You are on page 1of 43

Chapter 2

Methods
2.1 Introduction
In this thesis, two distinct methodological approaches are applied to the study of the
subsidence and uplift history and the thermo-mechanical evolution of the West Iberian
Margin (WIM) and its conjugate o Newfoundland. These consist of: 1, Process Oriented
Gravity Modelling; and 2, Thermo-mechanical rheological models for the formation and
development of passive continental margins.
Process Oriented Gravity Modelling, POGM, combines backstripping and gravity
modelling techniques (Watts [1988]). The method uses information on the present day
conguration of the margin (e.g. sediment thicknesses, volcanism and underplating) to
determine the contributions from dierent geological processes to the observed gravity
anomaly and determine the geometry of the margin at the time of rifting. In the process,
constraints may be placed on the long-term exural strength of the lithosphere, hence
T
e
, both during and following rifting (Watts [1988]; Stewart et al. [2000]).
In non-volcanic passive continental margins, such as the conjugates o West Iberia
and Newfoundland, backstripping involves the removal of the sediment and water loads
and the restoration of the basement depth in their absence (e.g. Watts & Ryan [1976]).
Since the sediments are often spatially and temporally well dened loads, which can be
sequentially backstripped, the subsidence/uplift history of the margin is revealed (Watts
& Ryan [1976]; Steckler & Watts [1978]; Watts & Steckler [1981]).
Sediment backstripping can be performed in one of two ways. At a point location
(1-D), commonly utilized to derive relatively complete tectonic subsidence curves at bore-
holes, where there is detailed information on the stratigraphy and paleoenvironmental
history of the margin (Watts & Steckler [1979]; Lin et al. [2003]; Stewart et al. [2000]). 1-
D backstripping assumes that the sediment loads are locally compensated (Airy isostasy).
In most margins, however, well data is only available along the continental shelf and on-
shore. In addition, many of these wells are drilled by commercial companies, and tend
to be located over structural highs, anticlines, and other sorts of oil traps.
Alternatively, sediment backstripping can be accomplished along a prole (2-D), or
23
Chapter 2. Methods 24
for an area of the basin, using spatial grids (3-D). In both these cases, a regional, rather
than local, model of isostatic compensation can be used (exural Isostasy). In contrast
to 1-D Airy backstripping, 2-D and 3-D exural techniques provide the means to e-
ciently access the subsidence history of large basin areas (e.g. Watts & Torne [1992];
Stewart et al. [2000]) but, in most cases, lack the detailed stratigraphic and paleoen-
vironmental information of 1-D well backstripping. Where data is available, therefore,
these techniques should be used together (e.g. Stewart et al. [2000]).
The main aspects of the formation and development of rift-type basins have been
successfully explained using simple kinematic thermal models which, following McKenzie
[1978], assume that rift basins evolve through extension and thinning of the continental
lithosphere and subsequent thermal relaxation (e.g. McKenzie [1978]; Royden & Keen
[1980]; Watts et al. [1982]; Cochran [1983]). Both during extension and thermal con-
traction the lithosphere subsides, creating space for sediments to accumulate. These,
in turn, will act as loads, forcing the basin to further subside (e.g. McKenzie [1978];
Watts et al. [1982]). By accounting for the long-term strength of the lithosphere, and
modelling the basement response to the sediment loading, predictions can also be made
on the stratigraphy of rift basins (e.g. Watts et al. [1982]).
A thermo-mechanical rheological modelling software has been developed in this thesis;
the Mrift code. The main dierence in relation to previous codes is that it computes the
T
e
s based on yield strength envelope (YSE) considerations; i.e. which take into account
the temperature structure of the lithosphere, the rheology of the crust and lithospheric
mantle and the bending stresses caused by the sediment loading. Moreover, the compiled
software also allows for nite rifting as well as multi-stage continental rift scenarios,
2.2 Backstripping, Flexure, Subsidence and Gravity Mod-
elling
2.2.1 1-D Well Backstripping
1-D well backstripping uses information on the lithologies, ages and depths of deposition
of the main stratigraphic units to determine the margins tectonic subsidence (TS) at
a point location, assuming that both sediment and water loads are locally (i.e. Airy)
compensated.
The rst step in the 1-D well backstripping is to decompact the sediment column
and reconstruct its original thickness and density, at the time of deposition (Figure 2.1a
and b). For this, exponential-type porosity-depth curves, such as those proposed by
Athy [1930], and later by Bond & Kominz [1984], or power-law curves, as suggested by
Chapter 2. Methods 25
Baldwin & Butler [1985], are commonly utilized. These curves are empirically derived
relationships which account for the eects of mechanical compaction associated with the
sediment burial. Other eects of sediment diagenesis, such as cementation, overpressur-
ing, and dissolution, which could alter the density of the formations in relation to their
thickness, will not be considered in this study.
Alternatively, the porosity-depth relationships can be constrained from the porosity
values measured in a particular well (e.g. Steckler & Watts [1978]), using dierent types
of geophysical logs (e.g. sonic and density logs; Schlumberger [1989]). In either case,
from the porosity-depth curves the decompacted thickness (S

) and average density (


s
)
of a particular sediment layer can be expressed in terms of the present-day layer thickness
(S) and the porosities of the compacted
s
and decompacted

s
layer,
S

= S
1
s
1

s
(2.1)

s
=
w

s
+
g
(1

s
) (2.2)
where
w
and
g
are the densities of the water and sediment grains, respectively.
Recovering the basement depth in the absence of the sediment and water loads is
now a simple exercise of balancing the pressures at the base of the decompacted and
backstripped columns (Figure 2.1b and c, respectively). We have,

w
gW
d
+
s
gS

+t
c

c
g = TS
w
g +
m
gZ
m
+t
c

c
g (2.3)
where W
d
and TS are the water depth at the time of deposition and the recovered tectonic
subsidence, respectively, g is the average gravity and
m
is the density of the lithospheric
mantle.
It follows from Figure 2.1 that the portion of mantle above the compensation depth
(Z
m
) is,
Z
m
= W
d
+S

+t
c
(TS +
sl
+t
c
) (2.4)
where (
sl
) is the mean sea-level height with respect to the present day.
Upon substitution, and solving for the TS, Equation 2.3 becomes (Steckler & Watts
[1978]),
TS = S

m

s

w
_
+W
d

sl

w
(2.5)
Chapter 2. Methods 26
Figure 2.1: Rational of the 1-D backstripping technique: 1 layer example (modied from
Watts [2001b]). See text for equations and symbols.
Equation 2.5 is known as the backstripping equation and includes three independent
terms. From left to right these are, the sediment loading term, the water-depth term and
the sea-level loading term. The water depth and sea-level terms are obtained from the
paleobathymetry and the curves of global sea-level changes for the geological past (e.g.
Pitman [1978]; Watts & Steckler [1979]), respectively.
In reality, sedimentary basins represent accumulations of more than one sedimentary
unit. Therefore, it is necessary to backstrip all these units in order to recover the basement
TS at dierent times during the basin evolution. For this, the thicknesses and porosities
of all the layers must be re-computed at each time step, decompacting the younger units
and compacting the older ones, as depicted in Figure 2.2a. At any particular time of the
process (Time 1, Time 2, ... Time N in Figure 2.2), the average density of the restored
sequence is given by summing the masses of all the restored units and divide by the total
thickness (Steckler & Watts [1978]),

s
=

n
i=1
[
w

si
+
g
i(1

si
)] S

i
S

(2.6)
where i is the i
th
layer and n the total number of stratigraphic layers at a particular
time-step.
Chapter 2. Methods 27
Figure 2.2: Multilayer 1-D backstripping. In a), the rst column represents an observed
(present day), fully lithied, stratigraphic sequence that comprises 3 compacted units.
At Time 1, Unit 1 is decompacted to its original thickness and density (see equations
2.1 and 2.2). At Time 1, the basement depth in the absence of Unit 1 is isostatically
restored (backstrip; Equation 2.5), recovering the TS at time-step 1. At Time 2 the initial
thickness of Unit 2 is restored and Unit 1 compacted according to its new depth. At Time
2, the fully uncompacted Unit 2 and the partially compacted Unit 1 are backstripped,
recovering the TS at time-step 2. This procedure is then repeated for sediment layer 3.
In b), the corresponding TS curve is plotted. Modied from Bond & Kominz [1984].

s
can then be substituted in Equation 2.5 and the tectonic subsidence/uplift calcu-
lated at a particular time (i.e. TS
i
in Figure 2.2b). Figure 2.2b shows a typical backstrip
from a rift-type basin, commonly referred to as the tectonic subsidence curve (e.g. (Steck-
ler & Watts [1978])). It shows an initial, rapid subsidence event (syn-rift), followed by a
long period ( 50 to 100 m.y.) of exponential-like subsidence (post-rift).
2.2.2 Flexural Methods
A widely accepted model to determine the deformation of the lithosphere associated with
long-term (> 1 m.y.) geological loads (e.g. sediments, volcanics) is that of a thin elastic
plate overlying an inviscid substrate; i.e. which acts at every point along the base of
the plate with a force that is proportional to the deection produced by the load (Watts
[2001b] and references therein). This is a classical problem in engineering know as the
Winkler foundation (Hetenyi [1979]).
The thin plate approximation model assumes that the thickness of the deformed
plate is small in relation to its width, and that only the stresses in the plane of the plate
are signicant (Timoshenko & Woinowsky-Krieger [1959]). Furthermore, by considering
Chapter 2. Methods 28
only small plate deections, i.e. defections much smaller than the width of the plate,
linear elasticity can be assumed (e.g. Turcotte & Shubert [1959]). For the described
mechanical setting, and neglecting the eects of the horizontal forces, the exural loading
of a thin elastic plate is described by the following dierential equation (Timoshenko &
Woinowsky-Krieger [1959]),
Dw + 2
D
x

