You are on page 1of 39

Geophys. J . R . astr. SOC.

(1981) 65, 291-329

Foreland basins

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


Christopher Beaumont Oceanography Department, Dalhousie University,
Halifax, Nova Scoria B3H 451, Canada

Received 1980 April 18

Summary. The characteristic pattern of the subsidence, deformation, uplift


and erosion of foreland basins, or exogeosynclines, has long been recognized
by geologists as an integral part of the geosynclinal cycle yet only descriptive
models have been proposed. This evolution is quantified in terms of a model
of regional isostatic adjustment of the lithosphere under the mass load of the
adjacent migrating fold-thrust mountain belt. It is shown that the foredeep,
which received the clastic wedge of detritus from the core of the orogen and
the fold-thrust belt, owes its existence to downward flexure of the litho-
sphere by the fold-thrust belt. Additional depression occurs in response t o
the infilling sediment. The scale of the basin therefore reflects the degree of
shortening in the orogen and the rheological properties of the underlying
lithosphere. Subsidence reflects mass that accretes at a faster rate than mass
wasting from the orogen, whereas regional uplift and erosion of the basin is
attributed to the large-scale erosion of the fold-thrust belt that has been
depressing the lithosphere. There is a complete coupling between the
evolution of the foreland basin and its adjacent mountain belt through the
lithospheric flexure. Lateral migration of the fold-thrust belt accounts for the
progressive overriding and disruption of some foreland basins.
The proposed model is tested by comparison with the structure of sections
across the Alberta Foreland Basin of Western Canada. It is shown that a
model in which mass loads continuously advance on to the craton in a series
of pulses between the Upper Jurassic and the Eocene is consistent with
observations. No successful models were found when the lithosphere was
modelled as a uniform thin elastic plate. However, when the plate was
allowed to relax stress according to a viscoelastic (Maxwell) rheology good
agreement between observed and theoretical cross-sections was obtained.
These results suggest that suitable values for the flexural rigidity, D,and
relaxation time constant, 7, are lOZ5Nmand 27.5 Myr. Models with signifi-
cantly different 0--7 parameter values failed to predict satisfactory results.
Predictions of basin erosion, coal and shale compaction, and thermal
metamorphism of coal are shown to be in agreement with observations. A
10
292 C Beaumont
significant result is that no change in geothermal gradient, with respect to
that presently observed, is needed to explain coalification. This result
confirms the view that subsidence of the basin was not thermally controlled,
unlike Atlantic type marginal basins, and explains the high value of the
flexural rigidity as that appropriate to an old, cool, thick lithospheric plate.

Introduction

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


Classical works have identified foreland basins as an integral part of the geosynclinal cycle,
yet have failed to provide more than descriptive models of their evolution. (See Aubouin
1965 for a comprehensive review of classical ideas on geosynclinal evolution.) The value of a
quantitative model for foreland basins lies partly in the recently reiterated observation (Price
1973; Laubscher 1978) that the foreland basin is mechanically coupled to the adjacent
orogen. As Price (1973) noted, it is the regional isostatic subsidence beneath the supracrustal
load of the fold-thrust belt that generates the foredeep in which sediments collect. Basin
subsidence, stored in the stratigraphic record, contains a measure of the history of the
orogen. The inference is that the timing and magnitude of orogenic events may be estimated
from a study of the basin once a deterministic model for its evolution has been established.
The intent of t h s paper is to place Price’s isostatic hypothesis in a detailed quantitative
framework, one in which the evolution of foreland basins in response to loads superimposed
on a thin plate flexural model of the lithosphere is investigated. It is further shown that, in
the case of the Alberta Foreland Basin, stratigraphy places severe constraints on the model
parameters, thereby providing evidence on the macro-rheological properties of the Earth’s
lithosphere and on mass movements in the Rocky Mountain fold-thrust belt.
Motivation for this study stemmed from a desire to characterize the lithospheric rheology
of shields, for these areas are believed to have the oldest, thickest and coldest thermal
boundary layer.

Definitions and a working hypothesis


Terminology for the structural components of the geosynclinal-orogenic cycle is often
confusing, particularly when plate tectonic and classical geologic descriptions are super-
imposed. The term foreland basin is used here in a manner that is descriptively synonymous
with exogeosyncline (Kay 1951), marginal basin (Krumbein & Sloss 1963), foredeep
(Aubouin 1965), foreland and foredeep trough (Price 1973), peripheral and retroarc
foreland basin (Dickinson 1974) and the marginal down-flexure of Russian authors, although
the genetic explanations provided by these authors are not necessarily adopted.
The scope of the discussion is limited to the depositional and erosional history of the
basin that is directly controlled by events in the orogenic belt; that is, the flysch and molasse
phases. Deposition of any underlying miogeosynclinal sequence takes place on a rifted
margin in a lithospheric plate spreading environment. Here, the concern is the response of
the continental margin to the accretionary events associated with plate convergence and
collision (Dewey & Bird 1970; Dickson 1974) and intraplate crustal shortening.
The working hypothesis is that foreland basins form at the site of downwardly flexed
lithosphere, and that t h s downward flexure is in response to passive loading by supra-
lithospheric mass loads superimposed during formation of the fold-thrust belt. This
hypothesis has more widespread application than foreland basins. Whenever there is lateral
transfer of a rock mass over an adjacent part of the lithosphere which responds by flexure,
a coupled trough is created in which sediments can accumulate. The form of this trough will
Foreland basins 293
depend on the timing and amount of the mass movements, the properties of the underlying
lithosphere, the amount of sedimentary infill, and the role that other processes, for
example, subduction, underthrusting and thermal cooling, play in warping the lithosphere
at the site of the trough. The influence of each of these factors, in regard to the formation
of foreland basins, can be estimated from descriptive models of the tectonic and thermal
environments in which foreland basins form.

Geological and plate tectonic framework

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


Geosynclinal evolution, when reinterpreted in terms of plate tectonics, may be considered
to occur in four basic distinct yet connected phases: (1) the geosynclinal phase of plate
separation and the formation of a passive continental margin with sedimentation in the
miogeosynclind-eugeosynclinal couple; (2) the early orogenic phase characterized by the
initiation of subduction and the underthrusting of the continent by an oceanic plate with
associated thermal updoming and thrusting in the zone between the craton and the
continental margin (Fig. la); (3) the late orogenic phase of ocean closure marked by
continent-island arc collision, or continent-continent collision (Fig. 1 b), and (4) when
suturing has ceased, the uplift and erosion phase prior to normal faulting and rifting which
mark renewed plate separation. Within this framework the contribution from each phase
varies greatly among orogenic belts or even along the length of a single orogen. As Dewey
& Bird (1970) emphasize, phases 2 and 3 may be considered archetypes of mountain
building. Phase 2 generates the island arclcordilleran type mountain belt, whereas phase 3
generates collision orogens of the Alpine-Himalayan type. They are considered successive
phases in this discussion because many orogens experience plate collision as a sequel to
subduction. No matter what the detailed form of phases 2 and 3, Dewey & Bird (1970)

FORE^^^ CORE FORELAND RETROARC


TRENCH BASIN ZONE FOLD-THRUST FORELAND

MIOGEOSYNCLINE

I
5 0 0 km

PERIPHERAL FOLD SUTURE


FORELAND THRUST ZONE
RASIN BELT I

MIOGEOSYNCLINE

Figure 1. Conceptual illustration of the formation of retroarc and peripheral foreland basins by the
flexural bending of cratonic lithosphere. Note that in each case the foreland basin is deposited uncon-
formably on an underlying Atlantic type margin. Retroarc basins are formed next to Cordilleran-type
orogens, whereas peripheral foreland basins occur as a result of continental collision.
294 C. Beaumont
envisaged the molasse basin (molasse facies of the foreland basin) to form on continental
lithosphere, either in a back-arc environment if phase 2 predominates, or, in phase 3, super-
imposed on the miogeosynclinal sequence of the plate whose oceanic limb was subducted.
Models of sedimentation in geosynclinal environments of Dickinson (1974, 1976)
reinforce the important tectonic distinction between the two classes of foreland basin
identified by Dewey & Bird (1970). Dickinson terms the cordilleran type retroarc foreland
basins. Those formed during suturing in continent-continent collisions (and presumably
continent-island arc collisions) when a rifted continental margin encounters the main sub-

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


suction zone are termed peripheral foreland basins (Fig. Ib).
The general stratigraphic records of the two subclasses of foreland basins are sufficiently
similar that they are difficult to classify without knowledge of the tectonic environment in
which they were created. Both types record similar transitions in sedimentation, the most
important of which marks the end of phase 1, the miogeosynclinal sequence. This is a wedge
of mainly carbonate-shale facies with underlying clastics that accumulate in littoral to neritic
environments, relatively shallow palaeowater depths, on a subsiding continental shelf. The
degree to which isostatic conditions prevail at the end of phase 1 depends on the age of the
margin at that time (Sleep 1971; Turcotte & Ahern 1977; Keen 1979). A margin that is in
excess of 200 Myr old can be assumed to be in isostatic equilibrium because the basin is
floored by continental, or modified continental, lithosphere that in later stages cools to
thermal equilibrium with an exponential time decay of approximately 50 Myr (Sleep 1971 ;
Parsons & Sclater 1977). Thus, a miogeosynclinal sequence that is old at the beginning of
the next phase is underlain by lithosphere which is at least as old, cool, and thick as the
oldest oceanic lithosphere created during phase 1. This is because it has had the whole of
phase 1 to cool.
The onset of phases 2 or 3, indicating the development of, or encounter with, an off-
shore subduction zone is marked by a radical change in sediment type and provenance.
The source of the marine clastic wedge of flysch is initially offshore at the site of the rising
orogenic dome, for the retroarc basin, or the approaching island arc or continent, for the
peripheral basin. Later an intervening highland source of the miogeosynclinal sediments
themselves dominates and sheds clastic detritus into basins of either type. Thus, in the upper
part of the sedimentary sequence, flysch gives way to thick molasse deposits typically con-
taining sandstone and shale facies with redbeds and paralic coals.
Evolution of the basin is diachronous both in sedimentation and in progressive disruption.
Disruption takes the form of a migrating fold-thrust belt that incorporates foreland basin
sediments with older sedimentary and even crystalline basement and transports them basin-
ward along listric thrust faults as nappes and thrust slices (Bally, Gordy & Stewart 1966;
Price & Mountjoy 1970; Hatcher 1972; Coney 1973; Wheeler et al. 1974). The whole
sedimentary wedge is seen to undergo lateral shortening of up to a factor of two while the
locus of the basin axis migrates further on to the craton. This pattern of progressive
incorporation of the foreland and miogeosynclinal sedimentary wedge into an allochthonous
fold-thrust belt is the most characteristic feature of foreland basins and marks the com-
pressional phases, 2 and 3 , of the geosynclinal cycle. By the culmination of these phases the
whole basin may have been extensively deformed, or remain relatively undisturbed,
depending on the extent of compression and the degree to which stress caused dkcollement
within the basin. If both compressional phases occur, the orogen has a final form with two
foreland basins that exhibit apparent mirror symmetry with respect to the suture.
During phase 4 all structural components of the orogen undergo erosion with greatest
uplift and unroofing in the core and fold-thrust belts. Erosion is so pervasive that the
combined effect of uplift of the basin to form a plateau and erosion of the highland zones
Foreland basins 29 5
M
*

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


(C) ?
L

Figure 2. Possible models for flexure of forelands in a retroarc environment. (a) Mutual lithospheric over-
thrusting and underthrusting in which the lithospheres retain sufficient integrity to exert bending stresses,
M . (b) Flexure induced by the loading, L , of stacked thrust sheets of the fold-thrust belt and other
allochthonous terranes. (c) Flexure induced by the Same mechanism as (b) but where the underlying plate
extends continuously across the orogen to the neighbourhood of the subduction zone. In (b) the over-
thrust plate is considered to have been decoupled beneath the orogen either by thermal thinning or strike-
slip motion. (d) Formation of a crush zone during continental collision in which the plates are driven
together in a high energy environment causing tectonic thickening, T , of at least the crustal part of the
downwarped lithosphere.

