earth and planetary science lettersdjs/donna_ldeo/publications_files/crosby_e… · the lithosphere...

13
Evolution of the NewfoundlandIberia conjugate rifted margins Alistair Crosby a, , Nicky White a , Glyn Edwards b , Donna J. Shillington c a Bullard Laboratories, Department of Earth Sciences, University of Cambridge, CB3 0EZ, UK b BP Exploration Operating Co. Ltd., Sunbury-on-Thames, TW18 7LN, UK c Lamont-Doherty Earth Observatory, Columbia University, Palisades, NY 10964, USA ABSTRACT ARTICLE INFO Article history: Received 7 November 2007 Received in revised form 19 June 2008 Accepted 20 June 2008 Available online 4 July 2008 Editor: C.P. Jaupart Keywords: continental margins stretching strain rate inversion paleobathymetry rheology It is accepted that mildly extended sedimentary basins form by largely uniform thinning of continental lithosphere. No such consensus exists for the formation of highly extended conjugate rifted continental margins. Instead, a wide range of models which invoke differing degrees of depth-dependent thinning have been proposed. Much of this debate has focussed on the well-studied NewfoundlandIberia conjugate margins. We have tackled the problem of depth dependency at this pair of margins in three steps. First, we have reconstructed water-loaded subsidence histories by making simple assumptions about changes in water depth through time. Secondly, we have used these reconstructed subsidence histories to determine the spatial and temporal variation of lithospheric strain rate. An inversion algorithm minimizes the mist between observed and predicted subsidence histories and crustal thicknesses by varying strain rate as a smooth function of distance across the margin, depth through the lithosphere, and geologic time. Depth- dependent thinning is permitted but, crucially, our algorithm does not prescribe its existence or form. Given the absence of signicant volumes of syn-rift magmatism, we have also applied a minimal melting constraint. Inverse modeling has yielded excellent ts to both reconstructed subsidence and crustal observations, which suggest that rifting occurred from 150135 Ma and at rates of up to 0.3 Ma 1 . Strain rate distributions are depth-dependent, suggesting that lithospheric mantle thins over a wider region than the crust. Beneath highly extended parts of the margin, crustal strain rates greatly exceed lithospheric mantle strain rates. Thirdly, we have tested our strain rate histories by comparing the total horizontal extension with the amount of extension inferred from normal faulting patterns. Both values agree within error. We freely acknowledge that there are important uncertainties in reconstructing the subsidence histories of deep-water margins. Nevertheless, stratigraphic records remain the only, albeit imperfect means of determining how crust and lithospheric mantle thin through time and space. © 2008 Elsevier B.V. All rights reserved. 1. Introduction Over the past 40 years, there has been much interest in the development of passive continental margins, which are a primary manifestation of plate tectonics. It is generally accepted that passive margins form by extension and cooling of continental lithosphere. Simple kinematic models have been developed which describe the growth of margins in terms of either uniform or depth-dependent stretching (e.g. McKenzie, 1978; Le Pichon and Sibuet, 1981; Hellinger and Sclater, 1983; Keen and Dehler, 1993). The temporal and spatial variation of crustal and lithospheric thinning across a margin control variations in temperature, heatow, subsidence, extensional faulting and decompression melting (White et al., 2003). Despite acceptance of the basic model, there is much disagreement about the spatial distribution of lithospheric extension. This controversy has particularly focussed on the NewfoundlandIberia conjugate margins which have been studied in detail using deep-seismic reection and wide-angle surveys, calibrated by deep-water boreholes. There are two substantive and unresolved issues. First, a signicant discrepancy between the amount of extension measured from upper crustal faulting and the amount of extension measured from crustal thickness is often reported (e.g. Davis and Kusznir, 2004). Secondly, broad tracts of exhumed and mostly unmelted lithospheric mantle occur at deep-water margins (e.g. Whitmarsh et al., 2001; Minshull et al., 2001). These two observations are mutually incompatible but if either or both are correct, they imply that various forms of depth dependency occur, perhaps at different stages of margin evolution. A closely related puzzle concerns the existence and importance of lithospheric detachment faults (e.g. Wernicke, 1985). These gently dipping detachment faults may offset stretching of the crust and lithospheric mantle, generating asymmetric patterns of syn- and post- rift subsidence. It is generally accepted that detachment faults do not control lithospheric extension in mildly extended basins (White, 1989). However, spectacular detachment faults are sometimes imaged on rifted margins and mapped in collisional mountain belts where the original rift architecture can sometimes be inferred (e.g. Hoffmann and Reston, 1992; Pérez-Gussinyé and Reston, 2001; Hölker et al., 2003). Symmetric Earth and Planetary Science Letters 273 (2008) 214226 Corresponding author. Tel.: +44 1223 337102. E-mail address: [email protected] (A. Crosby). 0012-821X/$ see front matter © 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.epsl.2008.06.039 Contents lists available at ScienceDirect Earth and Planetary Science Letters journal homepage: www.elsevier.com/locate/epsl

Upload: others

Post on 23-Oct-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

  • Earth and Planetary Science Letters 273 (2008) 214–226

    Contents lists available at ScienceDirect

    Earth and Planetary Science Letters

    j ourna l homepage: www.e lsev ie r.com/ locate /eps l

    Evolution of the Newfoundland–Iberia conjugate rifted margins

    Alistair Crosby a,⁎, Nicky White a, Glyn Edwards b, Donna J. Shillington c

    a Bullard Laboratories, Department of Earth Sciences, University of Cambridge, CB3 0EZ, UKb BP Exploration Operating Co. Ltd., Sunbury-on-Thames, TW18 7LN, UKc Lamont-Doherty Earth Observatory, Columbia University, Palisades, NY 10964, USA

    ⁎ Corresponding author. Tel.: +44 1223 337102.E-mail address: [email protected] (A. Crosby).

    0012-821X/$ – see front matter © 2008 Elsevier B.V. Adoi:10.1016/j.epsl.2008.06.039

    A B S T R A C T

    A R T I C L E I N F O

    Article history:

    It is accepted that mildly e

    Received 7 November 2007Received in revised form 19 June 2008Accepted 20 June 2008Available online 4 July 2008

    Editor: C.P. Jaupart

    Keywords:continental marginsstretchingstrain rateinversionpaleobathymetryrheology

    xtended sedimentary basins form by largely uniform thinning of continentallithosphere. No such consensus exists for the formation of highly extended conjugate rifted continentalmargins. Instead, a wide range of models which invoke differing degrees of depth-dependent thinning havebeen proposed. Much of this debate has focussed on the well-studied Newfoundland–Iberia conjugatemargins. We have tackled the problem of depth dependency at this pair of margins in three steps. First, wehave reconstructed water-loaded subsidence histories by making simple assumptions about changes in waterdepth through time. Secondly, we have used these reconstructed subsidence histories to determine thespatial and temporal variation of lithospheric strain rate. An inversion algorithm minimizes the misfitbetween observed and predicted subsidence histories and crustal thicknesses by varying strain rate as asmooth function of distance across the margin, depth through the lithosphere, and geologic time. Depth-dependent thinning is permitted but, crucially, our algorithm does not prescribe its existence or form. Giventhe absence of significant volumes of syn-rift magmatism, we have also applied a minimal melting constraint.Inverse modeling has yielded excellent fits to both reconstructed subsidence and crustal observations, whichsuggest that rifting occurred from ∼150–135 Ma and at rates of up to 0.3 Ma−1. Strain rate distributions aredepth-dependent, suggesting that lithospheric mantle thins over a wider region than the crust. Beneathhighly extended parts of the margin, crustal strain rates greatly exceed lithospheric mantle strain rates.Thirdly, we have tested our strain rate histories by comparing the total horizontal extension with the amountof extension inferred from normal faulting patterns. Both values agree within error. We freely acknowledgethat there are important uncertainties in reconstructing the subsidence histories of deep-water margins.Nevertheless, stratigraphic records remain the only, albeit imperfect means of determining how crust andlithospheric mantle thin through time and space.

    © 2008 Elsevier B.V. All rights reserved.

    1. Introduction

    Over the past 40 years, there has been much interest in thedevelopment of passive continental margins, which are a primarymanifestation of plate tectonics. It is generally accepted that passivemargins form by extension and cooling of continental lithosphere.Simple kinematic models have been developed which describe thegrowth of margins in terms of either uniform or depth-dependentstretching (e.g. McKenzie, 1978; Le Pichon and Sibuet, 1981; Hellingerand Sclater, 1983; Keen and Dehler, 1993). The temporal and spatialvariation of crustal and lithospheric thinning across a margin controlvariations in temperature, heatflow, subsidence, extensional faultingand decompression melting (White et al., 2003).

    Despite acceptance of the basic model, there is much disagreementabout the spatial distribution of lithospheric extension. This controversyhas particularly focussed on the Newfoundland–Iberia conjugatemarginswhichhave been studied in detail using deep-seismic reflection

    ll rights reserved.

    and wide-angle surveys, calibrated by deep-water boreholes. There aretwo substantive and unresolved issues. First, a significant discrepancybetween the amount of extensionmeasured from upper crustal faultingand the amount of extension measured from crustal thickness is oftenreported (e.g. Davis and Kusznir, 2004). Secondly, broad tracts ofexhumed andmostly unmelted lithosphericmantle occur at deep-watermargins (e.g. Whitmarsh et al., 2001; Minshull et al., 2001). These twoobservations are mutually incompatible but if either or both are correct,they imply that various forms of depth dependency occur, perhaps atdifferent stages of margin evolution.

