheric processes and near-surface tectonics of basins

224

North Sea Oil and Gas Reservoirs, II. Graham and Trot-

man, London.

Coffin, M.L. and Eldholm, 0. (Editors), 1991. Large Igneous

provinces: JOI/USSAC Workshop Report. Techn. Rep.

Univ. Texas, Austin Inst. Geophys., 114 pp.

Dunbar, J.A. and Sawyer, D.S., 1989. How preexisting weak-

ness zones control the style of continental breakup. J.

Geophys. Res., 94: 7278-7292.

Flemming, P.B. and Jordan, T.E., 1989. A synthetic strati-

graphic model of foreland basin development. J. Geophys.

Res., 94: 3851-3866.

Fuchs, K., Bonjer, K.-P., Gajewski, D., Luschen, E., Prodehl.,

C., Sandmeier, K.-J., Wenzel, F. and Wilhelm, H., 1987.

Crustal evolution of the Rhinegraben area, I. Exploring

the lower crust in the Rhinegraben rift by unified geophys-

ical experiments. Tectonophysics, 141: 261-275.

Gibbs, A.D., 1983. Balanced cross-section constructions from

seismic sections in areas of extensional tectonics, J. Struct.

Geol., 5: 152-160.

Gibbs, A.D., 1984. Structural evolution of extensional basins.

J. Geol. Sot. London, 141: 609-620.

Gibbs, A.D., 1987. Development of extensional and mixed

mode sedimentary basins. In: M.P. Coward, J.F. Dewey

and P.L. Hancock (Editors), Continental Extensional Tec-

tonics. Geol. Sot. London. Spec. Publ., 28: 19-33.

Gibbs, A.D., 1989. Structural styles in basin formation. In:

A.J. Tankard and H.R. Balkwill (Editors), Extensional

Tectonics and Stratigraphy of the North Atlantic Margins.

Am. Assoc. Pet. Geol. Mem., 46: 81-93.

Hirata, T., 1989. Fractal dimensions of faults systems in

Japan: Fractal structure in rock fracture at various scales.

Pure Appl. Geophys., 131: 157-170.

Hyndman, R.D., 1988. Dipping seismic reflectors, electrically

conductive zones and trapped water in the crust over a

subducting plate. J. Geophys. Res., 93: 13,391-13,405.

Jackson, J.A., 1980. Reactivation of basement faults and

crustal shortening in erogenic belts. Nature, 283: 343-346.

Jackson, J.A., 1987. Active normal faulting and crustal exten-

sion. In: M.P. Coward, J.F. Dewey and P.L. Hancock

(Editors), Continental Extensional Tectonics. Geol. SOC.

London Spec. Publ., 28: 3-17.

Jackson, J.A. and White, N.J., 1989. Normal faulting in the

upper continental crust: observations from regions of ac-

tive extension. J. Struct. Geol., 11: 15-36.

Jamieson, R.A. and Beaumont, C., 1989. Deformation and

metamorphism in convergent orogens: a model for uplift

and exhumation of metamorphic terrains. In: J.S. Daly,

R.A. Cliff and B.W.D. Yardley (Editors), Evolution of

Metamorphic Belts. Geol. Sot. London Spec. Publ., 43:

117-129.

Jenyon, M.K., 1986. Salt Tectonics. Elsevier, New York, NY,

191 pp.

Kakimi, T., 1980. Magnitude-frequency relation for displace-

ment of minor faults and its significance in crustal defor-

mation. Bull. Geol. Surv. Jpn., 31: 467-487.

Kamp, P.J., Green, P.F and White, S.H., 1989. Fissron track

analysis reveals character of collisional tectonics in New

Zealand. Tectonics, 8: 169- 195.

Keen, C.E., Boutilier, B., De Voogd, B.. Mudford, B. and

Enachescu, M.E., 1987. Crustal geometry and extension

models for the Grand Banks, eastern Canada: constraints

from deep seismic reflection data. In: C. Beaumont and

A.J. Tankard (Editors), Sedimentary Basins and Basin-fot--

ming Mechanisms. Can. Sot. Pet. Geol. Mem., 12: 101-l 16.

Kenyon. P.M. and Turcotte, D.L., 1985. Morphology of a

delta prograding by bulk sediment transport. Geol. Sot.

Am. Bull., 96: 1457-1465.

King, G.C.P., Stein, R.S. and Rundle, J.B., 1988. The growth

of geological structures by repeated earthquakes. I: Con-

ceptual framework. J. Geophys. Res., 93: 13,307-13,319.

Klein, G. de V. and Hsui, A.T., 1987. Origin of cratonic

basins. Geology, 15: 1094-1098.

KIemperer S.L. and Matthews D.H., 1987. Iapetus suture

located beneath the North Sea by BIRPS deep seismic

reflection profiling. Geology, 15: 195-198

Knipe, R., 1985. Footwall geometry and the rheology of thrust

sheets. J. Struct. Geol., 71: l-10.

Lie, J.E. and Husebye, ES., 1991. The structure of the crust

below Skagerrak: new results from deep reflection seismic

data. Terra Abstr., 6th EGU Meet., Strasbourg, p. 121

Luosto, U., Tiira, T., Korhonen, H., Azbel, I., Burmin, V.,

Buyanov, A., Kominskaya, I., Ionkis, V. and Sharov, N.,

1990. Crust and upper mantle structure along the DSS

Baltic profile in SE Finland. Geophys. J. Int., 101: 89-110.

Marrett, R. and Allmendinger, R.W., 1990. Kinematic analy-

sis of fault slip data, J. Struct. Geology, 12: 973-986.

McKenzie, D.P., 1978. Some remarks on the development of

sedimentary basins. Earth Planet. Sci. Lett., 40: 25-32.