eventually removes the contrasting topographic expressions of the basin and adjacent
mountains .
The pattern of lithospheric deformation beneath the core zone, foreland fold-thrust belt,
and foreland basin is illustrated in Fig. 1. ‘Cratonic’ is used to emphasize that the lithosphere
is stable, old and cold. At least two hypotheses exist for downwarping of the lithosphere.
One suggests that partial subduction, or underthrusting, of the lithosphere occurs in response
to horizontal stresses across the whole of the orogenic belt. In consequence, the cratonic
lithosphere is bent into a shape that is analogous to the shape of oceanic lithosphere at an
oceanic trench. The second hypothesis, and the one investigated in this paper, is that down-
warping is an isostatic response to the superimposed loads of the core zone and foreland
fold-thrust belt. These hypotheses differ in one important aspect, that of bending moments
applied by the underthrusting/overthrusting inherent in the partial subduction model. When
this occurs and the plates retain their structural integrity, as shown in the overstated diagram
(Fig. 2a), each plate exerts a bending stress, M , on the other. It is this force that warps the
lithosphere. In the second hypothesis (Fig. 2b) no such bending stress exists and the loads,
L , are deemed to be responsible for lithospheric flexure.
The loading hypothesis must be correct if no plate overthrusting and underthrusting
occurs. In this case, however, there is no net shortening at depth and an independent
mechanism must be found to peel the surface veneer from the plate and transport it
laterally, thereby achieving the observed supra-crustal shortening. Elliot (1976a) has
modelled the peeling of individual thrust sheets as a response to gravitational forces due to
palaeoslopes. Such slopes may reflect topography of a hot uplifted orogenic core zone.
29 6 C Beaumont
Thrust sheets travel down-slope with respect to surface topography but may move up-slope
on their basement detachment surface. This explanation, based on suggestions by Price &
Mountjoy (1970), probably does not apply to the Columbian Orogen of the Canadian
Cordillera because stretching of the core zone is insufficient to balance shortening of the
fold-thrust belt (Price, private communication). The loading hypothesis is also correct if the
overthrust plate cannot exert a bending stress because it has lost its structural integrity.
The model examined in detail in this paper (Fig. 2c) assumes that the bendingstress,M,
is neghgible and that any decoupling of the lithosphere (Fig. 2b) is sufficiently distant from

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


the fold-thrust belt and foreland basin as to have no influence on their formation, a point
that will be returned to later. A uniform plate model is therefore employed. It is to be
expected that a range of plate interactions occurs during the formation of peripheral and
retroarc foreland basins. The proposed model may not be appropriate for continental suture
when two plates are actively driven together to form a crush zone as in the Zagros thrust
belt. This interaction has been interpreted in terms of complete failure and thickening of at
least the crustal section of the continental plate (Bird 1978) (Fig. 2d). However, in low
energy collisions, where large-scale failure or underthrusting cannot be demonstrated, the
proposed model may be the most appropriate in explaining largely undisturbed peripheral
foreland basins.
Type examples of foreland basins are (Coney 1973; Dickinson 1974): the Ganga and
Indus Basins, formed south of' the suture between the Indian Plate and Euro-Asian Plate;
the Alpine and Pyrrenean Forelands, a complex series of foreland basins that resulted from
the accretion of several microcontinents to southern Europe; the Andean Foreland, where
the Andean fold-thrust belt is migrating eastward over the South American Plate; the Alberta
Basin, thought to have formed during accretion of island arc terranes to the North American
craton; and the Appalachian Foreland, as exemplified by the least disturbed sedimentary
sequences of New York State, and the Allegheny and Cumberland plateaus. The oldest relict
examples, from which the foreland basins have largely been stripped by erosion, may be the
suture zones of the Precambrian shields.
Retroarc foreland basins are the most likely to conform to the proposed model because
they form in an environment that is away from plate edge subduction and in relatively low
energy environments. This is why a cordilleran retroarc basin, the Alberta Foreland Basin,
was chosen for detailed comparison.

The flexural model


G E N E R A L PROPERTIES A N D CONSTRAINTS

The proposed model calculates the response of the lithosphere to loads placed on its surface
and can be described by reference to Fig. 3. The flexural response of the thin plate litho-
sphere to loading (Fig. 3a) is predicted using the Green function-convotution method
described by Beaumont (1978). The undeformed initial surface of the model is assumed to
be planar and is termed the depositional baseline. In its most simple form this surface
corresponds to a fixed sea-level. The loads, which correspond to the fold-thrust belt, are
termed specified loads to distinguish them from sedimentary loads. They are described by
their mass in terms of a load of height L , density pL that is assumed to be uniform over a
grid cell of size 50 x 50 km. The total grid covers an area of 1000 x 1000 km.
Solution to the flexure problem provides the downward displacement d(x, t ) of the
surface of the plate, with respect to the baseline, in compensation of the specified loads
(Fig. 3b). This displacement comprises a downwarped empty basin adjacent to the load and
a small amplitude upwarped peripheral area. Sediment, density ps, is assumed to fill the
Foreland basins 297
-50km-

Depositional Baseline P

Xi

(b)
n
.I 1 1
d(x,t) Empty Basin 1
-
Peripheral
Upworp

t Erosion of Peripheral
Sedimenf Influx

rv-

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


d(5.t) Sediment Filled Basin
Density Ps
Residuol Topography

Load
(d) . i
Basin for Loads la/

Figure 3. Elements of the flexural model in which the loads of the fold-thrust belt (specified loads) induce
a foredeep that is subsequently filled with sediment to form a foreland basin (see text for details).

basin linearly throughout each timestep, and the peripheral upwarp is linearly eroded in the
same way (Fig. 3c). This part of the problem is solved by iterative solution of the Fredholm
integral equation of the second kind that describes the response of the lithosphere to sedi-
mentary influx and erosion (Beaumont 1978). Filling and erosion are assumed to occur to
the depositional baseline, unless a more complex basin model is employed (see next section).
The model does not conserve mass. It is assumed that there is always a sufficient supply
of sediment to fill the basin, an appropriate sediment budget model for foreland basins
which normally remain overfiled or nearly filled throughout their history. Moreover, con-
servation of mass is unlikely within an area of 1000 x 1000 km. For example, much of the
sediment eroded from the Rocky Mountains was subsequently transported as far away as the
Gulf Coast of the United States.
That the overall response to an assemblage of surface loads is non-linear is illustrated by
Fig. 3(d). The response to specified loads is equal to the convolution of the loads with the
Green function that describes the response of the lithosphere to a point load (Beaumont
1978) and is therefore linear. The lateral extent and depth of the basin to be filled with
sediment, however, depends on the disposition of the specified loads (compare Fig. 3(d)
with the case where loads 1 and 2 are stacked vertically). Consequently, the final basin
configuration is not simply the superposition of simple basins that form in response to
archetypal specified loads. The physical reason for this non-linearity is that specified loads
occupy space in their own basin and therefore limit the space available for sediments.
The basic difference between the basin response to specified loads that are added
simultaneously and to those added over sequential timesteps is illustrated in Fig. 3(d and e).
In Fig. 3(e) specified load 1 was added during timestep 1 . Its basin is ‘Basin step 1’. A second
specified load that was added during timestep 2 creates its own sediment wedge, ‘Basin
step 2’. The main point is that specified load 2 is superimposed on the sediment that
accumulated during step 1. Even for an elastic model the basin that form in response to an
assemblage of specified loads that were added over a sequence of timesteps (Fig. 3e) is very
different from that formed had all the specified loads been added simultaneously (Fig. 3d).
29 8 C. Beaumont
This is a consequence of the same non-linearity that arises from specified loads occupying
space in their own basin.
Specified loads can also be removed by ‘synthetic’ specified erosion, in which case there is
isostatic uplift in the foreland basin. In simple models the same iterative solution is
employed t o erode uplifted sediments to the baseline and to calculate the overall isostatic
response of the basin. The main constraint on the removal of specified loads is that synthetic
erosion cannot erode mass that is below the baseline for deposition. There is no a przon way
to predict the maximum specified load that can be removed at any given time because this is

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


a function of basin evolution. Instead, the model resets synthetic erosion to the maximum
possible if the specified amount exceeds the maximum.
Specified loads that completely subside below the depositional baseline during basin
evolution become covered with sediment as though they are part of the basin. Once covered,
specified loads cease to be protected from the erosional processes that operate in the basin
should they be uplifted above the depositional baseline at a later time. Exposed specified
loads that have not been covered by sediment are protected from the erosional processes of
the basin and can only be removed by subtracting a specified load from the grid. This
technique partitions the erosional processes in the fold-thrust belt from those of the basin.
The migration of surface loads in a foreland fold-thrust belt is controlled by the physics
of thrust sheet movement (see, e.g. Elliot 1976a, b). No claim is made to incorporate such a
physical model. Instead, the specified loads represent empirical estimates of the net mass
transfer from one grid to another for each timestep as a result of the evolution of the fold-
thrust belt. Palinspastic reconstructions of the width of the orogen before compression are
used to estimate the areal extent of the grids which are to be loaded, but the size of the
loads and their disposition is mainly constrained by the requirement that the basin con-
figuration and residual topography of the fold-thrust belt agree with observations. A crude
check is then made that the specified loads so determined agree with the available geological
constraints. The constraints could certainly be improved with more geological information
but are inevitably limited by the loss of information with erosion.
Basin formation is considered to be ‘one sided’. Only a foreland basin to the right of the
specified loads in Fig. 3 is modelled. The justification is that pre-existing topography, from
uplift in the core zone, to the left of the loads is sufficiently high to prevent the accumula-
tion of sediment in a major basin even though downward flexure of the lithosphere (Fig. 3b)
also occurs in this region. This assumption is in agreement with the available palaeotopo-
graphic information. Small-scale intermontane and successor basins do form within the
orogen but these contain insufficient sediment to have had an effect on the foreland basin.
The model lithosphere is a uniform thin plate overlying an inviscid asthenosphere.
Justification for this model is given by Beaumont (1978). Nevertheless, the model may be
inadequate because a uniform plate does not necessarily reflect the mechanical properties
of the Earth’s thermal boundary layer. This is especially the case where the thermal age of
the cooling boundary layer varies as a function of position across the orogen from old cold
thickened lithosphere beneath the foreland basin to a core zone that may have had a very
thin lithosphere during the height of plutonic and metamorphic activity. This suggests that a
model in which the mechanical thickness of the lithosphere progressively thins across the
fold-thrust belt would be preferable. That reasonable results are achieved without this
thinning reflects the lack of sensitivity of the model foreland basin to loading from the far
side of the fold-thrust belt, this region being more than a flexural wavelength from the basin.
A thinner plate in this region would localize isostatic adjustment to loads in the core zone
and further reduce the sensitivity of the basin to these loads.
The choice of a linear model with its resulting mathematical simplicity limits the
Foreland basins 299
rheological complexity for the lithosphere to linear models. Only elastic and viscoelastic
(Maxwell) rheologies are employed. A major question posed is whether a more complex
rheology than elastic infinitesimal strain theory is required in order to model the
observations. The properties of a t h n viscoelastic plate are specified by two parameters
(Beaumont 1978), the flexural rigidity, D (where D = Ed 3/12(1 - o’), E = Young’s modulus,
d = plate thickness and u = Poisson’s ratio) and the viscous relaxation time, 7, which is
infinite for a purely elastic plate. These parameters and the densities determine the space-
time evolution of the Heaviside-Green functions for the plate response to a point load

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


(Beaumont 1978). The Green functions themselves are not used in these calculations.
Instead, influence function matrices, describing the effect of a uniform load on a 50 x 50 km
grid at all other grid field points, form the core of the numerical convolution.
In summary, the fundamental model parameters are:
(1) the rheology and density parameters; D, 7,pL, ps, pA ( p A = asthenospheric density);
(2) the specified loads as a function of spatial position and time; L ( x , t).
The model predictions are:
(1) the history of deformation, as evidenced by sediment deposition and erosion of the
model foreland basin; d ( x , t);
(2) the history of the topography of the specified loads; H ( x , t).
The following parameters that are necessary for comparison with observations can be
estimated from the model predictions:
(I) present stratigraphy of the basin;
(2) palaeo-stratigraphy of the basin;
( 3 ) present topography of the fold-thrust belt;
(4) palaeo-topography of the fold-thrust belt;
(5) evolution of the fold-thrust belt;
( 6 ) free air and Bouguer gravity anonalies associated with basin formation;
(7) deformation of the crust and MohoroviZib discontinuity associated with basin forma-
tion;
(8) maturation histories for oil that remains in its source beds and coalification estimates.
The way in which most of these parameters can be compared with observations is
obvious. However, (8) warrants explanation and is discussed in a subsequent section.