x
w + 2
D
y

y
w +Dw
(1 )
_

2
D
x
2

2
w
y
2
2

2
D
xy

2
w
xy
+

2
D
y
2

2
w
x
2
_
+p = q
(2.7)
where all variables may be function of spatial positioning (x,y), q and p represent the
force distribution due to the applied load and the restoring force exerted by the substrate,
respectively, = (
2
/x
2
) + (
2
/y
2
), and D is the exural rigidity of the plate, which
is related to T
e
by,
D =
ET
3
e
12(1 )
(2.8)
where E and are the Youngs modulus and Poissons ratio, respectively.
In the case of a lithospheric plate deformed under a sedimentary load, for example,
we have,
q(x, y) = (
s

w
)gh(x, y)
p(x, y) = (
m

infill
)gw(x, y)
where h(x,y) and w(x,y) are the amplitudes of the load and exural deformation of the
plate, respectively, and
infill
is the density of the material inlling the exural moat.
2-D Uniform Thickness Elastic Beams
The general equation for the two-dimensional exural bending of a thin elastic beam of
uniform rigidity (i.e. where D is constant) and innite extent perpendicular to the plane
of loading can be derived from Equation 2.7, by taking only the partial derivatives in the
x direction and substituting for the forces. We have,
D
d
4
w
dx
4
+ (
m

infill
)gw(x) = (
s

w
)gh(x) (2.9)
The solution to this equation is more eciently computed in the frequency domain,
Chapter 2. Methods 29
Figure 2.3: Flexural response function, (k). From Watts [2001b]
Dk
4
W(k) + (
m

infill
)gW(k) = (
s

w
)gH(k) (2.10)
where the upper case variables in bold represents the Fourier transform of the lower case
variables and k = 2/ is the wavenumber in the x-direction, where is the wavelength.
Upon rearranging Equation 2.10, the 2-D solution for the deection of a thin elastic
plate overlying an inviscid substrate in the frequency domain is,
W(k) = (k)
(
s

w
)
(
m

infill
)
H(k) (2.11)
The deection (w) can then be obtained by taking the inverse Fourier transform of
Equation 2.11. (k) is the exural response function, which is given by (e.g. Watts
[2001b]),
(k) =
_
1 +
Dk
4
(
m

infill
)g
_
1
(2.12)
and determines the range of wavelengths where the response to loading is exural (i.e.
the diagnostic waveband of exure of Watts [2001b], between 60 and 1000 km, approx-
imately; Figure 2.3).
For very long wavelengths, 1, which the Airy solution to Equation 2.11. On
the other hand, for short wavelength loads (i.e. 0) 0, which corresponds to
a Bouguer-type response to loading. Therefore, the lithosphere can be thought of as a
low pass lter between the loads and associated exural deformation, where the cuto
frequencies are determined by the rigidity of the plate (i.e. T
e
), as depicted in Figure 2.3.
Chapter 2. Methods 30
2-D Variable Thickness Elastic Beams
For a two-dimensional beam of laterally varying rigidity Equation 2.7 simplies to,
D(x)
d
4
w
dx
4
+ 2
dD
dx
d
3
w
dx
3
+
d
2
D
dx
2
d
2
w
dx
2
+p = q (2.13)
The code utilized in this thesis solves Equation 2.13 using a nite dierence approxi-
mation. The corresponding derivatives are fully described in Bodine [1980]. In geological
applications two situations might occur, which imply distinct boundary conditions. For a
continuous beam, zero deection and slope (built-in edge) are applied at great distances
from the edges of the load, i.e. at x and x ,
w = 0 ,
dw
dx
= 0
2.14
In the case of a broken plate model, commonly utilized to study the formation
and evolution of foreland basins, a built-in edge is applied at x , and boundary
conditions of zero curvature and shear (free edge) are applied at the end of the beam.
Consequently, at x = 0,
d
2
w
dx
2
= 0 ,
d
3
w
dx
3
= 0
2.15
2.2.3 Tectonic Subsidence and Stretching Factor
The McKenzie Rift Basin Model
In extensional settings, such as intracontinental rift basins and passive continental mar-
gins, the tectonic subsidence (TS) is commonly described as the vertical displacement of
the basement caused by the isostatic and thermal re-equilibrations of the stretched con-
tinental lithosphere (Sleep [1971]; McKenzie [1978]; Steckler & Watts [1978]). According
to the model of McKenzie [1978], the subsidence/uplift history of sedimentary rift basins
comprises two distinct stages. The rst stage corresponds to a rapid upward/downward
movement of the basement, which results from an isostatic re-equilibration of the masses
in the region of the rifted basin. The second stage consists of an exponential-like subsi-
dence pattern associated with the thermal relaxation of the lithosphere, which progres-
sively recovers its equilibrium (i.e. the pre-extension) thickness. This schema for the
formation and evolution of rift basins is illustrated in Figure 2.4.
Despite neglecting other eects of the rifting process, such as partial melting within
the mantle (e.g. Foucher et al. [1982]; Buck [2004]) and lateral migration of rheologically
weak lower crustal layers (Block & Royden [1990]; Bird [1991]), the McKenzie-type models
explain much of the observations related to formation and evolution of sedimentary rift
Chapter 2. Methods 31
basins, such as the subsidence and overall thermal structure, sediment thicknesses and
broad stratigraphic patterns (e.g. McKenzie [1978]; Watts et al. [1982]; Cochran [1983]).
Furthermore, they allow for simple relations to be established between the amount of
stretching, , and the TS and heat ow observed at a rift basin (McKenzie [1978]).
The initial (or isostatic) subsidence of a rift basin can be calculated by mass balancing
the columns of unstretched and stretched lithosphere down to a given reference level
(Figure 2.4a and b, respectively). The pressure at the base of the lithosphere before
stretching has occurred, P
un
, is given by,
P
un
= t
c

c
g + (L t
c
)
m
g (2.14)
where L and t
c
are the thicknesses of unstretched lithosphere and crust, respectively.
The densities of the mantle and crust,
m
and
c
, are average values which depend
on the temperature and are given by,

c
=
0c
_
1
T
m
t
c
2L
_

m
=
0m
_
1
T
m
2
_
1 +
T
c
L
__
(2.15)
where
0c
and
0m
are the densities of the crust and mantle at 0

C, respectively, is
the volumetric coecient of thermal expansion, and T
c
and T
m
are the average temper-
atures of the crust and mantle, which can be easily derived assuming linear temperature
gradients, as shown in the right panels of Figure 2.4.
Immediately after rifting has ceased, the pressure at the reference level underlying
the stretched lithosphere, P
st
, is given by (Figure 2.4b),
P
st
= S
i

w
g +
t
c


c
g +
_
L t
c

_

m
g +
_
L S
i

a
g (2.16)
where
a
is the density of the asthenosphere and S
i
the initial subsidence of the basin.
For the instantaneous stretching case considered here, the average densities of the
crust and lithospheric mantle are similar before and after stretching. By equating equa-
tions 2.14 and 2.16 we get,
S
i
= L
_
1
1

_
_
(
0m

0c
)
tc
L
_
1
Tmtc
2L
_


0m
Tm
2
_
[
0m
(1 T
m
)
w
]
(2.17)
After a relatively long period of time, the anomalous thermal gradient generated by
stretching will relax (Figure 2.4c). Since the average density of the asthenosphere is lower
than that of the lithosphere (
a
=
0m
[1T
m
]), thermal relaxation will result in further
Chapter 2. Methods 32
Figure 2.4: Two-stage rift model
proposed by McKenzie [1978] to
explain the formation and evolu-
tion of rift basins: subsidence (left)
and thermal structure of the litho-
sphere (right). a) Pre-stretched
lithosphere. b) After instantaneous
rifting event. c) After thermal re-
laxation of the lithosphere and re-
covery of the initial temperature
gradient. All the symbols are de-
scribed in the text, together with
the equations used to compute the
margins thermal and subsidence
evolution (see also Appendix A).
basin subsidence. The pressure at the reference level after the lithosphere has recovered
its equilibrium thickness, P
eq
, is now given by,
P
eq
= S
f

w
g +
t
c

g +
_
L S
f

t
c

_

m

g (2.18)
where S
f
is the nal basin subsidence, or the total tectonic subsidence (TTS), which
results from both the thinning and the thermal relaxation.
From Figure 2.4c it is also evident that the initial geotherm (before stretching has
occurred) is restored. However, the relative proportions of crust and mantle in the
lithosphere are now dierent and, consequently, the expressions for the average densities
of the crust and mantle have been altered,
Chapter 2. Methods 33

c

=
0c
_
1
T
m
t
c
2L
_

m

=
0m
_
1
T
m
2
_
1 +
T
c
L
__
(2.19)
Equating 2.14 and 2.18 and solving for the TTS holds,
TTS =
L(
m

m
) +t
c
(
m
+
c


c
)
(
m

w
)
(2.20)
For intermediate times since rifting, considerable thermal perturbations remain (i.e.
the geothermal gradients are highly non-linear) and the simple mass balances derived
above cannot be use to infer the subsidence/uplift history of the rift basins. The full
thermal treatment of the problem is considered in Appendix A.1, where it is shown
that for an instantaneous rifting event and vertical heat conduction (1-D heat ow) the
temperature, T, at a given time since rifting is given by (McKenzie [1978]),
T(z, t) = T
m
(1
z
L
) +
2T
m

sin
_

__
e
t/
sin
_
z
L
_
(2.21)
where z is depth, t is the time elapsed since rifting, and is the lithosphere thermal time
constant, which gives an indication of the length and time scale of the problem,
=
L
2

2
k
where k is the thermal diusivity.
The thermal subsidence of the rift basin can be derived from Equation 2.21. Following
McKenzie [1978], the elevation above the nal depth, E, to which the stretched lithosphere
subsides is,
E
4L
m
T
m

2
(
m

w
)
_

sin
_

__
(1 e
t/
) (2.22)
If an Airy isostatic model is assumed, where the masses of stretched and unstretched
lithosphere are balanced everywhere across the rift basin and the densities of the crust and
mantle do not vary with temperature, a simple expression can be derived from Equation
2.20, which directly relates the amount of stretching, , to the basins TTS,