    A closely related puzzle concerns the existence and importance oflithospheric detachment faults (e.g. Wernicke, 1985). These gentlydipping detachment faults may offset stretching of the crust andlithospheric mantle, generating asymmetric patterns of syn- and post-rift subsidence. It is generally accepted that detachment faults do notcontrol lithospheric extension in mildly extended basins (White, 1989).However, spectaculardetachment faults are sometimes imagedon riftedmargins andmapped in collisionalmountainbeltswhere theoriginal riftarchitecture can sometimes be inferred (e.g. Hoffmann and Reston,1992; Pérez-Gussinyé and Reston, 2001; Hölker et al., 2003). Symmetric

    mailto:[email protected]://dx.doi.org/10.1016/j.epsl.2008.06.039http://www.sciencedirect.com/science/journal/0012821X

  • Fig. 1. Sketches of two dynamical models redrawn from numerical experiments 8 and 9of Huismans and Beaumont (2003). In both cases, a complicated rheologicalstratification of the lithosphere was used which includes frictional-plastic strainsoftening. Model (a) was generated using a very slow strain rate and closely resembleshighly depth-dependent stretching of the lithosphere (Wernicke, 1985). Note that theform of depth dependency at each margin must be of opposite sense in order toconserve mass. Model (b) was generated using a moderate strain rate and closelyresembles uniform stretching of the lithosphere (McKenzie, 1978). Stippled shadingindicates crust, and grey shading lithospheric mantle.

    215A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    shear zones have also been invoked as a mechanism for extending,but not significantly melting, lithospheric mantle (Froitzheim et al.,2006).

    The existence and form of depth dependency have been exploredin two different ways. The first approach is observation-based andrelies uponmaking detailed measurements of upper crustal faults andcomparing themwith crustal thickness measurements. The extensiondiscrepancy is reported for highly extended margins and there ismuch debate about its cause. One school of thought argues that it isreal, which implies that the brittle upper crust barely extends at deep-water margins compared with the rest of the crust and lithosphericmantle (Davis and Kusznir, 2004). The unlikely inference is that thereis a major change in the depth distribution of extension at the brittle–plastic transition. A second school of thought argues that theextension discrepancy is not real but arises from the limitations ofseismic imaging of highly deformed rocks. Once stretching factorsexceed about 1.7, a second generation of faulting forms which dissectsthe older generation. Reston (2007) has shown convincing examplessimilar to those observed in the field but, in general, multiplegenerations of faulting are exceedingly difficult to image. Moreover,40–70% of brittle deformation occurs on scales which cannot beresolved by seismic imaging (Walsh et al., 1991; Marrett andAllmendinger, 1992).

    A second approach uses numerical experiments to investigate thedynamic evolution of conjugate margins. The starting point of thesethermo-mechanical models is an assumed rheological structure of thecontinental lithosphere. Typically, the upper crust is assumed todeform by frictional failure according to Byerlee's Lawwhile the lowercrust and lithospheric mantle deform by power-law creep (e.g. Braunand Beaumont, 1987; Chéry et al., 1992; Bassi et al., 1993; Tett andSawyer, 1996; Huismans et al., 2001; Huismans and Beaumont, 2003;Rosenbaum et al., 2005; Lavier and Manatschal 2006; Burov, 2007;Harry and Grandell, 2007). These models are intricate and it isimportant to bear in mind that rheological descriptions of thelithosphere are at best educated guesses. The inevitable uncertaintieshave led to a major debate about lithospheric strength (e.g. Jackson,2002).

    Fig.1 summarizes two endmembers from a recent set of numericalsimulations which were designed to investigate the effects offrictional-plastic and viscous strain softening (Huismans and Beau-mont, 2003). Strain softening promotes asymmetric extension of thatportion of the lithosphere controlled by the dominant rheology. Byincreasing or decreasing strain rate, Huismans and Beaumont (2003)showed that the locus of the dominant rheology moves, promotingvarying degrees of asymmetry. The first simulation was generatedusing an unrealistically slow extensional velocity of 0.06 cm/yr over400 km (i.e. ε̇xx∼0.002 Ma−1). Extension is highly asymmetric andnon-uniform with depth, closely resembling the lithospheric simpleshear model of Wernicke (1985). In contrast, the second simulationwas generated using a much faster velocity of 10 cm/yr over 400 km(i.e. ε̇xx∼0.3 Ma−1) and deformation is dominated by pure shear(McKenzie, 1978). These simulations may, or may not, be directlyapplicable to the Newfoundland–Iberia rift but it is clear that radicallydifferent lithospheric geometries can be obtained simply by varyingthe rate of deformation.

    Asymmetric extension is evidently an attractive way of producingtracts of exposed and largely unextended lithospheric mantle. Fig. 1aalsomakes a crucial point: if lithospheric mantle is largely unextendedon one margin, then the opposite must be true on the conjugatemargin. Thus, if the lithospheric mantle under the Iberian margin isrelatively unextended, the Newfoundland margin should showevidence consistent with a highly extended lithospheric mantle (e.g.a large ratio of post-rift to syn-rift subsidence and evidence forsignificant volumes of decompressionmelting). We all seek a dynamicunderstanding of conjugate margin formation but existing modelsshare three important drawbacks. First, the rheological descriptions

    upon which they are based rely upon extrapolation of laboratoryexperiments by up to ten orders of magnitude. Secondly, they are veryslow to run and so the more interesting inverse problem is not yettractable. Thirdly, dynamic models are seldom concerned with fittinggeologic observations in a formal sense. Instead, they are essentially‘thought experiments’ which permit the consequences of particularrheological frameworks to be explored.

    In order to determine the spatial and temporal evolution of deep-water margins, we argue that a kinematic approach is more fruitful.Our strategy is threefold. An obvious starting point is that conjugatemargins must be modeled simultaneously since their structuralevolution should be compatible. We then suggest that the best wayto identify the existence and form of depth dependency is to modelsubsidence histories. Although present-day crustal structure is animportant constraint, it cannot reveal how strain has been partitionedbetween the crust and lithospheric mantle. Subsidence histories, onthe other hand, contain valuable, albeit indirect, information about thedistribution of thermal anomalies and thus deformation throughoutthe lithosphere in both space and time. Finally, we seek smoothmodels which minimize the misfit between observed and predictedsubsidence and crustal histories. An important corollary is that noprior assumptions should be made about either the existence or theform of depth dependency. These aims are best achieved using aninverse model. We focus on the well-studied Newfoundland–Iberianconjugate margin pair, but our inverse strategy is generally applicable.

    2. A general strategy

    Heterogeneous sedimentary records are used to calculate thewater-loaded (i.e. tectonic) subsidence histories of basins andmargins. For many years, these tectonic subsidence histories havebeen investigated using forward models which assume eitherinstantaneous or finite-duration stretching.

  • Table 1Parameters used for inverse modeling

    Quantity Meaning Value

    T Temperaturet Timev VelocityG Strain ratex Offsetz Depthw Box widthtc Initial crustal thickness See Fig. 4a Initial lithospheric thickness See Fig. 7κ Thermal diffusivity 8×10−7 m2 s−1

    k Thermal conductivity 3.1 W m−1 K−1

    ρm Mantle density 3350 kgm−3 at 0 °Cρc Crustal density 2850 kg m−3

    ρw Water density 1030 kg m−3

    Model surface temperature 0 °CModel base temperature 1333 °CThermal expansivity 3.28×10−5 K−1

    Number of Fourier terms 20Number of spline heights 5Number of x-nodes where G specified 40Number of z-nodes where G specified 10Number of timesteps where G specified 16Misfit penalty if average stratigraphic error is 1 km 20Misfit penalty if average crustal thickness error is 1 km 1Misfit penalty if maximum dry solidus overstep is 1 °C 0.05Stratigraphic smoothing filter width 20 km

    Lithospheric thickness is chosen so that the reference column is in isostatic balancewitha mid-ocean ridge. Penalty components in the misfit function are cumulative. Values ofthe other components (which relate to strain rate smoothness and positivity) are givenin Edwards (2006).

    216 A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    White (1993) showed that one-dimensional subsidence data canbe inverted by varying lithospheric strain rate as a function of time.Strain rate inversion is easily applied to well-log information andmapped stratigraphic sections but inevitably many simplifyingassumptions must be made (e.g. Airy isostasy, one-dimensionalheatflow) and it provides only a partial view of lithosphericdeformation. This approach was extended into two dimensions byWhite and Bellingham (2002) and Bellingham and White (2002) whoshowed that the spatial and temporal variation of lithospheric strainrate can be recovered by inverting water-loaded stratigraphic profiles.First, a starting distribution of strain rate is calculated by assuming athermal expansion coefficient, α, of zero and by stacking a set of one-dimensional solutions. The misfit between observed and calculatedwater-loaded subsidence is then minimized by varying the strain ratein the forward model as a function of distance along the profile andgeologic time. The smoothest solution which yields the minimummisfit is found using a conjugate gradient search algorithm (Presset al., 1986). The resultant variation of strain rate through time andspace shows how sedimentary basins grow and provide importantdynamic insights (e.g. Jones et al., 2004).