Additional parameters included for the Alberta Foreland Basin model


Additional features were required in order to model the post-mid-Jurassic evolution of the
Alberta Basin. These properties were included only after it wu shown that agreement with
observations was not possible without them. The changes are, however, known to have
occurred and the failure of modelling without their .inclusion demonstrates the overall
sensitivity of sedimentary basin formation to what may be regarded as second-order effects
that perturb the basin surface.

Eustatic sea-level changes


The baseline for deposition in the simple version of the model was assumed to be a fixed
horizontal surface determined by a time invariant sea-level. That this surface had to be
changed during timestepping, thus allowing for enhanced or decreased sedimentation for a
300 C Beaumont
given mass load, is required by observational constraints. These include deposition over a
much wider basin than could be accounted for purely by flexure. A flexural rigidity that
implies a lithospheric thickness in excess of that seismologically permissible would be
necessary to produce a basin of equivalent width. Flexural rigidities of this magnitude also
failed to produce remotely acceptable results for the overall shape of the basin. Inclusion of
eustatic sea-level changes partitions the basin into three regions. The first is that adjacent to
the specified loads where the basin structure is dominated by the flexural response to
loading. The third is the most remote outer reaches of the basin where sea-level changes

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


determine the pattern of sedimentation. The second region is that intermediate to the others
at the edge of the flexural basin where load and sea-level effects are equally important.
Eustatic sea-level changes were not specified a priori. Instead, a sensitivity analysis was
carried out for the range of acceptable changes that gave a basin configuration in agreement
with observations. This sea-level curve was then compared with other independent estimates.

Sediment deposition above sea-level


The depositional environment throughout the evolution of the Alberta Basin varied from
marine (maximum outer neritic) through non-marine coastal plain. These changes reflect
-
departures of the baseline from sea-level of f 200 m (Jeletsky 1978). Underfilling was
judged to be unimportant, but an attempt has been made to allow for the extensive non-
marine coastal plain that existed during the Maestrichian and later. This sedimentation is
-
included by modifying the baseline from sea-level to a gradient of 66 m/100 km between
the specified loads and a hypothetical shoreline. The gradient was estimated from equivalent
modern depcsitional environments. Overfilling of the basin is based both on the extensive
non-marine deposition and the need for deep burial to produce the observed degree of coal
metamorphism (Hacquebard 1977).

Incomplete erosion
In the same way that sedimentation does not always occur to the depositional baseline or
sea-level, erosion in the basin is not always as efficient as that assumed in the basic model.
That foreland basins are uplifted to form raised plateaus is evidence for the lack of efficient
erosion. An erosion model that eroded to the modified depositional baseline also proved to
be excessive. Instead, a purely empirical model that ignored decreases in sea-level and
delayed erosion of 50 per cent of the uplifted sediments in any given timestep proved to be
quite acceptable. Certainly, a model based on the physics of erosion would be preferable. An
analysis of the sensitivity of model results to the percentage erosion at any given timestep
indicated that amounts from 30-50 per cent were acceptable but because erosion is a local
phenomenon no model duplicated the exact form of the observed erosional surface.
In summary, the additional model parameters are:
(1) eustatic sea-level changes S(t);
( 2 ) coastal plain depositional baseline C(x, t);
(3) basin erosion E(t).

Modelling philosophy and procedure


The intention of this paper is to demonstrate that laterally migrating loads on a thin plate
lithosphere provide an acceptable description of the evolution of foreland basins. For this
reason it was considered adequate to use mainly two-dimensional models and to compare
Foreland basins 301
predictions with cross-sections of one particular foreland basin, the Alberta Basin. No formal
inversion strategy has been employed to estimate model parameters, although it is believed
that inversion techniques could be used profitably with a three-dimensional model. The
modelling procedure was based on a parameter search-sensitivity analysis strategy. Model
parameters were partitioned into first-order variables, D , r and L ( x , t ) , that play a primary
role in the evolution of the model basin, and second-order variables, S ( t ) , C(x, t ) and E ( t ) ,
that were considered to be known for the initial parameter search. The formal parameter
search was therefore carried out in L-D-r space.

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


The method first sought to place bounds on D for a r = 00 model and to find the L(x, t )
that, within geological constraints, predicted a basin in best agreement with observations.
This is relatively easy because the timesteps are independent for an elastic lithosphere. An
approximate linear deconvolution was used to predict the specified load L(x, t ) responsible
for each sedimentary unit. Small AL(x, t ) adjustments, that result from non-linearities
during basin fdling, were estimated by full forward modelling. The results provide a lower
bound for D, DL, for all r < 00 models because all such models exhibit stress relaxation and a
lower apparent flexural rigidity than DL (Beaumont 1978).
The search was continued by choosing successively larger Ds,in decade steps, and posing
the question of whether there was a r for which an L ( x , t ) could be found to construct a
model in agreement with observations. This part of the search, although more difficult than
that for D , , proved to be efficient because the space-time Green functions, and
consequently the influence functions, have a time dependence that is linearly related to T.
The underlying linearity of the flexure in response to the specified loads also proved useful.
For each model the most acceptable L(x, t ) of the previous 0--7 combination was used as a
starting point in an iterative search. Few 0 - r combinations were necessary because a priori
bounds could be predicted on the basis of previous acceptable models. Iteration merely
refined these bounds. Bounds on L ( x , t ) would have been a more formidable task. It was
therefore decided that any acceptable L ( x , t ) would terminate the search for that 0--7
combination. The justification for this approach is the fact that the Alberta Basin is not
strictly two dimensional and that the observational constraints on L ( x , t ) are relatively poor.
For similar reasons initial comparisons between model results and observations were
restricted to one cross-section of the basin (Figs 4 and 6). This profile affords the most
complete stratigraphic section of the basin and, therefore, the strongest constraints. Having
determined bounds for acceptable 0--7 combinations for this section, a search was made for
acceptable values of L ( x , t ) for the other sections within these 0 - r bounds. This method is
logical because the premise of the whole modelling exercise is a laterally uniform litho-
sphere. Only when solutions for a two-dimensional model failed was L ( x , t ) allowed to vary
along the length of the basin.
Particular attention was given to the question whether T < w solutions were required by
the observational constraints. This is important because, in general, it determines whether
the lithosphere has exhibited an extrinsically, time-dependent rheology and, specifically
within the context of the model, whether a viscoelastic rheology adequately describes this
behaviour.

The Alberta Foreland Basin and its relationship to the Rocky Mountain fold-thrust belt
Only a brief review of the geological evidence on the structure and evolution of the Rocky
Mountains and their foreland is presented here. Detailed descriptions of the basin are
available from the following sources; McGrossan & Glaister (1966), Gussow (1962), and
Parsons (1973). Bally et d. (1966), Price & Mountjoy (1970) and Thompson (1979)
302 C. Beaumont

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


Figure 4. Map of the physiographic and structural elements of the Canadian Cordillera and Alberta Fore-
land Basin. T h e three cross-sections analysed are A, B and C, with C projected o n t o a normal t o the
structural trend of the Rocky Mountains, C'. 1 , 2 and 3 are the locations of the detailed sections of the
Rocky Mountains. The map also shows the spatial distribution of the Upper Jurassic and younger
sediments and t h e relationship between the Alberta and Williston Basins.

compiled similarly comprehensive palinspastic reconstructions and intepretations of the


development of the Rocky Mountains along cross-sections. Overviews of the tectonic hstory
of western North America, and western Canada, in particular, are provided by Monger,
Souther & Gabrielse (1 972), Coney (1972), Wheeler & Gabrielse (1 972), Dickinson (1976),
Griffiths (1977), Monger & Price (1979) and Price (1980) among others. Quantitative
estimates of the timing and amounts of deformation are given by Wheeler et al. (1974).
The main tectonic elements and cross-sections (Figs 4, 5, 6 and 7) show that the area to
be modelled has a fundamentally two-dimensional character with structural trend following
a north-west-south-east direction, The main concern in t h s paper is the interaction of the
Omineca core zone with the Rocky Mountain fold-thrust belt and the Alberta Foreland
Basin.
The evolution of the Canadian Cordillera and its extension into the northern United
States (Dickinson 1976; Monger & Price 1979; Price 1980) exhibits a preorogenic Late
Proterozoic to Middle Jurassic phase that was dominated by deposition in a rifted passive
continental margin environment, and a synorogenic late Jurassic to Tertiary stage of sub-
duction, collision and strike-dip tectonics. The main points to be made are: (1) that rifting
occurred to the west of the Purcell anticlinorium (see below), probably at the present
location of the Kootenay arc near the south-western boundary of the Omineca core zone
(Price 1980), and (2) that during the orogenic phase, the eugeosynclinal-miogeosynclinal
wedge, including the Purcell anticlinorium, presently located near the south-eastern
boundary of the m e c a core zone, and some crystalline basement, was thrust from its
marginal position on to the adjacent continental crust (Price 1980). The forces responsible
for this eastward thrusting are not understood in detail, but in a broad sense movement was
in response to offshore subduction and the accretion of island-arc terranes and lithospheric
fragments to the f m r Continental margin. An exact description of the forces is not
important from the point of view of this paper. What is important is that loads were trans-
ported from outboard of the continent, where they had been locally compensated on
Sweetgrass
Arch Wi IIiston
Basin

resent sea level


.-.-.-.-.-.-.-.-.-.