1
= 1
TTS
t
c
(
m

w
)
(
m

c
)
(2.23)
In the perspective of this thesis, and of POGM in general (Watts [1988]; Stewart et al.
[2000]), the TTS corresponds to the position of the basement in the absence of surface
(e.g. sedimentary, volcanics) and sub-surface (e.g. underplating) loading. Therefore, once
Chapter 2. Methods 34
these loads have been backstripped from the present day basin conguration, Equation
2.23 can be used to recover the geometry of the basin at the time of rifting, which is
dened by the TTS and the backstrip Moho.
Strength During Rifting and Depth of Necking, Z
neck
A limitation to the modelling procedures described above is that they imply Airy isostasy
during rifting whilst allowing for the sedimentary loads to be modelled assuming either
Airy or exural compensation. In passive continental margins, where an intense phase of
stretching is normally accompanied by widespread normal faulting and anomalously high
temperature gradients (e.g. Watts et al. [1982]; Fowler & McKenzie [1989]; Bechtel et al.
[1987]), a zero strength rifting model might be a reasonable approximation. Rheologically
dependent dynamical models of rifting, however, appear to indicate that during stretching
the lithosphere retains part of its strength (Braun & Beaumont [1989]; Govers & Wortel
[1999]; Burov & Poliakov [2001]). Braun & Beaumont [1989], for example, have shown
that during rifting the strain is not evenly distributed in depth, but concentrates around
regions of maximum strength in the crust and mantle. These horizons will then control
the redistribution of material within the lithosphere and, depending on their depth,
departures from an Airy-type model of extension might arise (Figure 2.5).
The inclusion of syn-rift rigidity also appears to be a necessary requirement of some
kinematic models, in order to explain the uplifted rift anks and/or the gravity anomalies
at passive continental margins (e.g. Beaumont et al. [1982]; Kooi et al. [1992]; Keen &
Dehler [1997]; Watts & Stewart [1998]; Fjeldskaar et al. [2004]), and intra-continental
rift basins (e.g. Artemjev & Artyushkov [1971]; Ebinger et al. [1991]). Weissel & Karner
[1989], for example, established a relationship between the depth of maximum strain
within the crust, or the depth of necking (Z
neck
; following the notation Kooi et al. [1992])
and the topography of the rift basin, S
t
,
S
t
=
_
1
1

_
Z
neck
(2.24)
The isostatic loads can then be obtained from the mass balance between the unde-
formed lithosphere and the extended region. In most cases, however, the geometry of the
margin at the time of rifting is not known a priori. To overcome this problem, Watts
& Stewart [1998] devised a method which inverts for the rift structure of the margin
directly from the TTS. In essence, the authors assume that the TTS can be expressed in
terms of the initial basin prole (S
t
), which is Z
neck
-dependent (Equation 2.24), subse-
quently deformed by isostatic restoring stresses, which are supported by a nite rigidity
Chapter 2. Methods 35
Figure 2.5: Introduction to the concept of lithospheric necking in extensional settings.
According to Braun & Beaumont [1989], departures from an Airy-type model of extension
will produce: Left - a net upward vertical force, if necking occurs at greater depths than
predicted by the Airy model; Right - a downward vertical force, if necking is shallower
than Airy predicted. The central panel shows the reference Airy rift model which, for
the parameters given in Table 2.1 corresponds to a depth of necking, Z
neck
, of 7.4 km
(Watts & Stewart [1998]). The solid pink lines in the bottom panels show the initial
basin geometry and calculated Moho for comparison. The Z
neck
is represented in all the
panels by the dashed black line. t

c
the thickness of the thinned crust.
plate, T
e
(rift). For a given T
e
(rift) and Z
neck
, the formulation of Watts & Stewart [1998]
predicts that,
=
Z
neck
(k) +q
0
Q
0
(k) TTS

(k)
(2.25)
where the bold upper case variables represent the fourier transform of the lower case
variables and k is the wavenumber; TTS

= TTSZ
neck
; q
0
= Z
neck
(
w

m
)+t
c
(
m

c
); = Dk
4
+ (
m

w
)g.
Figure 2.5 illustrates the concept of Z
neck
and associated syn-rift isostatic loads. In
the central panel, an Airy-type rift margin (reference model) is generated, which for the
parameters described in Table 2.1 (in Section 2.3.3) corresponds to a neutral Z
neck
at
7.4 km (see Watts & Stewart [1998] for derivation). For a deeper value of Z
neck
(left
panel) the margin will subside more than predicted by the Airy model, and the isostatic
loads will act upwards in the centre of the basin (i.e. the margin is in an upward state
of exure; Weissel & Karner [1989]; Kooi et al. [1992]). This results in the uplift of the
rift anks and downward bulging at the base of the slope (bottom left panel). On the
Chapter 2. Methods 36
other hand, when Z
neck
is shallower than 7.4 km, the isostatic loads will act downwards
in the centre of the basin (i.e. downward state of exure), resulting in subdued rift anks
(right panels in Figure 2.5). The amplitude of the exural deformation and nal rift
basin prole will be determined by the strength of the lithosphere at the time of rifting,
T
e
(rift), and by the magnitude of the isostatic restoring stresses.
Another interesting feature in Figure 2.5 is that deeper Z
neck
s are apparently asso-
ciated with a relatively smooth Moho relief, whilst a high Moho relief is observed in the
shallow Z
neck
case. This can be easily understood from Equation 2.24 which, for similar
variations in the depth of the initial basin prole (S
t
) predicts larger oscillations in the
values of for shallower Z
neck
s.
Physical Meaning of Z
neck
Despite the relative success of the depth of necking concept in explaining a number
of features at passive continental margins and intracontinental rift basins, its physical
meaning remains unclear. This is partly due to the discrepancies between the results from
kinematic modelling approaches and the predictions from dynamical models of rifting.
According to the model predictions of Braun & Beaumont [1989], the necking depth
corresponds to the level of strength maxima within the lithosphere, that underwent no
net vertical motion during extension. For a typical continental yield strength envelope
(YSE), this level is located in the upper portion of the olivine-dominated lithospheric
mantle (e.g. Brace & Kohlstedt [1980]).
Results from other dynamic modelling approaches are in good agreement with those
of Braun & Beaumont [1989], in that they predict lithospheric necking as a determinant
factor in the geometry of the rift basins (e.g. Bassi et al. [1993]; Lynch & Morgan [1990];
Govers & Wortel [1999]). Govers & Wortel [1999], however, argue that Z
neck
is not xed
throughout rifting, but changes depth laterally due to a rapidly evolving thermal and
mechanical structure of the lithosphere during stretching, at least for s < 1.5.
In contrast to the results from the dynamical models of rifting, most kinematic mod-
elling approaches, which use constraints on the basin stratigraphy, crustal geometry and
gravity data to invert for Z
neck
, predict xed and relatively shallow (within crustal
depths) levels of necking (Keen & Dehler [1997]; Stewart et al. [2000]; Fjeldskaar et al.
[2004]), which are dicult to conciliate. Spadini et al. [1995], for example, inferred a
depth of necking of 25 km, lying between maximum strength levels in the crust and man-
tle. To explain this result, the authors suggested that Z
neck
might represent an average
response of the lithosphere to the redistribution of material during rifting, a view also
supported by van der Beek et al. [1994] and Keen & Dehler [1997].
Chapter 2. Methods 37
2.2.4 Calculation of Gravity Anomalies
In this thesis, two distinct methods have been used to compute the gravity anomalies of
bodies with a given density contrast. The direct line integral method of Talwani et al.
[1959] is applied to 2-D bodies, which assumes a uniform and innite extension of the
bodies in the direction normal to the modelled prole. The equivalent direct method to
calculate the gravity signature of 3-D bodies, i.e. volume integrals, is computationally
intensive. In these cases, the method introduced by Parker [1972] was used, which is
an approximate Fourier domain solution of the volume integral for the gravitational
potential.
Direct integral method
A simplication of the line integral method for a semi-innite slab model is described in
Talwani [1973]. In this technique, the gravity anomaly of an arbitrary shape 2-D body is
obtained by adding the contributions from thin, horizontal, semi-innite layers dened
by the edges of the body. This is illustrated for a simple polygonal shape in Figure 2.6,
and the gravity contribution from any single interface is given by,
g = 2G([x
1
sin +z
1
cos ][sin log
_
r
2
r
1
_
+ cos (
2

1
)] +z
2

2
z
1

1
)
where G is the universal gravitational constant, the density contrast across the sur-
face, and x,z the spatial coordinates. Figure 2.6 shows r
1
, r
2
, ,
1
and
2
.
Since the gravity eld is sensitive to lateral density contrasts, the slope of the body
interfaces is important. In the representation of Figure 2.6, an increase in would result
in the thickening of Layer 3 and thinning of Layer 2. In the special case of an horizontal
gradient, where = 0, there will be no density contrast and, consequently, no gravity
signal for that interface.
The wavelength of the computed gravity anomalies is, however, primarily controlled
by the depth of the slab. This is because the strength of the gravity signal decreases as
a function of the wavenumber, k, by a factor of e
kh
, where h is the height above the
density interface. Shorter wavelength gravity anomalies, with steeper gradients, are then
associated with anomalous density bodies located closer to the observation surface (i.e.
where the gravity is being measured), normally within the upper crust and sedimentary
basins, whilst longer wavelength anomalies are caused by density contrasts at greater
depths, such as at the crust/mantle interface.
Chapter 2. Methods 38
Figure 2.6: Semi-innite slab model of Talwani [1973] for 2-D gravity calculations.
Parkers method
The Parkers method is an ecient Fourier domain solution to calculate the gravity
anomalies associated with density contrasts across uneven 3-D surfaces. As demonstrated
by Parker [1972], for a surface with a mean depth, Z
0
(see Figure 2.7),
F[g] = 2Ge
(kZ
0
)