    The two-dimensional algorithm makes several important assump-tions. First, inherently smooth solutions, which do not model thesmall-scale details of the faulting process, are sought. Secondly,Bellingham and White (2002) assume that lithospheric stretching isuniform with depth (i.e. a thin-sheet approximation with ∂u/∂z=0).This approximation is probably reasonable for mildly extendedsedimentary basins where β≤2. However, at highly extended deep-water margins, we cannot assume that uniform stretching applies. It isequally crucial that no particular form of depth dependency isimposed. Instead, we seek the smoothest distribution of lithosphericdeformation which optimizes the fit between observed and predictedwater-loaded subsidence. In this respect, our work appears to buildupon the one-dimensional, depth-dependent analysis of Keen andDehler (1993) and Dehler and Keen (1993). However, they state thattheir analysis does not conserve volume, which is a significant flaw.Instead we use algorithms developed and applied by Edwards (2006),Edwards et al. (in preparation) and Shillington et al. (2008) who use aspectral approach to determine the variation of lithospheric strain ratealong a profile as a function of distance, depth and geologic time suchthat volume is automatically conserved. The cornerstone of theiralgorithm is a forward model which calculates subsidence and crustalthinning from the variation of strain rate. This calculation is brokendown into three steps. First, the temperature history of the deforminglithosphere is calculated for a given distribution of strain rate. Usingthe following equation

    @T@t

    ¼ κ∇2T−v �∇T ð1Þ

    which has advective and diffusive terms in x and z. Relevantparameters are defined in Table 1. Strain rate G is related to velocity by

    G ¼ @υx@x

    ¼ − @υz@z

    ð2Þ

    Eq. (1) is solved using a combined Eulerian and Lagrangian tech-nique where the solution is decoupled from the grid movement. Theappropriate boundary conditions are

    @v@z

    ¼ 0 at x ¼ 0;w tð Þv ¼ 0 at x ¼ 0υz ¼ 0 at z ¼ a

    Secondly, the temperature history is used to calculate loads as afunction of time and space. Tracking the movement of major densityboundaries (e.g. the base of the crust) is not trivial when ∂u/∂z≠0.Edwards (2006) and Edwards et al. (in preparation) have used an

    ‘immersed boundary’methodwhich enables boundaries to be tracked ata higher resolution than is required for the thermal calculation itself.Thirdly, the load history is used to calculate the subsidence history givenan initially flat Moho by assuming either Airy isostasy or by assumingthat the lithosphere has a constant flexural rigidity. Since we areprimarily interested in thesmoothest (i.e. long-wavelength)deformationhistory, an Airy approximation suffices. Subsidence is calculated withrespect to a reference lithospheric columnwhich is prescribed using theparameters of Table 1. For simplicity, the crust is assumed to have auniform composition with no internal heat generation.

    The kinematic forwardmodel is simple and fast. It was designed sothat subsidence profiles can be calculated for all possible forms ofdepth dependency. It has also been carefully bench-marked againstinstantaneous and finite-duration analytical solutions. Edwards(2006) has used this forward model as the basis for solving the farmore challenging inverse problem. Strain rate inversion is relativelystraightforward when uniform stretching of the lithosphere isassumed. In general, however, strain rate, G(x, z, t), varies as a functionof x, the distance along the profile, z, the depth through thelithosphere, and t, geologic time. The principal difficulty is thatgeneralized strain rate distributions produce velocity fields which donot necessarily satisfy the boundary conditions. Thus G(x, z, t) must beparameterized in such a way which ensures that mass conservation isobeyed and that boundary conditions are satisfied.

    Edwards (2006) has shown that these conditions are automaticallyfulfilled provided G(x, z, t) is recast in spectral form. Thus the startingdistribution of G(x, z, t) can be rewritten as a set of Fourier series. Thecoefficients of this set of periodic functions are then systematicallyvaried until the misfit between observed and calculated subsidence isminimized. Linear splines are used for interpolation and inversemodeling is subject to the usual smoothness and positivity constraints.

    We are conscious that more sophisticated kinematic and dynamicdescriptions of lithospheric stretching exist. However, such descriptionsare slow to run and cannot yet be used as the basis of inversion. We areconcernedwith solving the inverse problemwhich requires the forwardmodel to be run many thousands of times. Our forward model contains

  • Fig. 2. Shaded relief bathymetric map of the North Atlantic Ocean between Newfoundland and Iberia overlain by free-air gravity anomalies which were low-pass filtered to excludewavelengths shorter than ∼700 km (GRACE satellite-only gravity model GGM02). These anomalies show that there is a difference of ∼40 mGal between each side. If the admittance,Z, is ∼30 mGal/km, then the difference in dynamic support is ∼750 m which reflects the spatial pattern of convective circulation beneath the lithospheric plate (e.g. (Crosby et al.,2006) and references therein). Thick annotated black lines show the wide-angle seismic lines discussed in text; medium black lines indicate plate boundaries; thin annotated blacklines delineate traces of the M0 (i.e. 121 Ma) magnetic anomaly. The inset shows line locations on a reconstruction of the Iberia–Newfoundland conjugate margins at M0 time (afterSrivastava et al. (2000)).

    217A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    the essential features needed to describe general depth-dependentdeformation along a two-dimensional profile. It was carefully designedso that it runs in less than about a second. Optimization is carried outusing Powell's algorithm (Press et al., 1986); typically, 2–4×105

    iterations of the forward model are required for convergence, whichtakes approximately 24–48hof CPU timeona 2.8GHz Intel P4Dual CorePC running a Linux operating systemwith an open source compiler. Asbefore, the resulting strain rate distributions are smoothon timescales of1–5 Ma and on length scales of 10–20 km. Stratigraphic profiles areusually smoothed beforehand.

    Edwards (2006) applied the depth-dependent strain rate algorithmto a number of sedimentary basins (including the North Sea andGulf ofSuez) and well-sedimented passive margins (the northern margin ofthe South China Sea and the Black Sea.). His results show that thesebasins and margins exhibit modest depth dependency. However,before applying this algorithm to the Newfound-Iberian conjugatemargins, wemust first reconstruct water-loaded stratigraphic profiles.

    2.1. Paleobathymetry at deep-water margins

    Sedimentary basins and passive margins are loaded by a combina-tion of water and sedimentary rocks of variable density. In one-dimensional analyses, this heterogeneous load is converted intowater-loaded subsidence of underlying basement rocks, Sw. At a given time,

    Sw ¼ Ss ρm−ρsρm−ρw

    � �þ dw ð3Þ

    where Ss is the thickness of a heterogeneous sedimentary column afterdecompaction, ρPs is themean density of this columnwhich is calculatedusing the method of Sclater and Christie (1980), and dw is the waterdepth at the time of deposition (i.e. paleobathymetry). In twodimensions, it is more convenient to model water-loaded stratigraphyrather than basement subsidence. Otherwise, two-dimensional sub-sidence analysis is similar (i.e. stratigraphic horizons are decompacted

    and water-loaded by the standard methods). Bellingham and White(2002) showed that uncertainties in compaction parameters have littleeffect on the water-loaded stratigraphy unless clastic and carbonaterocks are interchanged. The greatest source of uncertainty stems fromour sketchy understanding of paleobathymetry.

    In mildly extended sedimentary basins, there is considerable debateabout paleobathymetric constraints. In the North Sea basin, the deposi-tional water depth of chalk remains uncertain and estimates range from50–800m. In contrast, platformcarbonate deposits form in shallowwater,although a conservative range of 0–50 m is often used. Similar problemsafflict the paleobathymetry of clastic rocks. In coastal shelf deposits, thereare excellent paleontological and sedimentological indicators which canbe used to demonstrate gradual increases inwater depth from nearshoreto the shelf edge. However, once water depths exceed several hundredmeters, fine-grained sedimentary rocks generally do not contain sensitivepaleobathymetric indicators and it is difficult to determinewater depth towithin one order of magnitude. An improved understanding of the depthranges of benthic foraminifera would be of considerable help.

    These problems are exacerbated at continental margins where thehighly extended distal portions are often sediment-starved, poorlydrilled, and well below the range of depths where sedimentologicaland micropaleontological information are diagnostic of paleobathy-metry. We suspect that this apparently insurmountable problem hasled many to neglect subsidence records of highly extended margins.We acknowledge these concerns, but given that the only temporalinformation about margin evolution is contained in this sedimentaryrecord it is important to devise a strategy. We have reconstructed thewater-loaded subsidence history of margins by applying three simpleand transparent rules. The first rule is obvious: we know the present-day water depth across a margin. It is also reasonable to assume thatthewater depth was negligible at the start of continental rifting. In theabsence of any other information, we could assume that paleobathy-metry simply grades linearly through time. Fortunately, we know howwater-loaded oceanic lithosphere subsides from global depth-age studies(e.g. Parsons and Sclater, 1977; Crosby et al., 2006). This second rule gives

  • 218 A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    us a vital ‘pin’ at the distal end of a continental margin wherepaleobathymetric estimates are poorest. The third rule concerns thegeometry of sedimentary infill. If paleobathymetry varied smoothly withtime, sedimentary packages will consist of parallel-bedded strata as seenon theNewfoundland–Iberianmargins. In contrast, rapid changes inwaterdepth due to sediment influx are expected to bemanifest by developmentof easily identifiable clinoformal packages. These three rules provideimportant (and often the only) constraints onpaleobathymetric evolutionat margins, which subsequently evolve into oceanic basins.

    We have adapted this general strategy to reconstruct the pa-leobathymetry of the Newfoundland–Iberian conjugate margins(Figs. 2–4). First, an estimate of the present-day dynamic topographywas subtracted from the stratigraphic data. Paleobathymetry of theocean–continent boundary since initiation of seafloor spreading wascalculated using a combination of Eq. (3) and a thermal plate model.Finally, we assumed that the scaling function (i.e. water depth as afraction of present-day bathymetry) is constant across the margin.This assumption is reasonable since there is no evidence for periods ofrapid infill. In this study, mapped stratigraphic horizons above thebasement pick are younger than both the oldest ocean crust and themost significant period of stretching, so it is straightforward toconfigure the paleobathymetry of each horizon (Table 2). If olderstratigraphic horizons existed, it would be necessary to specifypaleobathymetry in some other way and to correct for layer strain.