Section A 3
- 2
FQ
Q-
8
Figure 5. Cross-section A (adapted from Gussow 1962) and topography of the adjoining Rocky Mountains. The topography is averaged in 50 km wide zones across 2'
strike and 500 km dong strike, but the most weight is given to that along a projection of section A. The two lines indicate upper and lower bounds for the
-
topography. The stratigrdphy is shown for undisturbed parts of the section and the numbers are estimates of the ages of stratigraphic horizons. The uplift of the Sweet-
grass Arch and assocuted erosion of the Oldman, Bearpaw, S t Mary and Willow Creek formations is clearly shown.

w
0
w
.c Sections I & 2-* Section 6 b
\ 0 200 km w
\ 0
\ A
\
\

2 n
----- 7

E .-.-.-Present sea level


.-.-.-.-.-.-.
W
y o
t
II
0,
.-
a
I -I

-2

-3

Figure 6 . Cross-section B (adapted from Gussow 1962) and topography of the adjoining Rocky Mountains. A-E are palinspastically restored estimates of the
thicknesses of now deformed units (from Price & Mountjoy 1970). F and G are estimates of the depositional baseline prior to erosion from coal moisture content of
near-surface coals (Hacquebard 1977) and shale compaction (Magma 1976). The distance from F and G to the present surface is an estimate of erosion since the
culmination of Laramide thrusting. H are upper and lower bounds on topography of the adjacent 300 km of the Cordillera averaged in the same way as that for cross-
section A. The numbers are estimates of the ages of the stratigraphic horizons.
3- Section 3
- - 1 - 1 I
Section C projected on to1-C
----- 0 200 km
I I I I
2 -
-
n
Pleistocene
E I - Smokey :9 97 100 /194 109
Y -- 4..1....L.
W
- -----
t
-----
I: Lrresenr
- . sea
- . - .ieve
-.-.---
m
I .-
.- 0- _----
a>
r -
-I -
i British Columbia I Alberta
u,
Figure 7. Cross-section C (adapted from Gussow 1962) projected on to C' and topography along an extension of C' across the Cordillera. The numbers are estimates of 8 w
the ages of the stratigraphic horizons. Gentle folding outboard of the foothills is approximately restored by the dashed lines. The upper dashed line is an estimate of the
2'
depositional baseline, prior to erosion, from coal moisture content of near surface coals.
__

Table 1. General classification of deposition and erosion of the northern, central and southern parts of the Alberta Foreland Basin and possible correlation with deformation in the
Rocky Mountains. The phases, 1-4, are those referred to in the text.
Central and southern Deposi- Maximum Inferred Northern Deposi- Maximum Inferred Orogeny Megacycle
basin groups and tional type isopach deformational basin groups tional type isopach defor- (Eisbacher
formations (m) zone and formations (m) mational eral. 1974)
zone

-143
Upper Jurassic Kootenay Forma- Shale then - 1300 West of Nikanassin Shale then - 2200 West of Coast range Megacycle 1
clastics with Rockies Formation minor clastics Rockies Nevadan 1
tion (foothills)
coal (molasse) (foothills) (shallow water
Minnes Group marine)
- 135
Early Lower Uplift and erosion - Hiatus and Uplift and - Hiatus and
Cretaceous erosion erosion erosion
-118
Barremanian Blairmore Group Mainly clastics -650 Western Bullhead and Minor clastics - 1100 Western Nevadan? Megacycle 1
Aptian Mannville but shales to Rockies Lower Fort St then shale Rockies Columbian?
early Albian south John Groups
- 103
Mid-Albian Upper and Lower Shale, marine -650 Hiatus? Fort St John Shale, marine - 1600 Hiatus? Mega ycle 2
early Campanian Colorado Group transgression and Smokey transgression
Lea Park Groups
- 74
Campanian PostColorado Super- Clastics, shale - 2000 Central Rockies PostColorado Clastics Largely Central Columbian
early Palaeocene group. (Belly River, coal (molasse) (main to front Supergroup eroded Rockies
Bearpaw, Edmonton, ranges) Wapiti, etc. (main to
St Mary’s River, front ranges)
Willow Creek)
- 65
Early Palaeocene Uplift and ? Hiatus and Uplift and ? Hiatus and
erosion erosion erosion erosion
-60
Mid-Palaeocene Paskapoo Clastics - 3000 Front ranges Eroded Eroded < 2000? Foothills? Laramide Megacycle 2
Eocene Porcupine Hills lignites and foothills
(molasse)
- 35
- Hiatus and Uplift and - - Hiatus and
Oligocene Uplift and erosion
Miocene at high rate erosion erosion erosion
- 15
- Hiatus and Uplift and Local con- - Hiatus and
Pliocene to Uplift and erosion Local con-
present at a reduced rate glomerates and erosion erosion glomerates and erosion
glacial deposits glacial deposits
Foreland basins 3 07
oceanic or transitional crust and lithosphere, and were placed on the edge of thermally
stable continental lithosphere which responded by flexure to form a foreland basin migrating
ahead of the advancing load.
A very much simplified classification of deposition within the Alberta Foreland Basin
(Table 1) indicates that there were four phases of molasse deposition in the southern part of
the basin but that further north, in the ‘northern basin’, molasse pulses 1 and 2 are very
much diminished in strength or absent (Eisbacher, Carrigy & Campbell 1974). Between the
molasse depositional phases the basin either underwent a marine incursion or was uplifted

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


and eroded.
Dickinson (1976) suggested that the inception of the fold-thrust belt may have been in
the Lower Jurassic. However, in phase 1 of the Alberta Basin the earliest clastic sediments
of western provenance are the late Jurassic marine Fernie Group and Passage Beds that grade
upward into the Tithonian-Berriasian non-marine and coal-bearing Kootenay Formation
and its northern equivalent, the nearshore marine Nikanassin Formation. The thickest of
these deposits occupy positions in the now deformed foothills. Undeformed deposits thin
basinward to a depositional or erosional edge no more than 200km beyond the edge of
disturbance. The source of these deposits was the uplifted miogeosynclinal wedge to the
west. An accurate estimate of the position of this source depends on a palinspastic recon-
struction to correct for the - 200 km of post-Jurassic shortening. A conservative
reconstruction (Monger & Price 1979) places the subduction zone, that marks the far side of
the orogen, 1200 km west of the present edge of disturbance.
There follows a period of non-deposition or erosion in the eastern part of the basin
before the influx of the widely transgressive Barremanian-Aptian basal conglomerate of
the Mannville Formation and its lateral equivalents, the mamly non-marine Blairmore and
Bullhead, and the mainly open marine Lower Fort St John Groups. This return to deposition
is interpreted by Eisbacher et al. (1974) as part of the same megacycle as the Jurassic
molasse deposition but there is no doubt that they represent distinct orogenic pulses. In the
now deformed western part of the basin the transition is gradational suggesting more
continuous subsidence. The basin achieved its largest lateral extent during this, the second,
molasse phase (Table 1). The dominantly clastic non-marine coal-bearing formations of the
central and southern basin extend far into the plains before interfmgering with marine
sediments. In the northern basin the deposits, although extensive and thick except on the
Peace River Arch, remained predominantly shallow water marine in nature and the molasse
phase is absent.
By approximately Middle Albian there was a widespread transgression from the north and
south to form the epicontinental Colorado Sea. The Alberta and Colorado Group sediments
and their northern equivalents, the Upper Fort St John and Smokey Groups are mainly
marine shales although brief regressions mark the Dunvegan and Cardium Formations. This
marine phase and the underlying molasse phase comprise depositional phase 2 (Table 1)
which is interpreted in the models in terms of a combination of basement downwarping,
c b t i c influx and eustatic sea-level rise.
I h e overlying Post-Colorado Supergroup, with the exception of the Bearpaw shale,
comprises non-marine clastics that were deposited as m o b phase 3 in response to the
beginning of the Laramide Orogeny. This phase of largely continental deposition continued
without interruption into the early Palaeocene over most of the basin. In the northern basin
the record of this sedimentation has been erased by erosion or extensive deposition may
never have occurred. In the central part of the basin there was a short precursory
depositional hiatus and some erosion in the early and mid-Palaeocene prior to the ehormous
influx of clastics associated with the Palaeocene culmination of the Laramide Orogeny.
308 C Beaumont
It is almost certain from independent evidence (Hays & Pitman 1973; Vail, Mitchum &
Thompson 1977) that sea-level fell during phase 3 but t h s had little effect on sedimentation.
It was the downwarping of the basin and clastic influx which dominated the sedimentation
pattern.
The exact timing of the culmination of Laramide deformation in the Rocky Mountains
is not known, although as will be shown on the basis of the modelling, a mid- to late-Eocene
date is preferred. The Palaeocene Paskapoo and Porcupine Hills Formations are the remnants
of the thick clastic wedge that was deposited during the Laramide Orogeny. It is assumed,

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


for modelling purposes, that uplift and erosion of the wedge was strongest during the
Oligocene and Miocene. This assumption is based on the return to isolated thin conglomerate
deposition during the Pliocene. Price (private communication and 1962) suggests that
normal faulting and uplift of the Rocky Mountains may have commenced as early as the
mid-Eocene.
The evolution of the Alberta Basin was in response to shortening across the Columbian
Orogen which translated thrust sheets to the east to form the Rocky Mountains. Shortening
of the supracrustal rocks along three sections (Fig. 4) has been estimated as: section 1,
120 km (BaIly et al. 1966), section 2, 200 km (Price & Mountjoy 1970) and section 3,
22 km (Thompson 1979). This resulted in supracrustal thickening of up to 8 km and
isostatic subsidence of the basement detachment surface of from 2-8 km during the Upper
Jurassic to pre-early Oligocene interval (Price 1973). Movements are believed to have
progressed in a north-easterly and easterly direction with early, Upper Jurassic to mid-lower
Cretaceous, deformation being confined to the neighbourhood of the core zone and the
Western Ranges. The Foothills were the last to undergo deformation during the Laramide
-
Orogeny. Average rates of motion during the main phases of deformation are 2 mm yr-',
an amount comparable with that for individual thrust slices, 3-7 mm yr-' (Wheeler et al.
1974). The shortening almost certainly occurred at a non-uniform rate and it is assumed that
pulses of activity coincide with the molasse phases in the basin. If this is true, structures
in the Main Ranges and western Front Ranges were probably initiated in the early Lower
Cretaceous and those in the rest of the Front Ranges by late Upper Cretaceous. Activity
was particularly high in the Aptian (Wheeler et al. 1974). The early Palaeocene reduction in
rate of deposition is interpreted as a lull in activity during the Laramide Orogeny prior to
the main phase of deformation in the Main and Front Ranges, and in the Foothills. Erosion
-
from the orogen has been estimated (Price & Mountjoy 1970) as l o a m 3for every metre-
wide strip along section 2, most of which came from the Main Ranges and western Front
Ranges.
Further to the north, section 3 and basin section C (Fig. 7) indicate a large-scale pattern
of development that parallels that to the south but with lower intensity (Thompson 1979).
The lack of a Lower Cretaceous molasse sequence is interpreted as an indication that de-
formation in the adjacent Rocky Mountains during this period was less than that to the
south. Similar comparisons for later phases are not possible because of the extensive basin
erosion. The 22 km shortening in the Foothills and easternmost Rocky Mountains, is
certainly much less than the equivalent 70 km further south (Thompson 1979). For these
reasons the models for basin section C assume that the timing of deformation (emplacement
of specified loads) is similar to that of other sections but that the size of the loads may be
considerably less.
In summary, detailed geological information is available on the structure of the Alberta
Basin where it has not been eroded. Equivalent information on mass movements in the fold-
thrust belt with a 10-20 Myr time-scale is not available except for a few thrust sheets
(Price, private communication). Consequently, the geological constraints on the specified
Foreland basins 309
model loads tend to be general rather than detailed, whereas constraints on the model basin
are much stronger.

Constraints from sediment compaction and thermal metamorphism


A major requirement for modelling is an estimate of the volume of sediments that has been
eroded from the basin since the Laramide Orogeny. Direct evidence is obviously lacking but
indirect evidence is available in the form of the effects that palaeo-overburden pressure and
palaeo-temperature during burial have had on the sediments that remain. Of these effects,

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


the maturation of oil and the coalification process are best understood, although some
evidence is also available from shale compaction. Coals and coaly particles (phytoclasts)
are the most useful because they are widespread and remain in situ. The constraints that
these sediments afford are mostly integral in nature. The sediment properties at present are :

M(x, t p ) = s, F [W),Wl d f
where M is a measure of the compaction, maturation, or degree of coalification and F [T(t),
P(t)] is a functional that describes alteration in response to temporal changes in tempera-
ture, T(t), and pressure, P(t). M can be evaluated for various locations within the basin
and the results compared with observations.
Hacquebard (1977) has provided a detailed discussion of the application of coalification
models to the Alberta Basin. His estimate of erosion (Fig. 6F) is based on an empirical model
of the decrease in coal moisture content with increasing depth of burial. The results are
based on the argument that percentage bed moisture content, FM,is solely an exponential
function of the maximum depth of burial for h g h volatile bituminous and lower rank coals.
Under these circumstances the present percentage bed moisture content, MM, reduces to,
= a exp (- bP,,)
= FM(prnax) = a' exp (- b'd,,)
where a, b , a' and b' are empirical constants, and P,,, and d, are the maximum pressure
-
and depth of burial. The constants a' 75 per cent and b' = 0.0009 m-', were estimated
from German coals that were mined from basins that had undergone little erosion or where
there was stratigraphic control on erosion estimates. Application of this model to the
moisture content of near surface coals in the Edmonton, St Mary's River, Eastend, French-
man, Foremost and Oldman Formations (Steiner, Williams & Dickie 1972) yields the
estimates of the eroded overburden (Hacquebard 1977). These results must be regarded with
some caution. They probably provide an upper bound for eroded overburden both because
moisture content measurements for high moisture coals, 5 25 per cent are difficult to make
accurately (Hacquebard, private communication) and because estimates of erosion from
shale compaction (Magara 1976) (Fig. 6G) suggest a lower value. Results from lignite coals
(Steiner et al. 1972) in the Paskapoo Formation that were not considered by Hacquebard
also suggest that the estimates are an upper bound.
For higher rank coals, buried at greater depths in the Mannville and contemporary
formations, percentage bed moisture is no longer a good measure of metamorphic grade.
For these coals Hacquebard (1977) has applied Bostick's (1973) modification of Kanveil's
(1956) coalification model. This model, which is based on reaction kinetics, contends that
the degree of organic metamorphism, as measured by the percentage vitrinite reflectance
(%R,) or percentage volatile matter (%VM) is dominated by the temperature history of the
coal particle. That is,

M V R = J1 FVR [T(t)l dt.