n=1
(k)
n1
n!
F[Z
n
(x, y)] (2.26)
where F[ ] represents the Fourier transform of the variable within brackets, g is the
gravity anomaly, k is the wavenumber, and Z(x,y) is the depth to the surface (positive
upwards) of density contrast ,
Although the Fourier series can be expanded innitely, only terms for n < 4 have been
considered in the calculations. For n > 4, the magnitude of the calculated anomalies is
negligible.
2.2.5 Process Oriented Gravity Modelling
Process Oriented Gravity Modelling, POGM, is a forward modelling procedure that uses
seismic constraints on the sediment structure and magmatic extrusive and intrusive bod-
ies to determine the contributions from distinct geological processes to the present day
gravity eld at passive continental margins (Watts [1988]; Watts & Marr [1995]; Watts
& Fairhead [1999]). In the process, the geometry of the margin at the time of rifting and
the basement subsidence history are recovered. The methodology followed in POGM is
illustrated in Figure 2.8, using a synthetic passive margin transect where only sediment
loads are present, and can be described as a sequence of steps:
Chapter 2. Methods 39
Figure 2.7: 3-D gravity anomaly calculation for an uneven, uniform layer [Z(x,y)] of mean
depth Z
0
(shaded yellow). Modied from Blakely [1996].
1. Backstrip the observed sediment loads to determine the basement TTS (dashed line
in Figure 2.8a), assuming either a xed or laterally varying T
e
structure;
2. Calculate the s and recover the geometry of the Moho at the time of rifting (i.e.
the backstrip Moho), assuming either an Airy or nite strength model of rifting
(Figure 2.8a), and compute the rift anomaly (water-crust + mantle-crust). This
resulting anomaly shows a characteristic edge-eect, with a high over the shelf
break and a low in the region of the continental slope and/or rise (Figure 2.8b);
3. Restore the sediment load to the margin and compute the exure for the T
e
struc-
ture used in step 1 (Figure 2.8c). The gravity anomaly associated with the sedi-
mentation process can thus be obtained from the sum of loading (sediment-water)
and exural (sediment-crust + crust-mantle) eects (Figure 2.8d);
4. Add the rift and sedimentation anomalies and compare the sum to the observed
gravity eld. By varying the T
e
structure used to backstrip the sediment loads
Chapter 2. Methods 40
Figure 2.8: Illustration of the POGM methodology.
dierent sum anomalies will be computed, and constraints can be placed on the
long-term mechanical structure across the margin transect (Figure 2.8f). Additional
constraints can be obtained by comparing the nal geometry of the gravity Moho
(i.e. the exed Moho in Figure 2.8e) to the seismic Moho constrained from wide-
angle and refraction data.
In this thesis, POGM techniques will be applied to closely space proles orthogonal to
the local trend of the WIM (Chapter 5). The results from POGM will provide the basis
to investigate the mechanical segmentation of the margin along strike, and its relation-
ship with the structure of the Variscan orogen onshore, as well as the eects of rifting on
the long-term (> 1 m.y.) strength of the continental lithosphere. A number of conjugate
transects along the West Iberia and Newfoundland margins will also be modelled (Chap-
ter 7), to study the symmetry of the rift system, in terms of the margins thermal and
mechanical structure, and how it relates to observations such as the unstretched crustal
structure, the amount of basement subsidence and the width of the rift.
Chapter 2. Methods 41
The advantage of the West Iberia and Newfoundland margins is that they have been
extensively targeted by seismic surveys during the last three decades, including a great
number of high quality multi-channel and several wide angle and/or refraction seismic
experiments, commercial drilling, particularly along the continental shelf and onshore,
and a number of DSDP and ODP legs. This data provides well constrained sediment
and crustal structures in dierent segments of the margins, which can be use to control
the results from the gravity modelling. For this study, new bathymetric and sediment
thickness datasets have also been compiled for the WIM. These will be discussed in
detail in Chapter 3. The sediment thickness dataset, in particular, is based on previously
compiled structural maps and on high quality, recently acquired commercial seismic data,
and represents a signicant improvement in relation to previous datasets in the region.
In addition to the contributions from rifting and sedimentation, the evolution and
present day setting of the West Iberia Margin are also marked by processes such as
upper mantle serpentinization and margin inversion (Pinheiro et al. [1996] and references
therein). One possibility, therefore, is that these processes might signicantly contribute
to the observed gravity anomaly. In this thesis, the discrepancies between the sum gravity
anomalies, calculated through the sequence of steps described above, and the observed,
will be discussed in terms of such contributions, using the available seismic and well data
constraints.
2.3 Thermo-Mechanical Rheological Models: T
e
Structure
and Stratigraphy of Rift Basins
2.3.1 Renements to the McKenzie Rift Model
The thermal model for the formation and evolution of rift-type sedimentary basins de-
scribed in Section 2.2.3 (i.e. the model proposed by McKenzie [1978]) assumes instanta-
neous and uniform lithospheric stretching and only accounts for vertical heat ux during
the post-rift thermal contraction stage (see Appendix A.1). A more accurate model to
describe the thermal structure and subsidence history of rift basins needs, however, to
include horizontal as well as vertical (i.e. 2-D) conduction of heat (Steckler [1981]; Watts
et al. [1982]). The thermo-mechanical modelling software compiled in this thesis, the
Mrift code, uses the solution to the 2-D heat ow equation implemented by Watts et al.
[1982]. In particular, these authors have shown that by subdividing a rift basin prole
into a nite number of xed width blocks, the temperature within a given block as a
function of depth and time is given by (see Appendix A.2 for full derivation),
Chapter 2. Methods 42
T(x, z, t) = T
m
z
L
+
1
2
N

n=i1
_
erfc
_
x x
i

4kt
_
erfc
_
x x
i+1

4kt
__

i
(z, t) (2.27)
where x
i
and x
i
+1 are the edges of block i, N is the total number of blocks in the prole
and (z, t) is the thermal transient of the vertical heat ow solution (Watts et al. [1982]).
Figure 2.9b compares the subsidence predicted for a simple basin geometry assuming
1-D (left) and 2-D (right) heat ow. The shape of the depression is determined by the
s shown in Figure 2.9a (note perfect symmetry around the centre of the basin), and
the model parameterisation is given in Table 2.1. The yellow shaded areas show the
amount of syn-rift subsidence. The subsidence is then calculated at n
2
m.y. since rifting,
where n = 0, 1 ,2 ,3 ,4... 13. Two main dierences are noticed between the 1-D and 2-D
scenarios. First, when both vertical and horizontal heat ow are considered the shape of
the basin does not remain constant through time. Second, while the initial subsidence
is identical, the lateral ux of heat during post-rift times produces a thermal bulge in
the basin anks. The maximum uplift occurs early in the history of the basin, where
it reaches 350 m. At later times, as heat diuses outwards, the uplift broadens and
becomes less prominent.
The uplift of the basin anks is accompanied by an increase in the amount of subsi-
dence near the base of the slope, where heat is lost more rapidly due to the higher thermal
gradients. Together, these two eects result in a slight upward bulging of the slope and
a bowl-shaped basin centre. After most of the anomalous heat has been diused, 50
m.y. following the cessation of the rifting event, the basin depression attens, but the
basement, in the 2-D case, remains deeper by a few hundred metres than in the simple,
1-D thermal model (Figure 2.9b).
In a subsequent study, Cochran [1983] investigated the combined eects of lateral heat
ow and nite rifting in the thermal structure and subsidence history of rift basins. This
analysis was motivated by evidences from several passive continental margins, namely
around the North Atlantic and the Red Sea (see Cochran [1983] and references therein),
which suggest that rifting events may often last between 20 and 50 m.y. This is signi-
cantly longer than the 20 m.y. period for which the instantaneous rift model of McKenzie
[1978] has been considered as a good approximation, according to the calculations of
Jarvis & McKenzie [1980].
As depicted in Figure 2.9c, nite rifting results in the increase of the syn-rift subsi-
dence, at the expense of the post-rift. The estimated reduction in the amount of post-rift
subsidence for the case shown in Figure 2.9c, which assumes 2-D heat ow and a 25
m.y. duration rifting event, is 22% in the centre of the basin. Towards the base of
the slope, the calculated reduction is even larger, approaching 28%. These results are in
Chapter 2. Methods 43
Figure 2.9: Predicted subsidence in a simple rift basin geometry. Eects of lateral heat
ow (a) and nite rifting (b). During the post-rift stages, the subsidence was calculated
at n
2
m.y. since rifting, where n = 0, 1 ,2 ,3 ,4... 13. During the syn-rift, the subsidence
was determined at 5 m.y. intervals.
good agreement with those of Cochran [1983], who also noticed that the eects of nite
rifting are particularly relevant when lateral as well as vertical heat ow is considered,
even for rift durations < 20 m.y.
In the kinematic models which allow for nite rift durations, some important aspects
of basins evolution and thermal structure need to be taken into account. The Mrift
code compiled in this study follows the model proposed by Cochran [1983], in that the
total amount of stretching associated with a rifting event is rst discretised into small
increments of extension. The amount of thinning at each block of the basin prole, and
correspondent horizontal stretching, is then computed for each time step assuming a
particular distribution throughout the syn-rift period (e.g. linear or exponential, see
Section 2.3.3). Finally, at the end of each step the temperature structure is computed
from Equation 2.27, and used as the new temperature structure in the next rift pulse.
As noticed by Cochran [1983], however, this poses a problem when calculating the
Fourier coecients which determine the vertical component of the temperature structure
Chapter 2. Methods 44
within each block (see Equation A.3), since the pre-rift temperature gradient, after the
rst pulse, is no longer linear. To overcome this diculty, a Fourier sine transformed is
performed on the temperature anomaly at the end of each time increment to determine
the new coecients; where the temperature anomaly is the dierence between the steady
state temperature gradient and the new temperature structure (T[x, z, t] T
m
z
L
).
The thermo-mechanical modelling software compiled in this study has also been
adapted so that it allows for multi-stage continental rifting, from where it acquired its
designation, Mrift. This is important since rift basins, in general, and passive continental
margins, in particular, often evolve through a sequence of extensional pulses, intercalated
by periods of thermal quiescence. Multi-stage continental rifting has been inferred from
the seismic stratigraphy and borehole data, for example, in the North China Basin (Shed-
lock et al. [1985]), the South China Sea Basin (Pigott & Ru [1994]), the Canning Basin,
Western Australia (Braun [1992]), the Vring Margin, o Norway (Reems & Cloetingh
[2000], the Rockall Trough, o NW Ireland (Hauser et al. [1995] and references therein),
and the conjugate margins of West Iberia (e.g. Wilson et al. [1989]; Pinheiro et al.
[1996] and references therein) and Newfoundland (Tankard & Welsink [1987]; Driscoll
et al. [1992]). The evolution of the WIM, for example, is characterized by at least three
extensional episodes (e.g. Wilson et al. [1989, 1996]; see Chapter 3 for discussion).
In order to account for a sequence of rifting events a couple of modications had to
be implemented into the previous versions of the Mrift code, namely:
1. The subroutines that compute the lateral heat ow had to be adapted to allow for
varying width blocks along the modelled prole during periods of thermal relax-
ation. This situation arises between two rifting stages, since the blocks full width
(which is constant throughout the prole) is only attained at the end of the last
rifting episode.
2. The temperature structure at the end of a rift stage needs to be stored, in order
to use it at the start of the next one, similarly to that proposed by Cochran [1983]
for discrete stretching pulses within the same rifting episode.
Figure 2.10 compares the subsidence history in a simple, symmetric rift basin geom-
etry predicted with the Mrift code for two distinct scenarios: 1 - a single, 169 m.y. old
rift event, lasting for 25 m.y. (grey lines in the left panel of Figure 2.10b); 2 - two nite
duration rifting episodes. The rst starting 169 m.y. ago and lasting for 25 m.y., followed
by a second, 95 m.y. old and lasting for 15 m.y. (orange and black lines in the right panel
of Figure 2.10b, respectively). Figure 2.10c shows the predicted subsidence history at
four dierent locations across the basin, which include the basin anks (Location A), the
Chapter 2. Methods 45
Figure 2.10: Predicted subsidence for single and multi-stage, nite duration continental
rift scenarios. a) Shows the s utilized to compute the basin subsidence. The grey lines
correspond to the single rift scenario, which starts at 169 Ma and lasts for 25 m.y. The
orange and black lines correspond to the rst and second events in the multi-rift scenario,
respectively. The rst rift starts at 169 Ma and last for 25 m.y., and the second episode
starts at 95 Ma and lasts for 15 m.y. The total amount of stretching is kept constant
throughout the basin in both cases. b) Shows the subsidence history for the s and rift
times and durations described above. In c), the tectonic subsidence curves are shown at
for 4 distinct locations.
upper slope (B), the lower slope (C) and the basin centre (D). The s used to simulate
the single and multi-stage rifting cases are shown in Figure 2.10a, where the total amount
of stretching is kept constant between the 2 scenarios.
In the two-rift scenario, a 130 km wide, relatively shallow basin is initially formed. The
s corresponding to this event are relatively small (< 1.2, orange lines in Figure 2.10a).
The lithosphere is then allowed to cool for a period of 49 m.y., before being subjected
to a second stretching pulse (black lines). This second rift stage is more localized, with
the s increasing abruptly from 1 to > 2 in less than 25 km width. As a consequence,
Chapter 2. Methods 46
a steep thermal gradient is developed and the slopes of the basin formed during the rst
rifting event are uplifted by more than 500 m (between 95 and 80 Ma in Site B of Figure
2.10c).
Although the total tectonic subsidence obtained with both rifting scenarios is ap-
proximately the same across the basin prole (< 250 m dierence; Figure 2.10b and c),
the predicted subsidence history varies considerably. By comparing the thermal model
predictions with the subsidence calculated from well data and detailed stratigraphic in-
formation, commonly available at intracontinental rift basins and along the shelf and
onshore regions of passive continental margins, constraints may be placed on the age and
duration of the stretching events, as well as on the evolution of the thermal structure of
the rifted lithosphere and, consequently, on its long-term mechanical properties.
The remaining of this chapter introduces and discusses briey the concepts underlying
the calculation of the T
e
s based on the thermal structure of the lithosphere and the
rhelogy of the crust and mantle (Section 2.3.2), describes the parameterisation and setup
of the thermo-mechanical rheological models (Section 2.3.3), and analysis the eects of
varying the thermal and mechanical settings (sections 2.3.4 and 2.3.5), as well as the
times and durations of the rifting events (Section 2.3.6), in the stratigraphy of rift basins
and passive continental margins.
2.3.2 Analytical Calculation of T
e
An important development of the Mrift code compiled in this study is that it computes
the lithospheric eective elastic thickness, T
e
, directly from the thermal structure inferred
from the kinematic rift models. For a given thermal state, T
e
will depend on the failure
properties of the rocks, and the state of stress generated by the bending of the plate
when acted by a geological load (e.g. the sediments).
In the crust and lithospheric mantle, plastic deformation is governed by two distinct
mechanisms, frictional resistance and dislocation creep (Brace & Kohlstedt [1980]). In
this study we adopt linear frictional failure relations of the form (Sibson [1974]),
=
_
(
_