McKenzie, D.P. and Bickle, M.J.. 1988. The volume and

composition of melt generation by extension of the litho-

sphere. J. Petrol., 29: 625-679.

Morton, A.C. and Parson, I.M., 1988. Early Tertiary volcan-

ism and the opening of the NE Atlantic. Geol. Sot.

London Spec. Publ., 39.

Nurmi, P. and Haapala, I., 1986. The Proterozoic granitoids

of Finland; Granite types, metallogeny and related crustal

evolution. Bull. Geol. Sot. Finl., 58: 241-261.

Okubo, P.G. and Aki, K., 1987. Fractal geometry in the San

Andreas fault system. J. Geophys. Res., 92: 345-355.

Pedersen, T. and Ro, H.E., 1992. Finite duration extension

and decompression melting. Earth Planet. Sci. Lett., 113:

15-22.

Quinlan, G.M. and Beaumont, C., 1984. Appalachian thrust-

ing, lithosperic flexure and the Paleozoic stratigraphy of

the Eastern Interior of North America. Can. J. Earth. Sci..

2 1: 973-996.

Royden, L. and C.E. Keen, 1980. Rifting process and thermal

evolution of the continental margin of eastern Canada

determined from subsidence curves. Earth Planet. Sci.

Lett., 51: 343-361.

Page 9: Relationship between deeper lithospheric processes and near-surface tectonics of basins

RELATIONSHIP BETWEEN DEEPER LITHOSPHERIC PROCESSES AND NEAR-SURFACE TECTONICS 225

Sassi, W., Livera, SE. and Caline, B.P.R., 1992. Reservoir

compartmentation by sub-seismic scale faults in Comorant Block IV, UK, Northern North Sea. In: R.M. Larsen, H. Brekke, B.T. Larsen and E. Talleraas (Editors), Structural and Tectonic Modelling and its Application to Petroleum Geology. Elsevier, Amsterdam, pp. 355-364.

Schwerdtner, W.M. and Osadetz, K., 1983. Evaporite di- apirism in the Sverdrup Basin: New insights and unsolved

problems. Bull. Can. Pet. Geol., 31: 27-36. Skogseid, J., Pedersen, T. and Larsen, V.B., 1992a. Voring

Basin: subsidence and tectonic evolution. In: R.M. Larsen,

H. Brekke, B.T. Larsen and E. Talleraas (Editors), Struc- tural and Tectonic Modelling and its Application to

Petroleum Geology. Elsevier, Amsterdam, pp. 55-82. Skogseid, J., Pedersen, T., Eldholm, 0. and Larsen, B.T.,

1992b. Tectonism and magmatism during the NE Atlantic continental break-up: The Voring margin. In: B.C. Storey, T. Alabaster and R.J. Plankhurst (Editors), Magmatism and the Causes of Continental Break-up. Geol. Sot. Lon- don Spec. Publ., 68: 305-320.

Stockmal, G.S., Beaumont, C. and Boutelier, R., 1986. Geody- namic models of convergent margin tectoncs: the transi- tion from rifted margin to overthrust belt and the conse- quences for foreland basin development. Am. Assoc. Pet. Geol. Bull., 70: 181-390.

Sundvoll, B., Neumann, E.R., Larsen, B.T. and Tuen, E., 1990. Age relations among Oslo Rift magmatic rocks: implications for tectonic and magmatic modelling. Tecton- ophysics, 178: 67-87.

Van Wees, J.D., De Jong, K. and Cloetingh, S., 1992. Two-di- mensional P-T-r modelling and the dynamics of exten- sion and inversion in the Betic zone (SE Spain). Tectono- physics, 203: 305-324.

Velde, B., Dubois, J., Touchard, G. and Badri, A., 1990. Fractal analysis of fractures in rocks: the Cantor’s Dust method. Tectonophysics, 179: 345-352.

Villemin, T. and Sunwoo, C., 1987. Distribution logarithmique

self-similaire des rejets et longueurs de failles: exemple du Bassin Houiller Lorrain. CR. Acad. Sci. Paris, 305: 1309-

1312. Vilotte, J.P., Melosh, J., Sassi, W. and Ranalli. G., 1993.

Rheology and sedimentary basins. In: S. Cloetingh, W. Sassi and F. Horvath (Editors), The Origin of Sedimentary Basins: Inferences from Quantitative Modelling and Basin

Analysis. Tectonophysics, 226: 89-96. Walsh, J.J. and Watterson, J., 1991. Geometric and kinematic

coherence and scale effects in normal fault systems. In:

A.M. Roberts et al. (Editors), The Geometry of Normal Faults. Geol. Sot. London, Spec. Publ., 56: 193-203.

Walsh, J.J., Watterson, J. and Yielding, G., 1991. The impor- tance of small scale faulting in regional extension. Nature,

351: 391-393. Warner M. and McGeary S., 1987. Seismic reflection coeffi-

cients from mantle fault zones. Geophys. J.R. Astron. Sot., 89: 223-230.

Watts, A.B., 1989. Lithospheric flexure due to prograding sediment loads: implications for the origin of offlap/onlap

patterns in sedimentary basins. Basin Res., 2: 133-144. Wernicke, B., 1985. Uniform sense normal simple shear of the

continental lithosphere. Can. J. Earth Sci., 22: 108-125. White, R.S. and McKenzie, D., 1989. Magmatism at rift

zones: The generation of volcanic continental margins and

flood basalts. J. Geophys. Res., 94: 7685-7729. Williams, G. and Vann, I., 1987. The geometry of listric

normal faults and deformation in their hanging walls. J. Struct. Geol., 9: 789-795.

Ziegler, P.A., 1990. Geological Atlas of Western and Central Europe. Shell Int. Pet. Mij./Geol. Sot. London, 2nd ed., 239 pp.