3 10 C Beaumont
The main limitations on interpretations of MVR in terms of burial history are the accuracy
with which the palaeo-geothermal gradient, dT(x, t)/dz and the alteration process,
F V R [T(t)], are known. It is initially assumed in thls paper for foreland basins whch form
on old cratonic lithosphere that the geothermal gradient has not changed except for a simple
thermal blanketing effect of the sediments. MVRcan then be used to estimate burial history.
Alternatively, MVR may be used to estimate palaeo-geothermal gradients (Hacquebard 1977)
o r to corroborate thermal evolution models (Keen 1979). Again, the results must be treated
with some caution because there is more than one unknown. Agreement with erosional

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


estimates from near-surface coals demonstrates internal consistency but does, not prove the
results t o be correct.
Fig. 8, adapted from Hacquebard (1977), illustrates the use of a coal sample to estimate
erosion from the Alberta Basin. The 107 Myr-old Mannville sample, from a depth of 1829m,
a depth at which the present temperature is 55”C, was found to have a Ro(max) of 0.74 per
cent. The form of FvR[T(t)] (Karweil 1956) is such that it is possible to achieve the
observed %Ro(max) from a range of paths in the temperature-time field; that is by ‘rapid
cooking in a hot oven’ or ‘slow cooking in a low oven’ (Bostick 1973, fig. 5 ) . For any choice
of path, e.g. 1, 2 or 3, M V Rcan be estimated using Bostick’s modification of Karweil’s
time-temperature-metamorphism nomogram for which the 10 Myr time path is reproduced
in Fig. 8. Bostick’s method for the approximate numerical integration of the histories was
followed. The average temperature IT;: = TSi+ 0.65AG (where = starting temperature and
A 7;. is the temperature change) for each successive 10 Myr interval, i, was found. The
amount of thermal metamorphism, Azi, for each of these steps was then determined from
the logarithmic z-scale of the nomogram. Cumulative metamorphism for the entire history is

z= 1 Azi.
i
z was then converted to vitrinite reflectance, %Ro(max), using the conversion scale estimates
provided by Bostick (1973). The discrepancies among the scales (Fig. 8) are a measure of the
uncertainty of the coalification functional; although Bostick prefers scale 2.
The calculated %Ro(max) for path 1 is too small, 0.54 per cent. A two-stage history of
deep burial followed by erosion requires a maximum temperature of 100°C and that this -
temperature should be achieved between 45 and 25 Myr BP. Such a history corresponds to
deepest burial at the end of the Laramide Orogeny followed by rapid and continuous uplift.
This is the preferred model but is by no means unique. Path 3 , for example, yields the same
%Ro(max).
The cumulative path dependence precludes a unique estimate of the maximum tempera-
ture and maximum depth of burial. Consequently, no a priori erosion estimate comparable

[Sub High Volatile M L


-
Lignite j Ell Bit V V S-A Anthracite
Buriol Time (Mvr)
- -
6 7 10 15 2 3 4 5 ( 6 I % ~ ; a x
Ill I51

100 0
Age (Myr) 2 Scale
Figure 8. Methodology for the calculation of thermal metamorphism of coals (see text for details).
Foreland basins 31 1
to that of Fig. 6 can be made. Instead, model predictions of age-depth histories are used t o
calculate %R,,(max) by integrating the histories based on a conversion of depth to tempera-
ture which assumes a constant geothermal gradient equal to that presently observed. Agree-
ment of observed and predicted vitrinite reflectances serves as a constraint on acceptable
models. Results that agree with erosion estimates from bed moisture content, yet fail to
predict vitrinite reflectances in agreement with observations, would indicate inconsistencies
in the assumed model, most probably either time changes in the geothermal gradient, or
inaccurate FVRand/or FM functionals.

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


Model results for cross-section B
ELASTIC MODELS

Simple one-step basin results (Fig. 9) illustrate the basins formed by 50 km wide, 1 km high,
infinitely long strip loads of density 2400 kg m-3 on an elastic lithosphere of flexural rigidity
lo2‘, lo2’ and Nm. The sediment that fills the depression to sea-level is assumed to have
the same density as the specified load. Isostatic adjustment is achieved by distributing the
downwarp over progressively larger length scales as the flexural rigidity increases. The
amplitude of the depression decreases with increasing flexural rigidity because the total
compensation, determined by Archmedes’ principle is approximately the same for all three
cases. Approximate scalings for distance (proportional to D”‘) and displacement (pro-
portional to D ” 2 ) are explained by Beaumont (1978).
Even these simple results indicate, from a comparison with observed basin widths and
depths, that the apparent flexural rigidity of the lithosphere beneath the Alberta Basin is in
the range 1024-1025Nm. The apparent flexural rigidity is the flexural rigidity of the best-
fitting elastic model.
The conclusion on the possible range of acceptable flexural rigidities is confirmed by a
comparison of cross-section B (Fig. 6) with elastic models having D = and loz5Nm
(Figs 10 and 11). These models have the same specified load history as the acceptable model
discussed in the next section. Similar models having a range of load histories also failed to
produce acceptable results when examined in detail, but confirmed that D F lo2’ Nm elastic
models gave basins that were too wide and that D 7 lo2‘ Nm elastic models gave basins that

0.1
Load , I km high

-.-.-.-.-

-
- 0
E
Y
v
c
1I 0 100 200 300 km
.!a
?l -0. I
I

Figure 9. Basins formed on an elastic lithosphere in response to loading by a 1 km high, 50 km wide two-
dimensional load of density 2400 kg m-3 and foredeep filling with &cnts of the same density. The
curves show the difference between the deformation of the basement for flexural rigidities of loa4,
lo2’and 10”Nm. The corresponding plate thicknesses are approximately 3 8 , 8 3 and 178 km.
W
P
h)

4
\
/ >' Section B , D=1024Nm, %=co
-2,ll -2,ll 0 200 km n
3 -
2
-€
X I
-
t
-c
.-
D O
a,
r
-1

-2
Figure 10. Model of cross-section B with a D = 1OZ4Nmelastic lithosphere (see text for a detailed explanation and compare with Figs 1 1 and 12). Note that although
the model correctly predicts the existence of all the stratigraphic units (compare with Fig. 6), many of the units have the wrong shape and thickness, and many of the
stratigraphic horizons dip in the wrong direction. Other major failings are that only Paskapoo and younger sediments outcrop at the surface and that the residual Rocky
Mountains are too high. Extra erosion would not improve the model, however, as it would serve to increase errors in the dip of stratigraphic horizons.
Section B , D=1Oz5N m x =OO
0 200 km

Figure 11. Model of cross-section B with D = lOZ5Nmelastic lithosphere (see text for a detailed explanation and compare with Figs 10 and 12). As in Fig. 10, the
model predicts the existence of all observed stratigraphic units (compare with Fig. 6) but the units have the wrong shape and thickness and that the dip on the strati-
graphic horizons is either in the wrong sense, or is too shallow. Only Paskapoo sediments outcrop at the surface and the Rocky Mountains are again too high.
w
*
w
P

Figure 12. Model of cross-section B with D = 102’Nm, 7 = 27.5 Myr viscoelastic lithosphere (see text for a detailed explanation and compare with Figs 10 and 11). This
model is in good agreement with the observed cross-section, Fig 6. Individual units have approximately the observed shape and thickness and stratigraphic horizons
have the observed dip. The model could be improved by reduced erosion at distances greater than 200 km from the edge of disturbance. This would preserve more of
-
the Belly River and older units and would prevent outcroppings of the Colorado Formation. The specified loads are two dimensional. A sensitivity analysis indicated
that tapering the loads to the north and south to conform to the observed along strike topography of the Rocky Mountains made only minor changes to the model.
This suggests that a series of two-dimensional models provides a good first approximation.
Foreland basins 315
were too narrow. The results shown in Figs 10 and 1 1 are not the best-fitting elastic models
that were sought in the parameter search. Those models had somewhat different specified
load hstories but were not included as figures because the results are not very different from
the ones shown and are similarly unacceptable in detail.

V ISCOE LASTIC M O D EL S

The best-fitting model with a two-dimensional specified load (Fig. 12) had D = 102’Nm,

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


7= 27.5 Myr. This result should be compared with the models of Figs 10 and 11, which
had the same specified load history, and cross-section B (Fig. 6 ) . The models have 1 1 time-
steps covering the period 140 Myr to the present. The ages of the stratigraphic horizons,
which were chosen to correspond to those of section B, are shown by the numbers listed
above the ‘Age’ label. Some simplifications were made, for example, the Colorado 100-
79 Myr sequence has not been subdivided to show explicitly the Cardium sandstone, nor
have the Lea Park and Belly River Formations been treated separately. It is important to
remember that ages of observed and model horizons may fail to correspond by -* 3 Myr
because there are dating inaccuracies and because all horizons are diachronous to some
extent. It should also be remembered that models assume a basin that is fded with sediment,
density 2400 kg m-3, at the end of each timestep, whereas stratigraphic horizons certainly
departed from sea-level, perhaps by as much as 200 m (Jeletsky 1978).
The specified load history is given by the figures assigned to the six 50 km wide load
columns at the left of the model and in Table 2. The numbers with plus or minus signs are
the heights of the specified loads in kilometres added or removed during the timestep
indicated by the boldface numbers beside or below them. The corresponding sedimentary
sequences for each timestep are labelled by the boldface numbers to the left of the
formation names. Changes in sea-level are given by the numbers listed above the ‘Sea-Level
Change’ label. Thus, for example, timestep 1 spanned the period 140-130 Myr during which
the Upper Jurassic Kooteney Formation was deposited in response to thrusting that added
a specified load of t 2 km to the first (far left) load column. There was no change in sea-
level. Similarly, the model Lea Park and Belly River Formations were deposited between 79
and 72 Myr during timestep 6 in response to specified loads of + 2 km on each of load
columns 3 and 4 and an increase in sea-level of t 20 m.
Timesteps 9, 10 and 1 1 are slightly more complex because they involve the effects of
erosion. During timestep 9, which spanned the period 65-35 Myr, J mmide thrusting added
loads of t 7 km to columns 4 and 5, and + 4 km to column 6. These loads and that of the
Paskapoo Formation, which filled the depression, were sufficient to downwarp the basin
to such an extent that the 35 Myr horizon coincided with the depositional baseline, DB,
thereby creating a very thick Paskapoo sedimentary wedge. Widespread erosion during time-
step 10 from load columns 3 to 6 and corresponding erosion from the basin removed a
substantial part of the Paskapoo, that part between the 35 and 15 Myr dashed lines. Had the
lithosphere been elastic with D = 1024Nm,deposition would have occurred from distances
-
greater than 200 km into the basin (Fig. 10). That no such deposition is observed
constitutes strong evidence against this model. Further erosion of specified loads from
columns 3 to 6 during the final timestep caused additional uplift in the basin. Half of the
uplifted sediments were eroded in an attempt to reproduce inefficient erosion. The results
suggest that even less erosion may have been more appropriate.
Of greater significance is the basic explanation of subsidence and uplift within the basin
afforded by the model. The flexural model demonstrates that the inevitable unroofing of the
fold-thrust belt unloads the lithosphere causing uplift in the basin and concomitant erosion.
316 C Beaumont
Table 2. Loading history for acceptable model (Fig. 12) and comparison elastic models
(Figs 10 and 11) for cross-section B . The dashed lines outline the advance of the fold-
thrust mountain belt. It should be remembered that the model cannot detect loads
that are replaced by an equivalent load. See text for details.
West East
Load column -f 1 2 3 4 5 6
Load Age Sea-level Load
step (Myr) change (km)