2
+ 1 )
2
1
_
gz(1 ) (2.28)
for the compression stress regime, and
=
_
1 (
_

2
+ 1 )
2
1
_
gz(1 ) (2.29)
to describe the extensional stress regime. Where is the sustainable dierential stress,
is the coecient of static friction, is the density of the lithostatic column, z is depth,
= P/gz, and P is pore pressure.
Chapter 2. Methods 47
Deformation in the brittle regime is then dependent upon the depth and the frictional
properties of the material, here assumed constant with = 0.65 and = 0.37 (Brace &
Kohlstedt [1980]).
Deformation by ductile ow, which describes the plastic behaviour of rocks at greater
temperatures and pressures, is given by a temperature dependent dislocation creep power
law,
=
_

A
_1
n
exp
_
E
nRT
_
(2.30)
where is the strain rate, A and n are experimentally derived material constants, R is
the gas constant, T is the temperature, and E is the activation energy of the material
(e.g. Goetze & Evans [1979]).
For a given rheology, the failure criterion at any depth is then determined by the
smaller of the sustainable dierential stresses, and commonly summarized via a yield
strength envelope (YSE) representation (Goetze & Evans [1979]). Figure 2.11 shows
a YSE for continental-type lithosphere, assuming a wet quartz rheology for the crust
(Kronenberg & Tullis [1984]) and dry dislocation creep for the olivine mantle (Hirth
& Kohlstedt [2003]; see Table 2.2), and using the lithospheric parameterisation given
in Table 2.1. The importance of quantifying the dierential stresses in the brittle and
ductile domains of plastic deformation is that these will impose limits on the portion of
the lithosphere deforming elastically (e.g. McNutt [1984]; Ranalli [1994]).
The exural rigidity, D, of an elastic plate is related to the bending moment, M,
generated upon loading by,
D = Mr (2.31)
where r is the radius of curvature of the exed plate.
_
1
r
_
=
_

2
z
x
2
_
(2.32)
and z is the vertical deformation imposed on the plate.
The bending moment, in turn, can be described in terms of the integrated bre
stresses,
f
, generated by the exure of the plate. We have,
M =
_
h/2
h/2

f
(z)(z z
n
)dz (2.33)
where h is the mechanical thickness of the plate (i.e. that supports the deviatoric stresses
on geological time scales), z is the depth, and z
n
is the depth of a neutral surface within
the plate, i.e. where neither extensional nor compressional deformation is occurring.
Chapter 2. Methods 48
Figure 2.11: The yield strength en-
velope (YSE) derived from equa-
tions 2.28 to 2.30 an using typical
continental crust and lithospheric
mantle rheologies (see text).
Assuming linear elasticity, the bre stresses generated due to exure are directly
related to the curvature of the plate by,