    The paleobathymetric scheme shown in Fig. 5 is simple andtransparent. It is also corroborated by water depth estimates whichwere inferred from nearby borehole information (e.g. ShipboardScientific Party, 2004; Péron-Pinvidic et al., 2007). Since we model thewater-loaded stratigraphy using an inverse strategy, it is straightforwardto investigate how paleobathymetric uncertainties affect our conclu-sions. The crucial point is that paleobathymetry is estimated indepen-dently rather than determined as part of the modeling process.

    Fig. 3. Summary of crustal structure along conjugate SCREECH and Iberian lines, redrawn fro(2006); SCREECH 3: Lau et al. (2006a); IAM-9: Dean et al. (2000); and GP-101: Whitmarsh eindicates transitional crust; blue shading indicates bona fide oceanic crust; stippled shadinstratigraphic horizons could be digitized from published seismic reflection interpretations (thocean transition, see Table 2 and text for details); finally, thin dashed lines show the crusta

    3. Application

    The Edwards (2006) algorithm was initially applied to intraconti-nental sedimentary basins. In order to model highly extendedmargins, a number of modifications are needed. The forwardcalculation, which is the cornerstone of the inverse algorithm, issolved spectrally and to avoid high frequency ringing, it is importantto model conjugate profiles or profiles which have been reflectedabout the ocean–continent boundary. At the landward margin, wehave added gradual tapers to incomplete stratigraphic profiles.Normally, the spatial and temporal variation of strain rate is calculatedfrom water-loaded stratigraphy alone. We have formally included themisfit between observed and predicted crustal thickness as anadditional constraint since the variation of crustal thickness acrossNewfoundland–Iberian margins is accurately known.

    The algorithm does not include decompression melting and we donot model the growth of oceanic crust. Since we are modeling ‘cold’margins, which are characterized by the absence of significantvolumes of basaltic magmatism, it is important that we ensure thatthe temperature of the lithosphere–asthenosphere boundary does notsignificantly overstep the solidus for dry peridotite (McKenzie and.Bickle, 1988). If overstep occurs during inversion, an appropriatepenalty is added to the misfit function (see Table 1 and Edwards et al.(in preparation) for details). For comparative purposes, we have alsorun each inversion without a melting constraint.

    3.1. Crustal structure and rift duration

    The Newfoundland–Iberia conjugate margins have been carefullystudied by acquisition programs over the last 10–15 years, whichinclude deep-reflection profiles, wide-angle seismic surveys anddeep-water drilling (Péron-Pinvidic et al., 2007). Here we model five

    mwide-angle models (SCREECH 1: Funck et al. (2003); SCREECH 2: Van Avendonk et al.t al. (1996), Zelt et al. (2003)). Red shading indicates continental crust; yellow shadingg indicates syn-rift and post-rift sedimentation; vertical lines delineate regions wheree landward linemarks the limit of the data, the seaward line is the estimated continent–l thickness variation predicted by inverse modeling of water-loaded subsidence record.

  • Fig. 4. Total water-loaded subsidence plotted against stretching factor, β, for each of lines shown in Fig. 1. In each case, estimates were made fromwide- angle models and dynamictopography was subtracted (see text). Red points are thinned continental crust; yellow points are transitional crust; and blue points are bona fide oceanic crust, which was includedfor comparison purposes only. The solid black line is the Airy isostatic prediction of total water-loaded subsidence as a function of β using a reference crust with average density of2850 kg m−3 and thickness as indicated (see expressions of McKenzie (1978) and parameter values of Table 1). The dashed black line is the Airy isostatic prediction, including theeffect of a 5 km thick layer of serpentinization which reduces the average density of sub-Moho mantle by 150 kg m−3. In each case, the inclusion of serpentinization improves the fit.Small deviations remain, which are probably consequence of pre-existing Moho topography, small but finite flexural rigidities and uncertainties in wide-angle models, particularlywhere crustal thickness gradients are steep. These deviations can impede our ability to simultaneously fit stratigraphy and crustal thicknesses. Finally, the dotted line in (e) is the Airyisostatic predictionwhen a more realistic reference crustal thickness of 36 km is used. (For interpretation of the references to colour in this figure legend, the reader is referred to theweb version of this article.)

    Table 2Regional seismic stratigraphic horizons fitted by inverse modeling (after Péron-Pinvidicet al. (2007) and Shipboard Scientific Party (2004))

    Horizon Boundary Assigned age(Ma)

    Common name

    Basement Latest Jurassic 150Top-A Aptian–Albian 112 UTop-B Turonian–Coniacian 90Top-C Eocene–Oligocene 34 Au

    Seabed Present day 0

    219A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    transects whose evolution is much debated (Figs. 2 and 3). It isgenerally accepted that at least two phases of rifting occurred(Tucholke et al., 1989). The older phase of rifting occurred in Permo-Triassic times and is largely confined to onshore sedimentary basinswhich fringe both continental margins. It is commonly observed onmany passive margins that an early phase of moderate rifting ends,giving way to a later phase which steps seaward; the Newfoundland–Iberian conjugate margins are no exception. This later phase began inLate Jurassic to Early Cretaceous times (150–135 Ma) and it is of fargreater significance since it gave rise to break-up and initially veryslow seafloor spreading by ∼127 Ma (i.e. magnetic anomalies M0–M4(Russell andWhitmarsh, 2003; Lau et al., 2006a). It is this rifting phasewhich has caused so much controversy and which we model in thefollowing sections.

    The structureof thecrust has been constrainedbyat leastfivemodernwide-angle seismic experiments andwe summarize the results in Figs. 3and 4. On the Newfoundland margin, the SCREECH experiments carriedout in 2000 yielded three high-quality profiles (Funck et al., 2003; VanAvendonk et al., 2006; Lau et al., 2006a). On the Iberian margin, profileIAM-9, which is conjugate to SCREECH 2, was acquired in 1995 (Deanet al., 2000). A wide-angle seismic profile, coincident with deep-

    reflection profile GP101 (Mauffret and Montadert, 1988) and conjugateto SCREECH 1, was acquired in 1988 (Whitmarsh et al., 1996) although amore recent coincident wide-angle survey also exists Zelt et al. (2003).The maximum crustal thickness is about 31 km landward of SCREECH 1and GP101, 33 km landward of SCREECH 2 and IAM-9, and about 35 kmlandward of SCREECH 3. Unstretched crustal thicknesses beneathNewfoundland and Iberia are about 35–40 km (Marillier et al., 1994)and 28–33 km (Paulssen and Visser, 1993; Pedreira et al., 2003;Fernández-Viejo et al. 1998), respectively. Continental crust on bothmargins has been dramatically and asymmetrically thinned over

  • Fig. 5. The paleobathymetry scaling function for each horizon in Table 2, calibratedusing the known water-loaded subsidence of the oldest ocean floor. Grey lines areillustrative since water-loaded stratigraphy is only specified at the agesmarked by blackcircles. The palaeo-water depth almost certainly increased more rapidly given the lowsyn-rift sediment thickness and high initial strain rates.

    220 A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    distances of 50–100 km. There is excellent evidence for pervasive normalfaulting, especially on the Newfoundland side where the ocean–continent transition is consequently indistinct. Sometimes, notably atODP sites 897, 899 (20 km north of IAM-9) and 1277 (at the ocean–continent transition of SCREECH 2), crustal thinning is complete andserpentinizedperidotites are exposed at the seafloor. TheP-wave velocityof the top 5 km of lithospheric mantle directly beneath highly thinnedcrust is often 0.5–1 km s−1 slower than normal, which is consistent withhydrothermal alteration promoted by pervasive faulting. Minor amountsof alkaline magmatism have been drilled on both margins but, crucially,there is scant evidence for voluminous basaltic volcanism typical ofmargins in the North Atlantic Ocean (e.g. seaward-dipping reflections,magmatic underplating), even though stretching factors exceed 3–4 overwide tracts (Jagoutz et al., 2007). Possible exceptions include portions ofthe Iberian margin where Russell and Whitmarsh (2003) haveinterpreted isolated magnetic anomalies as syn-rift igneous intrusions.In contrast, Sibuet et al. (2007) argue that linearmagnetic anomaliesM3–M17 (i.e. as old as 140 Ma) were acquired during amagmatic seafloorspreading and serpentinization.

    3.2. Dynamic topography

    Several unexpected problems arose whenwe began to analyze theconjugate margins. The first of these problems concerns the differencein dynamic topographic support on either side of the Atlantic Ocean.Dynamic topography is the long-wavelength component of topogra-phy which cannot be accounted for by crustal and/or lithosphericthickness variations but is generated by convective circulation be-neath plates. It is difficult to measure accurately, although inter-mediate (i.e. 500–4000 km) wavelength free-air gravity anomalies area useful guide. In the North Atlantic Ocean, there is a significant long-wavelength positive gravity anomaly over the Newfoundland marginand small negative anomalies over the relevant part of the Iberianmargin (Fig. 2). In the oceans, the ratio of the free-air gravity anomalyto the dynamic seafloor topography is 25–40 mGal km–1 (Crosby et al.,2006). Thus we estimate 0.6–1.2 km of dynamic uplift on theNewfoundland side and 0.25–0.5 km of dynamic subsidence on theIberian side. This differential support is consistent with the shiftrequired to align the two pairs of conjugatewater-loaded stratigraphicprofiles at the ocean–continent boundary (−0.6 km for SCREECH 2,+0.55 km for IAM-9; −0.6 km for SCREECH 1, +0.35 km for GP101).These values are broadly consistent with the subsidence of oceaniclithosphere which abuts both margins. The mean water-loaded depthof unperturbed ∼125 Ma oceanic lithosphere is 5.6±0.2 km when the

    mean crustal thickness is 7.1 km (Crosby et al., 2006; White et al.,1992). After applying an isostatic correction for unusually thin oceaniccrust, we calculated thewater-loaded subsidence of the oldest oceaniclithosphere and found that Newfoundland oceanic lithosphere is 0.6–1.0 km shallower than expected, while Iberian oceanic lithosphere atthe end of Line IAM-9 is 0.5–1.0 km deeper.