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


(m)

140
1
130
2 -
110 - _ -
3 + 1.0
103
4
100
5 -. + 2.0
79
6 + 2.0
72
7 --f + 0.25 + 0.25 -
70 I

8 + 1.0 + 3.0 + 2.0 ' -


65 1 - -
1
9 - 370 -+ +7.0 +7.0 +4.0 I

10

11
35

15

0
I -1.5

-1.6
-7.7

-2.0
-5.1

-2.0
-0.9

-2.0 I
I
I

As long as thrusting increases the net load in the fold-thrust belt there is subsidence within
the basin. When erosion dominates the foreland basin is uplifted. In this context the position
of the 35 Myr dashed line should be compared with the shale compaction and coalification
estimates of erosion (F and G , Fig. 6 ) .
The model structure beneath the fold-thrust belt cannot be directly compared with the
present structure because the model is shown as a pseudo-palinspastic reconstruction. That
is, the specified loads are pictured as though they come from an external source and then
retain their integrity. In fact, the load is created from reworked transported pre-existing
sediments that are thought to be 'bulldozed' by the spreading core zone or peeled during
underthrusting. For example, the models show approximately 400 m of Blairmore sediments
beneath load column 4. These sediments should be correlated with those shown in cross-
section B (Fig. 6 ) in a palinspastically restored position (Price & Mountjoy 1970, fig. 2.3).
The Blairmore sediments from this locations were later incorporated in the fold-thrust belt
and were moved basinward to a position in the foothills where they are now exposed by
erosion. Palinspastically restored sedimentary sequences, where available, are shown on
cross-section B (Fig. 6A to E) for comparison with theoretical predictions. The absolute
vertical position of these restored sections is, however, unknown and the diagram is arbitrary
in this respect. The palinspastic nature of the model results should also be remembered when
comparing the observed and model topographes and basal surfaces. The first two load
Foreland basins 3 17
columns, in the region adjacent to the international border, are now occupied by the Purcell
anticlinorium and the thlrd load column is the Rocky Mountain Trench transition region
(Fig. 4). This eastward migration reflects shortening in the model fold-thrust belt of 70
per cent. It is also important to remember that the specified loads represent a net mass
transport. The model cannot detect unloading and replacement with an equivalent load.
However, a net increase in load on a given column usually implies eastward transport of an
equivalent mass in each of the load columns to the west. This interpretation is not valid in
the case of a thrust sheet that is transported across one or more load columns without

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


disturbing the underlying loads. It is hoped that a detailed interpretation of the loading
history and comparison with improved geological estimates of mass movements can be
provided at a later date.
Sea-level changes cannot be constrained during time intervals for which the depositional
record was subsequently eroded. This leaves the present model sea-level as a free parameter
to be determined by comparison of observations with models in which it was assumed that
no sea-level changes occurred betwen 70 Myr and the present. Best results were given by a
-
present sea-level 200 m below that of 140 Myr. The results therefore suggest a sea-level rise
of - 170 m between 110 and 72 Myr followed by a fall of - 370 m between 72 Myr and the
present. This sea-level hlstory is in reasonable agreement with other estimates with the
exception that Hays & F’itman (1973) suggest a sea-level maximum at approximately 85 Myr.

PARAMETER SEARCH RESULTS

The results of the D - T - L(x, t ) parameter search (Fig. 13) are that acceptable models
were found for a restricted region of D - T space. The lower bound for D,D L = 10%Nm, is
from the elastic model results (Fig. lo), however, all elastic models were rejected because
they failed to reproduce the detailed form of the sedimentary wedge. For example, elastic
models predict a constant relationship between the location of mass movements and the

D( Nm)
lo22

Figure 13. The location of unacceptable and acceptable models in the 0-7 parameter space. D L is the
lower bound for the flexural rigidity. 1-5 are the models shown in Figs 10, 11, 12, 14 and 15. 3, the
acceptable model, is surrounded by an approximate error estimate. The stippled area outlines an estimate
of the bound for which acceptable uniform plate models may exist. Dots show some of the unacceptable
test models.
3 18 C Beaumont
depositional or erosional edge that is a direct consequence of the time invariant flexural
rigidity. Such models fail to expose successively older sediments as the basin surface is
crossed from west to east (Fig. 6). The models, shown in Figs 10 and 11, designated 1 and 2
in D - T space of Fig. 13, have only Paskapoo and younger sediments exposed at the surface.
Viscoelastic models, however, have the potential to reproduce the observations because the
net effect of stress relaxation is progressively to increase subsidence adjacent to the fold-
thrust belt and to uplift and erode the peripheral upwarp. The net effect is to tilt the sedi-
mentary wedge in the observed manner. However, for D = lo2' Nm only values of T between

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


20 and 35 Myr produced acceptable results. For D = 10Z6andlOZ7Nmno acceptable models
were found among the parameter pairs (dots, Fig. 13) that were tested. Two totally un-
acceptable parameter pairs 4 and 5 (Fig. 13) lead to the models shown in Figs 14 and 15.
For 4 the high flexural rigidity predicts a basin that is a factor of 3 too wide. Combination 5
attempts to compensate with a short relaxation time but only succeeds in predicting a very
narrow deep basin.
The stippled area surrounding 3 (Fig. 13) represents an empirical estimate of the
maximum trade-off between D and T. Acceptable models can probably be found within
these limits but it is unlikely that the L ( x , t ) for such models will be significantly different
from that for the model designated '3'.

AGE-DEPTH CURVES AND COALIFICATION RESULTS

A comparison of the 35 Myr model horizon (Fig. 12) with line F of cross-section B (Fig. 6)
demonstrates that the predicted and observed erosion estimates are in reasonably good agree-
ment if uncertainties in the empirical coal moisture versus depth of burial relation are
acknowledged. Conversely, the model predicts coal moisture levels that are in good agree-
ment with those observed in near surface coals (Hacquebard 1977).
The burial histories of selected Mannville, Blairmore, Edmonton and Paskapoo coals, as
predicted by the best fitting two-dimensional viscoelastic model (Fig. 12) are shown in Fig.
16. The two curves in each graph illustrate subsidence and the position of the surface with
respect to a datum of the initial sea-level. Their difference, the burial depth, was used to
compute vitrinite reflectances on the basis of assumed uniform geothermal gradients, surface
temperature and the Teichmuller (1971) and Bostick (1973) metamorphsm to vitrinite
reflectance conversion scales, 1 and 2.
To emphasize the need for deep burial followed by erosion and uplift graph 1 (Fig. 16)
postulates an alternate burial history 'b' that is consistent with the present stratigraphy.
This suggests that subsidence ceased with the deposition of the presently observed sedi-
mentary section. That 'b' predicts much lower vitrinite reflectances (Table 3) than those
observed or predicted by the geothermal models is additional evidence in favour of the deep
burial followed by uplift history, 'a'. Thermal models that have a gradient equal to that
observed, 0.025-0.03"C m-', and a surface temperature between 0" and 10°C predict
%R,(max) values in reasonably good agreement with sample 29 for hstory 'a'. Gradients
of less than 0.025"Cm-' certainly fail to predict sufficient thermal metamorphism no
matter which conversion scale is used.
The agreement between sample and predicted %R,(max)values is also good for Mannville/
Blairmore coal samples 2, 3 and 4 for the preferred thermal gradients. The marginally high
predictions for sample 3 are attributed to the deviation between theoretical and observed
depths of the samples. Sample 5 is predicted to be a lignite, in agreement with observations.
Graphs 6 and 7 are for Edmonton and Paskapoo samples positioned vertically above sample
1. The preferred model predicts that these will both be high volatile bituminous coals in
N
u)

II
II

w
0
t
7
c

m
E

a,
8

n
t'

0
Foreland basins

Figure 14. Model of cross-section B with a D = 1026Nmelastic lithosphere. Note that the model predicts a basin that is vastly too wide and that the residual topography
of the Rocky Mountains is much too high. Such a model may be appropriate to the Martian lithosphere.
3 19

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


320
C Beaumont

Figure 15. Model of cross-section B with D = 1026Nm,r = 1 Myr viscoelastic lithosphere. Rapid relaxation designed to compensate for a high flexural rigidity produces
too much relaxation. Such a model predicts a large amount of erosion, as the peripheral upwarp rises, and too much subsidence of the Rocky Mountains. This model
would be appropriate if the Earth had a thick lithosphere which was uniformly hotter. A planet with a high surface temperature, like Venus, may act in this way.

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


Foreland basins 32 1

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


100 0
Age (Myr)

Figure 16. Burial histories of selected coals predicted by the model shown in Fig. 12. The units on each of
the graphs are the same as those of panel 1. The two solid lines in each graph trace the depth of the
sample beneath the initial depositional baseline and the change in the surface with respect to the same
baseline. The difference between the two curves is the burial depth. The dashed line, b, in panel 1 is a
hypothetical burial history in which the basin is created without an erosional phase after Laramide
thrusting. Panel 8 shows two possible burial histories for Blairmore coals that were buried beneath the
fold-thrust belt before incorporation into a thrust sheet and transport to the foothills. These histories
are discussed in detail in the text.

agreement with the Edmonton Formation observations, but in disagreement with the lignites
found in the Paskapoo Formation. However, uncertainties in the reflectance conversion
scales (Fig. 8) for low levels of thermal metamorphism make such predictions potentially
unreliable.
Graph 8 (Fig. 16) represents an attempt to predict the metamorphism of Blairmore coals
that formed beneath load column 4 of the model but were later incorporated into a thrust
sheet and moved to the surface in the foothdls. The two suggested histories differ in that ‘a’
postulates lateral transport and uplift without deep burial under Laramide loads, whereas
‘b’ has a 10 Myr period of deep burial beneath Laramide loads before lateral transport and
uplift. Although both histories are speculative, they do indicate that transported Blairmore
and Kootenay coals, now mined in the foothills, will range in degree of thermal meta-
morphism from low volatile bituminous to anthracite. Lower grades ‘escaped’ early in the
Laramide Orogeny whereas higher grades suffered substantial burial before transport. That a
similar range of coal grades is observed and that thermal metamorphism was largely complete
before tectonic disturbance have been confirmed by studies of Jurassic and Cretaceous coals
(Hacquebard & Donaldson 1974).
In summary, moisture content and thermal metamorphism of coals on, or adjacent to,
cross-section B are explained by the flexural model if it is accepted that substantial erosion
of both the fold-thrust belt and the foreland basin has occurred. Alternative thermal models
in which apparently enhanced metamorphism with respect to present depth of burial is
explained by higher geothermal gradients in the past, as opposed to erosion, cannot be
totally rejected. However, these models fail to explain the coal moisture levels of near
surface coals; levels that are considered to be evidence of pressure associated with burial,
not enhanced geothermal gradients (Hacquebard 1977).
322 C.Beaumont
Table 3. Observed and calculated thermal metamorphism of Alberta Basin coals. The theoretical samples, 1-8,
are from the Mannville (MI, Edmonton (E) and Paskapoo (P) formations, and their present depth of burial is
given in the table at far right. Predicted maximum percentage vitrinite reflectance in oil was calculated by the
method explained in the text for differing geothermal gradients and surface temperatures, assumed to be
constant throughout the burial. The columns 1 and 2 refer to the conversion scales of thermal metamorphism
to vitrinite reflectance (Fig. 8). The difference in the corresponding predicted vitrinite reflectances is a measure
of the uncertainty of the method. The observed values are taken from Hacquebard (1977) where available.
Other estimates are from observed coal rank, but can be compared using the conversion scales of Fig. 8.