f
=
E
1
2
_

2
z
x
2
_
(z z
n
)dz (2.34)
where E and are the Youngs modulus and Poissons ratio, respectively.
If the entire thickness of the lithosphere behaves elastically, then T
e
corresponds to
the mechanical thickness (h) in Equation 2.33. This is the situation illustrated in Figure
2.12a, where the vertically integrated bre stresses are calculated within the grey shaded
area. However, if the plate bending stresses overcome those determined by the frictional
resistance and dislocation creep properties of the materials failure will occur, with the
consequent reduction in the thickness of the plate responding elastically (Figure 2.12b).
The examples shown in Figure 2.12 were built using a single rheology for the entire
crust. As shown by Burov & Watts [2006], however, the presence of a strong lower
crustal layer might result in a drastic increase of the overall rigidity of the lithosphere.
This is mainly due to the eects of coupling/decoupling between the mechanically strong
upper crust and upper olivine mantle layers (Figure 2.13). As demonstrated by Burov &
Diament [1992], the T
e
of a lithosphere composed of three mechanically decoupled layers
(i.e. upper-middle crust, lower crust and mantle) can be approximated by the following
expression,
Chapter 2. Methods 49
Figure 2.12: Relationship between the bre stresses calculated in a thin bending plate
a), and the YSE failure criterion for a crustal type rheology b). The light grey shaded
areas show the region where the plate behaves elastically. In b), the dark shaded areas
represent areas of the YSE where the elastic stresses exceed the strength of the material as
determined by its frictional resistance (above) and dislocation creep properties (below).
Consequently, the plate eective elastic thickness, T
e
, will be reduced. z
n
is the depth of
the neutral surface (see text). Modied from Lowry & Smith [1995].
T
e
=
_
Te
3
1
+Te
3
2
+Te
3
3
_1
3
(2.35)
This results in a signicantly lower T
e
then when vertically integrating the bending
moment along the entire thickness of the lithosphere (i.e. Equation 2.33). In the Mrift
code, the criterion for mechanical decoupling was established for the dislocation creep
power law (which asymptotically approaches zero) when < 20MPa.
Figure 2.13 shows the sensitivity of the T
e
calculated based on the YSE to variations
in the rheology of the lower crust, strain rate and amount heat generated by the decaying
of radioactive isotopes in the upper crustal layers. In Figure 2.13a, for example, changing
from an aggregate lower crust (50% dry quartz and 50% anorthosite) to stronger, more
mac, anorthosite and mac granulite compositions (see Table 2.2 for parameterisation
and references) results in a progressive increase in the T
e
, with a total variation of less
then 10 km. This variation, obtained for a crustal thickness of 31 km and a lower crust
of 10 km (see Table 2.1) is relatively small when compared to that estimated by Burov &
Watts [2006] for a considerably thicker lower crust layer (15 km) in a 40 km thick crust.
The eects of mechanical coupling and decoupling are well noticed in Figure 2.13b
Chapter 2. Methods 50
Figure 2.13: Sensitivity of the calculated T
e
s to variations in the lower crust rheol-
ogy (a), strain rates (b) and radiogenic heat production (c). For each YSE envelope
the reference model (see Section 2.3.3) is highlighted in red. The lithospheric thermal
parameterisation and rheologies are described in Tables 2.1 and 2.2, respectively. The
dashed grey vertical line at 20MPa represents the limit for mechanical decoupling (see
text). The greater oscillations in the calculated T
e
s occur when coupling/decoupling
between distinct rheological layers occurs.
and 2.13c, respectively. The strain rates control the thickness of a given rheological
layer that deforms by dislocation creep (Equation 2.30). For the values tested in Figure
2.13b, which are within the range predicted for relatively small amounts of extension in
rift basins ( < 1.2, e.g. White [1994]; Newman & White [1999]), signicant variations
in the calculated T
e
s are obtained, especially for the higher strain rates, due to the
mechanical coupling between the upper-middle and lower crust and the mantle.
The heat generated by the decaying of radioactive isotopes (mainly uranium, thorium
and potassium) in crustal rocks has long since been recognised as an important contri-
bution of the net heat ux measure at the surface of the continents (Roy et al. [1968];
Lachenbruch [1970]). The considered heat production follows an exponential decrease
with depth (z), which according to Lachenbruch [1970] is given by,
A(z) = A
0
e
(z/D
0
)
(2.36)
where A
0
is the surface radiogenic heat production and D
0
is the length scale parameter
that describes the depth-dependent decrease.
If the contribution from heat produced in the crust is taken into account an increase
in the Moho temperature between 50
0
and 100
0
C is expected (Morgan & Sass [1984]).
Pasquale et al. [1997], for example, estimated an increase of 80
0
C for a 28 km thick
Chapter 2. Methods 51
crust and values of A
0
similar to those suggested by Fern` andez et al. [1998] for the West
Iberia (see Table 2.1). As depicted in Figure 2.13c, accounting for the eects of radiogenic
heat production might have important consequences in the calculated T
e
. In this case,
the largest variation occurs due to the mechanical decoupling between the lower crust
and the mantle, even for a small value of D
0
.
2.3.3 Model Setup and Parameterisation
This section discusses briey the default parameterisation and setup of the thermo-
mechanical rheological models which will be presented in the following sections of this
chapter and in Chapter 6. All the models are based on a previously dened structure,
which determines the amount of thinning within a small segment, or block, of the basin
prole. As emphasized by Cochran [1983], an important limitation is imposed on the
size of the blocks, which must be short enough in order to approximate a continuous
rifting event, yet long enough so that the eects of lateral heat conduction are felt in the
adjacent block of the basin model (Equation 2.27). Throughout this thesis, a block size
of 5 km will be assumed, which for time increments of 0.5 and 1 m.y., used during the
syn-rift and post-rift stages, respectively, gives acceptable results (see Cochran [1983]).
For a given prole of s, the tectonic subsidence and temperature structure at each
block will then depend on the thermal and density parameterisation of the lithosphere,
and on some modelling options which are available in the compiled software. These are
described in Appendix B, together a list of the input and output les, and details on the
usage of the help les of the Mrift code.
The reference parameters used throughout this work are given in Table 2.1, and will be
assumed unless specied otherwise. The initial crustal thickness as well as the parameters
that control the amount of radiogenic heat production in the upper-middle crust (d and
A
0
in Table 2.1) were established according to published values for the WIM and adjacent
western Iberia continental areas (ILIHA et al. [1993]; Matias [1996]; Fern` andez et al.
[1998]). The average crustal and sediment densities were determined in this study, from
multi-channel stack velocities, wide-angle and refraction seismic data (see Chapter 3 for
details on the sediments and crustal velocity structure). The initial lithospheric thickness
and other thermal parameters were established so that the unstretched lithospheric prole
is isostatically balanced with a mid-ocean ridge system with a 5.5 km thick crust at a
depth of 2.6 km (Cochran [1981]).
From the temperature structure and the rheologies of the crust and upper lithospheric
mantle it is possible to constrain the YSEs used to calculate the T
e
s, as described in the
previous section, provided that both the strain rates and the bending stresses are known.
Chapter 2. Methods 52
Parameter Symbol Value
Initial lithospheric thickness L 125 km
Initial crustal thickness t
c
31 km
Initial lower crustal thickness t
lc
8 km
Average crustal density
c
2780 kg m
3
Mantle density (0

C)
0m
3330 kg cm
3
Water density
w
1030 kg m
3
Average sediment density
s
2230 kg m
3
Mantle potential temperature T
0
1333