    Oceanic lithosphere at the end of GP101 is 0.2–0.6 km deeper butcrustal thickness is poorly constrained. Our three different estimatesof dynamic support are similar but not identical. For simplicity, wehave chosen the correction required to match the two pairs ofconjugate water-loaded stratigraphic profiles. SCREECH 3 has noconjugate profile and was shifted by the same amount as SCREECH 2.We note that our estimates disagree in both sign and magnitude withthose of Conrad et al. (2004), who calculated 100–300m of subsidencefor the Newfoundland side and 200–300 m of uplift for the Iberianside. The calculation of dynamic topography is fraught withuncertainties and for the purposes of this paper we prefer to usedirect observations.

    4. Serpentinization

    Differential elevation across a margin can also be generated byintrinsic changes within the mantle lithospheric column. The mostobvious source of elevation change is serpentinization of lithosphericmantle rocks which has occurred on both margins. Reduction in theaverage density of a thickness of mantle, ts, by an amountΔρ causes anisostatic reduction in water-loaded subsidence, Δsw,where

    Δsw ¼ tsΔρρm−ρwð4Þ

    The age of serpentinization is not known although it is likely to haveoccurred within several million years of rifting (e.g. Sibuet et al., 2007;Skelton and Valley, 2000). It is easiest to correct for serpentinizationbefore inverse modeling. A practical method for estimating thiscorrection is to plot the total water-loaded subsidence as a function ofcrustal stretching (Fig. 4). On thermally and isostatically equilibratedmargins, we expect a simple relationship between total water-loadedsubsidence and crustal thickness. Visual inspection demonstrates thathighly extended continental and transitional crust have generallysubsided less than expected (Fig. 4). This difference is consistently0.3 km of water-loaded subsidence and cannot be caused by differentialstretching of the lithosphere which changes the ratio of syn-rift to post-rift subsidence but not the total subsidence. It is easily explained bydecreasing the density of the lithospheric mantle. If Δρ is −150 kg m−3

    (Escartín et al., 2001) and if the average P-wave velocity of theanomalous layer is 7.5 km s−1, we obtain ts∼5 km for crustal thinningfactors of greater than 4–6, which is consistentwithwide-anglemodels.

    Fig. 4 shows that serpentinization is not required when crustalthinning factors are smaller than 3 or 4. Deviations from an isostaticrelationship at lower stretching factors are harder to explain (e.g.SCREECH 3; Fig. 4e). We suspect that they partly arise from flexuralsupport of short-wavelength, fault-bounded blocks in the upper crustand pre-existing Moho topography, from short-wavelength variationsin crustal density, and from the inability of ray-based seismicalgorithms to model steep velocity gradients. Either way, thesedeviations present a problem when we simultaneously fit water-loaded stratigraphic profiles and crustal thicknesses. The biggestproblem occurred with SCREECH 3; if we use a realistic unextendedcrustal thickness (36 km), then the residual misfit for moderatelystretched crust is large. One explanation for this discrepancy is achange in lithosphere thickness across the line (Priestley andMcKenzie, 2006). As a result, we are forced to choose a referencecrustal thickness of 30 km, which is not strictly correct even though itreconciles observed crustal thinning and water-loaded stratigraphywhen βN3. On most profiles, water-loaded subsidence of oceanic

  • 221A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    lithosphere is consistent withmeasured crustal thickness. The obviousexception is SCREECH 1 where water-loaded subsidence is about0.6 km smaller than expected because serpentinization has pene-trated beneath oceanic crust where anomalously slow velocities areobserved (Funck et al., 2003).

    4.1. Stratigraphic records

    The temporal and spatial evolution of a stretched margin ispartially preserved by the stratigraphic record. The goal of our analysisis to use this incomplete stratigraphic record to reconstruct the water-loaded subsidence of a deep-water continental margin, which is then

    Fig. 6.Observed and predicted water-loaded subsidence for the five lines located on Fig. 2. (a)indicated by a vertical grey bar. Thin black lines show water-loaded subsidence calcuserpenitinization, and for dynamic topography (see text for details); open circles and dashedin Fig. 7a and b and obtained by inverse modeling. (b) Conjugate line pair SCREECH 2 and IAMthe continent–ocean boundary. Note the quality of fit between theory and reconstructed su

    used to recover the temporal and spatial history of deformation. Thedifficulties of reconstruction appear formidable but are tractable ifbroken down into a series of transparent steps.

    The regional seismic stratigraphy is summarized by Péron-Pinvidicet al. (2007). We have exploited their internally consistent stratigraphicrecords which are based upon a set of mappable seismic reflectionswhose ages are known from nearby boreholes (Table 2). The age of thebase of this stratigraphic record is uncertain and we have chosen theoldest likely age (150 Ma). On both margins, the stratigraphic recordforms sub-parallel packages and there is no evidence for the develop-ment of clinoformal geometries which suggests that paleobathymetryevolved smoothly with age. Digitized stratigraphic horizons were taken

    . Conjugate line pair SCREECH 1 and GP 101which join at the continent–ocean boundarylated from observed stratigraphy and corrected for changing paleobathymetry, forlines showwater-loaded subsidence calculated using the strain rate distribution shown-9. (c) SCREECH 3 does not have a conjugate and was modeled by reflecting data aboutbsidence in all three cases.

  • 222 A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    from the following publications: SCREECH 1: Hopper et al. (2004);SCREECH 2: Shillington et al. (2006); SCREECH 3: Lau et al. (2006b);IAM-9: Dean et al., 2000; GP-101:Whitmarsh et al. (1996) andMauffretand Montadert (1988). Where appropriate, we converted two-waytravel time into depth using interval velocities from the relevant wide-angle model. Internally consistent stratigraphic records were availablefor the central portions of the profiles 3.

    In each case, we calculated the water-loaded stratigraphy bydecompacting and backstripping the observed stratigraphic record inthe standard way (Eq. (2); Sclater and Christie (1980)). For simplicity,we assumed that sediment composition varied linearly from 100%sand at the proximal end of a profile to 100% shale at the distal end.Previous analysis has demonstrated that large changes in sedimentcomposition have minor effects. Our chosen paleobathymetry is

    Fig. 7. Strain rate functions which correspond to the best-fitting water-loaded subsidence splotted as a function of range, x, and age, t, for SCREECH 1 and GP-101 conjugate pair. Black crand their displacement is a measure of strain; white areas indicate masking of artificial taperocean boundary. (b) Contoured strain rate distribution plotted as a function of range, x, and deconjugate pair. (e) and (f) SCREECH 3. Note the form of depth dependency which concentrateof the corresponding regions of stretched mantle under less-stretched crust is a function of

    summarized in Fig. 5. The starting water-loaded depth of the oldestoceanic lithosphere is 3 km, which is deeper than average becauseoceanic crust here is unusually thin (Lecroart et al., 1997). The best-fitting thermal plate thickness is about 100 km and our resultantpaleobathymetry exceeds 80% of its present-day value by ∼110 Ma.Our paleobathymetric estimates are not easy to corroborate but wenote that deep-water benthic foraminifera recovered at ODP Site 1276imply that water depths at the SCREECH 2 ocean–continent transitionexceeded 2 km since Late Aptian–Early Albian times (115–110 Ma)(Shipboard Scientific Party, 2004). Where they have been drilled, post-rift carbonate rocks have evidently been deposited below the calcitecompensation depth (Péron-Pinvidic et al., 2007). Furthermore,sediments at Galicia Bank ODP sites 901, 1065 and 1069 of Tithonianage appear to have been deposited at moderate water depths, which

    hown in Fig. 6. (a) Contoured strain rate distribution at the surface of the lithosphereosses show positions in x–twhere the strain rate function is calculated during inversioning of subsidence profile at either end of conjugate pair; the solid line is the continent–pth, z, for first time stepwhen strain rates are greatest. (c) and (d) SCREECH 2 and IAM-9s highest strain rates within the crust toward the continent–ocean boundary. The widththe width of the box in the model.

  • Fig. 8. Total model extension factors (i.e. 1−1/β and 1−1/δ) for (a) SCREECH 1/GP-101,(b) SCREECH 2/IAM-9, and (c) SCREECH 3. Red lines refer to the crust and blue lines referto the original lithospheric mantle. The areas under the two curves are constrained to beequal. (For interpretation of the references to colour in this figure legend, the reader isreferred to the web version of this article.)

    223A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    implies rapid deepening immediately after the start of rifting. Thewater at these sites then continued to deepen in Berriasian andBarremian time (145–125 Ma) (Tucholke and Sibuet, 2007).

    5. Inverse modeling

    Fig. 6 shows the reconstructed water-loaded stratigraphies for bothpairs of conjugate margins and for the SCREECH 3 profile. In each case,total subsidence is consistentwith thewater-loaded estimates summar-ized in Fig. 4. Each stratigraphic profile was inverted using the Edwards(2006) algorithm. We sought the smoothest distribution of lithosphericstrain rate as a function of distance, depth and time, G(x, z, t), whichminimized the misfit between reconstructed and calculated stratigra-phies and between observed and calculated crustal thicknesses. Thecalculated stratigraphies and crustal thickness values are shown inFigs. 3 and 6, respectively. Fits are generally excellent, yieldingχ2 valuesof less than 2.