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


Subsistence Predicted % R Y a x Observed %R:ax
example from thermal gradient ("C m - l ) from Hacquebard
Depth (m) 0.02 0.025 0.03 0.025 (1977)
and and surface temperature ("C) Sample
formation 0 0 0 10 No. Depth %R;ax

1 2 1 2 1 2 1 2 (m)

la 2740M 0.66 0.70 0.80 0.90 1.0 1.2 0.90 1.1 29 2888 1.08
lb 2740M -0.56 -0.56 0.60 0.60 0.70 0.78 0.68 0.76
24 2115 0.81
2 2220 M 0.58 0.62 0.68 0.76 0.82 0.93 0.78 0.88 25 2086 0.80
26 2126 0.80
21 1358 0.67
3 1690M 0.57 0.58 0.60 0.65 0.71 0.80 0.68 0.76 22 1445 0.66
23 1224 0.59
7 1264 0.55
7A 1272 0.59
8 1299 0.66
4 1230M -0.53 -0.54 -0.57 0.59 0.58 0.63 0.59 0.64 9 1319 0.62
18 1262 0.58
19 1243 0.60
20 1285 0.62
5 430M <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 1 609 0.42
2 729 0.41
6 1220E -0.53 -0.54 -0.57 -0.57 0.60 0.65 0.60 0.65 HV bituminous
7 600P <0.5 <0.5 -0.53 -0.54 -0.56 -0.57 -0.56 -0.57 Lignites
8a OM - - -0.8 -0.9 - - - - LV bituminous
8b OM - - 0.5 0.6 - - - - To anthracite

Model results for cross-sections C and A


CROSS-SECTION C

The sediments of sections C and A have undergone proportionately more erosion than those
of section B. Therefore, they provide weaker constraints on both lithospheric rheology and
mass movements in the fold-thrust belt. Nevertheless, it can be shown that for D = 1OZsNm
and T = 27.5 Myr cross-section C can only be duplicated if the overall adjacent mass move-
ments are smaller than those for cross-section B.
Just how little sediment remains along section C can be seen from Fig. 7. Near the
foothills the gentle late stage folding that distorted the sedimentary wedge has been approxi-
mately restored to the position inhcated by the dashed lines. In the corresponding model
(Fig. 17) the Upper Jurassic first cycle of deposition is probably underestimated by the
single 2 km thick Columbian Orogeny load. The model agrees with the cross-section but fails
to predict the up to 2.5 km of deformed Upper Jurassic sediments in the foothills (Stott
1972). These may, however, have been deposited on a pre-existing tilted basement further
- 2 I-

Figure 17. Model of cross-section C with D = 102’Nm, T = 27.5 Myr viscoelastic lithosphere. The predictions of this model are in reasonable agreement with the g.
0)
observed cross-section, Fig. 7, once folding in the section has been removed. The model loads are two dimensional.

w
N
w
3 24 C Beaumont
to the west before eastward transport and do not provide strong constraints on the flexural
model. The following period of erosion, which provides the unconformity beneath the
Cadomin conglomerate, closely parallels events further south both in the orogen and basin.
The second cycle of clastic deposition which spans the Bullhead and Fort St John groups
and the Dunvegan formation is modelled by mass movements that are similar to those
required for section B. It therefore appears that thrusting during the Lower Cretaceous
and early Upper Cretaceous was approximately uniform along the length of the basin.
However, the need to erode so much of both the basin and the Rockies and yet produce

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


coals within the remaining sediments apparently requires that the northern part of the fold-
thrust belt experience reduced post-Santonian thrusting with a time sequence that differs
from that to the south. Modelling indicates that the major phase of thrusting started and
ended earlier than that to the south and was less intense. This is in agreement with the
22 km of shortening observed for cross-section 3 (Thompson 1979) in comparison with the
70 km estimated for cross-section 2 (Price & Mountjoy 1970) and Thompson’s result that
deformation in the foothdls was probably complete by the early Palaeocene (Thompson
1979). The only constraints from the basin are those from coal moisture levels (Fig. 7)
which suggest a substantially greater depth of burial at the end of the last major phase of
thrusting.
Strong post-Palaeocene erosion in the model unroofs both the Rockies and the basin. It
reduces the topography of the Rockies to the observed level but may slightly underestimate
erosion and uplift in the basin because a small wedge of Late Cretaceous--early Tertiary
sediments remains adjacent to the foothills and predicted erosion is low. Observational
constraints do not reject the model because the gentle folding adjacent to the foothills
(Fig. 7) would have uplifted and eroded the excess sediments and retained higher moisture
levels in the coals.
If the model is approximately correct, it indicates that no significant Eocene thrusting
occurred in the Rockies adjacent to section C. This major conclusion is in agreement with
observations (Thompson 1979) and is dictated by the need for a long period of erosion
between the last major thrusting event and the present.

CROSS-SECTION A

Cross-section A differs from those to the north in that it transects both the Alberta Foreland
Basin and the rim of the Williston Basin, an older intracratonic basin. An explanation of the
stratigraphy and, in particular, the development of the Sweetgrass Arch (Figs 4 and 5)
involves a yoking effect by the lithosphere between the two basins. An elastic model, in
which a circular subsidence with a Gaussian cross-sectional shape is added to the two-
dimensional foreland basin (Fig. 18a, b, c, d), illustrates how the two adjacent subsidences
combine to produce an intervening arch. This arch is most pronounced along transects
through the depocentres but can also be seen on the section that closely parallels A. The
simple model predicts the basic character of the yoked basins and serves to illustrate the
formation of broad arches that often occur between basins on continents. However, in this
case, observations require uplift of the Sweetgrass Arch and not merely that the arch be a
relative high between the two basins (compare crosssections A and B, Figs 5 and 6, with c
and d of Fig. 18). Such an uplift will occur if the lithosphere relaxes stress in a viscoelastic
or similar manner and the arch is located on the peripheral upwarp of the WillistonBasin.
Stress relaxation progressively uplifts the sediments on the upwarp and deepens the
depocentre as the isostatic balance evolves from a regional to a local equilibrium (Fig. 18e).
Regional equilibrium was probably achieved during the main phase of Williston subsidence
Foreland basins 325

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


*- Overlying
*lbertO
Viscoelastic
Underlying
Williston
I

Figure 18. An explanation of the growth of the Sweetgrass Arch between the Williston Basin and the
Rocky Mountains (compare with Figs 4 and 5). In the elastic model (a) sediment fills both a circular
Gaussian basin and the foredeep induced by the specified load. The crosssections (b, c, d) show the
formation of an arch, analogous to the Sweetgrass arch between the Gaussian basin and the specified load.
Both basins are considered to develop simultaneously. In the viscoelastic model (e) the Gaussian basin
predates the overlying foreland basin but continues relaxation toward local isostatic equilibrium. This
further uplifts the arch and deforms the foreland basin to form an arch like that of Fig. 5.

between the Upper Cambrian and the end of the Mississippian. The subsequent period of
slower subsidence, before mid-Jurassic reactivation, is evidence of stress relaxation which if
continued to the present would warp the overlying foreland basin sediments.
Modelling experiments predict an arch of the correct amplitude and position for a range
of lithosphere 0-7 values. Large D, low 7 combinations favour a large amplitude arch (see
Fig. 1 S), whereas arching is minimized for small D and T = 00. The model predictions are also
sensitive to the process that initiated the Williston basin. That this process is unknown
precludes narrow bounds on the choice of D and 7.However, it appears that a 7 ?: 5-10 Myr
is needed to give the arch sufficient amplitude. If this can be shown to be significantly
different from the 25-35 Myr believed to be appropriate for the other sections, it would
suggest thermal involvement in the formation of the Williston basin.
Under any circumstance arching would have peaked when Laramide loads in the Rockies
maximized the depth of the foreland basin. This placed the upper surface of the underlying
lithosphere under maximum tension relative to the undeformed state, a condition that
would favour release of magma from the lithosphere to the overlying sediments. The
intrusion of small igneous bodies into the Sweetgrass Hills at 48 Myr mzy therefore reflect
the culmination of thrusting in the adjacent Rockies, although similar volcanism also
occurred further south at the same time in different tectonic environments.
3 26 C. Beaumont
Discussion
A mechanical model of lithospheric flexure under laterally migrating loads in the fold-thrust
belt of orogens has been shown to provide an explanation of the formation of neighbouring
foreland basins. This model, which confirms the geologically observed coupling between an
orogen and its foreland, was applied in detail to the post-mid-Jurassic evolution of the
Alberta Basin of western Canada.
It was demonstrated that the largest scale structure of the basin is sensitive to mass
transport and mass wasting in the Rocky Mountains and these parameters were estimated by

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


requiring that the model predict a foreland basin in agreement with observations. Such
estimates may provide constraints on the evolution of the fold-thrust belt and the core zone
of orogens. A more detailed comparison of the estimates for the Rocky Mountains with
geological observations is obviously needed. The model also predicts the flexure of the
basement surface and other stratigraphic horizons which may become detachment surfaces
during thrusting. Insofar as thrusting is a gravitationally driven process, the slope of the
detachment surface is important in controlling the motion of thrust sheets.
It was also shown that the evolution of foreland basins is sensitive to lithospheric
rheology. The rheology was parameterized as that of a Maxwell fluid and the flexural
rigidity, D , and viscous relaxation time constant, 7, estimated by a parameter search-
sensitivity analysis. The results, D = lo2’ Nm and 7 = 20-35 Myr, and the parameterization
must be explained within the context of our understanding of lithospheric properties.
Viscosity, being a thermally activated material property, will be related to the geothermal
gradient and the thermal thickness of the lithosphere. If the lithosphere is the Earth’s
thermal boundary layer, flexural properties will also be related to the thermal age of the
lithosphere. Elastic flexural models estimate lithospheric thickness as the thckness of a
uniform elastic thin plate that has the same flexural characteristics as those observed. The
result may be interpreted as a measure of the thckness of that part of the lithosphere
which does not relax stress during the time of loading, but this estimate is truly equal to
rheological thickness of the lithosphere only in the unlikely circumstance that stress relaxa-
tion is an all or nothing process. An alternative parameterization of lithospheric flexure is
that of the uniform viscoelastic thin plate. This model merely adds a time parameter, 7,
in addition to the spatial parameter, D. In the same way that D is a parameter that measures
an effective lithospheric thickness for the time that the load has acted, 7 is a parameter that
measures an effective relaxation time, again for the time that the load has acted. The thermal
activation model of lithospheric viscosity predicts that D and T are functions of the thermal
age of the lithosphere and the loading age. It is my opinion that much of the confusion con-
cerning flexural models of the lithosphere can be attributed to the natural desire to give
physical interpretations of the estimates of D and T, interpretations that incorrectly con-
sidered these parameters to be constants. For example, it has often been claimed that the
viscoelastic model is wrong, particularly when 7 ? 1 Myr estimates are given, because these
are incompatible with regional, rather than local, isostasy that lasts for more than 100 Myr.
What has been ignored in this claim is that 7 = ~ ( tand ) will increase as ever cooler regions
of the lithosphere with higher viscosities are required to relax stress. Relaxation will stop
either when the cooling boundary layer freezes the system or when it attempts to spread
into a region that is already too cool.
When viewed in the context of the previous paragraph, the viscoelastic interpretation of
-
the Alberta Basin’s lithosphere implies that for the last 140 Myr the initial flexural rigidity
was 102’Nm and that stress relaxation has occurred with an effective time constant of
20-35 Myr. Were we privileged to view the next 140 Myr of evolution, 7 would increase
and D would apparently start from a smaller value because some stress has already been
Foreland basins 327
relaxed. What is more significant about the results is that r < 00, demonstrating that a finite
effective relaxation time parameter is apparently necessary and that a purely elastic para-
meterization would have been inadequate. The viscoelastic model is hardly better. That the
parameters are time dependent limits the usefulness of both models. A priori inclusion of
the temperature dependence of viscosity may result in a more generally applicable and
physically interpretable model.
The Alberta Basin was chosen for analysis in the belief that the underlying lithosphere
has been thermally stable over the basin’s lifetime. If this is true, time-dependent changes in