C
Volumetric coecient of thermal expansion 3.1 x 10
5
K
1
Thermal diusivity k 8.0 x 10
7
m
2
s
1
Mantle thermal conductivity K
m
3.138 W m
1
K
1
Crustal thermal conductivity K
c
2.700 W m
1
K
1
Surface heat production A
0
3 W m
3
Length scale parameter d 10
Table 2.1: Reference parameters for the thermo-mechanical models.
The composition of the deep crustal layers in continental areas is often deduced from
comparisons between seismic velocity models and measurements under laboratory condi-
tions (e.g Hoolbrook et al. [1992]; Christensen & Mooney [1995]). However, a considerable
amount of overlapping is observed between the velocity intervals measured over dierent
crustal type facies (Rudnick & Fountain [1995], and references therein).
In this study, wet and dry quartz rheologies are used for the upper-middle crust, and
anorthosite, aggregate (quartz+anorthosite) and mac granulite rheologies are tested
for the lower continental crust, in agreement with the layering and the range of seismic
velocities observed along several wide-angle and refraction seismic transects onshore West
Iberia (ILIHA et al. [1993]; Matias [1996] and references therein). For the mantle, the
rheological parameters were extracted from Hirth & Kohlstedt [2003] for wet and dry
dislocation laws. The parameters for the crustal and mantle rheologies, and corresponding
references, are summarised in Table 2.2.
The strain rate within each block gives the increment of extension in relation to its
previous width. Two options are available in the Mrift code to calculate the strain rates,
which correspond to a linear or exponential increase in the amount of extension with
time. Throughout this thesis, an exponential growth is assumed, which appears to be
in better agreement with the strain rates inferred from tectonic subsidence curves (White
[1994]), numerical models of fault growth and segment linkage (e.g. Gupta et al. [1998]),
Chapter 2. Methods 53
and 3-D analysis of the tectono-sedimentary evolution of rift basins (e.g. Gawthorpe &
Leeder [2000]).
Material n A (Pa
n
s
1
) E (KJ mol
1
)
Dry quartz
1
2.9 5 x 10
25
149
Wet quartz
2
2.4 1.3 x 10
20
134
Anorthosite
3
3.2 5.6 x 10-23 238
Polyphase Aggregate
4
3.04 4.9 x 10
24
192
Mac Granulite
5
4.2 1.4 x 10
4
445
Olivine dry dislocation
6
3.5 1.1 x 10
5
530
Olivine wet dislocation
6
3.5 1.600 x 10
3
520
Table 2.2: Rheological parameterisation of the crust and lithospheric mantle. Source
references are as follows: 1, Koch [1983]; 2, Kronenberg & Tullis [1984]; 3, Buck [1991];
4, Tullis et al. [1991]; 5, Mackwell et al. [1998]; 6, Hirth & Kohlstedt [2003]. The asterisks
means that A is given in MPa.
The bending stresses are calculated from the curvatures imposed by the sedimentary
loads (see Section 2.3.2). An initial, background value of 5x10
8
is assumed for the
rst calculations, which is below the minimum curvatures commonly observed at passive
continental margins (e.g. Wyer & Watts [2006]).
2.3.4 Subsidence, Flexure and Stratigraphic Modelling: The Steers
Head Basin
From the thermal structure and the subsidence, calculated using the kinematic models
described in sections 2.2.3 and 2.3.1, and the T
e
s, we can determine the sediment loads
and associated exural deformation at dierent stages of a rift basin evolution, and thus
model the basin stratigraphy (e.g. Watts et al. [1982]; Cochran [1983]; Watts & Thorne
[1984]; Reynolds et al. [1991]). The procedure is illustrated in Figure 2.14 for a simple
basin geometry and an instantaneous rifting event, and where the eects of lateral heat
ow and radiogenic heat production in the crust were neglected.
The stretching factors (s), temperature structure at the end of the rifting event
and predicted basin subsidence through time are shown in Figure 2.14a, 2.14b and 2.14c,
respectively. For a given interval of time, the sediment load is determine by the amount
of subsidence and the paleowater depths. For simplicity, the paleowater depths in the
models presented in this section are kept at the pre-deformation surface at all times; i.e.
zero paleowater depths. The dark grey shaded area in Figure 2.14d corresponds to the
Chapter 2. Methods 54
syn-rift sediment load. The way the sediment load is accommodated (light grey shaded
area) is determined by the exural rigidity of the basement. The T
e
s shown in Figure
2.14e were calculated from the YSE and using the thermal structure of the margin shown
in Figure 2.14b. For longer periods of time, the T
e
s are averaged over time increments of
1 m.y. The total sediment thickness for a given interval of time is the sum of the load and
the exural deformation. By repeating this process over a sequence of time increments a
synthetic stratigraphic basin prole can be generated.
The stratigraphic section shown in Figure 2.15 resembles the steers head basin
model of Watts et al. [1982]. In this model, sediments build up vertically by aggradation
in a perfectly symmetric basin where the subsidence decreases exponentially with time.
As the lithosphere cools and strengthens, during post-rift thermal relaxation, the base-
ment exural deformation associated with the deposition of younger strata will become
progressive longer wavelength (see Equation 2.11). As a result, the basin widens and the
younger sediment units will onlap onto older strata, or the basement, at the basin edges,
generating this typical steers head basin shape (see also Figure 2.16).
The subsidence histories and stratigraphic patterns for dierent models of isostasy,
rift durations and thermal models have been thoroughly evaluated for this type of basin
geometry (Watts et al. [1982]; Cochran [1983]; White & McKenzie [1988]). Therefore,
the steers head basin is a useful reference case to test the results from the modelling
approach proposed here.
In relation to previous studies, the main innovation introduced by the Mrift code
resides in the way the T
e
s are calculated. In the model of Watts et al. [1982], for exam-
ple, the T
e
s are computed as a function of age, and correspond to the depth of a given
isotherm, similarly to that commonly assumed to model the strength of oceanic litho-
sphere (e.g. Watts [1978]; Bodine et al. [1981]). However, as discussed in Chapter 1, the
values of T
e
estimated at rift basins and passive continental margins do not show a simple
correlation with the age of the lithosphere at the time of loading. Instead, it appears
that over stretched continental lithosphere, the evolution of the basement rigidity follows
a more complex pattern, which depends on non-linear relationships between the thermal
gradient, the lithostatic pressure, and the deformation imposed by the loads (Lowry &
Smith [1995]; Burov & Diament [1992]). These relationships can be parameterised via an
YSE representation of the failure properties of the rocks, as described in Section 2.3.2.
Figure 2.15 shows the evolution of the strain rates (blue lines) and the T
e
s (red lines)
through time at four dierent locations of the basin prole. During stretching, the strain
rates are high (between 10
16
and 10
14
), dropping to the background value of 10
18
at the end of the rifting event. This causes a sudden decrease in the calculated T
e
, as
depicted in locations B, C and D (see Figure 2.13 for the relationship between the strain
Chapter 2. Methods 55
Figure 2.14: Temperature structure, subsidence, T
e
and exure at rift basins. a) Stretch-
ing factors, s, used to generated the basin prole. b) Temperature structure at the end
of rifting, calculated from the s shown in (a), and assuming an instantaneous rifting
event. The eects of lateral heat ow and radioactive heat production in the crust have
not been taken into account. The model parameters are shown in Table 2.1. c) Predicted
margin subsidence at n
2
m.y. since rifting, where n = 0, 1 ,2 ,3 ,4... 13. d) Sedimentary
load corresponding to the syn-rift stage (dark grey shading), dened between the initial
subsidence (yellow area in c) and the paleowater depths (assumed at the pre-deformation
surface), and calculated exural deformation (light grey shading), from the T
e
structure
shown in (e). e) T
e
s calculated at the end of rifting from the thermal structure of the
margin and YSE considerations (see text).
Chapter 2. Methods 56
Figure 2.15: The steers head basin: T
e
s, strain rates and predicted stratigraphy.
rates, the T
e
s and the YSE).
For the case presented here, a minimum T
e
of 10 km is obtained in the centre of the
basin (Location C) at the end of rifting. As it cools, the lithosphere progressively recovers
its strength, in agreement with the predictions from dynamical models of rifting (Burov
& Poliakov [2001]). However, the T
e
s do not increase in a continuous way. Instead, some
major steps are observed, which are due to the mechanical coupling between distinct
rheological layers as the thermal gradients are attenuated (e.g. at 155 and 135 Ma in
locations C and D in Figure 2.15). At the basin edges, where the thermal structure has
not been signicantly perturbed (Figure 2.14b), the eects of stretching predominate,
and the T
e
s remain relatively constant between the end of the stretching event and the
Present (Location B).
Variations to this reference case, in terms of the computed stratigraphy and T
e
evo-
lution, are shown in Figure 2.16; the reference model of Figure 2.15 is shown is Figure
2.16a. In all the examples the rift occurs instantaneously, and zero paleowater depths are
assumed. When lateral heat ow is included in the model, the anks of the basin will be
uplifted and the post-rift subsidence in the centre of the basin increases, as depicted in
Figure 2.9b. This results in the thinning of the post-rift sediment units near the basin
edges and their thickenning towards the centre (Figure 2.16b).
In Figure 2.16c, the thermal structure of the lithosphere was calculated taking into
account the combined eects of lateral heat ow and radiogenic heat production. As
depicted in Figure 2.13c, including the amount heat produced by the decay of radioac-
tive elements in the crust results in the weakening of the lithosphere. This is observed
particularly near the basin anks, where the crust remains thick (e.g. Location B). As a
consequence, the width of the basin is reduced, with the exural bulge moving inwards by
approximately 150 km and forming a prominent rift ank, and the thickness of the syn-
Chapter 2. Methods 57
and post-rift units increases over the basin slope and anks. Furthermore, prominent
onlaps of the early post-rift units onto the syn-rift sediments are observed in the region
of the slope. In the centre of the basin, where the thickness of the crust has been reduced
to approximately 10 km (Location C; see Figure 2.15 for reference), the calculated T
e
s
through time are similar to those obtained in the previous cases (Figure 2.13a and b).
Figures 2.16d and 2.16e show variations to the model presented in Figure 2.16c. In
Figure 2.16d, a zero lithospheric strength was assumed during rifting (T
e
= 0; Airy rift).
This results in a narrower and deeper syn-rift basin, and the onlaping of the post-rift
units directly onto the basement. As the lithosphere recovers its rigidity, the onlaps
migrate from the slope to the basin anks ( 50 m.y. after rifting has ceased). Similar
stratigraphic patterns have been previously attributed to the eects of depth depending
stretching or high amplitude sea-level variations (White & McKenzie [1988] and references
therein). The lower post-rift T
e
s calculated in this model, in relation to the model of
Figure 2.16c, are due to the greater basement curvatures generated during rifting.
In the model of Figure 2.16e, a mac granulite lower crust rheology was utilized (see
Table 2.2). In relation to the reference quartz-anorthosite lower crust used in the other
models of Figure 2.16, this represents an increase in the T
e
of unstretched continental
lithosphere of 9 km (Figure 2.13a). Following rifting, however, the T
e
s in upper slope
and basin anks (Location B) can be greater by as much as 15 km, in relation to the
model Figure 2.16c. The higher T
e
values in this region are caused by the mechanical
coupling between the upper-middle crustal layer and the lower crust and mantle, which
occurs approximately 30 m.y. after the end of rifting (Location B in Figure 2.16e). The
main consequence of the increasing the basement exural rigidity is the increase in the
basin width, with the exural bulge being shifted by more than 50 km away from the
centre of the basin (in relation to the model of Figure 2.16c).
2.3.5 Subsidence, Paleowater Depths and Stratigraphy at Passive Con-
tinental Margins
Passive continental margins form by crust and lithosphere stretching, followed by conti-
nental rupture and the formation of new ocean basins. Their morphology is characterized
by a shallow continental shelf (up to 250 m depth at the latitude of the West Iberia and
Newfoundland margins), a slope, making the transition between the continental shelf
and the deep oshore, where gradients can reach up to 20%, and a gently oceanward
dipping continental rise, which can extend for more than 100 km over extended conti-
nental, transitional and oceanic crust, and where depths are generally greater than 3000
m (Figure 2.17a; see also Figure 1.2). The assumption of zero paleowater depths, or the
Chapter 2. Methods 58
Figure 2.16: The steers head basin: predicted stratigraphy (detail on the basin anks)
and calculated T
e
s for dierent thermal and rheological modelling settings. In all cases
an instantaneous rift event is assumed and the paleowater depths correspond to the pre-
deformation surface (i.e. zero paleowater depths). The reference model in a) is that of
Figure 2.15. In b), the eects of lateral heat ow have been included. In c) both lateral
heat ow and radiogenic heat production in the crust are accounted for. d) and e) are
variations to the model in c), where zero strength during rifting , i.e. T
e
= 0, or a strong
lower crust rheology (mac granulite; see Table 2.2) was assumed, respectively. The the
T
e
-strain rate graphics correspond to the locations B (upper slope) and C (basin centre)
of Figure 2.15.
Chapter 2. Methods 59
Figure 2.17: Synthetic passive continental margin transect: a) Stretching factors (s);
b) Subsidence; c) Paleowater depths. The subsidence was calculated for an instantaneous