    The corresponding strain rate histories are shown in Fig. 7 wherewe show two slices through each G(x, z, t) cube. In each case, the x–tslices show that a period of rapid strain rate (∼0.3 Ma−1) occurredbetween 150 and 140 Ma. Strain rate values increase towards thedeeper parts of each margin. Since syn-rift sediments are thin, weexpect the water to have deepened rapidly after the start of rifting.After ∼140 Ma, strain rates are much diminished and consistent withthe growth of unfaulted thermal subsidence. The x–z slices are moreinteresting since they are directly pertinent to the debate about depth-dependent stretching.

    Depth dependency takes two forms. Along the mildly extendedportions of the margins, lithospheric mantle has been extended over awider region than the crust. We have obtained similar resultselsewhere. Higher strain rates within the lithospheric mantle are adirect consequence of a lower than expected ratio of syn-rift to post-rift subsidence. As the lithosphere becomes increasingly stretched, wepass into a narrow zone, usually about 25–30 km wide, where strainrate is approximately uniform with depth. At the distal end of eachmargin, there is a significant switch in the form of depth dependencyand crustal strain rates are much higher than those in the lithosphericmantle (contra Dehler and Keen (1993)). This distribution is supportedby three arguments. First, the reconstructed ratio of syn-rift to post-rift subsidence is now significantly greater than expected for uniformstretching. This change is obviously contingent upon the chosenpaleobathymetric configuration. Secondly, the combination of a highlyextended crust and a lack of syn-rift magmatism must force strain tobe concentrated within the crust. At depth, the lithospheric mantlestrains at less than 0.1 Ma−1, which suppresses adiabatic decompres-sion and melting of the asthenosphere. Note that small amounts ofenriched melting are easily generated by decompressing fusiblematerial trapped against the wet solidus in the lower half of thelithospheric mantle. Thirdly, isostasy dictates that the crust cannothave extended as slowly as lithospheric mantle since we know thatwater depths increased very rapidly during syn-rift times; in any casethere is a short window of time available for rifting prior to sea-floorspreading. The cumulative extension for each line is shown in Fig. 8. Ineach case, although extension in the lithospheric mantle is clearlydistributed over a wider region, the total horizontal extensionaccumulated within the crust and within the lithospheric mantle arein good agreement.

    A supplementary figure shows the resultant strain rate functionswhen the misfit function is not penalized for overstepping the solidusof dry peridotite. As before, subsidence and crustal thinning observa-tions are well matched but peak strain rates are higher and there issubstantially less depth-dependent stretching (i.e. crustal and litho-spheric mantle extension factors are more similar). Under highlystretched crust, however, the dry solidus has been over-stepped by 30–160 °C for 10–20 km thickness of lithosphere. This overstep willgenerate substantial volumes of basaltic magmatism which have not

    beenobserved. Thus the depth-dependent stretchingobserved in Fig. 7is a direct consequence of restricting decompression melting.

    Are strain rate patterns well resolved? In order to investigate theextent towhich strain rate variation is sensitive to small uncertainties inthewater-loaded stratigraphy,we have invertedfifty perturbed versionsof the SCREECH 2/IAM-9 conjugate lines and analyzed the variance ofthe resultant strain rate distributions. Reassuringly, Fig. 9 shows that thevariation of strain rate with depth and distance is insensitive to theseperturbations (the standard deviation of the strain rate population iseverywhere less that 0.05Ma−1). As expected, sensitivity increases withdepth. By inference, Fig. 9 also shows that the choice of initial smoothingis unimportant. We acknowledge that systematic errors and/orasymmetries in the reconstructed water-loaded subsidence, which arenot ruled out by the sparse geologic data, could affect our inferred strainrate histories in important ways. We simply point out that ourreconstructions have the virtue of simplicity.

    6. Discussion

    Do these strain rate distributions yield ‘balanced’ models? TheEdwards (2006) algorithm automatically conserves mass at all timesand so the total amount of strain taken up within the crust is identicalto that taken up by the lithospheric mantle (Fig. 8). Thus

    ∫þ∞−∞ 1− 1β� �

    dx ð5Þ

    is conserved and equals the total amount of extension. It has oftenbeen argued that the total extension across basins and margins

  • Fig. 9. Bootstrap (i.e. Monte Carlo) modeling of water-loaded subsidence for SCREECH 2 and IAM-9 conjugate pair, with pseudo-topographic perturbations generated using a fractalmethod similar to that described by Turcotte (1999). (a) 50 perturbed versions of water-loaded subsidence with maximum perturbation of ±0.5 km. Perturbations decrease smoothlyto zero at distal end of each margin where paleobathymetry can be accurately determined from global age-depth models (Crosby et al., 2006). The grey vertical bar is the continent–ocean boundary. The crustal thickness was also perturbed in a similar way. (b) Strain rate map in x–z space. Solid blue lines are the 0.15Ma−1 strain rate contours obtained from the 50inversions; red shading defines the region within which the mean strain rate exceeds 0.15 Ma−1. The gross form of strain rate function is robust and relatively insensitive tostratigraphic perturbations. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

    224 A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    determined from subsidence and crustal observations greatly exceedsthe horizontal extension determined from normal faulting (e.g. Davisand Kusznir (2004)). If correct, this argument has dramatic con-sequences for the form of depth dependency particularly at highlyextended margins where the extension discrepancy is thought to beparticularly large. Fig. 10 shows fault interpretations of the fiveprofiles. Simplified versions of published interpretations are shownapart from SCREECH 2 and 3 which we interpreted ourselves. Theminimum amount of horizontal extension along each profile can be

    Fig. 10. Sketched interpretations of the five profiles (see Fig. 2 for location). Red, yellow andrespectively; grey shading is syn-rift subsidence with demonstrable fault-bounded stratigraplines are on-line and projected Ocean Drilling Program (ODP) drill holes; thin sub-vertical linconfidently identified). The total amounts of horizontal extension across each conjugate paredrawn from (Hopper et al., 2004) and (Whitmarsh et al., 1996). (b) SCREECH 2 and IAM-9difficulty in interpreting faults along portions of SCREECH 2. (c) SCREECH 3 has clearly iden

    gauged by summing heaves (i.e. horizontal components of displace-ment) across all visible faults. This simple method ignores unresolvedinternal deformation and multiple generations of faulting and isprobably a gross underestimate. A better estimate can be made if weassume that 50% of the extensional deformation occurs on small-scalefaulting which cannot be resolved (40–70% is thought to be areasonable range (Walsh et al., 1991; Marrett and Allmendinger,1992)). Our results are summarized in Table 3. Apart from SCREECH 1,we find good agreement between the predicted and observed

    blue shading indicate stretched continental crust, transitional crust, and oceanic crusthic growth; stippled shading is unfaulted post-rift subsidence; solid and dashed verticales are mappable normal faults (dashed lines delineate regions where faults could not beir are given in Table 3. (a) SCREECH 1 and GP-101 conjugate pair. Fault interpretationsconjugate pair. Fault interpretation for IAM-9 redrawn from (Dean et al., 2000). Notetifiable faulted syn-rift packages and is straightforward to interpret.

  • Table 3Estimates of upper crustal extension determined from summation of fault heaves

    Line Length, km Fault extension, km Model extension, km

    SCREECH 1 50 13–26 45SCREECH 2 150 37–74 65SCREECH 3 289 126–251 160GP-101 105 56–112 75IAM-9 208 109–218 200

    Lower bounds are calculated by assuming that all faulting is visible; upper bounds arecalculated by assuming that only 50% is visible (e.g. Walsh et al. (1991) and Marrett andAllmendinger (1992)).

    225A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    horizontal extension. Although our line of argument is different, weshare Reston's (2007) view that extension discrepancies at highlyextended margins are a consequence of poor acoustic imaging. Somezones of dramatically extended upper crust where closely spacedlistric faults detach into serpentinized mantle probably owe more tolarge-scale landsliding than to crustal extension (e.g. Clark et al.(2007)). We also suggest that multiple generations of normal faultingare sufficient to locally exhume mantle in regions where the crust hasbeen most thinned. Within the lithospheric mantle, it is difficult todetermine whether stretching occurs by plastic creep or whether it isaccommodated on sets of discrete shear zones (Tucholke et al., 1989).

    Do our results shed any light on the debate about lithosphericstrength? Does a strong crust or a strong lithospheric mantle governmargin evolution? Temperature is widely regarded as the keyparameter. The rheologic behavior of a material is principallycontrolled by τ, the ratio of its temperature to its melting temperature,both measured in Kelvin. We suggest that the evolution of the highlyextended Newfoundland–Iberia margins must be primarily controlledby the strength of lithospheric mantle. A key observation is theexistence of wide tracts of extremely thin and pervasively faultedcrust which overly lithospheric mantle which stretched very slowlybecause it did not produce substantial volumes of decompressionmeltprior to oceanic basin formation. If rapid stretching was mostlyfocussed within the crust, the temperature of the sub-Moho mantlemust have decreased rapidly. Serpentinization of ∼5 km of litho-spheric mantle is unlikely to have played a significant limiting rolesince 25 km of lithospheric mantle cools by ∼250 °C in our best-fitmodels.

    Finally, the particular form of depth dependency recovered byinverse modeling does depend upon the geological interpretation oftransition zone crust. If the transition zone consists of highly extendedcontinental crust and if syn-rift magmatism is largely absent then therecovered form of depth dependency is a logical consequence, giventhe short interval of time between initiation of continental stretchingand seafloor spreading: the crust must have been rapidly extendedand the lithospheric mantle must have either not extended orextended isothermally (Minshull et al., 2001). This conclusion isinsensitive to the choice of starting lithospheric template. If thetransition zone consists of serpentinized continental mantle, a similarform of depth dependency is required since the crust must have beenexcised. Finally, if the transition zone consists of exhumed mantlewith variable amounts of intruded gabbro and basalt, which is formedby ultra-slow seafloor spreading, then the Newfoundland–Iberianmargins could have been formed by largely uniform stretching.