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


viscosity are unimportant and the basin will reflect the relaxation of flexed old thickened
lithosphere. This explains the reason for the importance of mechanically controlled flexure
in basins that form on cratonic lithosphere without an accompanying thermal event.
Passively subsiding rifted continental margins are the antithesis of foreland basins because
most of their subsidence occurs on, and is controlled by, the cooling thermal boundary
layer. They are thermally controlled. By the time that the lithosphere has thickened
sufficiently to exhibit a flexural response most of the subsidence and loading have already
occurred. The remaining potential for loads that are sufficiently large to excite a significant
flexural response is small. Old sediment-starved margins that receive a large amount of
sediment late in their evolution are consequently most favoured for a flexural response. In
some ways the fold-thrust belt is this late load because the emplacement of thrust sheets
usually occurs on an old continental margin.
That elastic flexural models have proved adequate in many studies is probably not
because they are intrinsically correct but is more the result of weak observational con-
straints. Basin analysis is potentially powerful in this regard because it yields both temporal
and spatial constraints. In the Alberta Basin case, for example, it is certain that an elastic
model would have been acceptable were it not for the stratigraphc constraints. The dis-
advantage of basins is that the mechanism responsible for the tectonic subsidence is often
poorly known.
The weakest part of the analysis presented in this paper is that the lithosphere was
modelled as a laterally uniform plate (Fig. 2). Geophysical and geochemical evidence
indicates that the lithosphere to the west of the Cordilleran crystalline core zone has either
been thermally thinned or is exotic, having been accreted to the edge of the craton. Simple
‘worst case’ broken plate models show that lateral thinning of the lithosphere will have no
major influence on the stratigraphy of the foreland basin more than 300-600 km to the
east. It d l , however, increase the subsidence immediately beneath the fold-thrust belt. The
interpretation of loads for modified models wlll remain approximately correct but the
palaeotopography of the Rocky Mountains will be less. That the Rocky Mountains did not
have a sufficiently large rain shadow to prevent the growth of forests necessary for coals is
indirect evidence for lower topography than the present models predict, a point that is
reinforced by the fine grained nature of the molasse. A more sophisticated analysis in which
the lithosphere to the west of the fold-thrust belt is locally heated and then cools may
therefore be worthwhile.

Acknowledgments

I thank Ray Price, Terry Jordan and Carry Quinlan for their useful comments and criticism
of an earlier version of the manuscript. Ross Boutilier’s help with computer graphics for the
theoretical models is greatly appreciated. The work was supported by the Natural Sciences
and Engineering Research Council.
328 C Beaumont
References
Aubouin, J . , 1965. Geosynclines, Elsevier, New York.
Bally, A. W., Gordy, P. L. & Stewart, G. A., 1966. Structure, seismic data and orogenic evolution of the
southern Canadian Rockies, Bull. Can. Petrol. Geol., 14,337-381.
Beaumont, C., 1978. The evolution of sedimentary basins on a viscoelastic lithosphere: theory and
examples, Geophys. J. R. astr. Soc., 55,471-497.
Bird, P., 1978. Finite element modelling of lithosphere deformation: the Zagros collision orogeny,
Tectonophys., 50,307-336.
Bostick, N. H., 1973. Time as a factor in thermal metamorphism of phytoclasts (coaly particles), 7th

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


Cong. International Stratigraphie et GBologie Carbonif&re, Krefeld, 1971, Compte Rendu, 2,
183-193.
Coney, P. J., 1972. Cordilleran tectonics and North America plate motion, Am. J . Sci., 272,603-628.
Coney, P. J., 1973. Plate tectonics of marginal foreland thrust-fold belts, Geology, 1, 131-134.
Dewey, J . F . & Bird, J. M., 1970. Mountain belts and the new global tectonics, J. geophys. Res., 75,
2625-2647.
Dickinson, W. R., 1974. Plate tectonics and sedimentation, in Tectonics and Sedimentation, ed.
Dickinson, W. R., Spec. Publ. Soc. econ. Geol. Paleont., 2 2 , l - 2 7 .
Dickinson, W . R., 1976. Sedimentary basins developed during evolution of Mesozoic-Cenozoic arc-
trench system in western North America, Can. J. Earth Sci., 13, 1268-1287.
Eisbacher, G. H., Carrigy, M. & Campbell, R. B., 1974. Paleodrainage patterns and late-orogenic basins
of the Canadian CordiUera, in Tectonics and Sedimentation, ed. Dickinson, W. R., Spec. Publ.
Soc. econ. Geol. Paleont., 22,143-166.
Elliot, D., 1976a. The motion of thrust sheets, J . geophys. Rex, 81,949-963.
Elliot, D., 1976b. The energy balance and deformation mechanism of thrust sheets, Phil. Trans. R . Soc.,
283A, 289-312.
Griffiths, J . R., 1977. Mesozoic-early Cenozoic volcanism, plutonism, and mineralization in southern
British Columbia: a plate tectonic synthesis, Can. J . Earth Sci., 14, 1611-1624.
Gussow, W. C., 1962. Regional geological cross-sections of the western Canada sedimentary cover, Alberta
Soc. Petrol. Geol., Calgary.
Hacquebard, P. A., 1977. Rank of coal as an index of organic metamorphism for oil and gas in Alberta,
in The Origin and Migration of Petroleum in the Western Canadian Sedimentary Basin, Alberta,
eds Deroo, G., Powell, T. G., Tissot, B. & McGrossan, R. G., Bull. geol. Soc. Can., 2 6 2 , l l - 2 2 .
Hacquebard, P. A. & Donaldson, J. R., 1974. Rank studies of coals in the Rocky Mountains and inner
foothills belt, Canada, in Carbonaceous Materials as Indicators of Metamorphism, eds Dutcher,
R. R., Hacquebard, P. A., Schopf, J . M. & Simon, J. A., Spec. Pap. geol. Soc. Am., 153,75-93.
Hatcher, R. D. Jr, 1972. Developmental model for the southern Appalachians, Bull. geol. Soc. Am., 83,
2735-2760.
Hays, J . D. & Pitman, W. C., 1973. Lithospheric plate motion, sea level changes and climatic and
ecological consequences, Nature, 246, 18-22.
Jeletsky, J . A., 1978. Causes of Cretaceous oscillations of sea level in western and arctic Canada and some
general geotectonic implications, Geol. Soc. Can. Pap., 77-18.
Karweil, J., 1956. Die Metamorphose der Kohlen vom Standpunkt die physikalischen Chemie, Deursch
Geol. Gesell. Zeitschr., 107, 132-139.
Kay, M., 1951. North American geosynclines,Mem. geol. SOC.Am., 48, 1-143.
Keen, C. E., 1979. Thermal history and subsidence of rifted continental margins - evidence from wells
on the Nova Scotian and Labrador shelves, Can. J. Earth Sci., 16,505-522.
Krumbein, W. C. & Sloss, L. L., 1963. Stratigraphy and Sedimentation, 2nd edn, W. H. Freeman, San
Francisco.
Laubscher, H. P., 1978. Foreland folding, Tecronophys., 47, 325-337.
McGrossan, R. G. & Glaister, R. P., 1966. Geological History of Western Canada, 2nd edn, Alberta
Society of Petroleum Geologists, Calgary.
Magara, K., 1976. Thickness of removed sedimentary rocks, paleopore pressure, and paleotemperature,
southwestern part of the western Canadian basin, Bull. Am. Ass. Petrol. Ceol., 60,554-565.
Monger, J. W. H . & Price, R. A., 1979. Geodynamic evolution of the Canadian Cordillera - progress and
problems, Can. J . Earth Sci., 16, 770-791.
Monger, J . W. H., Souther, J . G. & Gabrielse, H., 1972. Evolution of the Canadian Cordillera: a plate-
tectonic model, Am. J . Sci., 272, 577-602.
Parsons, B. & Sclater, J. G., 1977. An analysis of the variation of ocean floor bathymetry and heat flow
with age, J. geophys. Res., 82, 803-827.
Foreland basins 329
Parsons, W. ff., 1973. Alberta, in Future Petroleum Provinces of Canada, ed. McGrossan, R. G., Mern.
Can. Soc. Petrol. Geol., 1, 73-120.
Price, R. A., 1962. Ferni map area, east half, Alberta and British Columbia 82GE?h, G.S.C. Paper, pp.
61-24.
Price, R. A., 1973. Large-scale gravitational flow of supra-crustal rocks, southern Canadian Rockies, in
Gravity and Tectonics, eds De Jong, K. A. & Scholten, R. A., pp. 491-502.
price, R. A., 1980. The cordilleran foreland thrust fold belt in the southern Canadian Rocky Mountains,
in Thrust and Nappe Tectonics, Spec. Publ. Geol. Soc., in press.
Price, R. A. & Mountjoy, E. W., 1970. Geologic structure of the Canadian Rocky Mountains between Bow

Downloaded from https://academic.oup.com/gji/article-abstract/65/2/291/580800 by guest on 02 December 2019


and Athabasca Rivers - a progress report, in Structure of the Southern Canadian Cordillera, pp.
7-26, ed. Wheeler, J . O., G.A.C. Spec. Pap. 6.
Sleep, N. H., 1971. Thermal effects of the formation of Atlantic continental margins by continental
break-up, Geophys. J. R . astr. Soc., 24,325-350.
Steiner, J., Williams, G. D. & Dickie, G . J., 1972. Coal deposits of the Alberta Plains, Prof. 1st Geol.
Conf. Western Canadian Coal, Res. Council Alberta, Inform. Ser., 60, 85-96.
Stott, D. F., 1972. Cretaceous stratigraphy, northeastern British Columbia, Proc. 1st Geol. Conf. Western
Canadian Coal, Res. Council Alberta, Inform. Ser., 60, 137-150.
Teichmiiller, M., 197 1. Anwendung kohlenpetrographischer Methoden bei der Erdol-und Erdgas
prospecktion, Erdol Kohle, 24,69-76.
Thompson, R. I., 1979. A structural interpretation across part of northern Rocky Mountains, British
Columbia, Canada, Can. J . Earth Sci., 16, 1228-1241.
Turcotte, D. L. & Ahern, J. L., 1977. On the thermal and subsidence history of sedimentary basins,
J. geophys. Res., 82,3762-3766.
Vail, P. R., Mitchum, R. M. Jr & Thompson, S. 111, 1977. Seismic stratigraphy and global changes of sea
level, part 4: global cycles of relative changes of sea level, in Seismic Stratigraphy -Applications
to Hydrocarbon Exploration, A . A. P. G. Mem., 26,83-98.
Wheeler, 1. O., Charlesworth, H. A. K., Monger, J. W. H., Muller, J. E., Price, R. A., Reesor, J. E., Roddick,
J. A. & Simony, P. S., 1974. Western Canada, in Mesozoic-Cenozoic Orogenic Belts, ed. Spencer,
A. M., Spec. Publ. Geol. Soc., 4,591-624.
Wheeler, I . 0. & Gabrielse, H., 1972. The Cordilleran structural province, G.A.C. Spec. Paper, 11, 1-81.

You might also like