rifting event and accounting for both vertical and lateral heat ow at n
2
m.y. since rifting,
where n = 0, 1 ,2 ,3 ,4... 13. The paleobathymetries are imaged from an hypothetical
present day bathymetry (thick blue line) and calculated in proportion to the amount of
syn-rift margin subsidence (see text).
pre-deformation surface, is therefore not valid to model the stratigraphy at these settings.
The paleobathymetries at passive continental margins are, in most cases, largely un-
known, or estimated with a large degree of uncertainty, particularly in the distal margin
(i.e. outer slope and rise). First, because most commercial drilling in the past focused
along the continental shelf and onshore basins, and DSDP and ODP drill sites are sparse
and often located over basement highs, where the stratigraphic sections are incomplete.
And second, because the gravity and current driven (e.g. turbidites, debris ows and
slides, contourites, etc), and the pelagic/hemi-pelagic deposits, which dominate the sed-
imentation in continental slope and rise environments (e.g. Winterer [1980]), are poor in
fossil benthic communities, or other sensitive paleowater depth indicators.
Figure 2.17a and 2.17b show the stretching factors (s) and predicted subsidence
along a synthetic passive continental margin transect, respectively. The prole comprises
two narrow basin depressions bordering the continent (up to 50 km width), a relatively
steep slope, and a wide continental rise region. The subsidence was calculated for an
Chapter 2. Methods 60
instantaneous stretching event and accounting for both vertical and lateral heat ow. As
discussed for the steers head basin, this results in uplifted rift anks ( 300 m) and an
excess in the amount of subsidence near the base of the slope (Figure 2.17b).
In Figure 2.17c, two sets of paleowater depths are represented, in black and red. These
have been created from an hypothetical present day bathymetry (thick blue line) and
taking into account the amount of post-rift subsidence, and will be used to test the eects
of the paleobathymetry on the predicted margin stratigraphy. The paleobathymetries
depicted in black assume that at the end of rifting the paleowater depths were 55% of
the present day bathymetry, whilst those in red were computed assuming 70%. Maximum
dierences between the two sets reach 300 m in the continental slope and 500 m in the
continental rise immediately after rifting has ceased, and progressively decrease to zero
at the present day.
These dierences are well within the uncertainties in the paleowater depths estimated
for the continental slope and rise of passive continental margins (commonly between a
few hundred metres and up to 2000 m) using calcite compensation depths (CCD) in the
geological past and dated erosional surfaces (e.g. Montadert et al. [1979]; Moullade et al.
[1988]; Kusznir et al. [2004]), or by comparing the subsidence predicted from the thermal
models with basement depths recovered from sediment decompaction and backstripping
(e.g. Royden & Keen [1980]; Kusznir et al. [1995]; Stewart et al. [2000]).
The eects of varying the paleowater depths on the predicted margin stratigraphy
are shown in Figure 2.18. Over the shelf and upper continental slope the dierences are
minimal (see inset boxes). Towards the lower continental slope and rise, however, the
syn-rift sediment loads calculated with the deeper paleobathymetries (red lines in Figure
2.17) are considerably smaller than when the shallower ones are used. This results in the
progressive reduction of the overall sediment thicknesses oceanward, and the onlap of the
syn-rift and the early post-rift units onto the basement (Figure 2.18b). A similar pattern
is observed in a number of Atlantic-type margins (e.g Pickup et al. [1996]; Lau et al.
[2006b]; Funk et al. [2004]; Chian et al. [1995]; dAcremont et al. [2005]; Contrucci et al.
[2004]). Since the sediment loads for each time-step are determined by the dierence in
the variations of the subsidence and the paleobathymetries, the post-rift sediment units
in the deep oshore are thicker when the deeper paleobathymetries are used.
Another interesting feature of the models presented in Figure 2.18 are the bowl-shaped
post-rift inlls draping the shallow basins that form near the continent. These result, in
part, due to the heat owing laterally away from the basin centres and uplifting their
anks. Similar geometries are observed in rift basins, where they have been associated
with immediate post-rift sedimentation (Prosser [1993]), as suggested in the models.
The calculated T
e
s along this hypothetical margin transect (red lines in the T
e
-strain
Chapter 2. Methods 61
Figure 2.18: Predicted stratigraphy, T
e
s and strain rates in a synthetic passive conti-
nental margin transect. The s and subsidence history are shown in Figure 2.17a and
2.17b, respectively. In a) and b) the stratigraphy is calculated using the shallow and
deep paleobathymetries shown in Figure 2.18, respectively. In the T
e
-strain rate graphics
the dotted lines are obtained without the contributions from lateral heat ow.
rate graphics shown in Figure 2.18) are similar for the shallow and deep paleobathymetry
cases, indicating that the dierences in the basement curvature caused by the variations
in the size of the sediment loads are not relevant for the overall strength of the lithosphere.
Over the shelf and upper slope, locations A and B in Figure 2.18, respectively, the T
e
s
vary between 15 and 25 km, and remain relatively low throughout the margins post-rift
evolution. This is due to the eects of radiogenic heat production in the upper-middle
crust, as described for the steers head basin (see Figure 2.16c). In the lower continental
slope and rise, locations C and D in Figure 2.18, respectively, T
e
increases following
rifting, as a result of cooling and lithospheric thickening, where the step-like increments
are caused by mechanical coupling between strong crust and mantle layers.
The dashed red lines in the T
e
-strain rate graphics correspond to the T
e
s calculated
when lateral heat ow is not included in the thermal model. These show a signicant
increase in the rigidity of the basement along the shelf and slope which, during post-rift
times reaches up to 38 km. As discussed for the steers head basin, this would result in
thinner post-rift sediment units and the widening of the basin anks.
Chapter 2. Methods 62
2.3.6 Multi-Stage Continental Rifting
All the models described above assume a single, instantaneous stretching event to describe
the subsidence history and predict the basins stratigraphy. As discussed in Section 2.3.1,
however, the evolution of some Atlantic-type margins, namely that of the West Iberia and
Newfoundland conjugate margins, is characterized by a sequence of rift pulses intercalated
with periods of thermal relaxation (e.g. Tankard & Welsink [1987]; Wilson et al. [1989];
Hiscott et al. [1990]; Driscoll et al. [1992]; Wilson et al. [1996]).
In Figure 2.19, we compare the stratigraphy and the calculated T
e
s for four dierent
scenarios: a) a single, instantaneous rifting episode; b) a single, nite duration (25 m.y.)
rifting episode; c) two nite duration rifting episodes (25 m.y. and 15 m.y., respectively);
and d), as in c) but assuming zero lithospheric strength (T
e
= 0) during the rifting events.
In all the cases, the eects of lateral heat ow and radiogenic heat production have been
taken into account. The total amount of stretching in the two-rift scenarios is the same as
in the single rift examples (see Figure 2.17a), and it was assumed that the narrow basins
under the continental shelf formed during the rst rift episode, whilst the second rift
episode was responsible for the development of the continental slope and deep margin.
In relation to the instantaneous rift example of Figure 2.19a, the nite rift case
shown in Figure 2.19b predicts thicker syn-rift and thinner post-rift sedimentary pack-
ages. This results from the increase in the syn-rift/post-rift subsidence ratio for longer
rifting periods, as discussed in Section 2.3.1. Another important dierence between the
instantaneous and nite rift scenarios relates to the way the T
e
s evolve through time in
the continental slope and deep oshore (locations C and D, respectively), which depends
on the complex interaction between dierent factors, such as the thinning and heating of
the lithosphere during rifting, strong lateral heat ow during and immediately following
rifting, the strain rates, and the eects of mechanical coupling between strong rheologic
layers both during the stretching and the post-rift thermal relaxation stages.
In Site C of Figure 2.19b, for example, an initial decrease in value of T
e
is observed
(thinning/heating), followed by a stepwise increment (crust-mantle mechanical coupling)
and a small continued increase (cooling due to lateral heat ow), followed by an abrupt
decrease at the end of rifting (strain rate drops and mechanical decoupling), and a pro-
gressive post-rift increase (thermal relaxation), interrupted by two stepwise increments
(mechanical coupling between lower crust and mantle and between upper-middle crust
and lower crust-mantle, respectively). The average rigidity of the lithosphere is greater
in the nite rift case, particularly during the syn-rift stages, when the strain rates are
higher. This results in the broadening of the syn-rift sedimentary units which drape the
continental rise.
Chapter 2. Methods 63
Figure 2.19: Finite rifting and multi-stage continental rifting passive margin scenarios:
predicted stratigraphy, T
e
s and strain rates. a) Single, instantaneous, rifting episode. b)
Single, nite duration (25 m.y.) rifting episode. c) Two nite duration (25 m.y. and 15
m.y.) rifting episodes; d) As in c) but assuming zero lithospheric strength during rifting.
In all the cases, both the eects of lateral heat ow and radiogenic heat production have
been taken into account. The total amount of extension along the basin is the same
between the single and the two-rift scenarios (see total s in Figure 2.17a). The yellow
and grey shaded areas highlight the syn-rift sediment packages deposited during the rst
and second rift events, respectively. The syn-rift periods and ages of the post-rift horizons
in relation to the end of the last rifting event are indicated, as in gures 2.9 and 2.10.
Chapter 2. Methods 64
The two-rift scenario of Figure 2.19c introduces a number of new features in the
stratigraphic framework of the modelled passive margin transect. In the region of the
shelf, an angular unconformity is observed between the sediment units deposited following
the rst rift event and the syn-rift units of the second rift episode (grey shaded). Towards
the continent, the margin is bordered by a prominent exural bulge (up to 500 m high; see
inset), which forms in response to the deposition of the thick (> 5 km) syn-rift packages
in the continental slope and rise during the second rift event and strong lateral heat ow.
The onlaping of the syn-rift and early post-rifts sequences in the distal margin results
from using deeper paleowater dephs, as in Figure 2.18b.
In average, the T
e
s calculated in the two-rift scenario are higher in the region of
the shelf and upper continental slope (locations A and B, respectively) and signicantly
lower (up to 20 km) in the distal margin (locations C and D). In the rst case, the overall
higher strain rates through time is the predominant eect, whilst in the distal margin
the lithosphere is still recovering its strength following the latest, more recent, stretching
event.
In the case of no strength during rifting (i.e. Airy rift; Figure 2.19d), the sedimen-
tary basins under the continental shelf become narrower and deeper, and a prominent
basement relief separates the shelf from the continental slope. Along the continental
slope and rise, both the syn- and post-rift sedimentary packages are thicker than in the
previous scenario, as a result of the lower lithosphere rigidity. As discussed for the steers
head basin (see Figure 2.16c-d), the lower post-rift T
e
s are determined by the higher
basement curvatures generated during rifting.
2.3.7 Modelling Strategy and Aims
In this thesis, the Mrift code will be used to model the subsidence/uplift history, the
thermal and mechanical structure, and the broad stratigraphy of the WIM (in Chapter
6). Proles orthogonal to the margins strike, where there are good constraints on the
crustal structure, and thus on the amount of stretching, between the continental shelf and
the distal margin (from seismic wide angle and/or refraction seismic data) will preferably
be utilized. The age and duration of the main rifting events that aected the WIM are
reasonably well constrained, from the stratigraphy, subsidence analysis and the age of
the seaoor spreading magnetic anomalies oshore (Pinheiro et al. [1996]; Wilson et al.
[1996, 2001] and references therein; see Chapter 3 for discussion). The results from the 1-
D subsidence modelling in a number of commercial wells distributed along the Portuguese
continental shelf, presented in Chapter 4 of this thesis, also reveal important aspects of
the margins thermal and subsidence evolution.
Chapter 2. Methods 65
The main aims of the thermo-mechanical rheological modelling approach developed
in this study are:
1. To model and compare the thermal structure and the subsidence/uplift history in
dierent segments of the WIM.
2. To investigate how the mechanical structure of the extended continental litho-
sphere evolved through time, both during and following the main rifting events.
Furthermore, by comparing the T
e
s calculated from the thermal structure of the
lithosphere and YSE considerations, with those estimated from combined back-
stripping and gravity modelling techniques (POGM), constrains may be placed on
the rheologies of the crust and mantle, and on the contributions from dierent
processes, such as normal faulting, exure, crust-mantle coupling/decoupling and
mantle serpentinization, to long-term strength of the lithosphere.
3. To model the broad stratigraphy of the WIM. By comparing the modelling results
to the observations further constrains may be placed on the long-term strength of
the lithosphere, as well as on the margin paleowater depths.

You might also like