    7. Conclusions

    The temporal and spatial distribution of lithospheric extensionbeneath the Newfoundland–Iberian conjugate margins is muchdebated. We have tackled this problem in two stages. First, we haveestimated the water-loaded stratigraphy along each margin. Theproblem of paleobathymetric variation through time has been tackledbyadopting a simple and smoothmodelwhich is tied to the subsidencehistory of adjacent oceanic lithosphere. Secondly, we have fitted

    stratigraphic and crustal measurements by varying lithospheric strainrate as a function of distance, depth and time with the condition thatthe dry peridotite solidus is not substantially exceeded. Our inversealgorithm allows for depth-dependent stretching but does notprescribe its existence or form. Excellent fits have been obtained andthe resultant distributions of strain rate suggest that the style of depthdependency changes as lithospheric extension increases. At theproximal end of each margin, the highest strain rates occur in thelithospheric mantle. Toward the distal, more highly stretched, end ofeach margin, the style of deformation switches and the highest strainrates are found in the crust. Our results conserve mass and precludegross lithospheric asymmetry. The total horizontal extension acrosseach margin is in good agreement with estimates of upper crustalextension obtained from summation of fault heaves on seismicreflection profiles.

    Acknowledgements

    This research is funded by BP Exploration and forms part of the BP-Cambridge Margins Research Programme. We are grateful to P.Carragher, R. Corfield, R. Gibson, S. Jones, H. Lau, D. Lyness, L. Mackay,S. Matthews, M. Thompson and R. White for their help. C. Jaupart, J.Hopper and two anonymous reviewers provided thorough and helpfulreviews. Department of Earth Sciences Contribution Number 9278.

    Appendix A. Supplementary data

    Supplementary data associated with this article can be found, inthe online version, at doi:10.1016/j.epsl.2008.06.039.

    References

    Bassi, G., Keen, C.E., Potter, P., 1993. Contrasting styles of rifting: models and examplesfrom the eastern Canadian margin. Tectonics 12, 639–655.

    Bellingham, P., White, N., 2002. A two-dimensional inverse model for extensionalsedimentarybasins2.Application. J. Geophys.Res.107, 2260. doi:10.1029/2001JB000174.

    Braun, J., Beaumont, C., 1987. Styles of continental rifting: results from dynamic modelsof lithosphere extension. In: Beaumont, C., Tankard, A.J. (Eds.), Sedimentary Basinsand Basin-Forming Mechanisms. Mem. Can. Soc. Pet. Geol., vol. 12, pp. 241–258.

    Burov, E., 2007. The role of gravitational instabilities, density structure and extensionrate in the evolution of continental margins. In: Karner, G.D., Manatschal, G.,Pinheiro, L.M. (Eds.), Imaging, Mapping and Modelling Continental LithosphereExtension and Breakup. Geol. Soc. Lond. Spec. Pub., vol. 282, pp. 139–156.

    Chéry, J.F., Lucazeau, M., Daigniéres, M., Vilotte, J.P., 1992. Large uplift of rift flanks: agenetic link with lithospheric rigidity? Earth Planet. Sci. Lett. 112, 195–211.

    Clark, S.A., Sawyer, D.S., Austin Jr., J.A., Christeson, G.L., Nakamura, Y., 2007.Characterizing the Galicia Bank-Southern Iberia Abyssal Plain rifted marginsegment boundary using multichannel seismic and ocean bottom seismometerdata. J. Geophys. Res. 112, B03408. doi:10.1029/2006JB004581.

    Conrad, C.P., Lithgow-Bertelloni, C., Louden, K.E., 2004. Iceland, the Farallon Slab, anddynamic topography of the North Atlantic. Geology 32, 177–180.

    Crosby, A.G., McKenzie, D., Sclater, J.G., 2006. The relationship between depth, age andgravity in the oceans. Geophys. J. Int. 166, 553–573.

    Davis, M., Kusznir, N.J., 2004. Depth-dependent lithospheric stretching at riftedcontinental margins. In: Karner, G.D. (Ed.), Proceedings of NSF rifted marginstheoretical institute. Columbia University Press, pp. 92–136.

    Dean, S.M., Minshull, T.A., Whitmarsh, R.B., Louden, K.E., 2000. Deep structure of theocean–continent transition in the southern Iberia Abyssal Plain from seismicrefraction profiles: The IAM-9 transect at 40 20'N.

    Dehler, S.A., Keen, C.E., 1993. Effects of rifting and subsidence on thermal evolution ofsediments in Canada's east coast basins. Can. J. Earth Sci. 30, 1782–1798.

    Edwards, G.R.H., 2006. Inverse modelling of extentional sedimentary basins andmargins, Ph.D. thesis, University of Cambridge.

    Edwards, G., White, N., Haines, J., in preparation. Inverse modelling of extentionalsedimentary basins and margins, J. Geophys. Res.

    Escartín, J., Hirth, G., Evans, B., 2001. Strength of slightly serpentinized peridotites:implications for the tectonics of oceanic lithosphere. Geology 29, 1023–1026.

    Fernández-Viejo, G., Gallart, J., Pulgar, J.A., Gallastegui, J., Dañiobietia, J.J., Córdoba, D.,1998. Crustal transition between continental and oceanic domains along the NorthIberian margin from wide-angle and gravity data. Geophys. Res. Lett. 25,4249–4252.

    Froitzheim, N., Pleuger, J., Nagel, T.J., 2006. Extraction faults. J. Struct. Geol. 1388–1395.Funck, T., Hopper, J.R., Larsen, H.C., Louden, K.E., Tucholke, B.E., Holbrook, W.S., 2003.

    Crustal structure of the ocean–continent transition at Flemish Cap: seismicrefraction results. J. Geophys. Res. 108, 2531. doi:10.1029/2003JB002434.

    http://dx.doi.org/doi:10.1016/j.epsl.2008.06.039http://dx.doi.org/10.1029/2001JB000174http://dx.doi.org/10.1029/2006JB004581http://dx.doi.org/10.1029/2003JB002434

  • 226 A. Crosby et al. / Earth and Planetary Science Letters 273 (2008) 214–226

    GRACE satellite-only gravity model GGM02, see http://www.csr.utexas.edu/grace/gravity/ for further documentation.

    Harry, D.L., Grandell, S., 2007. Adynamicmodel of rifting betweenGalicia Bankand FlemishCap during the opening of the North Atlantic Ocean. In: Karner, G.D., Manatschal, G.,Pinheiro, L.M. (Eds.), Imaging, Mapping and Modelling Continental LithosphereExtension and Breakup. Geol. Soc. Lond. Spec. Pub., vol. 282, pp. 157–172.

    Hellinger, S.J., Sclater, J.G., 1983. Some comments on two-layer extensional models forthe evolution of sedimentary basins. J. Geophys. Res. 88, 8251–8269.

    Hoffmann, H.J., Reston, T.J., 1992. Nature of the S reflector beneath the Galicia Banksrifted margin: Preliminary results from prestack depth migration. Geology 20,1091–1094.

    Hölker, A.B., Manatschal, G., Holliger, K., Bernoulli, D., 2003. Tectonic nature and seismicresponse of top-basement detachment faults in magma-poor rifted margins.Tectonics 22. doi:10.1029/2001TC001347.

    Hopper, J.R., Funck, T., Tucholke, B.E., Larsen, H.C., Holbrook, W.S., Louden, K.E.,Shillington, D., Lau, H., 2004. Continental breakup and the onset of ultraslowseafloor spreading off Flemish Cap on the Newfoundland riftedmargin. Geology 32,93–96.

    Huismans, R., Beaumont, C., 2003. Symmetric and asymmetric lithospheric extension:relative effects of frictional-plastic and viscous strain softening. J. Geophys. Res. 108.doi:10.1029/2002JB002026.

    Huismans, R.S., Podladchikov, Y.Y., Cloetingh, S., 2001. Transition from passive to activerifting: relative importance of asthenospheric doming and passive extension of thelithosphere. J. Geophys. Res. 106, 11271–11291.

    Jackson, J., 2002. Strength of continental lithosphere: time to abandon the jellysandwich? GSA Today 4–9 (September).

    Jagoutz, O., Müntener, O., Manatschal, G., Rubatto, D., Péron-Pinvidic, G., Turrin, B.D.,Villa, I.M., 2007. The rift-to-drift transition in the North Atlantic: a stuttering start ofthe MORB machine? Geology 35, 1087–1090.

    Jones, S.M., White, N.J., Faulkner, P., Bellingham, P., 2004. Animated models of extensionalbasinsandpassivemargins.Geochem.,Geophys.,Geosyst. 5.doi:10.1029/2003GC000658.

    Keen, C.E., Dehler, S.A., 1993. Stretching and Subsidence: rifting of conjugate margins inthe North Atlantic Region. Tectonics 12, 1209–1229.

    Lau, K.W.H., Louden, K.E., Funck, T., Tucholke, B.E., Holbrook, W.S., Hopper, J.R., Larsen,H.C., 2006a. Crustal structure across the Grand Banks–Newfoundland BasinContinental Margin — I. Results from a seismic refraction profile. Geophys. J. Int.167, 127–156.

    Lau, K.W.H., Louden, K.E., Deemer, S., Hall, J., Hopper, J.R., Tucholke, B.E., Holbrook, W.S.,Larsen, H.C., 2006b. Crustal structure across the Grand Banks–Newfoundland BasinContinental Margin — II. Results from a seismic reflection profile. Geophys. J. Int.167, 157–170.

    Lavier, L.L., Manatschal, G., 2006. A mechanism to thin the continental lithosphere atmagma-poor margins. Nature 440, 324–328.

    Lecroart, P., Cazenave, A., Ricard, Y., Thoraval, C., Pyle, D.G., 1997. Along-axis dynamictopography constrained by major-element chemistry. Earth Planet. Sci. Lett. 149,49–56.

    Le Pichon, X., Sibuet, J.C., 1981. Passive margins: a model of formation. J. Geophys. Res.86, 3708–3720.

    Marillier, F., Hall, J., Hughes, S., Louden, K., Reid, I., Roberts, B., Clowes, R., Coté, T., Fowler,J., Guest, S., Lu, H., Luetgart, J., Quinlan, G., Spencer, C., Wright, J., 1994. LithoprobeEast onshore-offshore seismic refraction survey — constraints on interpretation ofreflection data in the Newfoundland Appalachians. Tectonophysics 232, 43–58.

    Marrett, R., Allmendinger, R.W., 1992. Amount of extension on “small” faults: anexample from the Viking graben. Geology 20, 47–50.

    Mauffret, A., Montadert, L., 1988. Seismic stratigraphy off Galicia. Proc. Ocean Drill.Program Sci. Results 103, 13–30.

    McKenzie, D., 1978. Some remarks on the development of sedimentary basins. EarthPlanet. Sci. Lett. 40, 25–32.

    McKenzie, D., Bickle, M.J., 1988. The volume and composition of melt generated byextension of the lithosphere. J. Petrol. 29, 625–679.

    Minshull, T.A., Dean, S.M., White, R.S., Whitmarsh, R.B., 2001. Anomalous meltproduction after continental break-up in the southern Iberia Abyssal Plain. In:Wilson, R.C.L., Whitmarsh, R.B., Taylor, B., Froitzheim, N. (Eds.), Non- volcanicRifting of Continental Margins: A Comparison of Evidence from Land and Sea.Geological Society, London, Special Publications, vol. 187, pp. 537–550.

    Parsons, B., Sclater, J.G., 1977. An analysis of the variation of ocean floor bathymetry andheat flow with age. J. Geophys. Res. 82, 803–827.

    Paulssen, H., Visser, J., 1993. The crustal structure of Iberia inferred from P-wave coda.Tectonophysics 221, 111–123.

    Pedreira, D., Pulgar, J.A., Gallart, J., Díaz, J., 2003. Seismic evidence of Alpine crustalthickening and wedging from the western Pyrenees to the Cantabrian Mountains(north Iberia). J. Geophys. Res. 108. doi:10.1029/2001JB001667.

    Pérez-Gussinyé, M., Reston, T.J., 2001. Rheological evolution during extension atnonvolcanic rifted margins: onset of serpentinization and development ofdetachments leading to continental breakup. J. Geophys. Res. 106, 3961–3975.

    Péron-Pinvidic, G., Manatschal, G., Minshull, T.A., Sawyer, D.S., 2007. Tectonosedimen-tary evolution of the deep Iberia–Newfoundland margins: evidence for a complexbreakup history. Tectonics 26, TC2011. doi:10.1029/2006TC001970.

    Press, W.H., Teukolsky, S.A., Vetterling, W.T., Flannery, B.P., 1986. Numerical recipes inFORTRAN. Cambridge University Press.

    Priestley, K., McKenzie, D., 2006. The thermal structure of the lithosphere from shearwave velocities. Earth Planet. Sci. Lett. 244, 285–301.

    Reston, T., 2007. Extension discrepancy at North Atlantic nonvolcanic rifted margins:depth-dependent stretching or unrecognized faulting? Geology 35, 367–370.

    Rosenbaum, G., Regenauer-Lieb, K., Weinberg, R., 2005. Continental extension: fromcore complexes to rigid block faulting. Geology 33, 609–612.

    Russell, S.M., Whitmarsh, R.B., 2003. Magmatism at the west Iberia non-volcanic riftedcontinental margin: evidence from analyses of magnetic anomalies. Geophys. J. Int.154, 706–730.

    Sclater, J.G., Christie, P.A.F., 1980. Continental stretching: an explanation of the post-mid-Cretaceous subsidence of the central North Sea basin. J. Geophys. Res. 85,3711–3739.

    Shillington, D.J., Holbrook, W.S., Van Avendonk, H.J.A., Tucholke, B.E., Hopper, J.R.,Louden, K.E., Larsen, H.C., Nunes, G.T., 2006. Evidence for asymmetric nonvolcanicrifting and slow incipient oceanic accretion from seismic reflection data on theNewfoundland margin. J. Geophys. Res. 111, B09402. doi:10.1029/2005JB003981.

    Shillington, D.J., White, N.J., Minshull, T.A., Edwards, R., Edwards, G., Jones, S.M., 2008.Cenozoic evolution of the eastern Black Sea: a test of depth-dependent stretchingmodels. Earth Planet. Sci. Letts. 265, 360–378.

    Shipboard Scientific Party, 2004. Leg 210 summary. Proc. Ocean Drill. Program, Init. Rep.210, 1–78.

    Sibuet, J.C., Srivastava, S., Manatschal, G., 2007. Exhumed mantle-forming transitionalcrust in the Newfoundland–Iberia rift and associated magnetic anomalies.J. Geophys. Res. 112. doi:10.1029/2005JB003856.

    Skelton, A.D.L., Valley, J.W., 2000. The relative timing of serpentinization and mantleexhumation at the ocean–continent transition, Iberia: constraints from oxygenisotopes. Earth Planet. Sci. Lett. 178, 327–338.

    Srivastava, S.P., Sibuet, J.C., Cande, S., Roest, W.R., Reid, I.D., 2000. Magnetic evidence forslow seafloor spreading during the formation of the Newfoundland and Iberianmargins. Earth Planet. Sci. Lett. 182, 61–76.

    Tett, D.L., Sawyer, D.S., 1996. Dynamic models of multi-phase continental rifting andtheir implications for the Newfoundland and Iberia conjugate margins. Proc. OceanDrill. Program Sci. Results 149, 635–648.

    Tucholke, B.E., Sibuet, J.C., 2007. Leg 210 synthesis: tectonic, magmatic and sedimentaryevolution of the Newfoundland–Iberia Rift. Proc. Ocean Drill. Program 210.doi:10.2973/odp.proc.sr.210.101.2007 Sci. Res.

    Tucholke, B.E., Austen, J.A., Uchupi, E., 1989. Crustal structure and rift-drift evolution ofthe Newfoundland Basin. In: Tankard, A.J., Balkwill, H.R. (Eds.), Extensionaltectonics and stratigraphy of the North Atlantic Margins, pp. 247–263.

    Turcotte, D.L.,1999. Fractals and Chaos inGeology andGeophysics, 2nd edition. CambridgeUniversity Press.

    Van Avendonk, H.J.V., Holbrook, W.S., Nunes, G.T., Shillington, D.J., Tucholke, B.E.,Louden, K.E., Larsen, H.C., Hopper, J.R., Seismic velocity structure of the riftedmargin of the eastern Grand Banks of Newfoundland, Canada, J. Geophys. Res. 111(2006) B11404. doi:10.1029/2005JB004156. J. Geophys. Res. 105 (2000) 5859–5885.

    Walsh, J., Watterson, J., Yielding, G., 1991. The importance of small-scale faulting inregional extension. Nature 351, 391–393.

    Wernicke, B., 1985. Uniform-sense normal simple shear of the continental lithosphere.Can. J. Earth Sci. 22, 108–125.

    White, N., 1989. Nature of lithospheric extension in the North Sea. Geology 17, 111–114.White, N., 1993. Recovery of strain rate variation from inversion of subsidence data.

    Nature 366, 449–452.White, N., Bellingham, P., 2002. A two-dimensional inverse model for extensional

    sedimentary basins 1. Theory. J. Geophys. Res.107, 2259. doi:10.1029/2001JB000173.White, R.S., McKenzie, D., O'Nions, R.K., 1992. Oceanic crustal thickness from seismic

    measurements and rare Earth element inversions. J. Geophys. Res. 97, 19683–19715.White, N., Thompson, M., Barwise, T., 2003. Understanding the thermal evolution of

    deep-water continental margins. Nature 426, 334–343.Whitmarsh, R.B., White, R.S., Horsfield, S.J., Sibuet, J.C., Recq, M., Louvel, V., 1996. The

    ocean–continent boundary off the western continental margin of Iberia: crustalstructure west off Galicia Bank. J. Geophys. Res. 101, 28291–28314.

    Whitmarsh, R.B., Manatschal, G., Minshull, T.A., 2001. Evolution of magma-poorcontinental margins from rifting to seafloor spreading. Nature 413, 150–154.

    Zelt, C.A., Sain, K., Naumenko, J.V., Sawyer, D.S., 2003. Assessment of crustal velocitymodels using seismic refraction and reflection tomography. Geophys. J. Int. 153,609–626.

    http://www.csr.utexas.edu/grace/gravity/http://www.csr.utexas.edu/grace/gravity/http://dx.doi.org/10.1029/2001TC001347http://dx.doi.org/10.1029/2002JB002026http://dx.doi.org/10.1029/2003GC000658http://dx.doi.org/10.1029/2001JB001667http://dx.doi.org/10.1029/2006TC001970http://dx.doi.org/10.1029/2005JB003981http://dx.doi.org/10.1029/2005JB003856http://dx.doi.org/10.2973/odp.proc.sr.210.101.2007http://dx.doi.org/doi:10.1029/2005JB004156http://dx.doi.org/10.1029/2001JB000173

    Evolution of the Newfoundland–Iberia conjugate rifted marginsIntroductionA general strategyPaleobathymetry at deep-water margins

    ApplicationCrustal structure and rift durationDynamic topography

    SerpentinizationStratigraphic records

    Inverse modelingDiscussionConclusionsAcknowledgementsSupplementary dataReferences